<<

Impacts of on Salmon of the Pacific Northwest

A review of the scientific literature published in 2016

Lisa Crozier

Fish Ecology Division Northwest Fisheries Science Center National Marine Fisheries Service, NOAA 2725 Montlake Boulevard East Seattle, Washington 98112

July 2017

Highlights

Global annual mean temperature in 2016 was the warmest on record, continuing the streak of unprecedented temperatures that has characterized the past decade (NOAA 2017). The long-term global trend in annual temperature is now 0.07°C/decade since 1880, and 0.17°C/decade since 1970.

Rupp et al. (2016) analyzed differences across the latest round of global (GCM) projections and summarized these projections for the Columbia Basin. They reported that the global models that best represented the climatology of the Pacific northwest (PNW) also showed the largest temperature and precipitation changes, although some ranking metrics produced a different result. Consequently, ensemble means of all GCMs might be conservative projections for our region. Rana and Moradhkani (2016) also analyzed differences among GCM projections and statistical downscaling methods for the Columbia Basin for 10 GCMs.

New projections showing atmospheric rivers will become more intense are consistent with those previously reported, while those for changes in upwelling are somewhat more conservative. in the PNW was previously reported to be somewhat lower than the global mean. However, a new NOAA report found that sea level rise in the region is likely to exceed the global mean under moderate and high emission scenarios (Sweet et al. 2017).

Increasing numbers of studies attributed recent extreme events to anthropogenic emissions rather than natural climate or oceanographic variability. These events include the high temperatures seen in 2015 (Herring et al. 2016) and in the California Current (Feely et al. 2016; Turi et al. 2016).

Perturbation of the ocean soundscape is an interesting impact of ocean acidification (OA) not mentioned in previous literature reviews. Rossi et al (2016) reported that OA diminishes biological sounds and alters sound perception, which could affect fish dispersal, settlement, and other behaviors.

New results from an ecosystem model for the northern California Current found that intermediate levels of upwelling are optimal for marine productivity across four trophic levels (Ruzicka et al. 2016a). Shorter duration of upwelling also improves productivity. However, the impact of climate change on upwelling dynamics is still highly uncertain.

2 Keefer and Caudill (2016) showed that migrating Columbia River steelhead and fall Chinook experienced temperatures up to 3°C above the general river temperature (24 vs. 21°C) when traveling through fishways and into reservoirs. Benda et al. (2016) found that Willamette River Chinook showed significantly less prespawn mortality when held at cooler-than-river temperatures prior to outplanting, which is similar to previous reported results for Fraser River sockeye.

3 Contents

Highlights ...... 2

Objective and Methods ...... 5

Physical Changes in Climate and pH ...... 7 Annual state of the climate in 2016 ...... 7 Global and national climate projections ...... 7 Columbia Basin Climate Projections ...... 9 California Current ...... 10 Projections...... 10 Retrospective...... 11 Stream temperature and flow projections ...... 13 Ocean acidification ...... 14 Projections...... 14 Retrospective analyses ...... 14 Ecological impacts of ocean acidification ...... 15

Biological Responses to Climate Change ...... 16 Species-level impacts ...... 16 Projections...... 16 Retrospective analyses ...... 16 Stage-specific impacts in salmon and their ecosystems ...... 16 Juvenile freshwater stages: temperature and growth ...... 16 Adult freshwater stages: temperature and disease effects on upstream migrants...... 17 Marine survival ...... 17 Climate vulnerability assessments and Adaptive Capacity ...... 20

Literature Cited ...... 22

4 Objective and Methods

The goal of this review was to identify literature published in 2016 that is most relevant to prediction and mitigation of climate change impacts on Columbia River salmon listed under the Endangered Species Act. In our literature search, we elected to focus on peer-reviewed scientific journals included in the Web of Science database, although we occasionally included highly influential reports outside that database.

We sought to capture the most relevant papers by combining climatic and salmonid terms in search criteria. This excluded studies of general principles demonstrated in other taxa or within a broader context. In total, we reviewed over 500 papers, 80 of which were included in this summary.

Literature searches were conducted in February 2017 using the Institute for Scientific Information (ISI) Web of Science indexing service. Each set of search criteria involved a new search, and results were compared with previous searches to identify missing topics. We used specific search criteria that included a publication year of 2016, plus:

1) A topic that contained the terms climate,1 temperature, streamflow, flow, snowpack, precipitation, or2 PDO, and a topic that contained salmon, Oncorhynchus, or steelhead, but not aquaculture or fillet 2) A topic that contained climate, temperature, precipitation, streamflow or flow and a topic containing "Pacific Northwest" 3) A topic that contained the terms marine, sea level, hyporheic, or groundwater and climate, and salmon, Oncorhynchus, or steelhead 4) Topics that contained upwelling or estuary and climate and Pacific 5) Topics that contained ocean acidification and salmon, Oncorhynchus or steelhead 6) Topics that contained upwelling or estuary or ocean acidification and California Current, Columbia River, Puget Sound or Salish Sea 7) A topic that contained prespawn mortality

This review is presented in two major parts, with the first considering changes to the physical environmental conditions that are both important to salmon and projected to change with climate. Such conditions include air temperature, precipitation, snowpack, stream flow, stream temperature, and ocean conditions. We describe projections driven by global climate model (GCM) simulations, as well as historical trends and relationships

1 The wildcard (*), was used to search using "climat*" to capture all forms of the word "climate." 2 Boolean operators used in the search are shown in boldface.

