National Park Service U.S. Department of the Interior

Sequoia and Kings Canyon National Parks

A Climate-Smart Resource Stewardship Strategy for Sequoia and Kings Canyon National Parks Appendix C: Vulnerability Assessment

This report was prepared as part of Sequoia and Kings Canyon National Parks’ Resource Stewardship Strategy (RSS). The RSS is a long-range strategic planning tool that is informed by current, accurate science. For more information, go to the webpage: go.nps.gov/sekiRSS

On the Cover: View of and the valley below. Photo by Randy Morse, Golden State Images

A Climate-Smart Resource Stewardship Strategy for Sequoia and Kings Canyon National Parks Appendix C: Vulnerability Assessment

Koren Nydick Sequoia and Kings Canyon National Parks 47050 Three Rivers, CA 93271

Ginger Bradshaw Sequoia and Kings Canyon National Parks 47050 Generals Highway Three Rivers, CA 93271

October 2017

U.S. Department of the Interior National Park Service Sequoia and Kings Canyon National Parks

. Contributors

Project Management Team - Sequoia and Kings Canyon National Parks Dr. Koren Nydick, Science Coordinator/Ecologist - Project Manager Dr. Christy Brigham - Chief of Resource Management and Science Ginger Bradshaw – Ecologist

Sequoia and Kings Canyon National Parks Daniel Boiano – Aquatic Ecologist Tony Caprio – Fire Ecologist Athena Demetry - Restoration Ecologist Howard Eldredge – Archivist Annie Esperanza – Air Quality Specialist Erik Frenzel – Plant Ecologist Daniel Gammons – Wildlife Biologist Paul Hardwick – GIS and Data Coordinator David Humphrey – Cultural Resources Program Manager Erik Meyer – Physical Scientist Jessie Moore Russett - Archaeologist Larry Don Seale – Hydrologist Tom Warner – Forester

Sierra Nevada Network Inventory and Monitoring Sylvia Haultain – Program Manager Dr. Andrea Heard – Physical Scientist Dr. Jonathan Nesmith – Ecologist

U.S. Geological Survey – Sequoia Field Station Dr. Adrian Das - Ecologist Dr. Nathan Stephenson – Research Ecologist

Table of Contents

1 – Introduction ...... 1 2 - Stressor Exposure ...... 2 Air pollution ...... 2 Altered Fire Regimes ...... 3 Climatic Change ...... 4 Non-native plants ...... 5 Non-native animals ...... 6 Insects and Disease ...... 7 Fragmentation and Land Use ...... 8 Park Visitation / Human Use ...... 9 3 - Resource Vulnerability ...... 10 AIR RESOURCES ...... 11 WATER RESOURCES ...... 14 AQUATIC ECOSYSTEMS AND SPECIES ...... 17 CAVE AND KARST SYSTEMS ...... 21 WET MEADOW AND FENS ...... 23 FOOTHILLS TERRESTRIAL ECOSYSTEMS ...... 26 GIANT SEQUOIAS ...... 30 FORESTS ...... 33 ALPINE TERRESTRIAL ECOSYSTEM ...... 42 TERRESTRIAL WILDLIFE ...... 45 LANDSCAPE INTEGRITY AND BIODIVERSITY ...... 51 CULTURAL RESOURCES ...... 53 Referenced Literature ...... 57

1 – Introduction

Assessing vulnerability helps managers to understand why resources are at risk and how these resources may change in the future. Vulnerability refers to the extent to which a habitat, species, ecosystem process, or other resource is susceptible to harm (i.e., not attaining desired conditions) from climate change and other stressors. A vulnerability assessment evaluates what things are most vulnerable, why they are vulnerable, and what characteristics of the resource or its environment make it vulnerable. Vulnerability characteristics help us to identify adaptation activities that reduce vulnerability.

A vulnerability assessment includes three components: exposure, sensitivity, and adaptive capacity. Exposure describes how much change in climate or other stressors a resource is likely to experience. Sensitivity is to what degree and how a resource will be affected by a change in stressor exposure. Adaptive capacity is the ability to recover from or cope with the impacts of change with minimal disruption (Glick et al. 2011).

Vulnerability Assessment Terminology

Vulnerability: the extent to which a habitat, species, ecosystem process, or other resource is susceptible to harm (i.e., not attaining desired conditions) from climate change and other stressors.

Exposure: how much change in climate or other stressors a resource is likely to experience.

Sensitivity: to what degree and how a resource will be affected by a change in stressor exposure.

Adaptive Capacity: the ability to recover from or cope with the impacts of change with minimal disruption.

To start the vulnerability assessment, we reviewed results of the parks’ Natural Resource Assessment (NRCA), which provides detailed analyses and summaries of current resource conditions and stressors (NPS 2013). The Resource Stewardship Strategy (RSS) integrates NRCA results and then takes a step further to incorporate potential future stressor exposure, sensitivities, and adaptive capacity. The resulting vulnerability assessment informed the setting of RSS goals and the identification of management objectives and activities. This assessment was developed by park staff and partners using an expert opinion approach, including knowledge of the NRCA, a recent -wide vulnerability assessment (Kershner 2014), other scientific literature, unpublished data, and personal observations.

C-1

2 - Stressor Exposure

Climate change, altered fire regimes, invasive species, insects and disease, air pollution, and fragmentation and land use are major stressors that affect many of the parks’ priority resources and are likely to continue or intensify in the future.

Air pollution Sequoia and Kings Canyon National Parks experience some of the worst air pollution of any national parks in the U.S. The parks are downwind of many air pollution sources, including agriculture, industry, major highways, and urban pollutants from as far away as the San Francisco Bay Area. Air pollutants carried into the park can harm natural and scenic resources such as forests, soils, streams, fish, frogs, and visibility.

Ozone levels in the San Joaquin Valley and lower and mid-elevations in SEKI are a significant health concern for people, wildlife, and plants. High ozone concentrations mainly occur at low to mid elevations on the western side of the parks, below about 7,500 ft (2300 m). The human health metric, the annual 4th highest 8 hour concentration, for SEKI was 94.6 ppb from 2009- 2013 and is above standards. The National Ambient Air Quality Standard (NAAQS) for human health is 70 ppb, and the NPS benchmark is 60 ppb. From 2004 to 2013, SEKI’s trend in ozone concentration improved. The vegetation health metric W126 measures cumulative ozone exposure over the growing season. It was 54.0 ppm-hrs for SEKI from 2009-2013, compared to the NPS benchmark of less than 7 ppm-hrs. There was no trend in W126 from 2004-2013. The parks are at high risk for ozone injury to sensitive plants, which include 21 species such as Jeffrey pine, ponderosa pine, giant sequoia seedlings, quaking aspen, and blue elderberry. Significant ozone injury has been documented in Jeffrey and ponderosa pines, but other species have not been surveyed. Increases in light, temperature, and moisture increase plant uptake of ozone and increases the risk of ozone injury (NPS 2015, Panek 2013). With rising temperatures accompanying climatic change projections, ozone concentrations will increase (Perera and Sanford 2011).

Sulfur and nitrogen deposit into ecosystems from the atmosphere and may cause acidification, fertilization (eutrophication), and changes in soil and water chemistry that alter community composition and biodiversity. Estimated nitrogen and sulfur wet deposition were 3.6 and 1.1 kg per hectare per year, respectively for 2009-2013, respectively. No trends in concentrations of nitrate, ammonium, or sulfate in deposition were detected from 2004 to 2013, albeit over the longer-term sulfate concentrations in the last decade have been lower than those in the 1980s. Ecosystems in the parks are rated as having very high sensitivity to nutrient enrichment and acidification effects. Similar to spatial patterns in ozone, nitrogen deposition is highest along the western edge of SEKI at elevations below about 7,500 ft (2300 m). High-elevation ecosystems, while exposed to lower sulfur and nitrogen deposition than the lower elevations, may be especially sensitive. Nitrogen critical loads (i.e., the level of deposition below which harmful effects are not expected) have been exceeded for epiphytic lichens, phytoplankton in high- elevation lakes, and chaparral along the parks’ western edge; and critical loads have been C-2

approached - and exceeded in two areas - for mixed conifer forests. Shifts in lichen communities and eutrophication of high-elevation lakes have been documented in SEKI. (NPS 2015, Panek et al. 2013).

Atmospheric deposition is a major contributor of mercury and other toxic pollutants (such as pesticides, dioxins, polychlorinated biphenyls) to the parks, including remote lakes. These chemicals accumulate in the food chain and can affect both wildlife and human health by harming neurological, endocrine, and reproductive systems. Mercury converts to toxic methylmercury, which accumulates in organisms and is biomagnified up the food chain, especially in wetlands, lakes, and heavily forested areas. Mercury and pesticide concentrations in the parks are variable, but some fish exceed thresholds for consumption by human and/or wildlife. Air, vegetation, snow, and fish had among the highest concentrations for current-use pesticides, compared with other western U.S. national parks. Pesticides from the adjacent Central Valley could have contributed to the disappearance of the foothill yellow-legged frog, and may contribute to ongoing regional amphibian declines, increased physiological stress in reptiles, and abnormalities in peregrine falcon eggs (Davidson and Knapp 2007, Jarman 1994, Landers et al. 2008, Meyer et al. 2013, Panek et al. 2013, Meyer et al. 2014).

Visibility, as determined by the haze index, warrants significant concern in SEKI. Vistas in the park are sometimes obscured by pollution-caused haze. Average visibility at the parks from 2009-2013 was 9.7 deciview (dv) above the estimated natural conditions of 4.2 dv. From 2004 to 2013, visibility remained relatively unchanged on the 20% clearest days, but improved on the 20% haziest days. SEKI falls in the San Joaquin Valley Air Pollution Control District which does not meet fine particulate matter (PM2.5) public health standards. This air pollution district is an EPA-designated “nonattainment” area for both PM2.5 and ozone. Fine particle exceedance episodes at the parks are most often due to fires, primarily wildland fires. Wildland fire is a natural part of the Sierra Nevada ecosystem; however, its smoke may impair visibility and harm human health. Changing fire regimes as a result of climatic change may increase several airborne pollutants, including ozone and fine particles. According to the Interagency Monitoring of Protected Visual Environments (IMPROVE) network, SEKI’s Ash Mountain site ranked as 152 out 163 sites for highest average fine particle concentrations. The largest source identified was organic matter (vehicles, fires, wood stoves) (NPS 2015, Panek et al. 2013, University of Davis 2015).

Altered Fire Regimes Fire plays a critical role in Sierra Nevada ecosystems. Changes in fire frequency and severity due to management suppression of naturally occurring fires, reduced ignitions from no longer occurring Native American burning, and changes in climate have led to cascading impacts throughout many ecosystems. Lack of periodic low- and mixed-intensity fire in some lower and middle elevation montane forests has caused increases in overall forest density and fuels and shifts in forest composition toward more shade-tolerant species. These alterations can increase fire hazard while decreasing resistance and resilience of the forest to insects, disease, warming temperatures, and drought (Battles et al. 2013, van Mantgem et al. 2016). C-3

Fire return interval for a particular place on the landscape is defined as the average time between fires. Pre-settlement fire return intervals have been estimated for different vegetation types. The Fire Return Interval Departure (FRID) is a measure of the extent of alteration for fire frequency as an important component of the fire regime. It is calculated as the ratio of the time since last fire to the pre-historic fire return interval. While FRID in the parks is generally lower than the surrounding landscape, particularly to the south and west of SEKI, significant areas of the parks warrant concern because of high FRID. Most of these areas of concern are low-to mid-elevation vegetation types that historically experienced frequent fire, such as sequoia mixed conifer forests. These landscapes are concentrated in the Kaweah Watershed and lower elevations of Kings Canyon (Battles et al. 2013).

In the southern Sierra Nevada fire frequency, size, total area burned, and severity have increased over the past several decades. It is expected that more frequent and potentially severe fires will occur in future climate change scenarios. Biomass consumed by fires is modeled to to double or triple by mid-century and triple or quadruple by late century. The area burned, however, is only expected to increase by 20-65% by late century. A time lag between changes in climate and changes in vegetation is highly likely and not included in the model projections, making them uncertain (Safford et al. 2012, Geos Institute 2013).

Climatic Change In the Sierra Nevada, average annual temperatures have increased by around 1-2.5 °F (0.5- 1.4°C) over the last 75 to 100 years. Based on records from five meteorological stations in and near the parks, temperature rose by 1.0 °F (0.58 °C) from 1975 to 2011.The temperatures of the past 10-30 years were extremely warm (> 95th percentile) compared to the 1901–2012 historical range of variability. Rising temperatures have been implicated in several hydrological changes for the southern Sierras or western USA: declining snow-to-rain ratio, melting glaciers, earlier snowmelt, and declining spring snowpack below about 8,500 feet. Annual precipitation in SEKI has been highly variable with no recent temporal trends other than increase in snowpack at higher elevations. Ecological changes over the past several decades in the southern Sierra or western USA that have been influenced by changes in climate have included: increasing area burned in wildfires, doubling of mixed conifer tree mortality rates, earlier onset of spring leaf-out and blooming, shift in elevation ranges of some small mammals, and increasing fire frequency, size, total area burned, and severity except in the lower elevation foothills (Das and Stephenson 2013, Geos Institute 2013, Monahan and Fisichelli 2014, Monahan et al. 2016, Redmond and Fearon 2013).

Since 2012 California has been experiencing its most severe drought in the 120 year instrumented record and perhaps up to the last 1,200 years. The impacts of this drought are being made worse by the high temperatures of recent years. For the state of California as a whole, heating due to human climate change accounts for up to one-fifth of the drought’s severity. Hotter droughts are expected to be a significant driver of ecosystem change, including

C-4

tree mortality and forest die-off (Allen et al. 2015, Griffin and Anchukaitis 2014, Williams et al. 2015).

Average annual temperature in the southern Sierra Nevada is expected to rise about 4°F (2°C) by mid-century and 5-7°F (3-4°C) by late century. Summer temperatures are expected to rise slightly more (7-13°F; 4-6°C) than winter temperatures (5-7° F; 3-4°C) by the end of the century. Precipitation projections are more variable than for temperature, with both increases and decreases in precipitation, and changes in timing possible. Regardless of future precipitation, warming temperatures will increase the fraction of rain relative to snow, speed the onset of snowmelt, push the snowline uphill, and increase the potential for water to be evaporated into the atmosphere (i.e., evaporative demand). Under scenarios of 2-6˚C warming, Sierra Nevada snowpack is projected to decline by the end of century by about 70-90% of historical averages at elevations below 6,560 ft (2000 m) and decline up to 25% above 12,300 ft (3750m). Stream flows and soil moisture may increase during winter, but water availability in summer will more likely decline. Future climatic changes are expected to play out on top of natural variability, meaning that severe droughts and wet periods are expected in any scenario (Geos Institute 2013, Kershner 2014, Koopman 2014).

Non-native plants There are 219 non-native plant taxa known in SEKI. Gerlach et al. (2003) assessed 78 as having the ability to spread and to cause ecological impacts to native vegetation; these 78 species are considered “invasive.” The parks are able to actively manage 19 of these species. The invasive plant species of highest concern are those that spread rapidly, form persistent seed banks, are difficult to detect and identify, and/or cause severe ecological impacts (that is, they displace native species and habitats, reduce local diversity, form monotypic stands, or alter ecosystem processes such as hydrologic regimes, biogeochemical cycling, fire regimes, and other disturbance regimes). The most commonly observed invasive non-native plant species in the parks are Italian thistle, bull thistle, and cheatgrass. Invadedness in the Sierra Nevada and in these parks in particular increases with decreasing elevation and higher levels of disturbance. Campgrounds, pack stations, trails, dirt roads, other developed areas, pastures, and riparian areas show the highest level of invasion. The introduction of non-native plants through horticulture, presence of continuously disturbed habitats in developed areas, the import of aggregate or fill materials, and the movement of animals and people all contribute to invasion success. Sites where disturbance occurs together with a high availability of light, water, and nutrients – such as recent high severity fires, locations with past and current stock activity, gray- water spray fields, and high visitation meadows and stream crossings--are high-risk areas for invasion and are given high priority for both eradication and early detection efforts. Elevated nitrogen deposition in SEKI also could contribute to spread of non-native plants (Tu et al. 2013).

The Watershed has the highest number of non-native plants in SEKI, with generally higher invasion on the western side of the parks. Vegetation types with high threat from non-native plants (based on percent area invaded and average number of invasive

C-5

species) are blue oak woodland, mixed chaparral, montane hardwood, valley foothill riparian, wet meadow, and urban/agricultural. Vegetation types with medium threat from non-native plants are ponderosa pine, chamise-redshank chaparral, sierran mixed conifer, and giant sequoia mixed conifer. Vegetation types with low non-native plant threat are alpine dwarf scrub, aspen, barren, chamise, Jeffrey pine, juniper, lodgepole pine, montane chaparral, montane riparian, perennial grassland, pinyon-juniper, red fir, riparian, sagebrush, sierra white fir, and subalpine conifer. Annual grasslands are mostly composed of non-native species. Vegetation types that have been under-sampled for detecting non-native plants include subalpine conifer, barren, red fir, Jeffrey pine, and lodgepole pine (Tu et al. 2013).

Climate change is expected to worsen invasive plant problems. Many climate change scenarios predict alterations or disruptions in fire regime, hydrologic changes, precipitation or temperature, or an increase in extreme events. With these changes, many invasive plants are projected to benefit from these shifting conditions and colonize and spread to new sites. Amelioration of abiotic controls on non-native species at higher elevations is likely. Twenty-eight invasive plant species were recorded in SEKI at elevations beyond those cited in the literature, indicating possible elevation range expansion that might be a sign the climate-influenced migration of invasive plants has already begun (Tu et al. 2013). The CalWeed Mapper provides information on current and potential future distribution for many non-native plant species in California (http://calweedmapper.cal-ipc.org/).

Non-native animals There are over 30 non-native animal species in SEKI, and 13 of these are considered invasive. Among all non-native animals, fish currently pose the greatest threat to the parks’ resources. Non-native fish include those that are non-native to the region (such as brown trout), non-native to the parks (California ), and species native to the parks but now found in locations where they did not historically occur. Rainbow trout were historically restricted to elevations lower than 4,300 ft (1,300 m) for the Kaweah River; 7,200 ft (2,200 m) for the ; and 6,000-8,000 ft (1,800-2,400 m) for the (Moyle 2002). Little Kern golden trout historically were limited to the high-elevation headwaters of the . Fish stocking occurred in many aquatic habitats across the parks, including hundreds of historically fishless lake and ponds, and has dramatically altered the structure and functioning of these ecosystems. Endangered mountain yellow legged frogs (MYLFs: Rana muscosa, Rana sierrae) are particularly hard-hit and have disappeared from more than 90% of their historic range in the Sierra Nevada, in large part due to non-native trout. Other non-native animals that occasionally have been documented in SEKI include barred owl, brown-headed cowbird, trespassing cattle, turkey, and feral hogs. Additionally, American bullfrog, green sunfish, and black bullhead occur as persistent populations in about a one mile stretch of the North Fork of the Kaweah River (Austin et al. 2013). Two non-native birds (willow ptarmigan and chukar) occur in the alpine but little is known about their ecosystem effects (Haultain 2013).

C-6

Insects and Disease Forest pests are a natural aspect of forest ecosystems. Most are native bark beetles, wood borers, defoliators, and diseases. The population density, and consequently the impact, of some native beetles have increased in forests simultaneously stressed by drought and high tree density. Models of bark beetle population dynamics based on several climate warming scenarios document an increased risk of outbreaks and widespread tree mortality in western conifer forests. Drought and higher temperatures have triggered widespread eruptions of mountain pine beetle (Dendroctonus ponderosae) in lodgepole pine populations throughout western North America. Forests with a significant component of lodgepole pine cover more than 31,000 ha in SEKI (Battles et al. 2013b). Increased abundance and activity of bark beetles and defoliators has been observed in 2014 through 2016 during the ongoing California drought, contributing to a significant spike in tree mortality in lower elevation mixed conifer forest species (such as ponderosa pine, sugar pine, white fir, and incense cedar). This includes a native bark beetle that has been relatively rare in the past, but is now a significant cause of death for incense cedar branches and entire trees (USFS 2016; A. Das, USGS, personal communication).

Additionally, non-native pests have been increasing throughout California’s forests over time. White pine blister rust (Cronartium ribicola) ranks as one of the most destructive disease introductions in history. Its life cycle is complex with five different spore stages on two completely unrelated hosts: five needled pines (including sugar pine, western white pine, foxtail pine, limber pine, and whitebark pine) and shrubs in the genus Ribes (gooseberries and currants). These shrubs are ubiquitous in the understories of the Sierran conifer forests. The spores infect pines via the leaf stomata in the late summer/early fall under cool and very moist conditions. The infection spreads from the leaf to the branch or bole. The resulting cankers can kill the tree directly by girdling it or indirectly by predisposing an infected tree to other pests and pathogens. White pine blister rust was detected in SEKI in 1969 and has since spread throughout the range of some of its host species in the parks, including sugar pine and western white pine. Sugar pine is an important and widespread component of the montane forests in the parks and accounts for more than 7% of aboveground live-tree biomass on average. Western white pine is less abundant, but it is a dominant species in the western white pine-lodgepole pine forest alliance that covers more than 100 km2 (about 25,000 acres) in the parks. In 1997, blister rust infections rates were 22% of sugar pine trees sampled and 4% of western white pine trees in blister rust plots in the parks. Incidence was highest in the Kaweah River drainage. In 2013, a subset of higher-elevation plots was revisited. Forty-five percent of these plots were infected, up from 20% in 1997. On infected plots in 2013, mean incidence of infection was 7.5% of trees per plot with a range of 2-78%. No infections were found in limber, whitebark, or foxtail pine, but 96 infections were found in western white pine (sugar pine was not re-surveyed in 2013). Western white pine mortality was 18.2% in infected plots, compared to 2.5% in uninfected plots, and was strongly correlated with incidence of rust. Blister rust prevalence can be partially explained by elevation (more infection at lower elevations), watershed location, annual precipitation, and average minimum temperature (Battles et al. 2013b, Cahill 2013).

C-7

The amphibian chytrid fungus (Batrachochytrium dendrobatidis; Bd) is a highly virulent pathogen that is responsible for the extinction or decline of more than 200 amphibian species worldwide during the last few decades. In the Sierra Nevada, endangered mountain yellow- legged frogs are highly susceptible to Bd, and all populations are vulnerable to extirpation following Bd arrival at a site. Bd has been spreading across the Sierra Nevada for the last 40 years, and is now ubiquitous across most of the range. In SEKI only small portions of three watersheds remain Bd-free, and these areas harbor some of the largest mountain yellow-legged frog populations that remain. Based on patterns of Bd spread in SEKI, it is likely to invade these basins during the next decade, resulting in mass frog die-offs and potential extirpations. Currently research is ongoing to test effectiveness of disease intervention treatments (EGSPC and Schwartz 2013; R. Knapp, UC Santa Barbara, personal communication).