5 among these environmental conditions. In the second part, we summarize the literature on responses of salmon to these environmental conditions, both projected and retrospective, in freshwater and marine environments.

6 Physical Changes in Climate and pH

Annual state of the climate in 2016

In 2016, a new global annual temperature record was set for the third year in a row (NOAA 2017). Last year was the warmest in the 137-year series at 0.94°C above the 20th century average. The world’s oceans were 0.75°C above average. The highest ocean anomalies were due to a near-record El Niño at the beginning of the year. For 40 consecutive years, the annual temperature has been above average, with the five warmest years occurring since 2010. Land temperatures were 1.43°C above average. The long-term trend now shows a rate of warming of 0.07°C/decade since 1880, and 0.17°C/decade since 1970.

Annual average temperature for the contiguous United States was the second highest among all 122 years on record, behind only 2012, and was the 20th consecutive higher-than-average annual temperature. In the higher latitudes, Alaska had its warmest year since recording began in 1925, surpassing the previous record set in 2015. Many regions across western Canada also had the warmest summer on record in 2016.

Global and national climate projections

New analyses of the coupled model intercomparison project (CMIP5) ensemble global climate model (GCM) projections for the fourth U.S. National Climate Assessment have been completed, but are still in draft form with expected publication date in 2018. Other analyses characterized projected changes in the frequency of “mild weather,” which was defined as maximum temperatures between 18 and 30°C, precipitation not more than 1 mm, and dew point temperature not more than 20°C (van der Wiel et al. 2016).

The mean number of mild days per year declined by 11 in 2081-2100 projections, when weighted spatially by current human population densities. The frequency of mild days increased in mid-latitudes and decreased in the tropics and subtropics. The probability of mild weather increased along the U.S. West Coast in spring and fall (and winter in southern California) and decreased in summer. These projections showed the same trends observed in the recent past.

The U.S. Environmental Protection Agency produced their fourth report updating information from current indicators of climate change (EPA 2016). The report

7 summarizes most of the indicators mentioned previously in national and global reports using slightly different spatial extents and time periods. In addition, the report includes some new indicators especially relevant to topics in this review: river flooding, , stream temperature, and marine species distributions.

From 1940 to 2014, low (7-d) and high (3-d) streamflows tended to show minimal change or decreases up to 50% in the west, but large increases in the central and eastern U.S. Flow data from the USGS showed that flooding events have significantly increased at many sites in western Washington, decreased in many sites in western Oregon and California, and showed mixed trends in the interior Columbia Basin.

The EPA report showed widespread reductions in snowpack in the western states and increased area burned by , with Idaho showing the largest percent area burned of anywhere in the country. They highlight a 1.4°F increase in water temperature at an unidentified site near Nez Perce land in the Snake River between 1960 and 2015 and the potential impact on salmon. They also mentioned shifts northward and to deeper water averaged across 105 marine fish and invertebrates from 1982 to 2015 from the NOAA and Rutgers University OceanAdapt database.

Based on CMIP5 GCM projections, the risk of megadroughts, or 35-year periods equal to or more severe than the worst of the 20th century, will increase by 70-99% by the end of the century in the southwestern U.S. (Ault et al. 2016). Surprisingly, the increased risk is due to temperature increases and is not very sensitive to changes in precipitation. Therefore, the projections are much more confident than expected from the variation across GCMs in future precipitation.

Flooding on the U.S. West Coast is generally most severe during atmospheric river events, as quantified by Barth et al. (2016). Several papers in previous years concluded that global warming will cause more frequent and intense atmospheric rivers, and that there will be a shift in the latitudes most affected. Studies published in 2016 showed similar results, and refined our understanding of more specific impacts.

Shields and Kiehl (2016) studied the response produced by one specific GCM (CCSM4). They found that atmospheric rivers shifted equatorward, increasing specifically in southern California (32–35°N latitude bands), where they caused more intense precipitation events. Small-to-modest increases in rainfall rates are expected from northern California to Washington State.

8 Columbia Basin Climate Projections

Rana and Moradhkani (2016) compared two statistical downscaling methodologies: bias corrected and statistically downscaled (BCSD) and multivariate adaptive constructed analog (MACA). They characterized CMIP5 projections for the Columbia Basin for 10 GCMs. They focused on examination of the differences between downscaling methods and GCMs, rather than production of regional averages (which are unchanged from previous summaries). In future scenarios, they also found changes in the frequency of 2-year vs. 4-year precipitation events, but not heat waves.

Rupp et al. (2016) completed a regional analysis of 35 CMIP5 GCM projections averaged across the Columbia River Basin, exploring some nuances that were not addressed in previous summaries (Mote et al. 2013; Vano et al. 2015a; Vano et al. 2015b). Consistent with previous work, the multi-model ensemble for the 30-year mean annual temperature averaged over the Columbia Basin increases by 5.0°C in representative concentration pathway (RCP) 8.5 and by 2.8°C in RCP 4.5 in the late 21st century over the 1979–1990 baseline. More warming occurs in summer than in winter. Three-quarters of the projections show a decrease in precipitation in summer, and over 90% of projections show an increase in winter precipitation.