The transmission of diseases to Sierra Nevada Bighorn Sheep from domestic sheep is a concern (EGSPC and Schwartz 2013). While overlap between these groups is minimal, the risk of disease transmission occurs primarily as the result of extra-herd movements, either by bighorn or domestic sheep. For example, 23 domestic sheep that had wandered from the Bloody Canyon allotment into Dana Fork were discovered in Yosemite National Park in 1995 (USFWS 2007). Long-distance movements of bighorn rams is well documented, with disease transmission risk being exacerbated by the interest in and ability of bighorn rams to mate with domestic ewes. In the Southern Recovery Unit, which encompasses most of the bighorn that use SEKI, domestic sheep grazing allotments have been eliminated, but the Wheeler Ridge herd, which is one of the largest bighorn herds, has grazing allotments within 5 km. Non- commercial or hobby animals may pose a notable threat because the owners may less aware of the risk to bighorn sheep (USFWS 2007).

Fragmentation and Land Use The protected area centered ecosystem (PACE) that influences Kings Canyon, Sequoia, and Yosemite National Parks is 45,203 km2 and includes 45% U.S. Forest Service, 29% private, 14.5% NPS (54% of which is represented by SEKI), and 7.7% Bureau of Land Management lands. SEKI itself is primarily bordered by other federal lands with only 4% of the border adjacent to private holdings, which occur on southwest side of the parks. There are an additional 35 km2 of private inholdings within the parks’ boundaries (Thorne et al. 2013).

Despite the dominance of public lands in the southern Sierra Nevada, the region has experienced increasing development. Of the 10 counties in the PACE region, Fresno and Tulare Counties (in which SEKI is located) have the fastest and third-fastest growing populations. The foothills west of SEKI have experienced extensive growth and development over the past 40 years, impacting vegetation types typically found at lower elevations. From 1970-2010, 29% of private land in the PACE was converted from undeveloped to rural residential (< 7 dwellings/km2), 16% from rural to exurban (7-145 dwellings/km2), and 0.8% from exurban to suburban (Thorne et al. 2013). Overall, housing density from 1940-2000 has increased by 33% in the PACE. Projections of future housing density in the PACE suggest increases of 107% from C-8

2000-2030, 390% from 2000-2060, and 752% from 2000-2090 (Hansen et al. 2014). Areas in the Sierra Nevada that have experienced significant amounts of human settlement often have reduced canopy cover, higher proportion of non-native trees, and increased cover of impervious surfaces, such as pavement. These changes can reduce wildlife habitat, increase fire hazards, and alter hydrology (Thorne et al. 2013).

Landscape fragmentation is a measure of ecological disruption occurring on a landscape. Fragmentation in the PACE region was assessed based on the sizes and distribution of habitat patches separated by roads, which were buffered by 500 m to account for their indirect and cumulative impacts on habitat quality. In the regional analysis, 35% of the PACE was in the least fragmented category with patch sizes greater than 3,100 km2; 92% of SEKI was in this same least fragmented category. In contrast, 13.4% of the PACE was in the higher fragmentation categories with patch sizes of 36 km2 or less; only 0.2% of SEKI had this status. The majority of higher fragmentation occurred on the western side of the PACE. SEKI’s border mostly was characterized by very low fragmentation with moderate levels along the northwestern edge of Sequoia National Park and a very small area of more severe fragmentation adjacent to Grant Grove, Kings Canyon National Park. Because of the low fragmentation in SEKI compared to other areas of the PACE, the parks are likely critical to the remaining habitat and wildlife connectivity of the region, including California Essential Habitat Connectivity Least Cost Corridors and “Theobald Connectivity Corridors” (Thorne et al. 2013).

Within SEKI, fragmentation was calculated in a more detailed manner based on roads (500 m buffer) combined with maintained trails (100 m buffer) and unmaintained trails (50 m buffer). Thirty-six percent of the parks area was contained within six habitat patches that are each greater than 123 km2 in size and not even cut by a trail. Another 26 patches (42% of the parks’ area) are 34 to 122 km2 in size. The remaining 22% of the parks area contained 238 patches of less than 34 km2 size (Thorne et al. 2013). The majority of the more fragmented area of SEKI (i.e., smaller patch size) was found in the western and southern part of Sequoia National Park and a small area near Highway 180 along the western edge of Kings Canyon National Parks (Thorne et al. 2013).

Park Visitation / Human Use

Visitation was not studied in the Natural Resource Condition Assessment, but the Wilderness Stewardship Plan documents concerns in the wilderness regarding pack stock use, campsites, wood collection, water quality, etc. (NPS 2015). Recent visitor counts suggest growing visitation with warm, drought conditions - except when wildfires reduce access to the parks or cause major smoke impacts. Visitation in SEKI is positively correlated with temperature, at least on a monthly basis. Climate warming may boost visitation more in the future and this increase could occur disproportionately during the shoulder seasons of April-May and September-October (Fisichelli et al. 2015).

C-9

3 - Resource Vulnerability

Park staff synthesized information on sensitivity of resources to stressors and the adaptive capacity to cope with or recover from impacts. Confidence in the assessment is discussed at the bottom of each resource assessment with overall confidence rated as low, medium (med), or high. The vulnerability assessment was conducted as a step towards reconsidering goals, which also included reconsidering which resources were to be identified as “priority resources” in the RSS. Each section below begins with a table that summarizes vulnerability of the resource or component across a range of stressors followed by a narrative description of stressors, sensitivity, and adaptive capacity. The estimated level of vulnerability is depicted by red (higher vulnerability) to yellow (lower vulnerability) color-coding as follows:

Dark red Pink Orange Yellow

Highest vulnerability, High vulnerability Moderate vulnerability Slight vulnerability often due to interactions

C-10

AIR RESOURCES

Key vulnerabilities: Atmospheric deposition, ozone, light pollution, and machine generated noise that originates outside the parks, potentially interacting with increased smoke from fires and climatic changes.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Air Quality Nutrient Increased Increased Increased Air deposition, particulates particulates Emissions Pollution X contamin- & ozone Fire ants, ozone,

Visibility & Reduced Vegetation Reduced Altered Scenery Visibility Change Visibility Viewshed

Night Skies Visibility Light Pollution

Sound- Loss of Machine- Loss of scape natural generated natural sounds noise quiet

Air Quality, Visibility, and Scenic Resources

● Air pollution (see Section 2 above) is the main influence on the parks’ air quality. ● Scenic landscapes are vulnerable to air pollution, smoke from fires, development within and outside the parks, and changes in vegetation that can block valued scenic vistas. ● Vistas often are obscured by haze. Air pollution has reduced average visual range by 150 to 35 miles. On the haziest days, visual range has been reduced from 110 to below 20 miles. Episodic incidents can reduce visibility even lower. Over time, however, visibility has been improving (NPS-ARD 2016). ● Ammonium nitrates drive the worst visibility days. Organic carbon causes poor visibility days when nitrates are low (seasonal opposites). Ammonium sulfates are usually higher in spring and summer and are the third largest contributor on low visibility days. (NPS-ARD 2016). ● Climate change projections of increased fire frequency and size will result in reduced visibility from smoke. Warming temperatures will increase ozone formation (Panek et al. 2013). ● Most views from the park to outside the boundaries are not modified by development because of surrounding National Forests and Wilderness Areas. Continued development, where visible, will add structures and utilities (e.g. solar panels) to the existing views. ● Scenic resources within the parks can be protected by thoughtful management. Integrated vistas in SEKI were identified and discussed in the early 1980’s but no action was taken to develop management plans. Confidence: Medium - Interagency Monitoring of Protected Visual Environments (IMPROVE) monitoring has been continuous since the 1980’s and we have good trend data for those elements that contribute to reduced visibility. C-11

Night Skies

● Dark night skies can be degraded by lighting from park infrastructure or development outside the parks and from air pollution. ● Limited measurements suggest moderate concern of light pollution in SEKI. The Anthropogenic Light Ratio (ALR) is the total sky brightness averaged across the entire sky and compared to natural nighttime light levels. It can be directly measured or modeled when observational data is unavailable. Lower ALR reflects higher quality night sky conditions. Ground based measurements in SEKI were made from Buena Vista Peak, Buck Rock, Big Meadows, and Moro Rock. Opportunistic measurements were also made on Mt. Whitney. A range of 0.18 to 0.74 ALR (Mt. Whitney to Buck Rock) were measured in these parks. Parks with less than 0.33 ALR are considered to have minor impact from artificial light while parks with 0.33-2.00 are considered to have moderate impact (NPS-NSNSD 2015). ● An altered photic (light) environment can affect wildlife interactions and other vital ecological process such as predator/prey relationships, reproduction, navigation, and migration. ● The nighttime environment plays an integral role in human health and physiology. Loss of dark nights (and bright days) can cause disturbance to human circadian rhythms. ● SEKI is approximately 97% designated Wilderness. Loss of a dark night sky degrades wilderness qualities of solitude, naturalness, and untrammeled character. ● There is some management capacity for night skies both within the parks and locally and across the region. Within the parks, managers can reduce light pollution using established and tested means. Outside the parks, we can primarily influence individuals through interpretation, education and outreach such as evening programs, special events (e.g. Dark Skies Festival). Local communities may be engaged more directly through town meetings and speaker series. Confidence: Low to Medium - While data is limited, there is at least medium confidence in light pollution in the moderate range. We don’t have data on the actual impacts of this light pollution to wildlife and humans, however.

Natural Sounds

● The acoustic environment is important as a natural resource and as a cultural resource. Natural sounds are a critical component of wilderness character and play an important role in wildlife communication, behavior, and other ecological processes. ● Natural sounds may be degraded by human-generated noise from road and air traffic, developed areas, and construction. ● Impacts to sound volume are determined by the difference between the modeled natural ambient sounds levels and the predicted or observed existing sound levels. Via a geospatial sound model for SEKI, there is a range between 0 decibels (dBA) in some areas, to 9.7 dBA in others. The mean impact is predicted to be 1.5 decibels (dBA) above the natural ambient sound levels. Compared to NPS units across the country, this is a low number and shows a prominence of natural sounds that should be preserved and protected (Wood 2015). ● Natural sounds also may also be degraded due to other stressors, such as climate change, that may alter not only the volume but also the range and arrangement (biophony) of natural sounds

C-12

via changes in physical processes (such as flowing water) or biodiversity (such as insects and wildlife). ● Management capacity for protecting natural sounds is high when produced by human activities (reduction of noise during particular times of day or season and providing visitor education and outreach to local communities). Confidence: Medium to High - In addition to the national model, acoustical monitoring was conducted for SEKI. This research provides a more comprehensive description of the acoustic environment by characterizing existing sound levels, estimating natural ambient sound levels, and identifying audible sound sources. Therefore confidence is medium to high.

C-13

WATER RESOURCES

Key vulnerabilities: Reduced water quantity and shifts in flow timing due to reduced snowpack and warmer temperatures; air pollution effects on water quality; increased water temperatures; flooding; erosion; and sedimentation.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Water Nutrients, Warmer Runoff after Erosion, Warmer Personal Air pollution Quality contamin- water, more severe fires nutrients, water, care X climate ants, concen- - erosion, turbidity hydrologica products; change X acidity & trated nutrients, l changes illegal increased sensitive pollutants, turbidity marijuana fire X water eutroph- cultivation, increased bodies ication human and human use pack stock waste

Hydrology Low flow, Runoff after Changes in Low flow, Water for Drought X Climate altered flow fires runoff, altered flow human use beetles - change X timing, lower flows, timing, forest die- increased vegetation flooding vegetation offs fire X change change vegetation (die-offs) (die-offs) change

Water Quality

● Water quality may be degraded due to a variety of stressors, including climate change, changing fire regimes, atmospheric deposition from air pollution, and contaminants from personal care products, illegal marijuana gardens and flame retardant. In general, water quality in SEKI currently is thought to be good, but some impacts occur and many uncertainties exist. ● Nutrient concentrations in SEKI surface waters are typically better than federal regulatory criteria; however, elevated atmospheric deposition is a concern because mountain lakes and streams are naturally very dilute and small changes in nutrient and acid chemistry may increase productivity, decrease lake clarity, and cause shifts in algal species. Nitrogen deposition in the parks exceeds critical loads suggested for Sierra Nevada lakes and lake nitrate concentrations often exceed ecological assessment points, but local understanding of impacts is lacking for most water bodies (Heard and Sickman 2016, Saros et al 2011) . Lower elevation stream biota also could be affected by elevated nutrient deposition, especially seasonally, but studies are lacking. Recent evidence shows that some Sierran lakes are recovering from higher acidic deposition in the past (Day and Conklin 2013, Heard et al. 2014, Panek et al. 2013). ● Microbial counts (E. coli) from surface waters with varied uses (human, pack stock, wildlife) are typically better than federal water quality standards, except during storms when microbial counts can temporarily rise. At lower elevations, water quality standards for E. coli have been achieved throughout the Middle Fork Kaweah River and the , and during a 2015 study of streams in the Ash Mountain pasture of Sequoia National Park (Clow et al. 2013; SWAMP 2015; E. Meyer, NPS-SEKI, personal communication). ● Aquatic organisms in the parks are exposed to a broad list of current- and historic-use pesticides (CUPs and HUPs), heavy metals (Hg, Pb, Zn), combustion by-products (PAHs), C-14

industrial contaminants (PCBs, PBDEs, etc), and personal care products (medicine, DEET, etc). Atmospheric deposition distributes many contaminants broadly; however, geology, visitor use, and illegal marijuana cultivation also contribute. Whole sampling was limited, most fish specimens were better than wildlife and human health consumption thresholds for pesticides, but some exceed standards for HUPs, DDT, and dieldrin (banned insecticide and persistent organic pollutant). CUPs and HUPs are widespread in SEKI surface water, and detrimental neurological effects were detected in some aquatic species. According to an unpublished 2014 EPA study, the insect repellent DEET is the most commonly detected emerging contaminant in the parks. Stimulants, flame retardants, and allergy and asthma medication also are common (Day and Conklin 2013; Landers et al. 2008; Meyer et al. 2013; E. Meyer, NPS, personal communication). ● Mercury (Hg) was detected in precipitation, snowpack, surface water, lichen, vegetation, and fish in remote wilderness in SEKI and other western parks. In SEKI, average fish Hg concentrations exceeded wildlife and human health guidelines. While data are only for the parks higher elevations, effects of Hg may be more pronounced at lower elevation aquatic ecosystems, which are more productive and have higher potential for bioaccumulation of toxic methyl mercury. Western pond turtles from two of SEKI’s low elevation streams had relatively high blood Hg levels. Higher trophic levels are more sensitive to bioaccumulation. Sensitive species occur in organic-rich aquatic habitats including wetlands, riparian floodplains, and other areas prone to wetting and drying periods that mobilize mercury (Landers et al. 2008, Eagles- Smith et al. 2014, Meyer et al. 2014). ● Changes in water temperature are not well studied in SEKI, but are a concern for the future. Water temperatures are likely to rise as a direct result of warming air temperatures, but also could warm due to less snowmelt, lower flows, and loss of shading vegetation. Warmer water temperatures, alone or in combination with lower flows, could affect aquatic ecosystems in several ways including eutrophication and reduced dissolved oxygen. Shallow lakes and ponds and streams without deep groundwater sources, persistent snowpack, or shading topography are most sensitive to warming air temperatures (Hauptfeld et al. 2014, Herbst 2013). ● Small increases in stream nitrate, sulfate, and calcium have been measured following fires (Williams and Melack 1997, Heard 2005). Increased fire frequency and severity, which could occur as a result of droughts and climate change interacting with fire suppression, could increase runoff and erosion over what was experienced in the past. Erosional events during storms could increase nutrients and turbidity and impact aquatic ecosystems and species. Crown fires could also contribute to warming water temperatures due to loss of shading vegetation (Hauptfeld et al. 2014) Confidence: Low - While data on exposure to pollutants is fairly robust, their effects on aquatic ecosystems and species are uncertain. Future trends in airborne contaminants and personal use products are uncertain. Warming of water temperatures is very likely, in general, but the spatial pattern across the landscape and where cold refugia may persist is uncertain. The effects of warming on aquatic systems and species may be very significant in some cases, but which species are most at risk is uncertain. In general there is low confidence in ability to extrapolate from local studies up to the landscape scale. We know a lot about broad deposition patterns, but an understanding of effects is localized.

C-15

Hydrology - Water Quantity and Timing

● Changes in hydrology may result from climatic change, shifts in fire regime, and vegetation change, including forest die-offs. Effects of water withdrawals for human use are thought to be minor, but data are lacking. ● Hydrologic changes have already been observed and attributed to climate changes. Continued changes are anticipated and very likely to profoundly affect water quantity and timing in SEKI over the next several decades. Trends toward earlier snowmelt and melting glaciers already are apparent in the western US, including Sierra Nevada parks. In SEKI, spring snowpack generally has declined at elevations below about 8,500 feet and earlier snowmelt streamflow was detected in some rivers. Snowpack has increased at the higher elevations in the past several decades. Annual precipitation has been variable and lacking trends over time (Andrews 2013, Basagic and Panek. 2013, Geos Institute. 2013). ● While changes in precipitation are highly uncertain, warming alone will cause an increase in the rain-to-snow ratio, earlier snowmelt, and an increase in evapotranspiration. In the future, snowpack is expected to decline. Climatic water deficit is likely to increase, with greatest increase in summer and increase or decrease in winter. Because of more extreme storms and warmer temperatures, larger floods are expected. Annual runoff may increase or decrease depending on future precipitation, but increased runoff is more likely in winter with a decrease in summer. Future climatic changes will play out on top of natural variability, meaning that severe droughts and wet periods are expected in any scenario. Hotter droughts, including interactions with insects, may lead to forest-dieback and resulting changes in hydrology (Geos Institute 2013, Das et al. 2011). ● Changes in water quantity and timing will vary across the landscape due to topography, subsurface hydrology, and vegetation cover interacting with climate change. Some areas may remain moister than others (Kershner 2014). ● Following large severe fires, runoff could increase temporarily (Kershner 2014). ● Groundwater in SEKI is poorly understood and not actively monitored. Confidence: Medium -Based on past observations and model projections, changes resulting from warming alone have high confidence (earlier snowmelt, decreased snowpack, increased climatic water deficit), changes in vegetation are uncertain but will likely impact the relationships between precipitation, runoff, and river discharge. Future changes in precipitation and runoff are uncertain (Geos Institute 2013).

C-16

AQUATIC ECOSYSTEMS AND SPECIES

Key vulnerabilities: Interactions of non-native species, warming, habitat drying, and excess nutrients.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Low Nutrients, Warmer Erosion, Sediment- Decreased Species Introduced Increased/ Increased/ New Warming X Elevation contamin- water, fire ation, summer extirpations pollutants & new new disease? drying X Aquatic ants drying suppres- stream- water diseases invasions invasion invasives X habitat sion bank nutrients impacts alteration

High Nutrients, Drying of Drying and Loss of Introduced Non-native Chytrid Invasives X Elevation contamin- habitat freezing of connect- pollutants & fish fungus, drying X Aquatic ants (streams, ponds ivity due to diseases new warming X ponds, drying diseases? nutrients wetlands) habitat

Low Elevation Aquatic Ecosystems & Species

● Lower elevation streams have several legacy, current, and potential future stressors that are likely to interact with each other. Past pesticide-use, disease, and fragmentation may have played a role in the extirpation of the foothill yellow legged frog, as well as likely continuing impact on species such as western pond turtle. Non-native fishes, including historically introduced species such as brown trout as well as more recent introductions such as green sunfish, have impacted native fish assemblages (Meyer and Boiano 2010). The effects of elevated atmospheric deposition of nutrients and contaminants are relatively unstudied for low elevation ecosystems and species, but may play a role in the decline of native fishes and amphibians. Warming temperatures and altered flows from climate change are expected to further alter lower elevation aquatic ecosystems and some changes may already be occurring (Kershner 2014, NPS 2013). ● Water temperatures may rise as a result of warming air temperatures and/or lower flows. Low elevation streams lacking high elevation headwaters may be especially vulnerable to warming and drying as a result of climate change and drought. Some habitats, such spring-fed streams, may be relatively buffered from climate change, however, compared to the surrounding landscape (see above - water quality and water quantity and timing). These“climate change refugia” may help ecosystems and species persist in the face of climate change (Kershner 2014, Morelli et al. 2016). ● Native rainbow trout are generally restricted by natural and human barriers to low elevations (<1000-1500 m) in the Kaweah River and up to mid elevations (2200-2400 m) in the Kings River (Austin et al. 2013). Water temperatures of about 25℃ could kill trout. Warming and drying of rivers would restrict these fish to the bottoms of the coldest pools, which may be subject to low dissolved oxygen from groundwater inputs or stratification (Matthews and Berg 1997). ● Warming temperatures and lower flows may benefit and cause upward migration of existing non-native species (such as black bullhead, green sunfish, and bullfrogs) and may assist new non-native species (such as quagga mussels, New Zealand mud snails, or aquatic disease organisms or parasites) to colonize the parks and displace native species (Kershner 2014). C-17

Some undiscovered non-native aquatic species may already be present in the parks. Preventing or reducing low elevation aquatic invasive species is a challenge since many low-elevation systems border non-park lands and are interconnected habitats. ● Warmer water, alone or in combination with atmospheric deposition of nutrients, could increase eutrophication of streams resulting in algal blooms, reduced oxygen levels, and species composition shifts. Egg-hatching species and larval stages are very sensitive to temperature change (Kershner 2014). ● Native species dependent on aquatic systems for all life stages may be limited to stream reaches fragmented by natural or human-made barriers. Isolated aquatic ecosystems, like small seeps, may have endemic species or specific genotypes (invertebrates and plants) with low capacity to recover after drying because they cannot disperse (rare species or rare genotypes)(Kershner 2014). ● Some native species, such as western pond turtles, are not limited to year-around aquatic habitat and have capacity to survive drying streams. Sex in western pond turtles is determined by nesting temperature, however, and warming would lead to a female biased population. Warming temperatures also could reduce nest success of western pond turtles (Geist et al. 2015). ● Increased fire frequency and severity could temporarily increase water flow, benefitting some species and harming others. Intense storm events following large severe fires on steep terrain could strip away aquatic habitat, cause fish kills, and harm other aquatic species. Loss of riparian canopy would enhance warming of streams (Kershner 2014). ● Effects of fire exclusion on watersheds include overgrowth in riparian zones; changes in timing and duration of wetness/drying (see wet meadows); ephemeral streams drying up (before newt larvae metamorphose, for example) (E. Meyer, NPS, observation). ● Some native species like Sacramento pikeminnow are expected to occur in SEKI, but are found rarely or not at all in the parks (Austin et al. 2013). This may be because of artificial barriers such as the power station on the Middle Fork Kaweah River at the park boundary. Confidence: Medium - How climate change will affect aquatic communities, food webs, and individual species is much less certain, even if confidence in some degree of continued warming is high. Interactions with atmospheric deposition, including contaminants also are uncertain. Confidence is moderate that non-native aquatic species, in general, will do better with climate change, than native species - especially those that rely solely on aquatic habitat and cannot migrate upstream due to barriers. There is low confidence in the ability to identify refugia.