Importantly, the individual GCMs that projected more warming and precipitation are the same GCMs that more closely matched historical climate patterns in the PNW. Rupp et al. (2016) found a significant linear relationship between model rank and projected summer temperature and winter precipitation changes. However, this result did depend on the metrics they used to rank GCMs. Previous analyses have not found a systematic differences like this, so additional analyses to determine the robustness of these results is needed.

Rupp et al. (2016) also analyzed which climatological processes caused differences across seasons in the magnitude of climate change. Consistent with previous theories, wetter winters resulted from deepening and southward extension of the Aleutian low pressure system, which pushes more warm moist air into the PNW from the Pacific. Drier summers were caused by northward expansion of the North Pacific subtropical high pressure system. The combination of warmer and drier summers reflects land-surface interactions and a change in anticyclonic circulation. The difference between summer and winter warming rates can be attributed to a change in the ratio of latent and sensible heat fluxes, which follow from the differences in precipitation, cloud cover, soil moisture content, and seasonal differences in short- and long-wave radiation flux.

9 Rupp et al. (2016) further found that interannual variability decreases for temperature in winter but increases for temperature in other seasons and increases for precipitation in all seasons. Internal variability in GCM projections is high in both temperature and precipitation. However, relative differences in the importance of different types of uncertainty emerge over time and when comparing different variables. Variation among GCM temperature projections stems from different global climate sensitivities in each model to a greater extent than for precipitation. These systematic differences between models increase the differential in projections over time. Variation in precipitation among GCMs, on the other hand, stems from primarily internal model variability.

The third Oregon Climate Assessment Report has been published, summarizing the CMIP5 projections for Oregon State (Dalton et al. 2017). Climate projections are summarized for different geographic regions within the state. The report discusses the expected impacts on ecosystems, forests, crops, human health and tribal interests.

California Current

Projections NOAA produced a new report on global and regional sea level rise for the United States (Sweet et al. 2017). The report summarized historical and projected changes in sea level changes under six scenarios that range from 0.3 to 2.5 m by 2100. Sea level along the northern British Columbian and southern Alaskan coast rose less than the global average, and in southern California sea level rose more than the global mean in all scenarios. Sea level rise in the PNW was intermediate, ranging from slightly less than the global average in optimistic scenarios to somewhat more in moderate and business-as-usual scenarios. The report provided examples of risk-based analyses using these scenarios.

Sea level rise involves delayed responses to greenhouse gases, such that the ocean will continue thermal expansion for hundreds of years after short-lived gases have left the atmosphere (Zickfeld et al. 2017). Therefore, actions to limit short-lived gas emissions will have larger benefits than previously recognized.

Sea level rise has complex impacts on coastal zones and estuaries. A rising ocean has the benefit of mitigating temperature increases in lower estuaries, as demonstrated in a case study of the Yaquina estuary in Oregon (Brown et al. 2016). An increase in air temperature of 3°C caused only 0.7-1.6°C warming, on average, in the Yaquina Estuary. Nonetheless, the maximum average daily temperature of the upper estuary exceeded

10 18°C for 7 d under the 3°C air temperature scenario and for 40 d under a 4°C air temperature scenario.

In the Skagit estuary, sea level rise and projected changes in river flow will increase the inundation during a 100-year by 57% during the 2040s and by 74% during the 2080s over the 1970-1999 average, based on an ECHAM-5 GCM A1B scenario (Hamman et al. 2016). Salinity is expected to increase in the nearshore by 1 psu, and the salinity intrusion will extend 3 km further upstream in the Skagit River (Khangaonkar et al. 2016).

EPA conducted a quantitative analysis of temperature sensitivity of the South Fork Nooksack River under climate projections, and found an acceleration of the probability of exceeding water quality standards (Butcher et al. 2016). The found a high probability of exceeding lethal thresholds by the 2040s under low-flow conditions. They found that riparian shading becomes increasingly important over time to mitigate high temperatures.

EPA also completed a qualitative report on the impact of climate change on recovery actions for the Nooksack River (Grah et al. 2016). The report includes a detailed discussion of life-stage specific impacts and matches these impacts to specific action items. They found that although the right types of actions had already been identified, barriers to implementation needed to be addressed for them to be effective.

As described in previous reviews, the direction and magnitude of future changes in upwelling are uncertain. Tim et al. (2016) conducted a new analysis of the sensitivity of upwelling to climate forcing using two earth system models. They found no significant trends in the historical (900-1849) or future (2006-2100) climate, except for a slight weakening in the California Current under the RCP 8.5 scenario (Tim et al. 2016). However, the physical links between large-scale wind patterns, upwelling, and productivity are still an active area of research. Renault et al. (2016) showed that primary productivity can become decoupled from upwelling winds due to eddies produced by alongshore currents.

Retrospective A series of retrospective analyses examined causes and characteristics of the unusual climatic conditions of 2015. In a special supplement to the Bulletin of the American Meterological Society, 28 papers explored global climate anomalies. All heat-related phenomena were shown to be more intense or more likely due to anthropogenic influences on the climate (Herring et al. 2016).

11 For example, the 2014/2015 “snowpack ” in Washington State occurred during a near-normal year in precipitation. The drought stemmed from unprecedented warmth over winter, which caused more winter precipitation to fall as rain rather than snow, which is a temperature effect (Fosu et al. 2016). The extreme drought in western Canada in 2015 reflected the combination of anthropogenically-influenced warm spring conditions with natural variation producing dry summer weather (Szeto 2016) (similar to the California drought). The 5.1 million acres that burned in Alaska, the second most on record, was 34-60% more likely due to anthropogenic climate change (Partain Jr. et al. 2016).

Record-low sea ice extent in March 2015 would not have been as extreme without long-term warming (Fučkar et al. 2016). Daily rainfall extremes in 2015, on the other hand, could not be formally attributed to changes in atmospheric gases, but are consistent with that explanation (Kam et al. 2016).

Jacox et al. (2016) analyzed the influence of an extraordinarily strong El Niño on the California Current ecosystem during 2015. They found that the 2015 El Niño did not have impacts on the ecosystem that were comparable with those of the last two strongest ENSO events (1982-1983 and 1997-1998). The 2015 event had reduced impacts because it did not drive strong upwelling-favorable winds or strong equatorial Kelvin waves. Thus the large biological anomalies of 2015 were driven mostly by longer-term sea surface temperature anomalies. Sea surface temperatures in 2014-2015 were part of a record-breaking 15 months of negative anomalies in coastal wind, which were not associated with El Niño (Robinson 2016).

Biological impacts of “the blob” temperature anomaly in the eastern Pacific from 2013 to 2015, were described by Cavole et al. (2016). For example, they list species that displayed mass strandings (3 birds), shifts in distribution (12 fish, 3 invertebrates, and 1 reptile), shifts in abundance (3 fish and krill), and other unusual records (6 fish, 1 invertebrate, 2 whales, 1 reptile). They discuss mechanisms and economic impacts of the blob as well. Economic impacts included fishery closures due to harmful algal blooms. Negative effects on pollock and salmon fisheries were expected but not yet determined.

The blob could be considered an example of a marine . Scannell et al. (2016) conducted a formal analysis of marine heatwaves since 1950. They found a significant influence of anthropogenic warming on the occurrence of these heatwaves.

12 Stream temperature and flow projections

Stream temperature has a profound impact on biological processes. Stream temperature projections for Puget Sound, Washington (Cao et al. 2016), the Umatilla basin of northeastern Oregon (DeBano et al. 2016), and the Navarro River in California (Woltemade and Hawkins 2016) will all be useful for conservation planning. Changes in land use and riparian vegetation in the Puget Sound could potentially reduce local impacts of rising global temperatures, and stream temperature projections provide support for proactive conservation actions (Cao et al. 2016).

Analyses of stream flow have been conducted nationally (Dhungel et al. 2016), in the Salmon River Basin of Idaho (Gould et al. 2016), and in the Tucannon (Praskievicz 2016) and Skagit River Basins of Washington (Lee et al. 2016). In California, stream flow analyses have also been conducted for the Sacramento River (Stern et al. 2016) and for some mid-elevation catchments in the southern mountains (Jepsen et al. 2016).

The national analysis focused on hydrograph regime changes, finding more extensive changes in the central and eastern U.S. than the west, where many snow-fed streams maintained their current hydrograph classifications (Dhungel et al. 2016).

In model projections for the Salmon River Basin, severe exacerbated the advance in the timing of peak spring flows (Gould et al. 2016). Large sediment flows associated with severe fires may pose management challenges. However, another study found that the processes relating frequency and severity to drought are spatially heterogeneous, suggesting that in many existing projections of fire dynamics, key assumptions need to be examined more carefully (McKenzie and Littell 2017).

Sediment flows are also projected to alter the hydrograph of the Tucannon River, amplifying climate change-induced seasonal variation (Praskievicz 2016). Under a range of A2 global and regional model combinations, 4-34% more flooding occurred in the winter and 5-47% lower flows occurred in the summer (Praskievicz 2016). Changes in hydropower production were modeled for the Skagit River (Lee et al. 2016). Low-flows in the Skagit decreased by 5-20%, but changes in extremes were less predictable than changes in average low-flows (Stumbaugh and Hamlet 2016).

A hydrological model of Sierra Nevada dynamics explored the sensitivity of the snow/rain ratio and low flows to soil characteristics across elevation gradients (Jepsen et al. 2016). A model of the Sacramento River explored changes in sediment flow and flood risk with events (Stern et al. 2016).

13 Ocean acidification

Projections Eleven species groups in the California Current are consistently expected to respond negatively to ocean acidification (Busch and McElhany 2016). When concurrent effects of low pH and low dissolved oxygen were examined experimentally, most studies found additive negative effects, but synergistic negative effects were also demonstrated (Gobler and Baumann 2016). Further work is needed before we can predict when synergistic responses will occur, so these effects might be underestimated in current projections. Somero et al. (2016) provide a general overview of the interacting effects of multiple stressors.

A full scale end-to-end examination using the Atlantis model explored the impact of ocean acidification on the California Current ecosystem (Marshall et al. 2017). Species groups that showed direct negative effects of OA, such as epibenthic invertebrates (crabs, shrimps, benthic grazers, benthic detritivores, bivalves) experienced the largest declines. Their predators, including some demersal fish, sharks, and Dungeness crab also suffered because of declines of their prey.

With regard to human impacts, the crab fishery was most hurt in OA scenarios. The model was unable to reproduce realistic salmon biomass in calibration runs, so impacts on salmon were not reported. Relatively small effects were experienced by other functional groups.

In an examination of the effects of climate change in the Arctic found that although net benefits of warming are expected for fishermen, ocean acidification reduces the magnitude of this benefit (Lam et al. 2016).