High Elevation Aquatic Ecosystems & Species

● High elevation aquatic ecosystems are vulnerable to many stressors, including historic non- native fish introductions, disease (chytrid fungus), climate change, and atmospheric deposition of nutrients and contaminants. ● The widespread establishment of non-native trout in many high-elevation aquatic lakes and streams has fundamentally changed these ecosystems because the fish are both a predator and competitor of native species, and fragment native species’ populations. Changes include major shifts in aquatic food webs, species composition, and nutrient cycling. These changes affect terrestrial species, such as garter snakes, birds, and potentially bats that rely on lakes and streams for prey (Austin et al. 2013, EGPC and Schwartz 2013, NPS 2013).

C-18

● Non- native trout, coupled with the relatively recent emergence of chytrid fungus, is directly linked to the precipitous decline of the mountain yellow-legged frog and Sierra Nevada yellow legged frog (MYLFs), which today occupy a fraction of their historical range. All MYLF populations are susceptible to chytrid fungus, but a few populations appear to be developing persistence, indicating possible future genetic adaptation to resist the disease (Austin et al. 2013, EGPC and Schwartz 2013, NPS 2013). ● Management treatments to remove non-native trout have been highly successful at restoring MYLFs and other aquatic ecosystem components. The resulting increase in MYLFs also increases the likelihood for populations to survive subsequent chytrid fungus outbreaks. Management interventions using antifungal drugs appear to enable MYLFs to survive a chytrid outbreak, at least temporarily. Techniques, including translocation, captive rearing, immunization, and reintroduction of MYLFs to currently occupied and suitable historic habitat hold promise (Knapp et. al. 2007, NPS 2013b; R. Knapp, UC Santa Barbara, personal communication). ● High-elevation ponds may dry as a result of climate change. Conversely, lack of insulating snowpack could cause ponds to freeze through. Streams that link lakes, ponds, and wetlands may dry more often than historically, reducing cycling of nutrients and delivery of prey items to downstream ecosystems (Kershner 2014, Lacan et al. 2008), and potentially limiting the gene flow of frogs and other biota. ● Headwater streams in the southern Sierra are snowmelt-dominated (granite terrain) compared to more groundwater influenced streams in the north at similar elevation. Preliminary data suggest that the southern Sierra streams are more at-risk to drying and have lower invertebrate species diversity. Short headwater streams are more susceptible to drying. When streams dry, as seen during the recent drought, the stream environment shifted to a higher percentage of pool habitats. Pools harbored less invertebrate diversity than swift-flowing riffles due to more sediment deposition and less oxygen (Herbst 2015). ● Species that depend on aquatic systems for all life stages are limited to isolated lakes, ponds, and stream reaches fragmented by barriers (natural or human-made). Isolated aquatic ecosystems, like small seeps, may have endemic species or specific genotypes (invertebrates and wetland plants) that have low capacity to recover after drying because they have no method of dispersal (rare species or rare genotypes)(Kershner 2014). ● Water temperatures higher than about 20-25°C, which are lethal to some cold-water aquatic life, may occur as a result of warming temperatures alone or combined with lower flows. Warmer temperatures, alone or in combination with elevated atmospheric deposition of nutrients, could increase eutrophication of high elevation lakes and streams with accompanying shifts in species composition. Egg-hatching species and larval stages are very sensitive to changes in temperatures (Kershner 2014). ● Warming water temperatures and drying of aquatic habitat may further challenge the current and potential future restoration of threatened and endangered species (MYLFs, Yosemite toad, Little Kern golden trout) and persistence of the endemic Kern River rainbow trout, which are rated as highly vulnerable to climate change (Moyle et al. 2013). ● Some habitats, such a deep lakes and spring-fed streams, may be relatively buffered from climate change compared to the surrounding landscape (see above - water quality and water quantity and timing). These “climate change refugia” may help ecosystems and species persist in the face of climate change (Morelli et al. 2016).

C-19

● Effects of contaminants on high elevation aquatic species is a concern, although recent research did not find a negative association between MYLF population status and concentrations of 46 pesticide compounds (Bradford et al. 2011) (see above - water quality). Confidence: Medium to High - There is high confidence in understanding the effects of non- native trout and chytrid fungus on native species. There is high confidence in fish eradication to restore habitat and medium confidence in chytrid fungus treatment options. How climate change will affect aquatic communities, food webs, and individual species is much less certain. Interactions with atmospheric deposition, including contaminants also are uncertain. Due to the inventory of high lakes, there is high confidence in the ability to identify habitats that will buffer warming and drying (i.e., cold, deep, and/or large lakes; however, there is low to medium confidence in the ability to identify refugia (i.e., whether organisms will thrive there).

C-20

CAVE AND KARST SYSTEMS

Key vulnerabilities: Human use, warming and drying of habitat, contaminants, and diseases combined with a very sensitive resource.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Cave and Nutrients, Drying Erosion, Flooding, Drying of Limited Increased Increased/ White nose Warming X karst contam- habitat, fire erosion, surface and migration human new syndrome drying X systems inants reduced suppres- and subsurface among impacts invasion or other visitor speleothem sion sediments water flows isolated disease impacts generation impacts caves

● Cave formations and cave biota are sensitive resources that may be impacted by changes in environmental conditions and increased visitation. ● Some cave species, such as the harvestmen Calcina cloughensis, are endemic to SEKI and are known to occur in one or a few caves. Endemic invertebrates include scorpions, spiders, harvestmen, pseudoscorpions, diplurans, beetles, millipedes, isopods, and rockhopper (Panek and Despain 2013). ● Troglobionts (obligate cave species) are stenothermic (adapted to a narrow range of temperatures) and stenohygrobic (restricted to areas with nearly 100% humidity)(Barr and Kuehne 1971, Howarth 1980). ● Cave temperatures are closely related to mean annual temperature outside the cave. With increasing distance from cave entrance, however, temperature changes are increasingly attenuated and lagged relative to surface temperature (Toben et al. 2013). Changes in cave temperature will follow local trends with corresponding impacts to cave biota. Therefore, warming air temperatures are predicted to increase temperatures within caves. ● Decreased water quantity may lead to drying of cave environments, reduced speleothem formation or development, and reduction in habitat for cave biota, especially aquatic species such as, Bowmanasellus sequoia. Caves are vulnerable to local changes in groundwater (Tobin and Schwartz 2011). Passages may dry out or flood in response to local conditions. ● Cave vertebrates and invertebrates are sensitive to the direct effect of increased temperatures and the indirect effect of decreased summer water availability due to decreased snowpack. Most caves have cool temperatures and near-saturated humidity, to which the cave biota are adapted. As surface temperatures rise, internal cave temperatures respond accordingly. If temperatures continue to rise at their current rate, some biota may not adapt (Panek and Despain 2013). Observations in Clough Cave have demonstrated a change in the numbers of active animals throughout the summer season (Toben et al. 2013). It is presumed that animals migrate to seek more favorable conditions either within the cave or in another environment when possible. While many animals are present and active in the spring months when the cave is at its wettest with the highest humidity, later in the year animals are no longer active and very difficult to find (Darrell Ubick, California Academy of Sciences, pers comm.). ● Cave formations, such as speleothems, are sensitive to breakage or weathering and are irreplaceable physical resources. Human visitors have damaged some cave formations. Careless acts or wanton destruction is permanent. C-21

● Human impacts to cave invertebrates also have been recognized. Artificial light sources in tour caves can promote algal growth (Meyer et al. 2016). Human frequently deposit lint and food crumbs into caves. These subsidies may alter feeding behavior, diets, and distribution of cave biota. Recreational users in South Fork caves and Crystal Cave have trampled animals. ● White Nose Syndrome (WNS) is a disease that has killed millions of bats in eastern North America and has steadily spread westward from where it emerged in upstate New York in 2006. In 2016 it was detected on the West Coast in Washington (Sleeman 2016). WNS is associated with a fungus that thrives in cold and humid conditions characteristic of caves and primarily affects hibernating bats (WNS.org 2016). There is some evidence that SEKI bats do not commonly roost in caves and therefore may be less vulnerable to WNS. Acoustic monitoring in winter 2014-2015 detected 15 bat species, but only two species were found in caves. Out of 60 caves surveyed only 11 had bats and the largest aggregation was eight (Frick et al. 2015). There is anecdotal evidence that large maternal colonies might have roosted in Clough Cave prior to human disturbance. Despite more than a decade of gating in Clough Cave, bats have not re-occupied the site. ● Invasive non-native species entering caves may lead to disease, disrupted energy cycles, and altered predator-prey dynamics. ● Invertebrates may be susceptible to altered fire regimes, as fires causes chemical changes to surface streams, which then flow underground to become subsurface streams. Fires also cause increased organic matter in streams. Fire-retardants can be washed into streams after application and may be toxic to invertebrate species. However, evidence shows B. sequoiae to be robust to fire effects; the Yucca Creek and Redwood Canyon watersheds have seen extensive fires in recent decades, including fire retardant application in 2008 (Tobin et al. 2009), and B. sequoiae appear to persist in these watersheds without change in measured populations . Fires could also burn trees with roots that are food sources for invertebrates in caves, such as in Clough Cave, the invertebrate population could be adversely impacted (Panek and Despain 2013). ● The Generals Highway crosses both watersheds containing B. sequoiae and spills of hazardous materials along the highway could be dangerous to these animals. ● It is possible that airborne nitrogen fertilizers and pesticides from the heavily farmed upwind San Joaquin Valley and from illegal marijuana cultivation within the parks may enter water systems that enter the caves. Currently, the exposure and impact is unknown (Panek and Despain 2013). Confidence: Low - Detailed inventory of cave environments, features, and biota are limited to a small proportion of known caves and very few species have been studied in any detail. Therefore confidence is low.

C-22

WET MEADOW AND FENS

Key vulnerability: Loss of wetland area due to warming and drying of habitat and interactions with other stressors that may speed erosion (including gully formation), lowering of water table, and loss of wetland soils.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Wet Nutrient & Shrinkage, Watershed Runoff, Decreased Historic Trampling, Increased/ Increased/ Yosemite Drying X meadows contam- total loss of effects, tree gully summer livestock new roads new new toad other and fens inant wetland migration formation water grazing, and trails invasion invasion sensitive to stressors deposition pack stock chytrid (peat fires, use fungus? soil loss, gully formation)

 Exposure to stressors for wet meadows and fens include roads and trails; historical livestock grazing, present-day pack stock grazing, and foot traffic; invasion by non-native plants; atmospheric deposition of pollutants; altered fire regimes; and climate change. o Roads and trails can impact meadows by concentrating water flows in meadows. The extent of impacts is believed to be small (although severity may be high), and there are ongoing efforts to re-route trails out of wetlands, redesign drainage during road construction and maintenance, and to restore meadows impacted by roads and trails. o Livestock grazing is eliminated in the parks, although some impacts persist, particularly where grazing was most intense and most recently abandoned (Neuman 1996, Wolf et al. 2015). Total administrative and recreational pack stock grazing has been declining over time and is increasingly concentrated in fewer, more tightly regulated locations (Hopkinson et al. 2013). Potential effects of pack stock use include physical changes to wetlands, such as soil compaction or erosion, as well as impacts to vegetation structure and small changes in macroinvertebrate communities (Ostoja et al. 2015). o High elevation wet meadows and fens in the Sierra Nevada have a low incidence of invasive non-native plant species. Lower elevation wetlands (particularly drier systems) often support introduced pasture grasses such as Kentucky bluegrass (Poa pratensis) and redtop (Agrostis gigantea). Several montane wet meadows in SEKI and Yosemite National Park have been invaded by velvet grass (Holcus lanatus) and reed canary grass (Phalaris arundinacea). In SEKI, 39 non-native plant species were documented in wet meadows and up to 17% of the wet meadow area may be infested (D’Antonio et al. 2004, Gerlach et al. 2003, Tu et al. 2013). o Changes in fire regimes due to management actions or climate may impact wet meadows and fens through changes in water availability, sediment, and impacts to soils. Fire suppression can aid tree and shrub migration into meadows due to lack of periodic burning in and around meadows. Fire suppression may indirectly reduce soil moisture in downstream meadows if upstream forests become dense and increase their evapo- transpiration rate (Kershner 2014). Fires upstream of a meadow may cause temporary surface and groundwater increases for a few years following a fire. Large fires can also increase the amount and alter the type of sediments delivered to a meadow.

C-23

o Since the hydrology of wet meadows and fens is the driving force behind their community structure and function, significant alteration of meadow hydrology from climate change or other stressors is considered a potential major impact on these systems. Alteration of hydrologic regimes due to climate change, including changes in the amount, or timing of precipitation, snowmelt, or runoff could alter the species composition, productivity, and overall function of wet meadows and fens (Hopkinson et al. 2014, Kershner 2014, Viers et al. 2013, Wolf et al. 2015). Reduction in spring snowpack or prolonged drought could reduce the amount of time that soils are inundated, allowing the establishment of upland plant species including trees and shrubs. Wetter conditions could cause expansion of wet meadows and fens in some places. o Warmer temperatures could increase productivity in moderately wet meadows (Moore et al. 2012).  Sensitivity: Sensitivity of wet meadow and fens is expected to vary widely across the landscape. Climate change may worsen impacts to meadows already impacted by other stressors due to direct effects on plant species and indirect effects on hydrology (Kershner 2014, Viers et al. 2013). o Some meadows (particularly those already affected by channel and bank instability, incision, and decreased water tables) are sensitive to flashy runoff events, which can exacerbate stream incision and downcutting that lower the water table. Significant lowering of the water table leads to drying of wet meadow soils, shifts toward non- wetland species, decomposition of accumulated organic soils, and loss of wet function. o While we don’t know the degree of impact to vegetation, soils, and hydrology in most of SEKI’s wet meadows and fens, there are examples of places that have been severely degraded due to impacts from road building (e.g., Halstead Meadow) and historic livestock grazing (e.g., Cahoon Meadow). In both cases, water flow was significantly altered, which led to erosion, gully-formation, lowering of water table, drying of soils, reduction in wetland plant species, and loss of wetland function (Wolf et al. 2015). o Where available, monitoring data for meadows in SEKI suggest that current levels of pack stock use do not adversely impact vegetation composition or productivity (Hopkinson et al. 2013). o A warmer climate may remove abiotic barriers to the establishment and growth of non- native species and make meadows and fens more sensitive to impacts from those species. o Fire frequency and risk of large, high severity fires may increase with climate change (Hauptfeld 2014c). Under warmer, drier conditions meadows may become more sensitive to fire; if they burn when soils are dry, long-sequestered organic matter may be consumed (DeBenedetti and Parsons 1984), releasing additional greenhouse gasses and dramatically altering soil properties. o Individual meadows and their watersheds will be exposed to different degrees of climate- driven hydrologic change (Morelli et al. 2016). Meadow response to changing hydroclimatic conditions also will vary depending on several factors, such as the elevation, topography, geology, soils, vegetation composition, and existing degree of human alterations to the meadow and its watershed. Meadows above the snow-rain transition and those with stable groundwater sources may be less sensitive to climate- driven hydrologic change. C-24

 Response of meadows to past climate and disturbance events provides an indication of adaptive capacity. o Establishment of modern montane wet meadows began about 2,500 years ago under climate conditions that were cooler than the present. If future temperatures are equal or greater than that of the middle Holocene, sites currently occupied by wet meadows may be occupied by upland forest vegetation as they were in the past (Wood 1975). o Impacts from intensive livestock (cattle and sheep) grazing particularly can be seen in places where heavy grazing persisted into the 20th century on USFS or private lands that later entered NPS ownership, such as Cahoon Meadow, Sugarloaf Meadow, and other meadows in the Roaring River and Sugarloaf Creek watersheds. These impacts have not returned to pre-disturbance conditions (Wolf and Cooper 2016). Less intense impacts from pack stock grazing appear to be more transient. Large gully formation is a catastrophic event only reversible through extremely costly restoration work.  Meadows and fens are palustrine wetlands with direct connections to aquatic systems. Impacts and changes in wetlands are likely to have consequences for connected aquatic systems and habitat for certain species. o Atmospheric contaminants have the potential to adversely impact aquatic biota through estrogenic effects and bioaccumulation (see water quality and aquatic ecosystems; Panek et al. 2013, Landers et al. 2008). Wetland vegetation is capable of assimilating nutrients from atmospheric deposition, although high amounts of nitrogen can affect carbon cycling; loss of wetlands may increase the amount of these nutrients entering aquatic systems (Verhoeven et al. 2006). o Meadows represent approximately 1% of the land base in the Sierra Nevada and 3% in the parks. The non-uniform distribution and lack of connectivity between meadows may reduce the adaptive capacity of meadow-dependent species to cope with climatic changes, depending on their dispersal ability. Some meadows are more connected than others, however, and may serve as dispersal hubs for connecting large populations of meadow-dependent species and for species migration to optimal habitats (see high elevation aquatic systems and species).  Meadows serve as primary habitat for Yosemite toads, which are listed as threatened under the U.S. Endangered Species Act. This species is endemic to the Sierra Nevada with the southern portion of its range located in the northern third of SEKI where it was documented by USGS surveys as occupying 42 meadows. Yosemite toad occupancy is fairly constant year to year at the watershed scale and is strongly related to snowpack. Occupancy at the site scale is more variable year to year, however, and is not well understood. Yosemite toad is vulnerable to the stressors discussed above that reduce meadow and surface water habitat. Grazing and recreation may be of greatest concern to the Yosemite toad because their effects are widespread, frequent, persistent, can be locally intense, and are potentially irreversible. Additionally, the species may be sensitive to disease (chytrid fungus) (NPS 2013a).

Confidence: Low - There is high confidence in the current distribution of many stressors. However, we only know the current severity of human alteration to vegetation, soils, and hydrology for a small subset of meadows and fens. Understanding expected changes in hydrology due to climate change is key to understanding vulnerability and adaptive capacity of these features. Confidence is high that impacts from climate will vary spatially, but tools to predict which wetlands will be most vulnerable are still in development.

C-25

FOOTHILLS TERRESTRIAL ECOSYSTEMS

Key vulnerability: Drought, warming, and changing fire regimes. Prolonged, hot drought currently is causing decline in oaks, chaparral, and manzanita and may further depress seedling regeneration. Altered fire regime, including increases in fire severity may cause contraction in oak woodlands (in favor of shrublands) and increased fire frequency may lead to conversion of shrublands to annual grasslands.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Foothills Nutrient Decreased Too much/ Loss of Increased Type Increased/ Increased Heat X terrestrial deposition soil too little fire. connectivity risk of conversion new damage Drought ecosystem moisture Suppression priority invasion from impacts species insects and introduction disease

Oak Woodlands

● According to the 2013 Natural Resource Condition Assessment (NRCA), foothills floristic integrity is variable by vegetation type: while herbaceous types (and layers) exhibit high diversity, they are also the most invaded by non-native species; hardwood diversity is low to moderate. The NRCA also noted concern that blue oak does not have enough regeneration to balance mortality (Rodriguez-Buritica and Suding 2013). ● The Sierra Nevada Ecosystem Vulnerability Assessment (Kershner 2014) lists the overall vulnerability of oak woodlands as low- moderate due to moderate-high sensitivity to stressors, moderate-high adaptive capacity, and low exposure. For blue oak ecosystems, vulnerability was rated as moderate due to moderate sensitivity to stressors, moderate adaptive capacity, and low-moderate exposure. ● Climate and drought ○ Though oak species are tolerant to high temperature and drought, the recent prolonged, hot drought is taking a significant toll in mortality of mature trees, particularly blue and live oaks (USFS 2016). ○ Sensitivity to climate change varies by species. Oak species distribution is predicted by precipitation, so a generally drying climate would be expected to result in shifts in dominant species of oaks. Warming and drying is expected to cause increased seedling mortality (Kershner 2014). Decreased soil moisture may further depress acorn germination in blue oak (Rodriguez-Buritica and Suding 2013). ● Altered Fire Regime, Herbivory, and Nonnative Plants ○ Sensitivity to fire varies by species: canyon live oak is very sensitive to fire; blue oak is less sensitive than interior live oak (Kershner 2014). Mean maximum fire return interval (MFRI) is 17 and 23, years for foothills hardwood/grasslands and mid-elevation hardwood, respectively (Caprio and Lineback 2002). Blue oak resprouts after fire, accounting for observed increases in post-fire regeneration in past studies. Resprouting is negatively correlated with diameter, with 20 cm diameter at breast height being a general threshold, so larger blue oak may be disproportionately impacted by intense fire (Rodriguez-Buritica and Suding 2013). C-26

○ Grazing by pack stock and occasional trespass cattle can impact oak woodlands. Cattle and deer have been observed to graze more under blue oak canopies where there is richer forage, and to eat acorns. Intense summer grazing may favor non-native grasses (Rodriguez-Buritica and Suding 2013). Managers can adjust timing and intensity of pack animal grazing to minimize impacts and deer abundance is well below what it was 30-50 years ago. Acorn consumption and seedling herbivory may interact with drought to further reduce regeneration due to low production of acorns and seedlings. ○ Seed-caching birds and mammals also facilitate blue oak regeneration by burying acorns, which increases the germination rate compared to acorns on the surface (Kershner 2014). This interaction may confer some adaptive capacity where wildlife populations are present within healthy food webs. ○ Dominance in the understory of non-native annual grasses has been shown to reduce blue oak seedling establishment, likely through reducing soil moisture availability. Thick litter produced by grasses also prevents the acorn radicle from reaching the soil under experimental conditions (Borchert et. al. 1989). Although type conversion to dominance by nonnative annual grasses has already occurred in herbaceous understory and grassland communities, these types are still vulnerable to invasion by transformer species such as yellow starthistle (Centaurea solstitialis) (Rodriguez-Buritica and Suding 2013). ● Pathogens and Insects ○ Species including live oaks, black oaks, California buckeye, and whiteleaf manzanita, known to be hosts to Sudden Oak Death (SOD; Phytophthora ramorum) are abundant in the southern Sierra Nevada. However, Meentemeyer et. al. (2004) predicted that there is very low risk of SOD becoming established in the parks due to warm, dry climate at low elevations and cold climate at high elevations. A strong El Nino year could increase the odds of spread if the pathogen is introduced (Kershner 2014). Other pathogens, including others in the Phytophthora genus are showing up in native plant nursery stock and restoration sites (in a few CA locations, not SEKI) and warrant watching (M. Baer-Keeley, NPS, pers. comm. 2015). ○ Gold-spotted oakborer may also present a threat to host species black oak and canyon live oak that are abundant in the parks (NPS 2013). ● Air Pollution - Exposure to pollutants ozone and excess nitrogen is highest in the foothills compared to higher elevation areas of the parks (Panek et al 2013), but the sensitivity of foothills flora to these pollutants may be comparatively low. Ozone can damage vegetation, but damage has not been detected in foothills vegetation types. Impacts of elevated nitrogen deposition appear to be limited in the foothills to accelerated nitrogen cycling rates, leaching of nitrate from soils, loss of sensitive epiphytic lichen species, and enhancement of non-native plant species (e.g. Italian thistle that tends to proliferate under tree canopies) (McGinnis and Keeley 2011, Fenn et al. 2010). ● Fragmentation and Land Use - The majority of lower elevation oak woodlands in the parks ecoregion are found on private property and thus they are more subject to fragmentation, potentially damaging summer grazing, and loss of groundwater to domestic use. This distribution has implications for the spread of fire and nonnative species and wildlife movement/behaviors (e.g. bears that get food-conditioned outside parks) into the parks. ● Adaptive Capacity

C-27

○ Oaks have innate adaptive capacity due to their broad distribution, ecosystem diversity (e.g. small patch size), and adaptations to fire (e.g. stump sprouting after top-kill) and drought. Innate adaptive capacity of oaks is limited by slow growth: functional replacement of mature trees may take 50-100 years (Kershner 2014). ○ Blue oak regeneration has been extensively studied and there is general agreement that mortality is greater than regeneration. (Rodriguez-Buritica and Suding 2013). Lack of baseline information (e.g. herbaceous species composition, disturbance regimes) and monitoring in foothill ecosystems reduces management capacity due to a resulting inability to detect change as it occurs and to respond before changes are irreversible. Direct management with fire is limited due to lack of funding, proximity to developed areas, and difficulty of finding an appropriate prescription window to obtain desired effects and minimize risk.