Retrospective analyses An analysis of ocean acidification in the California Current found that from 1979 through 2012, pH decreased by -0.02/decade and aragonite saturation decreased by -0.12/decade in the top 60 m, averaged across the whole spatial domain-wide (Turi et al. 2016). Acidification at the domain scale was attributed largely to the increase in

atmospheric CO2 levels. However, spatial heterogeneity in acidification rates within the California Current ecosystem was caused by wind anomalies, which affected upwelling rates. The nearshore areas of northern California and Oregon acidified much more than elsewhere in the domain because of the changes in along-shore winds.

Aragonite saturation occurred at shallower depths than it has historically, with changes in aragonite saturation depth of up to -50 m per decade in the nearshore ocean.

14 This change compresses suitable biotic habitat into a much smaller area than was the case 40 years ago. A separate analysis of ocean acidification in the California Current produced similar results, adding that pteropod shell dissolution increased approximately 19-26% in both nearshore and offshore waters since pre-industrial times (Feely et al. 2016).

Ecological impacts of ocean acidification Multiple literature reviews outlined the mechanisms by which ocean acidification affects ocean ecosystems (Mostofa et al. 2016) and found little support for a simplistic conceptual model such as aerobic scope limitations for all stressors (Lefevre 2016). Other reviews summarized current information on species-specific responses to pH.

One novel result found an impact of pH on hearing and the marine soundscape

(Rossi et al. 2016). Biological sound production was diminished in natural CO2 vents, which were explored as an analogue for future conditions. Furthermore, fish did not

respond normally to those sound levels after they had been exposed to high CO2 conditions. Larval fish use sound to orient during dispersal and find suitable habitat in which to settle, so OA might interfere with these key recruitment processes.

For the California Current, negative effects of low pH on kelp growth rates (Britton et al. 2016) and krill metabolic rates (Cooper et al. 2016) are relevant, as are numerous studies of negative effects of low pH to fish such as silversides (Miller et al. 2016), cod (Stiasny et al. 2016), and sand smelt (Lopes et al. 2016; Silva et al. 2016).

Various behavioral studies found effects of low pH on fish schooling (Lopes et al. 2016), olfactory, and predator avoidance behavior (Nagelkerken et al. 2016). However, as pointed out by Wahl et al. (2016), the utility of laboratory studies for predicting effects in the natural environment is limited when natural variation in pH is ignored.

15 Biological Responses to Climate Change

Species‐level impacts

Projections Sharma et al.(2016) expect Pacific lamprey and Pacific eulachon to be negatively impacted by climate change. Specifically, higher late fall and winter stream flows are expected to have a negative impact on lamprey but possibly a positive impact on eulachon. Higher ocean temperatures and reduced upwelling would have negative impacts on both species. Cumulative effects will be most severe for lamprey in the interior Columbia.

Retrospective analyses Climate-related local extinctions the affect all taxonomic groups across the globe have been documented by Wiens (2016). Of 976 species surveyed, 47% exhibited evidence of climate-related losses. Extinctions were most common at the warm edges of the range, causing range contractions or shifts, while just 6% of species examined showed range expansions. Effects have been more severe in freshwater than terrestrial or marine habitats (74 vs. 46 and 51%, respectively). Freshwater habitats might be more vulnerable due to their greater sensitivity to drought or variability in precipitation and limited options for vertical range shifts relative to marine species. Approximately 60% of the 69 fish species examined included populations that went locally extinct with a climate attribution.

Stage‐specific impacts in salmon and their ecosystems

Juvenile freshwater stages: temperature and growth Temperature and flow affect freshwater juvenile growth rates, which affect survival in later stages through size-selective mortality and behavioral consequences, and juvenile and adult survival. Studies published in 2016 focused on the importance of temperature at critical stages for egg development and juvenile growth. They identified effects measured at later life stages that could be pinpointed to the timing of development and growth.

Specifically, Thompson et al. (2016) found that for hatchery-reared winter steelhead smolts, spawn date was a significant predictor of smolt size at the time of release one year later. This smolt size effect was attributed to slower development of

16 early spawned eggs, which occurred because the hatchery uses temperature to synchronize emergence date across families. The significance of this finding for climate change analyses is that warmer temperatures will drive faster development, which is more akin to the later-spawned eggs here.

This experiment implies that even under similar juvenile growth conditions, the history of warmer winter/spring development temperatures will result in smaller smolts. This result is consistent with previous experimental work showing effects of development temperature on yolk absorption rates and fry size. However, in this case, the implications extend all the way to the smolt stage, one year later.

By examining growth patterns in different age groups of Skagit River steelhead in the wild, Thompson and Beauchamp (2016) found evidence for size-selective survival and identified how improved growth opportunities at critical stages could improve survival.

Adult freshwater stages: temperature and disease effects on upstream migrants A major concern for upstream migrant survival in the Columbia and Snake Rivers is the risk incurred from high water temperature. Based on data from archival temperature tags, Keefer and Caudill (2016) summarized thermal exposure for summer steelhead and fall-run Chinook through four dams in the lower Snake River. They found that annual maxima were higher at a site just downstream from Ice Harbor Dam (21.4-22.4°C) compared with Lower Granite Dam (19.9-20.2°C).

Some fish experienced temperatures higher than those recorded at either temperature monitoring site (24°C). Most logger-tagged fish from both species encountered the warmest temperatures inside dam fishways or in reservoirs just upstream from fishway exits. They demonstrated that both species used a cool-water site near a tributary to cool their body temperatures, but Chinook stayed in these sites for less than one day, whereas steelhead might stay for weeks. Steelhead typically traverses this route more slowly than Chinook, which corresponds to higher average heat loads.

In an experiment designed to evaluate the potential benefit of cool-water holding for upstream migrants, Benda et al. (2016) found that experimental holding in cold water lowered prespawn mortality in Willamette spring Chinook, but resulted in a different pathogen profile compared with in-river fish. This result is consistent with generally higher risks of disease exposure when fish are constrained, yet there remained an overriding benefit of reduced mortality.

Marine survival

17 Marine survival has been a major “black box” for salmon life cycle modeling because of the complexity of ocean ecosystems. A variety of approaches have been employed to understand and model this life stage. A useful review of modelling approaches and the extent to which they integrate environmental drivers was published by Koenigstein et al. (2016).

In life cycle models of Columbia River salmon, we have primarily relied on statistical analyses of the correlation between standard indices of abiotic conditions (e.g., sea surface temperature, upwelling indices, etc.) and smolt-to-adult return rates. This is the first annual literature review in which we found no new correlation analyses of this sort specifically for salmon, although we include some studies on salmon prey, such as rockfish and anchovies.

Rockfish larvae are an important prey item and indicator of juvenile salmon survival (Daly et al. 2013), making the physical factors that influence rockfish indirectly relevant for salmon. An analysis of black rockfish and kelp greenling found that cool basin-scale conditions of indices such as the PDO, NPGO, and ENSO favored rockfish growth in the California Current (von Biela et al. 2016). However, alternating warm and cool years were more beneficial for benthic-feeding kelp greenlings. Warm conditions were more favorable for both species in the Alaska Coastal Current system. These conditions are similar to the physical conditions that are beneficial for salmon.

Another analysis explored the mechanistic basis of the well-known climate-driven oscillation between sardine and anchovies. Species-interaction models showed that differences in the prey base shared by the two species could explain these fluctuations (Canales et al. 2016). Sardines are better at foraging on smaller prey, so they have an advantage when the prey base consists of plankton with smaller body sizes. However, anchovy thrive when larger plankton are available, which ends up hurting sardines because they become prey to the anchovies (Canales et al. 2016).

Two studies measured salmon responses to oceanographic and community characteristics directly. Sabal et al. (2016) related juvenile catch data to hatchery release information and oceanographic conditions for Central Valley fall Chinook. They found that the best ocean predictors of both growth and survival included diatoms, predatory zooplankton, temperature and currents. However, each of these factors had opposite effects on growth and survival, such that growth was highest during periods of low survival.

This result indicates either a density-dependent relationship between growth and survival (fish grow better when there are fewer of them), or stronger size-selective mortality during periods of low survival than high survival. Highest survival occurred

18 during oceanographic relaxation events, which occur in between upwelling events. These periods might be the most favorable for salmon because of pre-conditioning by upwelling, which brings nutrients to the surface, followed by weaker currents, which retain prey nearshore, reduce salmon energetic requirements, and produce warmer temperatures that favor salmon growth.

Consistent with these results, Wells et al. (2016) published a conceptual model supported by an individual-based-model of the bottom-up drivers of central Californian Chinook salmon growth and survival during the early marine stage. Wells et al. (2016) further link the relaxation conditions identified by Sabal et al. (2016) with the strength and location of the North Pacific High pressure system in winter.

The second salmon study found a significant negative relationship between jellyfish biomass (specifically, the sea nettle Chrysaora fuscescens) and smolt-to-adult returns of coho and Chinook salmon off Oregon and Washington during 1999-2012 (Ruzicka et al. 2016b). The spatial distribution of sea nettles and juvenile salmon catch within years was also correlated, and salmon stomachs were less full at locations with greater overlap with sea nettles. These results are consistent with indirect competition for zooplankton between jellyfish and salmon.

Consistent with most previous literature, both salmon-specific papers in 2016 reported a combination of bottom-up and top-down regulation of marine survival in salmon. Ecosystem models have the benefit of incorporating both types of pressures. Ruzicka et al. (2016b) were essentially testing one of the predictions from their ecosystem model, which was that among all trophic interactions within the Pacific Northwest coastal food web, juvenile salmon are particularly sensitive to jellyfish blooms. Here we include a discussion of ecosystem models of the California Current published in 2016.

End-to-end models attempt to mechanistically link atmospheric and physical drivers through ocean ecosystem processes all the way to top predators and fisheries. End-to-end models for the California Current explored how different upwelling scenarios affected productivity (Ruzicka et al. 2016a), how environmental variability affected sea lion foraging ecology (Fiechter et al. 2016), and how ocean acidification changed fisheries revenue (Marshall et al. 2017).

Ruzicka et al. (2016a) described results from an end-to-end model of the Northern California Current Ecosystem. The model integrates physical, trophic, and nutrient cycling processes. They found that productivity of four indicator groups (copepods, forage fishes, invertebrate epifuana, and semi-demersal fishes) showed an intermediate optimum upwelling rate. Productivity increased from low to moderate upwelling rates

19 because higher nutrient supply increased primary and consumer production.