Confidence: Low-High - Low for understanding of fire and grazing history. In general, we lack spatially complete information about foothills ecosystems. High for current impacts of drought.

Chaparral

● According to the 2013 Natural Resource Condition Assessment (NRCA), foothills floristic integrity is variable by vegetation type: shrublands exhibit the lowest diversity of the foothills systems overall, but shrub diversity varies from low to high within the different watersheds. Shrublands were found to have the lowest overall herbaceous diversity, and the lowest native grass abundance: this is probably due to dense shrub canopies excluding most herbaceous plants (Rodriguez-Buritica and Suding 2013). ● The Sierra Nevada Ecosystem Vulnerability Assessment lists the overall vulnerability of chaparral systems as moderate due to moderate sensitivity to stressors, moderate adaptive capacity, and moderate exposure. Chaparral systems may give way to grasslands in some locations, but may expand into former woodlands in response to warming temperatures and after fires (Kershner 2014). ● Altered Fire Regime and Non-native Plants ○ Mean maximum fire return interval (MFRI) is 60 and 75 years for foothills chaparral and montane chaparral, respectively (Caprio and Lineback 2002). Keeley et. al. (2005b) found few negative impacts of very long fire return interval (~150 years) on shrub regeneration, though these stands were found to be more highly invaded by non-native species than mature stands that had burned more recently. ○ Less frequent fire tends to prevent invasion of non-native annual grasses and tends to favor shade tolerant species (Keeley et. al. 2005b). Fire that is too frequent, estimated at two fires within 10 years by Keeley et. al. (2005a), may lead to conversion to grasslands dominated by non-native annual grasses due to woody species regeneration inhibition (Rodriguez-Buritica and Suding 2013, Kershner 2014). ○ Among chaparral species, obligate seeders such as Ceanothus cuneatus (buckbrush) and Arctostaphylos viscida (whiteleaf manzanita) need fire to open the canopy and reproduce, but they are more susceptible to very frequent fire. Obligate resprouters such as Cercocarpus betuloides (birchleaf mountain mahogany) may be favored by more frequent fire (Keeley et. al. 2005b, Kershner 2014).

C-28

● Drought ○ Though chaparral are thought to be tolerant to high temperature and drought, the recent prolonged, hot drought appears to have increased death rates of chaparral in some areas. Manzanitas also appear to be in decline in lower elevations (manager observations). ○ Shrub species that resprout after fire tend to be less drought-tolerant, while post-fire seeding species are more resistant to drought (Keeley et. al. 2005b and others cited in Kershner 2014). ● Pathogens - Whiteleaf manzanita and California buckeye are known hosts of sudden oak death (SOD). The warm dry climate of the southern Sierra Nevada doesn’t favor SOD, but a shift to warmer wetter conditions may allow SOD to establish if introduced. ● Adaptive Capacity ○ Chaparral species exhibit high adaptive capacity due to life history traits (heat and drought tolerance, long generations, and long persistence of seed banks (Kershner 2014). Ecosystems exhibit moderate adaptive capacity due to broad distribution with high natural and human-caused fragmentation (Kershner 2014). Chaparral may be able to expand into previously wooded habitat after large fires (Kershner 2014). ○ Lack of baseline information and monitoring in foothill ecosystems reduces management capacity due to a resulting inability to detect change as it occurs and to respond before changes are irreversible. Confidence: Medium - Moderate to high for general sensitivity, exposure, and adaptive capacity information, but lower for applicability to local conditions.

C-29

GIANT SEQUOIAS

Key vulnerabilities: Interaction among hotter droughts, fire, and unknown future insects or pathogens.

Air Warming/ Altered Fire More Loss of Fragmentat Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ion and visitation plants animals disease Stressor weather land use Interactions

Giant Ozone Moisture Lack of fire; Increased Moisture Isolated Soil Increased/ Increased/ Unknown Hot drought sequoia stress on Stress large high wind Stress groves, low compaction new new insects and X fire X seedlings severity damage? genetic - seedling invasion invasion disease in insects/ fires diversity, regener- the future? pathogens limited ation dispersal

Mature giant sequoia trees

● Mature giant sequoias currently appear to resist, cope with, or recover from stressors. Biotic mortality agents (insects, disease) able to kill giant sequoias currently are minor to non-existent in the species’ native range (York et al. 2013). Death rates are thought to be relatively stable based on park staff staff observations, although monitoring data are lacking. ● In the future, however, synergies among rising temperatures, drought, high severity fire, insects, and disease (such as root rot) may exceed their sensitivity tolerances (York et al. 2013). The paleoecological record supports this hypothesis; pollen records suggest that giant sequoias may have been near extinction during a period of slightly warmer temperatures from about 10,000 to 4,000 years ago (Davis 1999, Anderson 1999, Anderson and Smith 1994). ● Giant sequoia groves occupy a narrow band of elevation from 1400-2150 m (4600-7000 ft) in the Sierra Nevada where they experience a range of temperatures from average summer highs of about 29 °C (84 °F) to average winter lows of about -5 °C (25 °F) (Harrison 2013). Giant sequoia is a water-demanding species with a single mature individual using about 2000 liters on a summer day (Ambrose et al. 2016). ● Moisture deficit and other stressor thresholds are difficult to define for this long-lived species, and recent “baseline” climate conditions may not represent survival thresholds. With these caveats, changes in hydroclimate conditions (hydroclimate envelope) project potential relative exposure to climate changes. Under two climate change scenarios, <3% of current locations of sequoia groves in the parks are projected to be outside of the 99th percentile baseline hydro- climate envelope by 2010-2039. In the much warmer and drier scenario (GFDL-A2), 74% of current locations of sequoia groves in the parks are projected to be outside of the 99th percentile baseline hydroclimate envelope by 2070-2099. In the moderately warmer and similar precipitation scenario (PCM-A2), 25% of current locations of sequoia groves in the parks are projected to be outside of the 99th percentile baseline hydroclimate envelope by 2070-2099. Compared to all the grove area in the southern Sierra Nevada, groves in SEKI have lower projected climate exposure (M. Schwartz, University of California Davis, unpublished data). ● Hotter droughts, which combine the warming effects of climate change with natural variability in precipitation, are expected to be a significant driver of tree mortality and forest die-off (Allen et al. 2015). Observations of foliage dieback during the recent warm drought suggest that about half of the observed mature sequoia were under stress and about 10% more significantly so (N. Stephenson, unpublished data). Water potential measurements of 50 mature sequoias in late C-30

summer 2015 indicate variable response to drought, with some trees healthy and some trees the most stressed ever measured. One of the 50 trees has since died during the drought (A. Ambrose, University of California Berkeley, unpublished data). Dieback of foliage may be a mechanism that allowed most of the stressed giant sequoia trees to survive the drought. ● Giant sequoia groves have the highest Fire Return Index Departure (FRID) among SEKI’s forest vegetation alliances, as they experienced fire about every 5-30 years prior to EuroAmerican fire exclusion, and fire and fuels management treatments have been focused on the most accessible groves. Giant sequoia may be especially prone to uncharacteristically severe fires and sensitive to climate change in areas where fire exclusion has increased fuel loads, ladder fuels, tree density, and biomass (Battles et al. 2013b, York et al. 2013). A preliminary analysis supports the hypothesis that prescribed fire promotes resistance to drought in low elevation forests in the parks (van Mantgem et al. 2016), and this may be true of sequoia groves. ● Low genetic diversity, together with competition in an environment to which giant sequoia is likely already poorly adapted, will pose major constraints on its success in the face of increasing aridity (Dodd and DeSilva 2016). ● Park staff hypothesize that giant sequoia may be more sensitive in areas of marginal environmental conditions, such as lower elevation grove areas, grove edges, small groves, or areas with low subsurface moisture storage. ● Accessible giant sequoia groves are the more heavily-visited areas within the parks, leading to soil compaction, loss of topsoil, and reductions in soil organic matter for the small areas of groves affected. While there is concern about impacts on mature sequoia, the more relevant impact of concern is on local giant sequoia regeneration because of the likely loss of adequate rooting substrates for seed germination on heavily compacted soils (Harrison 2013, NPS 2013, York et al. 2013). Confidence: Medium - High confidence in rising temperatures and moisture stress, medium confidence in more large high severity fire, low confidence in insect and pathogen exposure, low confidence in predicting thresholds of sensitivity to these stressors, but medium confidence that the interaction of these stressors will negatively impact giant sequoias.

Giant sequoia seedling regeneration

● Giant sequoia seedling regeneration is vulnerable because it requires a combination of specific fire, moisture, and light conditions. Seedling regeneration is sensitive to lack of fire, but also to increased moisture stress. Warming and the resulting loss in available moisture could halt sequoia regeneration even when fires are sufficient to promote reproduction. Pines, firs, incense cedar, and oaks are tolerant to a broader range of conditions for reproduction compared to giant sequoia. Sequoia seedlings also are somewhat sensitive to ozone pollution and may be affected by soil compaction in local areas of high visitor use. These effects may interact with other stressors (NPS 2013, York et al. 203, Harrison 2013). ● Giant sequoia are reproductive for thousands of years, so a few hundred year gap in reproduction might be tolerated. Seed dispersal is limited to 500 m (Harvey 1980), so large highly disturbed areas may not be naturally reseeded and ability for natural migration to new suitable areas is low. In 2015 after four years of drought, sequoia seedlings appear to mostly be tolerating low soil moisture; however, foliage dieback is occurring on sequoia seedlings in some areas. Water potential measurements of 30 sequoia seedlings in late summer 2015 indicate variable C-31

response to drought, with some seedlings healthy and some the most stressed ever measured. The impact of July 2015 rainstorms to soil moisture at different depths in not known, however, and may have aided seedling survival (A. Ambrose, University of California Berkeley, unpublished data).

Confidence: Medium - Relationship of seedling regeneration to fire forest gap creation and the seed dispersal distance are well documented. The moisture tolerance of giant sequoia seedlings is less known as is the likelihood of seed source destruction from large very severe fires. It is not known how long a gap in reproduction can be tolerated by giant sequoia before long-term sustainability is threatened.

Giant sequoia grove ecosystems

● Tree species that co-occur with giant sequoia in the groves are more sensitive to many stressors than giant sequoia; however, specific stressors often affect certain species. For example, sugar pine is sensitive to white pine blister rust; Jeffrey and ponderosa pine are sensitive to ozone pollution; and a variety of these species are susceptible to bark and ips beetles. Increasing mortality rates of pines and firs over recent decades have been attributed to warming temperatures (van Mantgem and Stephenson 2007). ● Giant sequoia may prove to be climate change refugia for montane forest ecosystems. Moisture in sequoia groves is, in general, greater than adjacent non-grove areas, which may provide some degree of drought resistance for tree species co-located with giant sequoia. In 2015-2016, death rates of pine, firs, and incense cedars appear lower within sequoia groves compared to death rates of these species at similar or slightly lower elevations outside of groves (park staff, anecdotal observations). In contrast, moist conditions in sequoia groves may promote white pine blister rust, which is killing sugar pines (J. Nesmith, NPS, and A. Das, USGS, unpublished data). ● Both giant sequoias and other co-occurring tree species may be more sensitive to stressors in areas where fire exclusion has increased fuel loads, ladder fuels, tree density, and biomass (van Mantgem et al. 2016). Park staff also hypothesize they also may be more sensitive in areas of marginal environmental conditions, such as lower elevation grove areas, grove edges, small groves, or areas with low subsurface moisture storage. Confidence: Medium - Sensitivity to current exposure levels of these stressors has been documented and they are projected to increase with climatic change. Variable responses are expected across grove areas and confidence

Giant sequoia management capacity

 Capacity of management to reduce vulnerability is mixed. Tools such as prescribed fire, managed wildfire, mechanical thinning, planting, and irrigation exist to reduce vulnerability but constraints such as budget, poor air quality, risk to human safety and property, wilderness constraints, and access make it unlikely that managers can apply these tools at large scales. Because sequoia are highly valued and some of the grove areas are accessible, management capacity is higher than for other forest types.

Confidence: Medium - Ratings for this category are based on expert judgment, experience, and recent observed trends.

C-32

FORESTS

Key vulnerability: Interactions among hotter droughts, current/future insects and diseases, and altered fire regimes. Currently, the exposure to and impacts of these stressors are more severe in lower elevation montane forests. In the long-term, subalpine forests may be more vulnerable, however, due to less adaptive capacity to cope with or recover from impacts.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Montane Ozone Moisture Fire Moisture More Increased/ Increased/ Weakened Hotter forests injury stress; exclusion X stress ignitions? new new or killed drought X upslope climate Introduced invasion invasion trees insects/ movement change = species? pathogens of suitable larger, X fire range more severe fires

Subalpine Contraction Increased Moisture (occurs in More Increased/ Increased/ Weakened Hotter forests of suitable fire stress more ignitions? new new or killed drought X range frequency fragmented Introduced invasion invasion trees insects/ stands than species? pathogens montane) X fire

Whitebark Contraction Increased Moisture (occurs in Weakened Hotter pine of suitable fire stress more or killed drought X range frequency fragmented trees beetles X stands than (blister rust blister rust montane) X beetles)

Montane Forests

● According to the 2013 Natural Resource Condition Assessment (NRCA), the parks’ intact forests generally have high ecological integrity. These forests provide resilience to large disturbances; maintain forest function across broad swathes of landscape; provide good habitat for forest-dependent species; have good representation of forest structural stages; have sufficient supply of snags; and have higher big-tree density and live-tree biomass in the parks compared to the region (Battles et al. 2013b). ● The NRCA also concluded that some areas of SEKI forest are exposed to high level of stressors including fire exclusion, air pollution, white pine blister rust, and climatic warming. Currently, these stressors are more intense for lower elevation forests versus those at higher elevations. These stressors and their interactions could affect forest ecological integrity in the future (Battles et al. 2013b). ● The Sierra Nevada Ecosystem Vulnerability Assessment (Kershner 2014) lists the overall vulnerability of yellow pine/mixed conifer forests as moderate-high to high due to moderate-high sensitivity to stressors, moderate-high adaptive capacity, and moderate exposure. For red fir ecosystems, vulnerability was rated as moderate due to moderate-high sensitivity to stressors, moderate adaptive capacity, and moderate exposure. ● Fire ○ For more than 40% of SEKI’s montane forest, fire has been absent for two or more cycles of the historic fire return interval, placing much of the forest at risk of severe fire. High C-33

FRID levels were concentrated in the Kaweah River Basin. Forests in other SEKI watersheds were much closer to the historic fire regime (Battles et al. 2013b). ○ Lack of periodic low- and mixed-intensity fire in some lower and middle elevation montane forests has caused increases in overall forest density and fuels and shifts in forest composition toward more shade-tolerant species. These alterations can increase fire hazard while decreasing resistance and resilience of the forest to insects, disease, warming temperatures, and drought (Battles et al. 2013b, Kershner 2014). ○ Changing disturbance regimes (e.g., increases in fire frequency and burned area, and, in some forest types, fire severity) are projected in future climate change scenarios and are likely to be the most significant influence on changes in vegetation types and distributions (Safford et al. 2012). ○ Shade-tolerant species that currently dominate forests because of fire suppression may be less drought and fire tolerant than the pines and oaks they are supplanting (Kershner 2014). ○ Intense storms following large, severe fire events or tree mortality episodes would expose soils on steep slopes to accelerated erosion. Vulnerable locations may type-convert to shrublands if seed sources are lost and/or the local climate no longer supports forest species (Kershner 2014). ● Climate change and drought ○ Rising temperatures and an increase in climatic water deficit were linked to a doubling of background tree death rates from 1983-2004 for pines and firs in the parks montane forest (10% and 3% per year for pines and firs, respectively). The increase in mortality rate was greater for lower elevation forests (van Mantgem and Stephenson 2007). ○ Tree mortality rates further increased during the ongoing (2012-2016) warm drought. The greatest mortality has occurred the lower elevation forest band (e.g., ponderosa pine, sugar pine, incense cedar, white fir) with up to 25% drought mortality observed in USGS long-term monitoring plots. Beetles are attacking both stressed and healthy trees and generally appear to be killing the larger pines, smaller incense cedars, and firs of all size- classes. A previously rare bark beetle is attacking small diameter cedars and a previously rare weevil is impacting pines (N. Stephenson and Adrian Das, USGS, unpublished observations). In 2016 USFS reported more than 35 dead trees per acre for much of SEKI’s lower-elevation forest band that was aerial surveyed (USFS 2016). ○ For some forest types, prescribed fires conducted prior to a drought can reduce the forest’s sensitivity to drought conditions. Lower drought mortality was documented in lower-elevation mixed conifer forest for plots that had been burned compared to unburned areas (van Mantgem et al. 2016). ○ Models project that in the Sierra Nevada deciduous forest will replace conifer dominated forests at lower and middle elevations. Future increases in temperature and fire may result in higher importance of broadleaf trees (especially oak species). Oaks may shift and mix with conifers creating temperate mixed forest areas. A drier future may present an increase in grasslands and shrublands in historically mixed conifer forest areas (Jiang et al. 2013, Gonzalez 2012, Kershner 2014, Lenihan 2008). ○ Component species of Sierra Nevada montane forests have varying sensitivities to changes in precipitation, temperature, and climatic water deficit due to differences in environmental tolerances, variability in physiological traits (such as foliar gas exchange, C-34

water use efficiency, leaf water potential, needle retention, etc.), and susceptibility to fire, insect, and disease interactions (Grulke 2010, Kershner 2014). . White fir exhibited the greatest variability in key physiological traits of four conifer species (e.g., white fir, Jeffrey pine, ponderosa pine, and sugar pine) studied in the western US, including Sequoia NP. White fir also had dynamic stomatal responses to changes in moisture availability (Grulke 2010). . Jeffrey pine also exhibited significant variability in key traits, as well as high needle elongation growth at the margins of its range limits. However, the susceptibility of Jeffrey pine to drought, as detected by reduced needle elongation and an association of bark beetle attack with lower needle elongation, belies its vulnerability (Grulke 2010). . Sugar pine exhibited the least variability in physiological traits. This may be due to prior strong, selective pressure by white pine blister rust infestations experienced California-wide (Grulke 2010). . Ponderosa pine had intermediate variability in physiological traits (Grulke 2010). . Upper montane red fir forests occupy relatively cool and wet conditions (mainly in the upper montane forest), making them sensitive to increase in climatic water deficit, particularly due to reduction in winter snowpack. Red fir is sensitive to competition with lodgepole pine and especially white fir, which may be less sensitive to climatic changes than red fir. Red fir occupying riparian zones may be somewhat buffered from rising temperatures (Kershner 2014). ● Insects and disease ○ The National Insect and Disease Forest Risk Assessment predicts that 30% and 38% of the treed area of Sequoia and Kings Canyon National Parks, respectively, could be at risk for accelerated tree mortality representing 24-29% of the basal area (Table 7, Krist et al 2014). ○ Models of bark beetle population dynamics based on several climate warming scenarios document an increased risk of insect outbreaks and widespread tree mortality in western conifer forests. In other areas of the western US, drought and other stressors have triggered widespread eruptions of mountain pine beetle (Dendroctonus ponderosae)(Battles et al. 2013b). ○ Observations in 2015 and 2016 suggest beetle activity has increased during the drought (see drought section above) and appear to be most severe in lower elevation forests, including sugar pine, ponderosa pine, white fir, and incense cedar (N. Stephenson and A. Das, USGS, unpublished). ○ Observations of beetle infestations and forest die off from other western US locations suggest that stands dominated by lodgepole pine at higher elevations may be at risk too. Forests with a significant component of lodgepole pine cover more than 31,000 ha of the parks (31% of the intact forest)(Battles et al. 2013b). The National Insect and Disease Forest Risk Assessment predicts that the basal area loss of lodgepole pine from mountain pine beetle and dwarf mistletoes could be >25% in much of the Sierra Nevada by 2027 (Figure 26, Krist et al 2014). ○ Agents of red fir and white fir mortality and/or growth reductions include fir engraver beetle, dwarf mistletoe, annosus root disease, broom rust, trunk rot, and the Douglas fir-

C-35

tussock moth. Dwarf mistletoe infection and decay of red fir stands may be important in creating habitat for wildlife that den or nest in snags (Kershner 2014). ○ White pine blister rust is a non-native pathogen that infects five needle pines. Sugar pine is highly vulnerable to blister rust and beetle mortality, and this tree species has been decreasing steadily in SEKI. Blister rust is especially virulent at lower elevations and in more moist micro-sites, such as Giant Forest. A single dominant gene exists in sugar pine that confers virtual immunity to blister rust, although the gene has low frequency in natural populations. Distribution or frequency of rust-resistant sugar pine is not known in SEKI (Battles et al. 2013b). ○ Over the past 15 years, SEKI has experienced a dramatic increase in sugar pine and western white pine populations infected by white pine blister rust, as well as in the severity of these infections. An initial survey of white pine blister rust infection within SEKI conducted in the late 1990s found moderate rates of infection among sugar pine (21%), low infection among western white pine (3%), and virtually no infection among foxtail, whitebark, or limber pine. During the summer of 2013 a subset (45 of 154) of these plots was re-measured. White pine blister rust was found in 14 of 37 plots where it was previously undetected and the percentage of western white pine plots with blister rust increased from 3% to >50%. Within re-visited plots where infection had been recorded in the initial survey, the mortality rate among white pines was 66%. This suggests that there has been a substantial increase in the aerial extent of blister rust infection throughout the parks (Battles et al. 2013b; Cahill, NPS, unpublished data; Duriscoe and Duriscoe 2002). ● Non-native plants ○ Ponderosa pine was the most highly invaded forest type within SEKI with 53 non-native plant species documented in total. Survey plots in this vegetation type had an average of 1.7 non-native plant species per plot. Thirty percent (30%) of the plots were infested by non-native plants, but this is likely an overestimate due to sampling bias. Other forest vegetation types had lower infestation percentages (Sierran white fir - 11%, red fir - 11%, Sierran mixed conifer - 9%, Jeffrey pine-8%, sequoia big trees - 7%, lodgepole pine-3%) (Tu et al. 2013). ○ Montane forest types are vulnerable to non-native plant invasions, which could increase as rising tree mortality makes resources more available to non-native plants, or if forests type-convert to shrublands or grasslands. Invasive grasses could lead to a major shift in fire regime to grassland type fires (manager opinion). ○ Many climate change scenarios predict alterations or disruptions in fire regime, hydrologic changes, precipitation or temperature, or an increase in extreme events. With these changes, many invasive plants are projected to benefit from these shifting conditions and colonize and spread to new sites (Tu et al. 2013). ● Air pollution ○ Ozone exposure to forests in the Kaweah drainage exceeded levels that lead to observable foliar injury in sensitive species such as Jeffrey pine and ponderosa pine. While ozone levels do not kill these species outright, it could weaken trees and make them more susceptible to other stressors. Outside of the Kaweah, most of the montane forest was exposed to moderate levels of ozone (Battles et al. 2013b). ○ Across the parks, nitrogen deposition was generally below the threshold considered damaging to forests. Only the North Fork Kaweah watershed area and the ponderosa pine

C-36

woodland and giant sequoia vegetation types received nitrogen deposition above this threshold. Elevated nitrogen deposition may affect species composition by favoring some species over others and may increase vulnerability to non-native plant invasions (Battles et al. 2013b). ● Adaptive Capacity ○ The adaptive capacity of lower-elevation montane forest (e.g. yellow pine/mixed conifer) to accommodate changes in climate may be supported by its wide distribution across the Sierra Nevada and its diverse topographical, structural, and biological characteristics, particularly with the old-growth and late-successional features protected in SEKI. Broad distribution and wide elevational range may aid elevational shifts and facilitate access to potential climate refugia that might occur in canyons or north facing slopes, as well as favorable soil and microclimate conditions (Morelli et al. 2016). Tree species that have the highest variability of key ecophysiological traits may have the highest ability to adapt to climate change. Therefore, while significant changes in forest composition and structure are very likely by 2100, it also is likely that the parks will continue to benefit from a large area of forest providing critical functions (e.g., wildlife habitat, hydrologic regulation, nutrient cycling, carbon sequestration, and recreation)(Kershner 2014, SEKI managers’ opinion). ○ The adaptive capacity of upper-elevation montane forest to accommodate changes in climate may be limited by its fragmented distribution and restricted space to shift upslope. As snow retreats to higher elevations and topographically shaded areas, red fir forests may shift to cooler northern-aspect and uphill slopes, although shifts may be restricted if adequate soils do not align with new climate bands. White fir and lodgepole pine components of red fir system may be better able to adjust to climatic changes than red fir. Effects of fire on red fir forest are poorly understood. Red firs may have limited capacity to adapt to increased frequency of fire due to low recruitment and retarded seed production (Kershner 2014).