However, when upwelling rates increased the offshore transport of plankton faster than they could be eaten, productivity declined. Compared to the 9-d mean duration of upwelling events observed during 1967-2014, shorter events increased productivity of all groups by reducing export of plankton, with an optimum duration of 3 d. Ruzicka et al. (2016a) also found a shift in plankton community composition toward larger diatoms (compared with flagellates) with longer upwelling events.

Climate vulnerability assessments

NOAA-Fisheries published results from the first regional vulnerability assessment of the impact of climate change in the northeastern U.S. (Hare et al. 2016). Atlantic salmon scored very high in both climate exposure and species sensitivity, as did just one other species (bay scallop). Other anadromous fish (American and hickory shad, sturgeon and smelt also scored very high in climate exposure, and high in sensitivity.

Recent work in the general ecological literature has encouraged consideration in vulnerability assessments of the potential for adaptive capacity to alter predicted effects of climate change. Adaptive capacity can include either plastic (non-genetic) or genetic responses to climate change. Theoretical analyses suggest that some general principles might determine the relative roles of plasticity and adaptation along latitudinal clines due to systematic variation in the variability and predictability of environmental conditions (Boyd et al. 2016). Invertebrate assemblages from Singapore to Antarctica showed the predicted patterns in heat tolerance and acclimation ability (Morley et al. 2016).

The general expectation is that adaptive capacity will increase resilience beyond what is assumed by niche models using observed distributions (Beever et al. 2016). However, the opposite result might also be true, and might be more typical for salmon. Genetic variation has already been depleted in most listed populations, so adaptive capacity is more likely to reduce resilience than increase it. For example, local adaptation to climate at the southern range edge of steelhead has been greatly depleted by introgression from introduced hatchery rainbow trout (Abadia-Cardoso et al. 2016).

However, consistency in scientific results regarding adaptive capacity is still low. Wade et al. (2016) compared the rankings of steelhead and bull trout in the Columbia River Basin, and found that the specific methodology used to quantify adaptive capacity significantly affected rankings.

20 In Columbia Basin steelhead, climate variables are correlated with genetic differentiation within metapopulations (Hand et al. 2016). However, different metapopulations had different environment-genetics relationships, so it is not possible to extrapolate particular relationships to larger spatial scales (Hand et al. 2016). Rapid evolution to environmental conditions within lineages is well established in salmon. This was further supported by an analysis showing the summer-run ecotype of steelhead evolved many times in parallel in Oregon and N-Cal (Arciniega et al. 2016).

For marine ectotherms, metabolic demands of a warmer climate favor a faster life cycle characterize by earlier age at maturation and smaller body sizes (Waples and Audzijonyte 2016). This is qualitatively the same direction of response already imposed by the fishing industry (fishery-induced evolution), and many fish populations have altered their size structure and age at maturity in recent decades (Audzijonyte et al. 2013).

Thus, heavily-fished populations already display the characteristics expected to become much more widespread in marine species. Some of these populations have failed to recover after fishing has stopped, possibly due to evolutionary effects. It will be important for fishery management to account for the increased boom-bust population cycles associated with faster life histories.

21 Literature Cited

Abadia-Cardoso, A., and coauthors. 2016. Population genetic structure and ancestry of steelhead/rainbow trout (Oncorhynchus mykiss) at the extreme southern edge of their range in North America. Conservation Genetics 17(3):675-689. Arciniega, M., and coauthors. 2016. Parallel evolution of the summer steelhead ecotype in multiple populations from Oregon and Northern California. Conservation Genetics 17(1):165-175. Audzijonyte, A., A. Kuparinen, and E. A. Fulton. 2013. How fast is fisheries-induced evolution? Quantitative analysis of modelling and empirical studies. Evolutionary Applications 6(4):585-595. Ault, T. R., J. S. Mankin, B. I. Cook, and J. E. Smerdon. 2016. Relative impacts of mitigation, temperature, and precipitation on 21st-century megadrought risk in the American Southwest. Science Advances 2(10). Barth, N. A., G. Villarini, M. Nayak, and K. White. 2016. Mixed populations and annual flood frequency estimates in the western United States: The role of atmospheric rivers. Water Resources Research:257-269. Beever, E. A., and coauthors. 2016. Improving conservation outcomes with a new paradigm for understanding species' fundamental and realized adaptive capacity. Conservation Letters 9(2):131-137. Benda, S. E., G. P. Naughton, C. C. Caudill, M. L. Kent, and C. B. Schreck. 2016. Cool, pathogen-free refuge lowers pathogen-associated prespawn mortality of Willamette River chinook salmon. Transactions of the American Fisheries Society 145(6):1159-1172. Boyd, P. W., and coauthors. 2016. Biological responses to environmental heterogeneity under future ocean conditions. Global Change Biology 22(8):2633-2650. Britton, D., C. E. Cornwall, A. T. Revill, C. L. Hurd, and C. R. Johnson. 2016. Ocean acidification reverses the positive effects of seawater pH fluctuations on growth and photosynthesis of the habitat-forming kelp, Ecklonia radiata. Scientific Reports 6:10. Brown, C. A., D. Sharp, and T. C. M. Collura. 2016. Effect of climate change on water temperature and attainment of water temperature criteria in the Yaquina Estuary, Oregon (USA). Estuarine Coastal and Shelf Science 169:136-146. Busch, D. S., and P. McElhany. 2016. Estimates of the direct effect of seawater pH on the survival rate of species groups in the California Current Ecosystem. Plos One 11(8):28. Butcher, J. B., M. Faizullabhoy, H. Nicholas, P. Cada, and J. T. Kennedy. 2016. Quantitative assessment of temperature sensitivity of the South Fork Nooksack River under future climates using QUAL2Kw. U.S. Environmental Protection Agency. Canales, T. M., R. Law, and J. L. Blanchard. 2016. Shifts in plankton size spectra modulate growth and coexistence of anchovy and sardine in upwelling systems. Canadian Journal of Fisheries and Aquatic Sciences 73(4):611-621. Cao, Q., N. Sun, J. Yearsley, B. Nijssen, and D. P. Lettenmaier. 2016. Climate and land cover effects on the temperature of Puget Sound streams. Hydrological Processes 30(13):2286-2304. Cavole, L. M., and coauthors. 2016. Biological impacts of the 2013-2015 warm-water anomaly in the Northeast Pacific. Oceanography 29(2):273-285. Cooper, H. L., D. C. Potts, and A. Paytan. 2016. Metabolic responses of the North Pacific krill, Euphausia pacifica, to short- and long-term pCO(2) exposure. Marine Biology 163(10):13.