Confidence: Medium - Confidence in exposure to stressors, now and in the future, is high or at least moderate. More uncertain is how these stressors will interact with resource sensitivities and adaptive capacity to cause shifts in species composition, forest structure, and ecosystem function. Change in composition and structure are very likely, as is a reshuffling across the landscape, but if, how, and where these changes will contribute to alteration in forest functioning is unsure.

Subalpine forest and whitebark pine

● The Sierra Nevada Ecosystem Vulnerability Assessment (Kershner 2014) lists the overall vulnerability of alpine/subalpine systems as moderate to high due to high sensitivity to stressors, low adaptive capacity, and moderate exposure. Loss of subalpine pines, in particular, could result in substantial changes in ecosystem structure and function, including reduced soil development, increased erosion, and reduction in food sources and physical shelter for wildlife, such as the Clark’s nutcracker. Subalpine ecosystems will be affected by direct impacts of climate change, such as reduced snowpack and increased growing season length, as well as indirect effects, such as changes in beetle population dynamics and white pine blister rust abundance and distribution. Changing fire dynamics due to altered fuel availability, fire season length, and climatic conditions may also impact the subalpine forest. C-37

● The ongoing decline of whitebark pine, a foundational species in subalpine ecosystems, from interacting stressors has led the U.S. Fish and Wildlife service to determine that whitebark pine is warranted for listing as a Threatened or Endangered species under the Endangered Species Act, but that this listing is precluded by higher priority actions (Federal Register 76, no.138 [July 19, 2011]: 42631-42654). ● Climate change ○ Sensitivity – Although sensitivities vary by species and growth morphology, alpine/subalpine systems appear to be sensitive to several climate and climate-driven changes, including increased temperature, decreased water supply (e.g. precipitation and snowpack), and increased climatic water deficit. Some studies indicate that responses of high elevation forests may be largely dictated by water supply (Lloyd and Graumlich 1997; Fites‐Kaufman et al. 2007), and evidence suggests that the warming, plus higher precipitation [that has occurred since the 1930s] may have improved growing conditions for some tree species in the subalpine zone (Bouldin 1999; Dolanc et al. 2013). Although steady or increased precipitation and warming temperatures have led to less stressful conditions for recruitment and survival of small trees, these changes may also contribute to increased mortality of large subalpine trees (Dolanc et al. 2013. The paleoecological record suggests that future warming is unlikely to cause an expansion of subalpine extent if accompanied by a reduction in water supply (Lloyd and Graumlich 1997). However, understanding of limiting factors such as temperature means and extremes, and moisture availability in species establishment and survival in subalpine habitats remains poor (Graham et al. 2012). ○ Sensitivity - Many models of climate change in the Sierra Nevada predict uphill migration and a restriction of the distribution of alpine/subalpine plant communities (Hayhoe et al. 2004; Lenihan et al. 2006; Van de Ven et al. 2007). Models predict all species to shift upslope and decrease in range, with some species predicted to shift from south to north facing slopes, and experience fragmentation (Van de Ven et al. 2007), suggesting high sensitivity to climatic change. However, upslope migration of subalpine forests into the alpine will be constrained by factors such as dispersal and lack of soil. Overall understanding of future range expansion or contraction of subalpine forests is poor. ○ Exposure – High elevation forests have seen pronounced increases in temperature over the past century (Dolanc et al. 2013). Over the next century, annual temperatures in the Sierra Nevada are expected to rise between 2.4‐3.4˚C varying by season, geographic region, and elevation (Das et al. 2011; Geos Institute 2013). On average, summer temperatures are expected to rise more than winter temperatures throughout the Sierra Nevada region (Hayhoe et al. 2004; Cayan et al. 2008). Associated with the rising temperatures will be an increase in potential evaporation (Seager et al. 2007). Projections for future precipitation in the Sierra Nevada vary among models; in general, annual precipitation is projected to exhibit only modest changes by the end of the century (Hayhoe et al. 2004; Dettinger 2005; Maurer 2007; Cayan et al. 2008), with decreases in summer and fall (Geos Institute 2013). Frequency of extreme precipitation, however, is expected to increase in the Sierra Nevada between 18-55% by the end of the century (Das et al. 2011). Despite modest projected changes in overall precipitation, models of the Sierra Nevada region largely project decreasing snowpack and earlier timing of runoff (Miller et al. 2003; Dettinger et al. 2004b; Hayhoe et al. 2004; Knowles and Cayan 2004; Maurer 2007; Maurer et al. 2007; Young et al. 2009), as a consequence of early snowmelt events and a greater percentage of precipitation falling as rain rather than snow (Dettinger et al. 2004a, 2004b; Young et al. 2009; Null et al. 2010). Annual snowpack in the Sierra C-38

Nevada is projected to decrease between 64-87% by late century (Thorne et al. 2012; Flint et al. 2013), with declines of 10‐25% above 3750 m (12,303 ft) (Young et al. 2009). ○ Adaptive capacity – The adaptive capacity of alpine/subalpine systems to changing climate is largely constrained by the limited opportunity for expansion and vertical migration (Hayhoe et al. 2004), slow establishment and recovery. It may be further constrained by the fact that many of these species are poor competitors and could be out- competed by shade-tolerant species such as lodgepole pine. Adaptive capacity can be enhanced by management actions such as assisted migration, selective breeding and planting, or fuels manipulation. However, these options are logistically challenging given the remote location of much of the subalpine habitat within the parks, and existing policy restrictions related to actions allowed in wilderness. ● Mountain pine beetle ○ Sensitivity – Mountain pine beetle outbreaks (Dendroctonus ponderosae) have caused major mortality events in recent decades for subalpine tree species, such as whitebark pine and limber pine, in other parts of western North America. Rising minimum temperatures, combined with drought, contribute to bark beetle infestations and can aggravate climate-driven mortality (Millar et al. 2012). Subalpine forests in the parks may be similarly susceptible to insects and pathogens. Large outbreaks are not unprecedented in the Sierra Nevada and have recently occurred in a whitebark stands in the southern Sierra Nevada (Millar et al. 2012). Beetles preferentially target larger trees within denser stands, and therefore can have a substantial impact on reproductive potential and live biomass, resulting in cascading effects on ecosystem processes (Edburg et al. 2012). Whitebark pine populations within SEKI often grow in a Krummholz growth form, however, which appears to be less susceptible to large-scale beetle attacks. The patchy distribution of subalpine forests in SEKI compared to other impacted areas, such as the greater Yellowstone region, may also make these systems less susceptible to large-scale mortality events due to beetles. ○ Exposure – To date, the subalpine white pines in SEKI appear relatively unexposed to widespread outbreaks of mountain pine beetle compared to what has occurred in other areas of North America over the past several decades (Keane et al. 2012). Beetle activity in the subalpine forests of the Sierra Nevada has been increasing, however, and has led to isolated patches of significant mortality, especially in the June Mountain area (Millar et al. 2012). Increasing temperatures and climatic water deficit have been shown to increase beetle-cause mortality in whitebark pine (Millar et al. 2012), and exposure to beetles is likely to increase in the future. The National Insect and Disease Forest Risk Assessment predicts that the basal area loss of whitebark pine from mountain pine beetle and white pine blister rust could be >25% in much of the Sierra Nevada by 2027 (Figure 24, Krist et al 2014). ○ Adaptive Capacity – Subalpine pine systems are highly vulnerable to beetle outbreaks and have already experienced dramatic declines in portions of their range. Management actions including planting, selective thinning of competing species, the reintroduction of fire to areas that have experienced long periods of fire suppression, and the use of anti- aggregation pheromones have been suggested as actions that can enhance the resilience of subalpine ecosystems to disturbances and enhance adaptive capacity (Keane et al. 2012).

C-39

● White pine blister rust ○ Sensitivity – All white pine species, which are a key component of many subalpine ecosystems, are susceptible to Cronartium ribicola, the introduced pathogen that causes white pine blister rust. This pathogen can kill trees by girdling branches and stems (Hoff 1992) and has reduced stand density by >90% in some areas of its range (Keane et al. 2012). Some level of genetic resistance has been identified in all subalpine species of white pine, but in general, resistance among white pines is low (Schoettle and Sniezko 2007). ○ Extent – White pine blister rust remains rare in subalpine forests within the southern Sierra Nevada (Maloney 2011). During the 1990s, an extensive survey of white pine species within SEKI was conducted, and no blister rust infections were found in limber, whitebark, or foxtail pine monitoring plots, and infection among western white pine was rare (Duriscoe and Duriscoe 2002). However, the extent of this pathogen appears to be increasing. During a recent re-survey of 45 blister rust monitoring plots within western white pine stands, incidence of blister rust had almost doubled in ~15 years (Cahill 2014). In addition, blister rust on whitebark pine within long-term monitoring plots has been discovered for the first time in both Yosemite and SEKI in the last several years. It is likely that white pine blister rust will continue to spread into subalpine forests within SEKI over the next several decades, leading to increased mortality and reduced cone production. ○ Adaptive Capacity – Adaptive capacity of white pines to blister rust is low to moderate, as there is likely some genetic resistance to the pathogen within existing populations, though individuals that display resistance may be relatively rare. Management potential related to white pine blister rust is lower for subalpine than montane forest due to their remote locations, which are almost entirely within designated wilderness. In addition, whitebark pine is dependent on Clark's Nutcracker for seed dispersal, and if populations decline substantially, the birds will not continue to visit the stands (McKinney et al. 2009). Most conservation and restoration efforts to enhance adaptive capacity to blister rust are focused on breeding and planting rust-resistance seedlings (Keane et al. 2012). Other restoration strategies include using fire and mechanical thinning to remove competition from other tree species and enhance natural regeneration (Tomback and Auchuff 2010). ● Altered fire regimes ○ Sensitivity – Historic fire return intervals for subalpine forests in the sierra Nevada are highly variable, but are generally considered to be on the scale of hundreds of years (Caprio and Lineback 2002; van Wagtendonk and Fites-Kaufman 2006). Because of the long return interval, the effects of anthropogenic fire suppression that have caused substantial changes in the lower montane forest, have had relatively minor impacts on subalpine systems (Miller and Safford 2008). However, changes in fire frequency may have particularly acute effects in ecosystems that are not adapted to frequent fire, such as subalpine forests (Schwartz et al. 2015). Fire severity in these systems may also increase due to warming temperatures that reduce fuel moisture (Fried et al. 2008). Increased fire severity may be particularly acute following periods of drought as trees are more likely to die from fire injury (van Mantgem et al. 2013) ○ Extent – While the current impact of altered fire regimes in subalpine systems in the Sierra Nevada is low, increased fire frequency and extent of area burned is predicted. Indeed, fires in subalpine forests have already increased in frequency and extent over the past 100 years (Schwartz et al. 2015). In fact, the subalpine forest type exhibited the largest proportional increase in area burned of any forest type in the Sierra Nevada (Mallek et al. C-40

2013). This may be due to increased fuels and tree density that is hypothesized to have resulted from warming temperatures (Dolanc et al. 2013), changes in fire management policy that has allowed more fires to burn in high elevation forests (Schwartz et al. 2015), or other factors such as increased ignitions ○ Adaptive Capacity – It is unclear how subalpine forests will respond to altered fire regimes. Little is known about the resilience of these systems to more frequent fire and associated effects on the persistence of species adapted to longer fire intervals. One consequence of changing fire regimes could be that it will allow increased migration of other upper montane species into the subalpine, increasing fuels and competition, though there is no evidence that this has occurred to date (Dolanc et al. 2013). Confidence - The combination of climate change, beetles, and blister rust seem likely to reduce the extent of subalpine trees species, as has occurred in other regions of North America. Evidence for this prediction is still developing in the southern Sierra Nevada.

Stressor Sensitivity Exposure Adaptive Capacity Vulnerability

Climatic Moderate High–increasing Low-Moderate Moderate-High Change (low confidence) (high confidence) (moderate (moderate confidence) confidence)

Bark Moderate-High Low –increasing Low-Moderate Moderate Beetles (moderate (high confidence) (moderate (moderate confidence) confidence) confidence)

White Moderate- High Low – increasing Moderate Moderate-High Pine (high confidence) (moderate-high (low confidence) (moderate Blister confidence) confidence) Rust

Altered Low-Moderate Low –increasing Unknown Low-Moderate Fire (low confidence) (moderate (low confidence) Regimes confidence)

C-41

ALPINE TERRESTRIAL ECOSYSTEM

Key vulnerability: Shifts in alpine species composition and loss of alpine habitat due to the interaction of warming, changes in snowpack dynamics, nitrogen deposition, and increased spread of non-native plants.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Alpine Nitrogen - Shifts in Drying of Alpine is Introduction Increased Increased New Warming X Terrestrial shifts in plant habitat, isolated of non- or new or new diseases? altered Ecosystem plant species, loss of “sky island” native invasion invasion snowpack species range insulation plants; soil X nitrogen contraction compaction, X invasions erosion, social trails

Alpine Terrestrial Ecosystems

● The Sierra Nevada Ecosystem Vulnerability Assessment lists the overall vulnerability of alpine/subalpine systems as moderate to high due to high sensitivity to stressors, moderate exposure, and low adaptive capacity (Kershner 2014). Interaction of stressors such as climate change, atmospheric nitrogen deposition, and species invasion make alpine environments some of the most vulnerable in the region. Although the alpine environments of SEKI have high diversity, many taxa are tightly constrained by extreme environmental conditions and these species have limited migration potential if climate change makes habitat unsuitable (Haultain 2013). ● Sensitivity ○ Dominated by slow-growing perennial plants which are adapted to the extreme climatic conditions that characterize high elevations, alpine vegetation is thought to be particularly sensitive to shifts in temperature and snowpack dynamics predicted under anticipated climate change scenarios. The snowpack is an insulating layer that protects both animals and plants from extreme cold and drying, and prevents deep-freezing of water bodies (Haultain 2013). ○ Hypothesized (and in some cases realized) impacts of climatic warming include an expansion of subalpine species into alpine environments, an increase in native species that colonize disturbed environments (ruderal species), expansion of the ranges of woody taxa (both shrubs and trees), and increased potential for non-native species to establish in these previously un-invaded environments (Haultain 2013). ○ Alpine plants evolved under low-nitrogen conditions and are particularly at risk from the fertilizing effects of atmospheric nitrogen deposition, which favor some plants (such as some invasive and non-native species) over others (such as some wildflowers). The interaction of nitrogen deposition with climatic warming was modeled for the Rocky Mountains, and simulated changes in plant biodiversity from warming X nitrogen deposition were more severe as nitrogen inputs rose. In particular, nitrogen deposition facilitated the invasion of tree species at deposition of 5.5 kg N/ha/yr and above (Porter et al. 2011).

C-42

○ A study of treeline dynamics over the past 3,500 in the southern Sierra Nevada suggest that warming is unlikely to cause an expansion of subalpine forests if it is accompanied by a reduction in available moisture, changes in treeline are usually more complex than simple upward or downward movements, and that changes in treeline position may lag climate changes by decades to centuries (Lloyd and Graumlich 1997). ○ Recent experiments in the alpine have shown high capacity for small mammals to eat conifer seedlings before they can grow large enough to withstand herbivory (R. Klinger, USGS, unpublished data). ○ If woody species from lower elevations are able to expand their ranges into the alpine, this could result in feedbacks that further alter the structure of alpine communities through mechanisms such as changes in soil properties or trees or shrubs trapping snow. ○ Fire has played a limited role in the parks’ alpine environment due to low fuel loads. How future changes in climate may impact fire regimes in the alpine is uncertain due to the interaction of ignitions, fuel loads, and fuel moisture (Haultain 2013). ○ Historic livestock grazing, particularly by sheep, is assumed to have shaped the terrestrial alpine communities we see today. Widespread overgrazing and its impacts in the Sierra Nevada, including severe erosion, were well-documented during the early and mid-1900s (Haultain 2013). This legacy of livestock grazing may affect the sensitivity of these communities to other stressors. ○ Some species, such as meadow invertebrates, may be sensitive to trails that fragment the alpine, including unmaintained social trails (Haultain 2013). Alpine soils and vegetation also may be sensitive to compaction and erosion. ● Exposure ○ Temperatures in the parks have been rising since about 1950 and there is some evidence that warming has been greater at higher elevations, at least in the springtime ((Redmond and Fearon 2013). Recent trends indicate increasing snowpack above about 8,500 ft. (2,600 m) ft in the southern Sierra Nevada (Andrews 2012), but some models project that alpine areas could see decreased snowpack of -10% to -25% by the end of the century. Loss of snowpack could lead to diminished water availability during the growing season ( Kershner 2014). In contrast, if precipitation increases, snowpack in alpine areas could remain similar or increase in the future, even with rising temperatures. An increase in summer monsoonal activity could also mitigate a loss of moisture delivered by the melting snowpack. ○ Introduction of non-native plants (via seeds carried on gear and clothing) or disease organisms may increase due to increased visitation in the wilderness at the same time that climate change may make conditions favorable for establishment of new species. Introduction of non-native fish have had cascading effects on terrestrial species (such as mountain garter snakes and gray crowned rosy finches) that rely on food subsidies from aquatic food webs. Currently, two non-native birds (willow ptarmigan and chukar) occur in the alpine, but little is known about their ecological effects. No non-native mammals, amphibians, or reptiles have been reported in the parks (Haultain 2013). ○ The parks’ alpine environments receive a stream of nutrients and contaminants deposited from atmospheric pollution. Nitrogen deposition at high elevations in SEKI is on average 3- 4 kg N/ha/yr and up to about 6-7 kg N/ha/yr (Sickman et al. 2001). These rates are above the critical load of 1-2 kg N/ha/yr that is hypothesized to protect terrestrial alpine

C-43

communities from a 5% change in plant species coverage based on a model parameterized for the Rocky Mountains (Porter et al. 2011). Little is known about the effects of these pollutants on alpine terrestrial environments in the Sierra Nevada. ○ Shifts in the elevational ranges of small mammals have been documented for some alpine species over about the last 100 years and are likely linked to climate change. When shifts occurred, high-elevation species typically contracted their lower limits upslope, whereas lower elevations species had variable responses. For the high-elevation species, site- specific change in temperature was a better predictor the direction of shifts than precipitation (while for lower-elevation species, temperature or precipitation did not predict direction of range shifts). A contracted lower range has been documented for the following species: mountain vole (Microtus montanus), Lodgepole chipmunk (Tamias speciosus), American water shrew (Sorex palustris), Belding's ground squirrel (Urocitellus beldingi) and alpine chipmunk (Tamias alpinus) (Rowe et. al. 2014). ● Adaptive capacity ○ Adaptive capacity of the alpine is limited because alpine habitats are essentially “sky islands” of habitat. Alpine taxa that are displaced by subalpine species have nowhere else to go (Haultain 2013). ○ Some amount of adaptive capacity may be conferred by the alpine’s topographic complexity, which creates many microclimate areas that could ameliorate climate change impacts. One example is the cooler environments created by rock glaciers, which might be considered ‘micro-refugia’. Pika in particular may benefit from climate change refugia in a warming world (Morelli et al. 2016).

Confidence: Low - Confidence is high that alpine will warm, but combined effects of temperature and precipitation on moisture availability and snowpack dynamics may be more important for some responses. There is moderate to high confidence that alpine species are sensitive and that interactions among stressors are likely to amplify changes to alpine structure and function. Projections from different climate-vegetation models and results of observational studies tend to be contradictory (or species and site specific), with some projecting up-slope migrations. Recent work suggests that the microtopography could lead to the protection of small-scale refugia for alpine species. The alpine is likely to change significantly but confidence in specifics is low.

C-44

TERRESTRIAL WILDLIFE

Key vulnerability: varies by species

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Sierra Nitrogen Changes in Fire Changes in Small Changes in Disease Small Nevada deposition - habitat and suppres- habitat and population, habitat and from populations bighorn changes in food sion food limited food domestic X predation sheep habitat and resources reduces resources distribution, resources sheep, new X disease food open low genetic disease? X loss resources habitat diversity habitat /food

American Nitrogen Heat Extreme Loss of Isolated Vegetation Disease? Warming X pika deposition - stress, heat winter high change loss of changes in habitat loss insulation, elevation snowpack habitat and vegetation habitat X loss of food change habitat resources

Pacific Loss of Habitat loss Loss of Small, Illegal use - Small fisher habitat - large habitat, but isolated poisoning isolated severe fires increased populations populations mobility? , low X fire X genetic poison X diversity habitat loss

Birds Bioaccumul Habitat Too much Habitat Local, Collisions Non-native West nile Varies by ation, loss, range or too little loss, range regional, with cars, birds virus, etc. species respiratory shifts, fire shifts, and global recreational rates phenology phenology impacts impacts mismatch mismatch

American Loss of Develop- Increased Loss of More black bear acorns, ment conflict and sugar pine visitors X sugar pine outside food (beetles/bli less natural seeds, and parks rewards ster rust), food other fall wildlife sources foods disease?

Sierra Nevada Bighorn Sheep

● The Sierra Nevada Bighorn Sheep (Ovis canadensis sierrae) was listed as a federally endangered species in 2000. The prehistoric population likely exceeded 1,000 individuals but declined to 125 animals in 1999 due to competition and diseases from domestic livestock, unregulated hunting, predation, and changes in habitat. Since its listing under the Endangered Species Act and following active reintroduction efforts, the population has grown to over 600 in 2014. Of the 16 areas in Sierra Nevada identified as suitable habitat, 13 are currently occupied. A new herd unit, not identified in the Recovery Plan, Cathedral Range (in Yosemite), is also now occupied (CDFW 2016). ● The Sierra Nevada Ecosystem Vulnerability Assessment rates the overall vulnerability of Sierra Nevada bighorn sheep as moderate-high, due to their moderate-high sensitivity to climate and non-climate stressors, low-moderate adaptive capacity, and high exposure (Kershner 2014).

C-45

● According to the parks’ 2013 Natural Resource Condition Assessment (NRCA), Sierra Nevada bighorn sheep were rated in relative worse condition in three park watersheds, intermediate in one park watershed, and better in two park watersheds compared to historic conditions (NPS 2013). ● Sierra Nevada bighorn sheep use habitats ranging from the highest elevations along Sierra Nevada crest to winter ranges at the eastern base of the range as low as 4,760 ft. (1,450 m). These habitats range from alpine to Great Basin sagebrush scrub. Primary elements of preferred habitats are visual openness and close proximity to steep rocky terrain used to escape from predators. Of particular importance is the nutrient content of forage. Nutrient quality of diets varies greatly with season and elevation and is limited primarily by effects of temperature and soil moisture on plant growth and density. Low-elevation winter ranges provide an important source of high quality forage early in the growing season (USFWS 2007). ● Factors limiting Sierra Nevada bighorn sheep recovery include disease, predation, low population numbers and limited distribution, availability of open habitat, and potential further loss of genetic diversity due to small population sizes and inadequate migration between them. Since the vast majority of Sierra Nevada bighorn sheep habitat is publicly-owned land, loss of habitat has not been a limiting factor. However, management in bighorn sheep habitat (e.g., fire suppression) can result in habitat alterations and loss of key dispersal corridors connecting herds, which could be limiting factors (USFWS 2007). ● Climate change models predict changes in temperature and precipitation in the range of Sierra Nevada bighorn sheep. In particular, these include milder winters, less precipitation falling as snow over a shorter winter period, and possibly increased total annual precipitation. These changes could potentially cause earlier seasonal drying of high-elevation meadows, conifer trees invading higher elevations, and changes in summer- and winter-range quality and accessibility in the Sierra Nevada. Additional factors may include the interaction of invasive alien plant species with climate change and increased atmospheric nitrogen fertilization from air pollution. These effects would alter the distribution of high-quality bighorn summer and winter habitats, which would influence recovery (USFWS 2007). ● Predation by mountain lions and interaction with domestic sheep are thought to be low within SEKI, but may impact Sierra Nevada Bighorn Sheep when they migrate outside park boundaries to their winter range. ● Despite their fragmentation, bighorn sheep have the ability to disperse long distances provided suitable habitat remains connected. Although they are dependent on alpine areas near escape terrain during portions of the year, they also occupy a range of habitats, and have considerable plasticity in relation to other subspecies of bighorn sheep. Utilization of a range of systems, from alpine to pinyon-juniper and sagebrush steppe at lower elevations during the seasonal upslope and downslope migration may indicate a level of plasticity that will aid them in adapting to climate changes (USFWS 2007).

Confidence: Medium - Confidence in species information and herd unit occupation is high due to 30-40 years of active management and research. Confidence in projected changes to habitat is lower.

C-46

American pika

● The current condition and future predictions for American pika (Ochotona princeps) in the Sierra Nevada is a subject of disagreement and ongoing research. ● Pikas are sensitive to heat stroke and death if exposed to temperatures exceeding 78 °F (25.6 °C) (Smith 1974). Pikas stay active all winter long traveling in tunnels under the rocks and snow. If the snow is deep enough it provides insulation for the pikas living below. Too little snow and the pika risk freezing to death. ● Millar and Westfall (2010a) found that pikas in the Sierra Nevada and southwestern Great Basin appear to be thriving and tolerating a wide range of thermal environments. Wolf (2010) challenged these conclusions and Millar and Westfall (2010b) defended them. ● The species was petitioned for listing under the Endangered Species Act in 2007, largely because of potential climate change impacts. The US Fish and Wildlife Service (2010) concluded that although the American pika could potentially be impacted by climate change, the species as a whole will have enough high elevation habitats to ensure its long-term survival despite higher temperatures in a majority of its range. ● Jeffress et al. (2013) suggested that “idiosyncrasies of place” (e.g., unique combinations of environment, biota, and history) may result in weak relationships between climate and pika distribution. ● The model developed by Stewart et al. (2015) projected that, depending on the climate scenario considered, by 2070 pikas will be extirpated from 39% to 88% of the 67 historically occupied sites studied in California. ● While it is commonly thought that climate change will force upslope distribution shifts for many alpine species, at least one study suggested that counter to species distribution model predictions, seemingly unsuitable low elevation sites may be a better refugia for pika under climate change (Varner and Dearing 2014). ● The relatively large and intact nature of SEKI’s high-elevations suggests that SEKI could be an important stronghold for American Pika. Confidence: Low - There is contradictory information regarding the future of pika in the Sierra Nevada, but relative to other locations in the western U.S. it seems reasonable that SEKI could continue to provide suitable habitat for the species longer than many other places.

Pacific fisher

● The West Coast Distinct Population Segment (DPS) of the Pacific fisher was proposed as federally threatened in 2014 based on potential threats to its habitat from wildfire, timber harvest practices, and indiscriminate and illegal use of pesticides. These threats were subsequently found to be not as significant as previously thought, and in 2016 the U.S. Fish and Wildlife Service concluded that it does not face the risk of extinction now or in the foreseeable future and therefore does not require the protection of the Endangered Species Act (USFWS 2016). ● The Sierra Nevada Ecosystem Vulnerability Assessment lists the overall vulnerability of Pacific fisher as moderate due to moderate-high sensitivity to non-climate and climate change stressors, moderate exposure, and moderate adaptive capacity (Kershner 2014). ● As a habitat specialist, the Pacific fisher’s sensitivity to climate change likely will be driven by changes that remove resting and denning habitat in late-successional mixed forests; these C-47

changes include increased climatic water deficit and altered wildfire regimes. However, predicted reductions in future snowpack could provide competitive benefit to fishers, whose movement ability, in relation to martens, is currently limited by deep snow (Kershner 2014). ● Predicted shifts of conifer dominated forests to mixed woodland and hardwood forest types may benefit fishers; however, the expectation is for an overall decrease in the availability of fisher habitat as changes in fire regime and loss of late seral habitat, and decreases in the density of large conifer and hardwood trees and canopy cover (Kershner 2014) ● Pacific fisher is sensitive to several non-climate stressors including: predation pressure, habitat conversion (timber harvest and exurban development), disease, road mortality, and poisoning. These non-climate stressors can cause direct mortality, reduced fitness, and amplify the effects of climate-driven changes. For example, habitat conversion reduces canopy cover, a habitat element that will likely be further reduced by wildfire regimes in the future (Kershner 2014). ● Mortality and reduced fitness may be a consequence of rodenticides and other pesticides used at marijuana grow-sites. Direct and secondary exposure to rodenticides and pesticides represent a significant risk to isolated California populations, and appear to cause both mortality and decreased fitness by increasing fisher susceptibility to hypothermia, parasites, pathogens and predation. Poisoning risk has the potential to shift a population from positive to negative growth rate (Kershner 2014). ● The capacity of the Pacific fisher to adapt to future changes in climate is limited by its small population size, low reproductive rate, geographic and genetic isolation, and low genetic diversity (Kershner 2014). ● In California, fishers occur in two isolated populations. The southern Sierra Nevada population is completely isolated and has an estimated 100 to 400 individuals. Pacific fisher populations are considered at high risk of local extinction from stochastic events such as disease or wildfire (Kershner 2014). ● Current condition information for Pacific fisher in the parks is lacking. The only research conducted on fishers in these parks is a carnivore survey in 2002-2004 which found that fishers occupy the same habitat associations in which they would be expected (NPS 2013). ● While we don’t have information on population dynamics for fisher in SEKI, there is information for the California spotted owl, a state species of conservation concern under petition for listing under the Endangered Species Act. Like the fisher, this owl relies on old growth or otherwise high biomass forest habitat for nesting and roosting. SEKI’s California spotted owl population is stable or increasing while populations in Lassen and Sierra National Forests are more likely in decline (Conner et al. 2013). Confidence: Low – Little is known about the current status of the fisher population (increasing, decreasing, or stable), but we think that the fisher in SEKI are doing better relative to the surrounding national forests because we have more intact habitat (less altered by past human activities like extractive logging or road building). The assumption is that they like park habitats more—but they might prefer the heterogeneity of USFS habitats, because for example, there may be more prey density.

Birds

● The parks’ low-lying southwestern region has the highest bird diversity, associated with montane hardwoods, montane riparian habitats, and water (Schwartz et al. 2013).

C-48

● Conservation Status: Of 145 bird species evaluated for the NPS Sierra Nevada Network parks (including SEKI), three species are listed as threatened or endangered at the state level (bald eagle, great gray owl, willow flycatcher) and 16 are considered Species of Special Concern (state), Birds of Conservation Concern (federal), or imperiled (NatureServe rank) (harlequin duck, northern harrier, northern goshawk, peregrine falcon, flammulated owl, California spotted owl, long-eared owl, black swift, Vaux’s swift, calliope hummingbird, Lewis’s woodpecker, Williamson’s sapsucker, olive-sided flycatcher, American pipit, yellow warbler, and Cassin’s finch) (Steel et al. 2013). SEKI’s California spotted owl population is stable or increasing while populations in Lassen and Sierra National Forests are more likely in decline (Conner et al. 2013). ● Climate change: Of 145 species, 18 seem most likely to benefit from climate change, while a warmer climate seems likely detrimental to 77 species. Progressively earlier springs are associated with changes in the timing of cyclical events, such as earlier migration, earlier breeding, and changes in clutch size. Some Sierra Nevada birds have already adjusted their breeding ranges to compensate for shifting climate conditions. More than half moved northward. Bird species currently found in lower elevations of SEKI may shift upwards. Species currently limited to alpine areas (including horned lark, American pipit, and gray-crowned rosy finch) may lose most or all suitable habitat and perhaps cease to occur in the parks. Shifts northward to colder latitudes may bring new species into the parks from the south. Other species at least moderately likely to suffer major declines due to climate change include Vaux’s swift, Calliope hummingbird, black-backed woodpecker, Clark’s nutcracker, MacGillivray’s warbler, pine grosbeak, and Cassin's’ finch (Steel et al. 2013). ● Fragmentation and human use: Of 145 species, 19 may benefit from fragmentation and human alteration of the landscape, while 100 species are likely negatively affected, including Harlequin duck, sooty grouse, Northern Harrier, red-shouldered hawk, California spotted owl, Vaux’s swift, some woodpeckers (hairy, black-backed, pileated), some flycatchers (olive-sided, Hammond’s, willow), etc. Exurban and agricultural development within the foothills may affect short-distance migrants, such as owls. Forest fragmentation in the region may affect species with large home ranges like many raptors. Habitat degradation of wintering grounds or migratory routes of neo- tropical migrants is another threat (Steel et al. 2013). ● Fire regimes: Habitat alteration can occur due to altered fire regimes, which change the structure of forests. Fire suppression is detrimental to species, such as the northern goshawk, adapted to forests with varied age structure, but may benefit species associated with dense, late-successional habitat (e.g., California spotted owl and Hammond’s flycatcher). More frequent and higher-severity fire would benefit several species, such as hairy and black-backed woodpeckers and lazuli bunting, but harm others like Hammond’s flycatcher and golden- crowned kinglet. With increased fire, 67 of 145 species would generally benefit, but 38 species are typically fire-adverse (Steel et al. 2013). ● Disease: West Nile virus is transmitted by mosquitoes that can be contracted by birds and humans. It arrived in California in 2003. It is especially virulent in the crow family and birds of prey. Two of the four most affected species in the state, the western scrub-jay and the house finch, occur in the parks. Avian influenza (bird flu) is highly virulent to humans and birds but has not been reported in North America (Steel et al. 2013). ● Non-native species: Non-native bird species compete with native species for resources and nesting sites. The brown-headed cowbird is a non-native species with expanding range; however it’s not thought to be a major stressor in SEKI. The non-native barred owl has recently been detected in SEKI, and is a threat to California spotted owls. The raven, a native species,

C-49

has expanded its range in the Sierra Nevada including Kings Canyon NP. (Steel et al. 2013). Introduced trout are detrimental to some species, such as harlequin duck and gray-crowned rosy finch (Epanchin et. al. 2010), that rely on aquatic subsidies. ● Together, invasive species and disease are likely to negatively impact 87 species (4 with major negative impact: Anna’s hummingbird, willow flycatcher, western scrub jay, Clark's nutcracker) and four species may slightly benefit (Steel et al. 2013). ● Other impacts that generally have small effects on populations include collisions with cars, structures or wind turbines; disturbance from aircraft; predation from feral pets; electrocutions; and recreational impacts. Of 145 species, 101 are likely to be negatively affected (12 with major negative impact) by human use activities; twenty are likely to be positively affected (5 with major benefit)(Steel et al. 2013). ● Air pollution may affect birds with high respiratory rates. Fish-eating species, such as the belted kingfisher, are likely to ingest heavy metals and pesticides, which bio-accumulate in the food they eat. Some bird species may be indirectly threatened by ozone injury to ponderosa and Jeffrey pines if it degrades bird habitat or food sources (Steel et al. 2013). Confidence: Low to high - There is high confidence that bird species will shift somewhat as some species benefit and others decline due to changing conditions. Which specific species will suffer or benefit is more uncertain. The NRCA Bird Assessment (Appendix 17)(Steel et al. 2013) suggests impact amplitude and confidence ratings for stressor impact on individual species.

American black bear

● Black bears are generalist omnivores, are intelligent, and can travel long distances. It is likely that at the population level this species will persist in the face of habitat fragmentation and human land use, climate change, altered fire regimes, and other stressors (manager’s opinion). ● Sugar pine seeds and oak acorns are both important fall food sources for black bear (and a variety of other wildlife species) in the Sierra Nevada, but both have variable mast production and are in decline. In the fall, black bears make heavy use of both sugar pines and oaks. Although they prefer acorns to sugar pine seeds, the loss of either food source would lead to an increased dependence on the other. When both are unavailable due to continued decline or simultaneously low mast crops, an increase in human-bear incidents is likely (Mazur et al. 2013). ● Observations of some bears in poor (starved) condition and increased attempts of bears at obtaining human food in 2015 after four years of warm drought suggest that individual animals will suffer as climate warms and dries, probably because of food shortages. Therefore, the bear population in SEKI may shrink somewhat if years with low food availability are not buffered by periodic food abundances. This dynamic is suggested by reduced human-bear incidents in 2016 despite the continuation of drought conditions. Yosemite NP has not observed the same effects of drought as we have in SEKI, however, which suggests that other factors are important to explain the 2015 observations in SEKI (managers observations and opinions). ● If climate change motivates bears to move more in search of food and habitat, more collisions with motor vehicles may result (manager opinion). Confidence: Medium - Confidence in knowledge of bear habits is high due to decades of management and research. However, we still do not understand the mechanisms behind rises and falls in human-bear conflict incidents. C-50

LANDSCAPE INTEGRITY AND BIODIVERSITY

Key vulnerability: Climate change, drought, changing disturbance regimes, and invasive species interactions combined with barriers for species to adapt to these changes.

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals and Stressor weather land use disease Interactions

Landscape Species Biome Extreme Species Biome Natural and More Shifts in Shifts in Forest Climate integrity shifts, shifts, large, extirpation, shifts, human- develop- ecosystem ecosystem and change X and favors non- species crown/high flooding or species caused ment structure or structure or woodland drought X biodiversity natives extirpation severity erosion of extirpation barriers to pressure function, function, die-off landscape fires; sensitive dispersal species species fragmentation accelerate locations and extirpations extirpations that limits biome migration adaptive shifts, capacity species extirpation

Landscape integrity and biodiversity

● Fragmentation in the parks could increase with widening or extension of existing roads or trails in response to increased visitation. Widening may occur unintentionally due to increased traffic and lack of parking. ● Fragmentation is increasing in the ecoregion due to development (Thorne et. al. 2013, Hansen et al. 2014), threatening critical connectivity corridors for species migration. Large fires, floods, and widespread tree mortality caused by insect outbreaks and hot drought contribute to larger gaps (Battles et. al. 2013b), reducing connectivity for species such as Pacific fisher (Kershner 2014). See Forest and Wildlife vulnerability assessments for more specific information. ● Biodiversity is threatened by altered fire regimes; warming and drying climate; floods, erosion, and mass wasting; invasion of non-native species (Tu et. al. 2013), and by interactions among these stressors. ● Biome shifts are likely as conifer species are lost from their lower elevation limits due to drought and fire, and are replaced by oaks and chaparral; as shrubs are lost in the foothills and replaced by herbaceous communities dominated by non-native annual plants; as pines expand into subalpine areas, replacing firs; and the alpine climate becomes more suitable for subalpine trees. One-fifth to four-fifths of SEKI may be vulnerable to biome shift by 2100 based on MC1 dynamic model results (Gonzalez 2012). A time lag between changes in climate and changes in vegetation is likely, however, and not included in the MC1 model projections, making vegetation projections uncertain (See Appendix B). ● There is inherent adaptive capacity conferred by landscape diversity in elevation and topography that introduce variability in temperature, precipitation, humidity and insolation and may result in climate change refugia that are buffered to change compared to the surrounding landscape (Morelli et al. 2016). ● Management capacity is good in some areas where we are able to apply prescribed fire to influence forest structure, and in areas where intact ecosystems can be managed by restricting use. Management can also decrease fragmentation or increase connectivity by removing infrastructure from key areas and through ecological restoration. Management interventions are C-51

limited in spatial extent due to terrain, budget, and by law and policy; and management capacity is low for combatting climate change effects. Confidence: Medium - Low confidence in assessment of fragmentation vulnerability; high confidence that changes, including connectivity and biome shifts will occur on moderate to large spatial scales; low confidence in predicting where changes will occur; moderate confidence that we can detect and measure biome shifts.

C-52

CULTURAL RESOURCES

Key vulnerabilities: Fire, flood, erosion, vandalism, lack of information, and limited preservation effort

Air Warming/ Altered Fire More Loss of Fragment- Increased Non-native Non-native Insects and Multiple Pollution Drying Regime extreme snowpack ation and visitation plants animals disease Stressor weather land use Interactions

Historic Increased Fires burn Increased Faster More New or Lack of buildings and freeze/thaw, structures, freeze/ decay (non- vandalism, increased document- structures faster smoke thaw, wind, freezing erosion destructive ation X deterioration damage, flooding, winters) organisms? stressors erosion erosion Less structural damage

Cultural Shifts in Shifts in Flooding, Shifts in Washout of More Threat to New or Lack of landscapes resource resource erosion resource roads or vandalism, culturally increased document- setting, setting setting trails erosion significant destructive ation X phenology native organisms? stressors species

Archaeology Changes in Heat Flooding, Melting Develop- More Lack of vegetation damage, erosion glaciers ment may looting, document- can expose erosion, and damage trampling, ation X or hide vegetation snowfields artifacts erosion stressors artifacts shifts may expose artifacts

Ethnographic Loss of Loss of Flooding, Develop- More non- Threat to Lack of resources culturally culturally erosion ment traditional culturally document- significant significant obliterates use of significant ation X plants and plants and cultural cultural native stressors animals animals significance sites (e.g. species river recreation)

Museum Increased Fire/smoke Heat, Manage- collections difficulty and damage or flooding, ment and archives expense to destruction humidity attention maintain (funding) X constant stressors temperature

 For cultural resources, limited documentation of the existence or significance of resources is a major vulnerability. Approximately 2.3% of the parks have been surveyed for cultural resources (19,859 of 865,964 acres). Fifty-five percent (55%) of the 629 archaeological sites listed in ASMIS are in good condition. Less than two percent (2%) of sites are known to have adequate National Register documentation. In ASMIS, two sites are listed and seven are determined eligible for the National Register. Most (95%) of sites have not been evaluated for the National Register. Only about 4% of sites have been dated within a specific time frame. Many sites do not yield diagnostic material.

 The vulnerability assessment presented below was developed from a generic assessment of potential impacts to cultural resources prepared by the NPS Climate Change Response Program (Morgan et al. 2016) and manager opinion. C-53

Historic Buildings and Structures

● Historic buildings and structures are subject to deterioration over time. This deterioration can be hastened, both directly and indirectly by changing climate conditions. The greatest threats are likely to be increased uncontrolled fires, flooding, increased erosion, and new or increased destructive organisms. ● Fire may damage or destroy historic structures. Other impacts of fire and fire management might include discoloration caused by smoke or heat; damage from tree-fall; damage due to fire retardants; damage due to post-fire erosion. Areas with projected high fire vulnerability that coincide with historic structure occurrence will be particularly vulnerable. ● Depending on location, flooding may damage or destroy historic structures. ● Erosion may cause undermining of structures and/or loss of historic setting. ● Insects and other destructive organisms may damage or destroy historic structures at an accelerated pace. ● Wind might become more of an issue with loss of vegetative cover due to increased scouring by blown particles. Warming could expose some structures to longer periods of wetness due to warmer (not frozen) winters. ● Increased freeze/thaw cycles could cause surface cracking, flaking, and sugaring of building stone and spalling of brick; and damage to foundations. ● Extreme heat will increase the natural deterioration of wood and adobe. ● Drier conditions could deteriorate masonry and porous stone (due to salt deposits) and increase cracking and splitting of wood/organic features. ● Wetter conditions could cause swelling/distortion of wood materials; increased risk of rot and fungal/insect attack; erosion of supporting ground around structures; sewage backup and overflow damage; accelerated decay of masonry due to increased extremes of wetting and drying; cracks and destabilization caused by ground heave and subsidence and shrink/swell of soils. ● Increased visitor use (correlated with warming temperature) could lead to more vandalism of structures, and greater erosion that destabilizes structures. ● Species changes may include increased (or in some instances decreased) growth of destructive organisms (termites, bacteria, mold, algae, fungi); increased or decreased use of structures as wildlife habitat; and change in the historic landscape context due to vegetation shift. ● Adaptive reuse of historic buildings will require efforts to maintain historic fabric. Lack of complete inventory, evaluation, and documentation does not directly cause deterioration, however the lack of complete baseline information hinders efforts to discover and evaluate impacts.

Confidence: Medium - We are fairly confident in general vulnerabilities, but not in the direction or magnitude of changes or the impacts of those changes. While the correlation of visitation and increasing temperature is positive and strong, possible impacts (looting, vandalism) from increased visitation are purely speculative. Positive impacts such as reaching more and new populations are just as likely, if not more likely assuming that the parks can increase outreach to match visitation growth.

C-54

Historic Districts & Cultural Landscapes

● Threats to historic districts and cultural landscapes are similar to the threats to historic structures and buildings, with the addition of other elements that contribute to the integrity of a site. ● Changes in fire regime, increased erosion, vegetation shifts, loss of culturally significant species, and phenological/seasonal shifts would be particularly significant in altering the historic or cultural context of the area. In addition, wash out or damage to roads, trails, and landscape features may result from increased flooding. ● Areas with projected high fire vulnerability that coincide with cultural landscape occurrence will be particularly vulnerable. Confidence: Medium - We are fairly confident in general vulnerabilities, but not in the direction or magnitude of changes or the impacts of those changes. While the correlation of visitation and increasing temperature is positive and strong, possible impacts (looting, vandalism) from increased visitation are purely speculative. Positive impacts such as reaching more and new populations are just as likely, if not more likely assuming that the parks can increase outreach to match visitation growth.

Archaeology

● Archaeological resources are subject to deterioration over time, albeit at mostly longer time spans than for historic structures. This deterioration can be hastened by visitor impacts, uncontrolled fires, erosion, and climate change. ● Increased visitation may result in more extensive maintenance of current trails and facilities. Additional foot traffic may result in deterioration of currently known archaeological resources adjacent to existing trails and facilities. Increases in both vandalism and looting may also result from increased visitation. The creation of new trails and facilities could impact buried archaeological resources. ● Climate change effects (direct or indirect) likely to be most significant are increased fire activity, increased erosion, increased visitation, and vegetation shifts. If conditions become wetter, deterioration due to faster decomposition and cracking may be a factor. ● Increases in fire intensity may directly damage archaeological resources due to heat, combustion, or firefighting activities. However, increases in fire frequency may increase protection of some archaeological resources via fuel reduction. ● Increased extreme rain events, rain on snow, accelerated snowmelt, and/or higher severity fires may increase water runoff and erosion (especially with post-fire storms). ● Climate and/or fire-driven vegetation loss also could contribute to increased erosion. Erosion may result in physical damage, loss of integrity, and loss of spatial coherence if materials are re- deposited away from their original location. ● Additionally, melting of glaciers and snowfields may expose previously buried archeological resources. ● Increased vegetation growth could hide archaeological resources from view or cause more root damage to these resources. Changes in burrowing animals, micro-organisms, insects (termites), and fungi could change rates of decay.

C-55

● Lack of information on the location, historic or pre-historic context, and condition reduces managers’ abilities to protect archaeological resources. Confidence: Low - Most of SEKI has not been surveyed so we don’t know the location or condition of many archaeological resources. Future vulnerability is uncertain.

Cultural Anthropology & Ethnographic Resources

● The greatest climate-related threats to ethnographic resources are fire, drought, flooding, and species shifts. ● Fires may scar or obliterate rock images and may change the character of significant sites. ● Drought may cause decreases in acorn harvest (especially mast years) and increased mortality of locally-favored species of oak. ● Flooding may alter the character of significant sites. ● Species shifts may also alter the character of significant sites and may mean that important plant species are no longer associated with the same locations. Invasion by non-native species may have similar effects. ● Ethnographies for the area are outdated (Gayton 1948). Lack of information on the type and condition of resources reduces managers’ abilities to protect ethnographic resources. Confidence: Low - Low confidence on identify, current condition, and vulnerability of ethnographic resources.

Museum Collections & Archives

● Museum and archival collections are entirely dependent on active management. Without an adequate and consistent level of funding, collections cannot be made available to researchers, cannot be cared for, and may not persist. ● For electronic data, we depend on changing technologies in software and hardware. This can constitute a threat of total loss. ● Storage environments require monitoring and control. With rising temperatures a given, the natural deterioration of museum holdings will accelerate. Any shift in the relative humidity of the ambient environment of museum storage spaces will also pose specific threats. Managing museum environments will require a greater dependence on utility systems that will likely become more expensive and less reliable. ● Climate change and its threats to the parks' biotic resources may indirectly contribute to a loss in perceived value of the parks' cultural resources. ● The increasing probability of high-intensity and long-lasting fires threatens collections environments with smoke inundation and potential destruction. The need to evacuate museum collections will place increasing stresses on park resources. ● New collections will require additional climate-controlled space. Confidence: Low to high - Certainty about deterioration is high, but whether or not rates of loss will increase or decrease is low confidence due to uncertainty in future funding and environmental variables. C-56

Referenced Literature

Allen, C.D., D.D. Breshears, and N.G. McDowell. 2015. On underestimation of global vulnerability to tree mortality and forest die-off from hotter drought in the Anthropocene. Ecosphere. 6(8):129. Ambrose, A.R., W.L. Baxter, C.S. Wong, S.S.O. Burgess, C.B. Williams, R.R. Naesborg, G.W. Koch, and T.E. Dawson. 2016. Hydraulic constraints modify optimal photosynthetic profiles in giant sequoia trees. Oecologia, DOI 10.1007/s00442-016-3705-3 Anderson, R.S. 1994. Paleohistory of a Giant Sequoia Grove: The Record from Log Meadow, Sequoia National Park. USDA Forest Service Gen. Tech. Rep.PSW-151. Anderson, R.S. and S.J. Smith. 1994. Paleoclimatic interpretations of meadow sediment and pollen stratigraphies from California. Geology 22:723-726. Andrews, E.D. 2013. Hydrology of the Sierra Nevada Network national parks: status and trends. Natural Resource Report NPS/SIEN/NRR-2013/500. National Park Service, Fort Collins, CO. Austin, J., D. Boiano, D. Gammons, E. Meyer, and H. Werner. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 19 - native and non-native vertebrate species. Natural Resource Report NPS/SEKI/NRR- 2013/665.19. National Park Service, Fort Collins, CO. Barr T.C. and Kuehne R.A. 1971 - Ecological studies in the Mammoth Cave system of Kentucky. II. The ecosystem.Annales de Spéléologie, 26: 47-96. Basagic, H. and J. Panek. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 4 - glaciers. Natural Resource Report NPS/SEKI/NRR-2013/665.4. National Park Service, Fort Collins, CO. Battles, J.J., T. Moody, and D. Saah. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 21 - altered fire regimes. Natural Resource Report NPS/SEKI/NRR-2013/665.21. National Park Service, Fort Collins, CO. Battles, J.J., D.S. Saah, T. Robards, S. Cousins, R.A. York, and D. Larson. 2013b. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 12 - intact forests. Natural Resource Report NPS/SEKI/NRR-2013/665.12. National Park Service, Fort Collins, CO. Borchert, M.I., Davis, F.W., Michaelsen, J., and Oyler, L.D. 1989. Interactions of Factors Affecting Seedling Recruitment of Blue Oak (Quercus Douglasii) in California. Ecology, Vol. 70, No. 2 (Apr., 1989), pp. 389-404, Ecological Society of America. http://www.jstor.org/stable/1937544 Accessed: 09-10-2015 17:51 UTC Bouldin, J. 1999. Twentieth Century Changes in Forests of the Sierra Nevada Mountains. PhD Thesis, University of California, Davis. Bradford, D.F., R.A. Knapp, D.W. Sparling, M.S. Nash, K.A. Stanley, N.G. Tallent-Halsell, L.L. McConnell, and S.M. Simonich. 2011. Pesticide distributions and population declines of California, USA alpine frogs, Rana muscosa and Rana sierrae. Environmental Toxicology and Chemistry 30:682-691.

C-57

Cahill, M. 2013. An investigation into the range expansion of white pine blister rust (Cronartium ribicola) in subalpine forests in Sequoia and Kings Canyon National Parks. Investigator Annual Report. https://irma.nps.gov/rprs/Iar/Search Cahill, M. 2014. Changes in white pine blister rust (Cronartuim ribicola) infections in high- elevation white pines in Sequoia and Kings Canyon National Parks, USA. Masters Thesis. Department of Plant Biology, University of Vermont. Burlington, VT. California Division of Fish and Wildlife. 2016. Sierra Nevada Bighorn Sheep Recovery Program website, accessed February 12, 2016. http://www.dfg.ca.gov/snbs/RecoveryHome.html Caprio, A.C. and P. Lineback. 2002. Pre-twentieth century fire history of Sequoia and Kings Canyon National Parks: a review and evaluation of our knowledge. In: Proceedings: Fire in California Ecosystems: Integrating Ecology, Prevention, and Management, Nov. 18- 20, 1997, San Diego, CA. Association for Fire Ecology Misc. Publ. No. 1:180-199. Cayan, D. R., E. P. Maurer, M. D. Dettinger, M. Tyree, and K. Hayhoe. 2008. Climate change scenarios for the California region. Climatic Change 87:21-42. Clow, D.W., H. Forrester, B. Miller, H. Roop, J.O. Sickman, H. Ryu, and J.S. Domingo. 2013. Effects of stock use and backpackers on water quality in wilderness in Sequoia and Kings Canyon National Parks. Environmental Management. doi:10.1007/s00267-013- 0166-x Cole, D. N. 2009. Ecological Impacts of Wilderness Recreation and Their Management. Pgs. 395-438 in Dawson, C. P. and J. C. Hendee. Wilderness Management. Fulcrum Press. Golden, CO. Conner, M.M., J.J. Keane, C.V. Gallagher, G. Jehle, T.E. Munton, P.A. Shaklee, and R.A. Garrard. 2013. Realized population change for long-term monitoring: California spotted owl case study. The Journal for Wildlife Management 77(7):1449-1458. D’Antonio, C. M., E. L. Berlow, and K. L. Haubensak. 2004. Invasive exotic plant species in Sierra Nevada ecosystems. USDA Forest Service Gen. Tech. Rep. PSW-GTR-193. 175- 184. Das, A.J. and N.L. Stephenson. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 22 - climatic change. Natural Resource Report NPS/SEKI/NRR-2013/665.22. National Park Service, Fort Collins, CO. Das, T., M.D. Dettinger, D.R. Cayan, and H.G. Hidalgo. 2011. Potential increase in floods in California’s Sierra Nevada under future climate conditions. Climatic Change 109:S71- S94. Davidson, C. and R.A. Knapp. 2007. Multiple stressors and amphibian declines: duel impacts of pesticides and fish on yellow-legged frogs. Ecological Applications, 17:587–597. Davis, O.K. 1999. Pollen analysis of , CA: Great-Basin-like vegetation in Central California during the full glacial and early Holocene. Review of Paleobotany and Palynology 107:249-257. Day, J.P. and M. Conklin. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 6 - water quality, Natural Resource Report NPS/SEKI/NRR-2013/665.6. National Park Service, Fort Collins, CO. DeBenedetti, S.H. and D.J. Parsons.1984. Postfire succession in a Sierran subalpine meadow. American Midland Naturalist, 118-125. C-58

Dettinger, M. D. 2005. From climate‐change spaghetti to climate‐change distributions for 21st Century California. San Francisco Estuary and Watershed Science 3(1): Article 4. Dettinger, M. D., D. R. Cayan, N. Knowles, A. Westerling, and M. K. Tyree. 2004a. Recent Projections of 21st‐Century Climate Change and Watershed Responses in the Sierra Nevada, USDA Forest Service. Gen. Tech. Report PSW‐GTR‐193. Dettinger, M. D., D. R. Cayan, M. K. Meyer, and A. E. Jeton. 2004b. Simulated Hydrologic Responses to Climate Variations and Change in the Merced, Carson, and American River Basins, Sierra Nevada, California, 1900–2099. Climate Change 62:283-317. Dodd, R.S. and R. DeSilva. 2016. Long-term demographic decline and late glacial divergence in a Californian paleoendemic: Sequoiadendron giganteum (giant sequoia). Ecology and Evolution. 6:3057–3439. doi: 10.1002/ece3.2122 Dolanc, C. R., J. H. Thorne, and H. D. Safford. 2013. Widespread shifts in the demographic structure of subalpine forests in the Sierra Nevada, California, 1934 to 2007. Global Ecology and Biogeography 22(3):264‐276. Duriscoe, D. M., and C. S. Duriscoe. 2002. Survey and monitoring of white pine blister rust in Sequoia and Kings Canyon National Parks: Final report on 1995-1999 survey and monitoring plot network. Science and Natural Resources Management Division Sequoia and Kings Canyon National Parks, Three Rivers, CA. Eagles-Smith, C.A., Willacker, J.J., and Flanagan Pritz, C.M. 2014, Mercury in fishes from 21 national parks in the Western United States—Inter and intra-park variation in concentrations and ecological risk: U.S. Geological Survey Open-File Report 2014-1051, 54 p. http://dx.doi.org/10.3133/ofr20141051. Ecology Graduate Student Project Collective (EGSPC) and M.W. Schwartz. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 15a - animals of conservation concern. Natural Resource Report NPS/SEKI/NRR-2013/665.15. National Park Service, Fort Collins, CO. Edburg, S. L., J. A. Hicke, P. D. Brooks, E. G. Pendall, B. E. Ewers, U. Norton, D. Gochis, E. D. Gutman, and A. J. Meddens. 2012. Cascading impacts of bark beetle-caused tree mortality on coupled biogeophysical and biogeochemical processes. Frontiers in Ecology and the Environment 8:416-424. Epanchin, P.N., R.A. Knapp, and S.P. Lawler. 2010. Nonnative trout impact an alpine-nesting bird by altering aquatic-insect subsidies. Ecology. Vol. 91, No. 8 (August 2010), pp. 2406-2415 http://www.jstor.org/stable/27860805?seq=1#page_scan_tab_contents Fenn, M. E., Allen, E.B., Weiss, S.B., Jovan, S., Geiser, L., Tonnesen, G.S., Johnson, R.F., Rao, L.E., Gimeno, B.S., Yuan, F., Meixner, T., and Bytnerowicz, A. 2010. Nitrogen critical loads and management alternatives for N-impacted ecosystems in California. Journal of Environmental Management, 91, 2404–2423. Fisichelli, N.A., G.W. Schuurman, W.B. Monahan, and P.S. Ziesler. 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 Fites‐Kaufman, J. A., P. Rundel, N. Stephenson, and D. A. Weixelman. 2007. Montane and Subalpine Vegetation of the Sierra Nevada and Cascade Ranges. Terrestrial Vegetation of California. Berkeley, University of California Press: 456‐501.

C-59

Flint, L. E., A. L. Flint, J. H. Thorne and R. Boynton. 2013. Fine‐scale hydrologic modeling for regional landscape applications: the California Basin Characterization Model development and performance. Ecological Processes 2: 25. Frick, W.F., A. Shawler, S. Topp, and P.A. Heady III. 2015. Surveys for winter activity and cave use by bats in Sequoia-Kings Canyon National Park. Draft Report to Sequoia and Kings Canyon National Parks, Three RIvers, CA. June 2015. Geist, N.R., Z. Dallara, and R. Gordon. 2015.The role of incubation temperature and clutch effects in the development and phenotype of head-started western pond turtles (Emys marmorata). Herpetological Conservation and Biology 10(Symposium):489–503. http://herpconbio.org/Volume_10/Symposium/Geist_etal_2015.pdf Geos Institute. 2013. Future Climate, Wildfire, Hydrology, and Vegetation Projections for the Sierra Nevada, California: A climate change synthesis in support of the Vulnerability Assessment/Adaptation Strategy (VAAS) process.Geos Institute, Ashland, OR. http://www.geosinstitute.org/climatewise-program/completed-projects/1044-future- climate-wildfire-hydrology-and-vegetation-projections-for-the-sierra-nevada- california.html Gerlach, J. D. J., P. E. Moore, B. Johnson, D. G. Roy, D. M. Whitmarsh, D. M. Lubin, D. M. Graber, S. Hultain, A. Pfaff, and J. Keeley. 2003. Alien Plant Species Threat Assessment and Management Prioritization for Sequoia-Kings Canyon and Yosemite National Parks. Open File Report 02-170. U.S. Geological Survey, Carson City, NV. Glick, P., B.A. Stein, and N.A. Edelson, editors. 2011. Scanning the Conservation Horizon: A Guide to Climate Change Vulnerability Assessment. National Wildlife Federation, Washington, D.C. Gonzalez, P. 2012. Climate change trends and vulnerability to biome shifts in the southern Sierra Nevada. National Park Service. Climate Change Response Program. Report for Sequoia and Kings Canyon National Parks. Graham, E.A., P. W. Rundel, W. Kaiser, Y. Lam, M. Stealey, and E.M. Yuen. 2012. Fine‐Scale Patterns of Soil and Plant Surface Temperatures in an Alpine Fellfield Habitat, White Mountains, California. Arctic, Antarctic, and Alpine Research 44(3):288-295. Griffin, D., and K. J. Anchukaitis. 2014. How unusual is the 2012–2014 California drought?, Geophys. Res. Lett.,41, 9017–9023, doi:10.1002/2014GL062433. Grulke, N.E. 2010. Plasticity in physiological traits in conifers: implications for response to climate change in the western U.S. Environmental Pollution 158(6): 2032-2042. http://www.fs.fed.us/psw/publications/grulke/psw_2010_grulke001.pdf Hansen, A.J., , N. Piekielek, C. Davis, J. Haas, D.M. Theobald, J.E. Gross, W.B. Monahan, T. Olliff, and S.W. Running. 2014. Exposure of U.S. National Parks to land use and climate change 1900-2100. Ecological Applications 24(3):484-502. Harrison, R.W. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 11b - Giant sequoia literature review. Natural Resource Report NPS/SEKI/NRR--2013/665.11b. National Park Service, Fort Collins, Colorado. Haultain, S. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 8 - alpine environments. Natural Resource Report NPS/SEKI/NRR-2013/665.8. National Park Service, Fort Collins, CO.

C-60

Hauptfeld, R.S. and J.M. Kershner. 2014. Sierra Nevada Ecosystem Vulnerability Assessment Briefing: Alpine/Subalpine. Version 1.0. EcoAdapt, Bainbridge Island, WA. Hayhoe, K., D. Cayan, C.B. Field, P.C. Frumhoff, E.P. Maurer, N L. Miller, S. C. Moser, S. H. Schneider, K. N. Cahill, E. E. Cleland, L. Dale, R. Drapek, R.M. Hanemann, L.S. Kalkstein, J. Lenihan, C.K. Lunch, R.P. Neilson, S.C. Sheridan, and J.H. Verville. 2004. Emissions pathways, climate change, and impacts on California. Proceedings of the National Academy of Sciences 101(34):12422‐12427. Heard, A. M. 2005. Effects of landscape scale prescribed fire on hydrology and stream chemistry. MS Thesis. Colorado State University, Fort Collins, CO. Heard, A.M., J.O. Sickman, N.L. Rose, D.M. Bennett, D.M. Lucero, J.M. Melack, and J.H. Curtis. 2014. 20th century atmospheric deposition and acidification trends in lakes of the Sierra Nevada, California, USA. Environ. Sci. Technol. 48:10054−10061. dx.doi.org/10.1021/es500934s Heard, A.M., and J.O. Sickman. 2016. Nitrogen assessment points: development and application to high-elevation lakes in the Sierra Nevada, California. Ecosphere 7(11):e01586. 10.1002/ecs2.1586 Herbst, D.B. 2015. Climate change in the Sierra Nevada. The Herbst Lab, Sierra Nevada Research Institute, University of California Santa Barbara. Website accessed 10/2/15. http://herbstlab.msi.ucsb.edu/research.html#climatechange Hoff, R. J. 1992. How to recognize blister rust infection on whitebark pine. Research Note INT- 406, USDA Forest Service, Intermountain Research Station, Ogden UT. Hopkinson, P., M. Hammond, J. Bartolome, M. Brooks, E.L. Berlow, R. Klinger, J.R. Machet, P. Moore, S. Ostoja, C.Soulard, L. Joppa, R. Williams, O.Alvarez, Q. Guo, S. Haultain, E. Frenzel, and S. Saah. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 13 - meadows. Natural Resource Report NPS/SEKI/NRR-2013/665.13. National Park Service, Fort Collins, CO. Howarth F.G., 1980 - The zoogeography of specialized cave animals: a bioclimatic model. Evolution, 34: 394-406. Jiang, X., S.A. Rauscher, T.D. Ringer, D. M. Lawrence, A.P. Williams, C.D. Allen, A.L. Steiner, S.M. Cai, and N.G. McDowell. 2013. Projected future change in vegetation in western North America in the Twenty-First Century. Journal of Climate 26:3671-3687. DOI: 10.1175/JCLI-D-12-00430.1 Jarman W. M. 1994. Levels and trends of DDE in California peregrines. U.S. Fish and Wildlife Service Report. 16 pp. http://webapp1.dlib.indiana.edu/virtual_disk_library/index.cgi/4259133/FID1247/R12/R12 MB.PDF Jeffress, M.R., T.J. Rodhouse, C.Ray, S. Wolff, and C.W. Epps. 2013. The indiosyncrasies of place: geographic variation in the climate-distribution relationships of the American pika. Ecological Applications 23:864-878. Karvonen A, P. Rintamäki, J. Jokela, and E.T. Valtonen. 2010. Increasing water temperature and disease risks in aquatic systems: climate change increases the risk of some, but not all, diseases. Int J Parasitol 40:1483–1488.

C-61

Keane, R. E., D. F. Tomback, C. A. Aubry, A. D. Bower, E. M. Campbell, C. L. Cripps, M. B. Jenkins, M. F. Mahalovich, M. Manning, S. T. McKinney, M. P. Murray, D. L. Perkins, D. P. Reinhart, C. Ryan, A. W. Schoettle, and C. M. Smith. 2012. A range-wide restoration strategy for whitebark pine (Pinus albicaulis). Gen. Tech. Rep. RMRS-GTR-279. Fort Collins, CO: U.S. Department of Agriulture, Forest Service, Rocky Mountain Research Station. 108 p. Keeley, J. E., M. Baer-Keeley, and C. J. Fotheringham. 2005a. Alien plant dynamics following fire in Mediterranean-climate California shrublands. Ecological Applications 15:2109- 2125. Keeley, J. E., A.H. Pfaff, and H.D. Stafford. 2005b. Fire suppression impacts on postfire recovery of Sierra Nevada chaparral shrublands. International Journal of Wildland Fire 14: 255–265. Kershner, J.M., editor. 2014. A Climate Change Vulnerability Assessment for Focal Resources of the Sierra Nevada. Version 1.0. EcoAdapt, Bainbridge Island, WA. http://ecoadapt.org/data/documents/EcoAdapt_CALCC_SierraNevadaVulnerabilityAsses sment_26Feb2014.pdf Knapp, R. A., D. M. Boiano, and V. T. Vredenburg. 2007. Removal of nonnative fish results in population expansion of a declining amphibian (mountain yellow-legged frog, Rana muscosa). Biological Conservation 135:11-20. Knowles, N. and D. Cayan. 2004. Elevational dependence of projected hydrologic changes in the San Francisco Estuary and Watershed. Climate Change 62:319-336. Koopman, M.E. 2014. Future Climate, Hydrology, Vegetation, and Wildfire Projections for the Southern Sierra Nevada, California: A Climate Change Synthesis in Support of Integrated Regional Water Management Planning. Geos Institute, Ashland, OR. http://www.southernsierrarwmg.org/documentsstudies.html Lacan, I. K. Matthews, and K. Feldman. 2008. Interaction of an introduced predator with future effects of climate change in the recruitment dynamics of an imperiled Sierra Nevada yellow-legged frog (Rana sierrae). Herpetological Conservation and Biology 3:211-223. Landers, D.H., S.L. Simonich, D.A. Jaffe, L.H. Geiser, D.H. Campbell, A.R. Schwindt, C.B. Schreck, M.L. Kent, W.D. Hafner, H.E. Taylor, K.J. Hageman, S. Usenko, L.K. Ackerman, J.E. Schrlau, N.L. Rose, T.F. Blett, and M.M. Erway. 2008. The Fate, Transport, and Ecological Impacts of Airborne Contaminants in Western National Parks (USA). EPA/600/R-07/138. U.S. Environmental Protection Agency, Office of Research and Development, NHEERL, Western Ecology Division, Corvallis, Oregon. https://www.nature.nps.gov/air/studies/air_toxics/WACAPreport.cfm Lavoie, K. H., & Northup, D. E. 2006. Bacteria as indicators of human impact in caves. In 17th National Cave and Karst Management Symposium, Proceedings: 40-47. Lenihan, J. M., D. Bachelet, R. P. Neilson and R. Drapek. 2008. Response of vegetation distribution, ecosystem productivity, and fire to climate change scenarios for California. Climatic Change 87(S1): 215-230. Lloyd, A.H. and L.J. Graumlich. 1997. Holocene dynamics of treeline forests in the Sierra Nevada. Ecology 78: 1199-1210. Lowry, C. S., S. P. Loheide II, C. E. Moore, and J. D. Lundquist. 2011. Groundwater controls on vegetation composition and patterning in mountain meadows. Water Resour. Res., 47 C-62

Mallek, C., H. Safford, J. Viers, and J. Miller. 2013. Modern departures in fire severity and area vary by forest type, Sierra Nevada and the southern Cascades, California, USA. Ecosphere 4:1-28. Maloney, P. E. 2011. Incidence and distribution of white pine blister rust in the high-elevation forests of California. Forest Pathology 41:308-316. Maurer, E. P. 2007. Uncertainty in hydrologic impacts of climate change in the Sierra Nevada, California, under two emissions scenarios. Climatic Change 82(3‐4):309-325. Maurer, E. P., I.T. Stewart, C. Bonfils, P.B. Duffy, and D. Cayan. 2007. Detection, attribution, and sensitivity of trends toward earlier streamflow in the Sierra Nevada. Journal of Geophysical Research 112(D11). Matthews, K.R. and N.H. Berg. 1997. Rainbow trout responses to water temperature and dissolved oxygen stress in two southern California stream pools. Journal of Fish Biology 50:50–67. http://www.fs.fed.us/psw/publications/matthews/psw_1997_matthews003.pdf Mazur, R., A.P. Kimley, and K. Folger. 2013. Implications of the variable availability of seasonal foods on the home ranges of black bears, Ursus americanus, in the Sierra Nevada of California. Animal Biotelemetry 1:16. DOI: 10.1186/2050-3385-1-16. McKinney, S. T., C.E. Fiedler, and D.F. Tomback. 2009. Invasive pathogen threatens bird-pine mutualism: implications for sustaining a high-elevation ecosystem. Ecological Applications 19:597-607. McGinnis, T. and J. Keeley. 2011. Effects of eradication and restoration treatments on Italian thistle (Carduus pycnocephalus). Madrono. 58(4): 207–213. Meentemeyer R., D. Rizzo, W. Mark, and E. Lotz, 2004. Mapping the risk of establishment and spread of sudden oak death in California. Forest Ecology and Management 200: 195– 214. https://www.researchgate.net/profile/Ross_Meentemeyer/publication/222522620_Mappi ng_the_risk_of_establishment_and_spread_of_sudden_oak_death_in_California/links/5 52fce4d0cf20ea0a06e6f36.pdf Meyer, E. and D. Boiano. 2010. Effects of non-native fish species on native fish assemblages in the Kaweah River drainage of Sequoia National Park. National Park Service. Sequoia and Kings Canyon National Parks, Three Rivers, CA. Meyer, E., D. Sparling, and S. Blumenshine. 2013. Regional inhibition of cholinesterase in free- ranging western pond turtles (Emys marmorata) occupying California mountain streams. Environ Toxicol Chem. 32(3):692-8. Meyer, E., D. Seale, B. Permar, and A. McClary. 2016. The effect of chemical treatments on lampenflora and a Collembola indicator species at a popular tour cave in California, USA. Manuscript submitted for publication in Environmental Management. Millar, C.I., and R.D. Westfall. 2010a. Distribution and climatic relationships of the American pika (Ochotona princeps) in the Sierra Nevada and western Great Basin, U.S.A; periglacial landforms as refugia in warming climates. Arctic, Antarctic, and Alpine Research 42: 76-88. Millar, C. I., R. D. Westfall, D. L. Delany, M. J. Bokach, A. L. Flint, and L. E. Flint. 2012. Forest mortality in high-elevation whitebark pine (Pinus albicaulis) forests of eastern California,

C-63

USA; influence of environmental context, bark beetles, climatic water deficit, and warming. Can. J. For. Res. 42:749-765. Miller, N. L., K. E. Bashford, and E. Strem. 2003. Potential impacts of climate change on California hydrology. Journal of American Water Resources Association 39(4):771‐784. Miller, J. D. and J. Safford. 2008. Sierra Nevada fire severity monitoring 1984-2004. R5-ECOL- TP-003. Vallejo, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Region. 102 p. Monahan, W.B. and N.A. Fisichelli. 2014. Climate exposure of US national parks in a new era of change. PLoS ONE. 9(7):e101302. http://journals.plos.org/plosone/article?id=10.1371/journal.pone.0101302 Monahan, W.B., Rosemartin, A., Gerst, K. Fisichelli N.A., Ault, T. Schwartz, M.D., Gross, J.E., and Weltzin, J.F. 2016. Climate change is advancing spring onset across the US national park system. Ecosphere 7(10):e01465. 10.1002/ecs2.1465. http://onlinelibrary.wiley.com/doi/10.1002/ecs2.1465/epdf Moore, Peggy E.; van Wagtendonk, Jan W.; Yee, Julie L.; McClaran, Mitchel P.; Cole, David N.; McDougald, Neil K.; Brooks, Matthew L. 2013. Net primary productivity of subalpine meadows in Yosemite National Park in relation to climate variability. Western North American Naturalist. 73(4): 409-418. Morelli TL, Daly C, Dobrowski SZ, Dulen DM, Ebersole JL, Jackson ST, et al. 2016. Managing Climate Change Refugia for Climate Adaptation. PLoS ONE 11(8): e0159909. doi:10.1371/journal.pone.0159909 Morgan, M, M. Rockman, C. Smith, and A. Meadow. 2016. Climate Change Impacts on Cultural Resources. Cultural Resources Partnerships and Science. Washington, DC: National Park Service. https://www.nps.gov/subjects/climatechange/impactsonculturalresources.htm Moriarty, K.M., W.J.Zielinski, A.G. Gonzales, T.E. Dawson, K.M. Boatner, C.A.Wilson, F.V. Schlexer, K.L. Pilgrim, J.P. Copeland, and M.K.Schwartz. 2009. Wolverine confirmation in California after nearly a century: native or long-distance immigrant? Northwest Science 83: 154-162. Moyle, P. 2002. Inland Fishes of California, 2nd Edition, Berkeley, University of California Press. Moyle P.B., J.D. Kiernan, P.K. Crain, and R.M. Quinones. 2013. Climate Change Vulnerability of Native and Alien Freshwater Fishes of California: A Systematic Assessment Approach. PLoS ONE 8(5): e63883. doi:10.1371/journal.pone.0063883 Neuman, M. 1996. Past and present conditions in backcountry meadows, Sequoia and Kings Canyon National Parks, 2nd edition. Unpublished report. Three Rivers, CA. National Park Service. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks. Natural Resource Report NPS/SEKI/NRR-2013/665. National Park Service, Fort Collins, Colorado. NPS 102/120974, June 2013. http://www.nature.nps.gov/water/nrca/assets/docs/SEKI_NRCA_Final.pdf National Park Service. 2015. Wilderness Stewardship Plan and Final Environmental Impact Statement, Sequoia and Kings Canyon National Parks, Department of the Interior, National Park Service, April 2015.

C-64

https://parkplanning.nps.gov/document.cfm?parkID=342&projectID=33225&documentID =65018. National Park Service - Natural Sounds & Night Skies Division. 2015. Night Skies and Photic Environment Resource Summary for Sequoia & Kings Canyon National Parks. Natural Resource Stewardship and Science, Natural Sounds & Night Skies Division. National Park Service. 2016. Restoration of Native Species in High Elevation Aquatic Ecosystems Plan and Final Environmental Impact Statement, Sequoia and Kings Canyon National Parks, California. Department of the Interior, National Park Service, June 2016. https://parkplanning.nps.gov/document.cfm?parkID=342&projectID=17157&documentID =73190 National Park Service - Air Resources Division. 2016. Air quality conditions and trends website, http://www.nature.nps.gov/air/data/products/parks/index.cfm, (select Sequoia & Kings Canyon National Parks) and Air Pollution Impacts - Sequoia & Kings Canyon National Parks, http://www.nature.nps.gov/air/Permits/aris/seki/impacts.cfm, accessed August 24, 2016. Null, S. E., J. H. Viers, and J. F. Mount. 2010. Hydrologic response and watershed sensitivity to climate warming in California's Sierra Nevada. PLoS One 5(4). Ostoja, Steven M. , M.L. Brooks, P. E. Moore, E. L. Berlow, R. Blank, J. Roche, J. Chase, and S. Haultain. 2014. Potential environmental effects of pack stock on meadow ecosystems of the Sierra Nevada, USA. The Rangeland Journal 2014(36): 411–427. Panek, J., D. Saah, and A. Esperanza. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 2 - air quality, Natural Resource Report NPS/SEKI/NRR-2013/665.1. National Park Service, Fort Collins, CO. Panek, J. and J. Despain. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 18 - ace invertebrates, Natural Resource Report NPS/SEKI/NRR-2013/665.1. National Park Service, Fort Collins, CO. Perera, E.M. and T. Sanford. 2011. Climate change and your health: rising temperatures, worsening ozone pollution. Union of Concerned Scientists. Cambridge, MA. http://www.ucsusa.org/sites/default/files/legacy/assets/documents/global_warming/climat e-change-and-ozone-pollution.pdf Porter, E., H. Sverdrup, and T.J. Sullivan. 2011. Estimating and mitigating the impacts of climate change and air pollution on alpine plant communities in national parks. Park Science 28(2):58-64. http://www.nature.nps.gov/parkscience/Archive/PDF/Article_PDFs/ParkScience28(2)Su mmer2011_58-64_Porter_et_al_2815.pdf Redmond, K and M. Fearon. 2013. Southern Sierra Nevada Temperature Reconstruction. Draft report to Sequoia and Kings Canyon National Park. Western Regional Climate Center. Desert Research Institute, Reno, NV. http://www.wrcc.dri.edu/nps/ssnv/index.html Rockman, M. 2015. An NPS Framework for Addressing Climate Change with Cultural Resources. The George Wright Forum, vol. 32, no. 1, pp. 37-50. http://www.georgewright.org/321rockman.pdf Rodriguez-Buritica, S., and K. Suding. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 10 – foothills vegetation. Natural C-65

Resource Report NPS/SEKI/NRR—2013/665.10. National Park Service, Fort Collins, Colorado. Rowe, K.C., K.M.C. Rowe, M.W. Tingley, M.S. Koo, J.L. Patton, C.J. Conroy, J.D. Perrine, S.R. Beissinger, and C. Moritz. 2014. Spatially heterogeneous impact of climate change on small mammals of montane California. Proc. R. Soc. B. 282:20141857. http://dx.doi.org/10.1098/rspb.2014.1857 Safford, H.D., M. North, and M.D. Meyer. 2012. Climate change and the relevance of historical forest conditions. In Managing Sierra Nevada Forests, edited by M. North.Albany, CA. US Department of Agriculture, Forest Service. Pacific Southwest Research Station. http://www.fs.fed.us/psw/publications/documents/psw_gtr237/ Saros J.E., D.W. Clow, T. Blett, and A.P. Wolfe. Critical nitrogen deposition loads in high- elevation lakes of the Western US inferred from paleolimnological records. Water, Air, & Soil Pollution. 2011;216(1–4):193–202. doi: 10.1007/s11270-010-0526-6. Schoettle, A. and W. Sniezko. 2007. Proactive intervention to sustain high-elevation pine ecosystems threatened by white pine blister rust. Journal of Forest Research 12:327- 336. Schwartz, M. W., N. Butt, C. R. Dolanc, A. Holguin, A. A. Moritz, M. P. North, H. D. Safford, N. L. Stephenson, J. H. Thorne, and P. J. van Mantgem. 2015. Increasing elevation of fire in the Sierra Nevada and implications for forest change. Ecosphere 6(7):121. Schwartz, M.W., J. Thorne, and A. Holguin. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 20a - biodiversity. Natural Resource Report NPS/SEKI/NRR-2013/665.20a. National Park Service, Fort Collins, CO. Seager, R., M. Ting, I. Held, Y. Kushnir, J. Lu, G. Vecchi, H. P. Huang, N. Harnik, A. Leetmaa, N. C. Lau, C. Li, J. Velez, and N. Naik. 2007. Model projections of an imminent transition to a more arid climate in southwestern North America. Science 316(5828):1181‐1184. Sickman, J.O., A. Leydecker, and J.M. Melack. 2001. Nitrogen mass balances and abiotic controls on N retention and yield in high-elevation catchments of the Sierra Nevada, California, United States. Water Resources Research 37:1445-1461. Sleeman, J. 2016. White-Nose Syndrome Updates for the 2015/2016 Surveillance Season. U.S. Geological Survey, National Wildlife Health Center. Wildlife Health Bulletin 2016-05. July 15, 2016. http://www.nwhc.usgs.gov/publications/wildlife_health_bulletins/WHB_2016- 05_WNS_updates.pdf Smith, A.T. 1974. The distribution and dispersal of pikas: Influence of behavior and climate. Ecology 55(6):1368-1376. Steel, Z.L., M.L. Bond, R.B. Siegel, and P. Pyle. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 17 - birds. Natural Resource Report NPS/SEKI/NRR-2013/665.17. National Park Service, Fort Collins, CO. Stewart, J.A.E., J.D. Perrine, L.B. Nichols, J.H. Thorne, C.I. Millar, K.E. Goehring, C.P.Massing, and D.H.Wright. Revisiting the past to foretell the future: summer temperature and habitat area predict pika extirpations in California. Journal of Biogeography 42:880-890.

C-66

Surface Water Ambient Monitoring Program (SWAMP). 2015. California Environmental Protection Agency. State Water Resources Control Board. http://www.waterboards.ca.gov/water_issues/programs/swamp/ Thorne, J., W.B. Monahan, A. Holguin, and M.W. Schwartz. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 1 - landscape context. Natural Resource Report NPS/SEKI/NRR-2013/665.1. National Park Service, Fort Collins, CO. Thorne, J. H., R. Boynton, L. Flint, A. Flint, and T.-N. G. Le. 2012. Development and Application of Downscaled Hydroclimatic Predictor Variables for Use in Climate Vulnerability and Assessment Studies, Prepared for California Energy Commission, Prepared by University of California, Davis. CEC-500-2012-010. Tobin, B., Krejca, and J. Despain. 2009. Final Report: Hidden Fire Suppression Damage Research Projects. Sequoia and Kings Canyon National Parks, Three Rivers, CA. 18 pp. Tobin, B. and B. Schwartz. 2011. Quantifying concentrated and diffuse recharge in two marble karst aquifers: Big Spring and Tufa Spring, Sequoia and Kings Canyon National Parks, California, USA. Journal of Cave and Karst Studies, v. 74, no. 2, p. 186–196. DOI: 10.4311/2011JCKS0210. Tobin B.W., Hutchins B.T. and Schwartz B.F. 2013. Spatial and temporal changes in invertebrate assemblage structure from the entrance to deep-cave zone of a temperate marble cave. International Journal of Speleology, 43 (3), 203-214. Tampa, FL (USA) ISSN 0392-6672 http://dx.doi.org/10.5038/1827-806X.42.3.4 Tomback, D. F. and P. Auchuff. 2010. Blister rust and western forest biodiversity: ecology, values, and outlook for white pines. Forest Pathology 40:186-225. Tu, M., A. Demetry, and D. Saah. 2013. A natural resource condition assessment for Sequoia and Kings Canyon National Parks: Appendix 23 - non-native plants. Natural Resource Report NPS/SEKI/NRR-2013/665.23. National Park Service, Fort Collins, CO. University of California, Davis. 2015. Interagency Monitoring of Protected Visual Environments (IMPROVE) Annual Site Report. http://airquality.crocker.ucdavis.edu/files/1114/3931/1693/SiteReports.pdf. Date accessed August 14, 2015.US Fish and Wildlife Service. 2007. Recovery plan for the Sierra Nevada Bighorn Sheep. https://nrm.dfg.ca.gov/FileHandler.ashx?DocumentID=27634 U.S. Fish and Wildlife Service. 2007. Recovery Plan for the Sierra Nevada Bighorn Sheep. California/Nevada Operations Office, U.S. Fish and Wildlife Service, Sacramento, CA. https://nrm.dfg.ca.gov/FileHandler.ashx?DocumentID=27634&inline=true U.S. Fish and Wildlife Service. 2010. Endangered and threatened wildlife and plants; 12-month finding on a petition to list the American pika as threatened or endangered. Federal Register 75:6438-6471. U.S. Fish and Wildlife Service. 2016. West Coast Population of Fisher Will not be Listed Under Endangered Species Act, press release, April 14, 2016. https://www.fws.gov/news/ShowNews.cfm?ref=west-coast-population-of-fisher-will-not- be-listed-under-endangered-&_ID=35559 U.S. Forest Service. 2016. Forest health protection special aerial detection survey, Sierra Nevada foothills, Traverse, and Coastal Ranges, May 16-19, 2016. Pacific Southwest C-67

Region. Forest Health Protection Aerial Survey Program. http://www.fs.usda.gov/Internet/FSE_DOCUMENTS/fseprd509278.pdf Van de Ven, C. M., S. B. Weiss, and W. G. Ernst. 2007. Plant Species Distributions under Present Conditions and Forecasted for Warmer Climates in an Arid Mountain Range. Earth Interactions 11(9):1‐33. van Mantgem, P. J., J. C. B. Nesmith, M. Kiefer. E. E. Knapp, A. Flint, and L. Flint. 2013. Climatic stress increases forest fire severity across the western United States. Ecology Letters 16:1151-1156. van Mantgem, P.J., and N.L. Stephenson. 2007. Apparent climatically-induced increase of tree mortality rates in a temperate forest. Ecology Letters 10:909-916. van Mantgem, P.J., A.C. Caprio, N.L. Stephenson, and A.J. Das. 2016. Does prescribed fire promote resistance to drought in low elevation forests of the Sierra Nevada, California, USA? Fire Ecology 12:13-25. doi: 10.4996/fireecology.1201013. van Wagtendonk, J. W. and J. Fites-Kaufman. 2006. Sierra Nevada bioregion. In: N. G. Sugihara, J. W. van Wagtendonk, J. Fites-Kaufman, K. E. Shaffer, A. E. Thode, eds. Fire in California’s ecosystems. University of California Press, Berkeley, CA. Chapter 12. Varner, J. and M.D. Dearing. 2014. The importance of biologically relevant microclimates in habitat suitability assessments. PLoS ONE 9(8): e104648. doi:10.1371/journal.pone.0104648 Verhoeven, J. T., Arheimer, B., Yin, C., & Hefting, M. M. (2006). Regional and global concerns over wetlands and water quality. Trends in ecology & evolution, 21(2), 96-103. Viers, J.H., S.E. Purdy, R.A. Peek, A. Fryjoff-Hung, N.R. Santos, J.V.E. Katz, J.D. Emmons, D.V. Dolan and S.M.Yarnell. 2013. Montane Meadows in the Sierra Nevada: Changing Hydroclimatic Conditions and Concepts for Vulnerability Assessment. Center for Watershed Sciences Technical Report (CWS-2013-01), University of California, Davis. 63 pp. White-Nose Syndrome.org. 2016. Frequently Asked Questions (FAQs). https://www.whitenosesyndrome.org/faqs, accessed 9/230/2016. Williams, M. R., and J. M. Melack. 1997. Effects of prescribed burning and drought on the solute chemistry of mixed-conifer forest streams of the Sierra Nevada, California. Biogeochemistry 39:225-253. Williams, A.P., R. Seager, J.T. Abatzoglou, B.I. Cook, .J E. Smerdon, E.R. Cook. 2015. Contribution of anthropogenic warming to California drought during 2012–2014. Geophysical Research Letters, 42, 6819-6828, doi:10.1002/2015GL064924. http://www.ldeo.columbia.edu/~williams/ca_drought_2015_grl.html Winter, T. C. 2000. The vulnerability of wetlands to climate change: a hydrologic landscape perspective. Journal of the American Water Resources Association 36: 305-311. Wolf, E. A. Demetry, D. Cooper, and J. Wagner. 2015. Existing condition assessment and preliminary stabilization/restoration alternatives for Cahoon Meadow, Sequoia & Kings Canyon National Parks. Wolf. S. G. 2010. Distribution and climatic relationships of the American pika (Ochotona princeps) in the Sierra Nevada and western Great Basin, U.S.A; periglacial landforms as

C-68

refugia in warming climates. Comment. Arctic, Antarctic, and Alpine Research 42: 490- 492. Wood, L. 2015. Acoustic Environment and Soundscape Resource Summary, Sequoia & Kings Canyon National Parks. Natural Sounds & Night Skies Division. Wood, S.H. 1975. Holocene Stratigraphy and Chronology of Mountain Meadows, Sierra Nevada, California. PhD Dissertation. California Institute of Technology, Pasadena, CA. Young, C. A., M. I. Escobar-Arias, M. Fernandes, B. Joyce, M. Kiparsky, J. F. Mount, V. K. Mehta, D. Purkey, J. H. Viers, and D. Yates. 2009. Modeling the Hydrology of Climate Change in California’s Sierra Nevada for Subwatershed Scale Adaptation. Journal of American Water Resources Association 45(6):1409‐1423.

C-69

National Park Service U.S. Department of the Interior

Sequoia and Kings Canyon National Parks

The National Park Service preserves unimpaired the natural and cultural resources and values of the National Park System for the enjoyment, education, and inspiration of this and future generations.