22 Dalton, M. M., K. D. Dello, L. Hawkins, P. W. Mote, and D. E. Rupp. 2017. The third Oregon climate assessment report. Oregon Climate Change Research Institute, Oregon State University. Daly, E. A., T. D. Auth, R. D. Brodeur, and W. T. Peterson. 2013. Winter ichthyoplankton biomass as a predictor of early summer prey fields and survival of juvenile salmon in the northern California Current. Marine Ecology Progress Series 484:203-217. DeBano, S. J., D. E. Wooster, J. R. Walker, L. E. McMullen, and D. A. Horneck. 2016. Interactive influences of climate change and agriculture on aquatic habitat in a Pacific Northwestern watershed. Ecosphere 7(6):27. Dhungel, S., D. G. Tarboton, J. Jin, and C. P. Hawkins. 2016. Potential effects of climate change on ecologically relevant streamflow regimes. River Research and Applications 32(9):1827-1840. EPA, U.S. Environmental Protection Agency. 2016. Climate change indicators in the United States. U.S. Environmental Protection Agency. Feely, R. A., and coauthors. 2016. Chemical and biological impacts of ocean acidification along the west coast of North America. Estuarine Coastal and Shelf Science 183:260-270. Fiechter, J., L. A. Huckstadt, K. A. Rose, and D. P. Costa. 2016. A fully coupled ecosystem model to predict the foraging ecology of apex predators in the California Current. Marine Ecology Progress Series 556:273-285. Fosu, B. O., S.-Y. S. Wang, and J.-H. Yoon. 2016. The 2014/15 snowpack drought in Washington State and its climate forcing. Bulletin of the American Meteorological Society 97(12). Fučkar, N. S., and coauthors. 2016. Record low northern hemisphere sea ice extent in March 2015. Bulletin of the American Meteorological Society 97(12). Gobler, C. J., and H. Baumann. 2016. Hypoxia and acidification in ocean ecosystems: coupled dynamics and effects on marine life. Biology Letters 12(5):8. Gould, G. K., and coauthors. 2016. The effects of climate change and extreme wildfire events on runoff over a mountain watershed. Journal of Hydrology 536:74-91. Grah, O., and coauthors. 2016. Qualitative Assessment: Evaluating the impacts of climate change on endangered species act recovery actions for the South Fork Nooksack River, WA. Hamman, J. J., A. F. Hamlet, S. Y. Lee, R. Fuller, and E. E. Grossman. 2016. Combined effects of projected sea level rise, , and peak river flows on water levels in the Skagit floodplain. Northwest Science 90(1):57-78. Hand, B. K., and coauthors. 2016. Climate variables explain neutral and adaptive variation within salmonid metapopulations: the importance of replication in landscape genetics. Molecular Ecology 25(3):689-705. Hare, J. A., and coauthors. 2016. A vulnerability assessment of fish and invertebrates to climate change on the Northeast U.S. Continental Shelf. Plos One 11(2):e0146756. Herring, S., and coauthors. 2016. Explaining extreme events of 2015 from a climate perspective. Bulletin of the American Meteorological Society 97(12). Jacox, M. G., and coauthors. 2016. Impacts of the 2015-2016 El Nino on the California Current System: Early assessment and comparison to past events. Geophysical Research Letters 43(13):7072-7080. Jepsen, S. M., T. C. Harmon, M. W. Meadows, and C. T. Hunsaker. 2016. Hydrogeologic influence on changes in snowmelt runoff with climate warming: Numerical experiments on a mid-elevation catchment in the Sierra Nevada, USA. Journal of Hydrology 533:332-342. Kam, J., T. R. Knutson, F. Zeng, and A. T. Wittenberg. 2016. Multimodel assessment of anthropogenic influence on record global and regional warmth during 2015.

23 Bulletin of the American Meteorological Society 97(12). Keefer, M. L., and C. C. Caudill. 2016. Estimating thermal exposure of adult summer steelhead and fall Chinook salmon migrating in a warm impounded river. Ecology of Freshwater Fish 25(4):599-611. Khangaonkar, T., W. Long, B. Sackmann, T. Mohamedali, and A. F. Hamlet. 2016. Sensitivity of circulation in the Skagit River Estuary to sea level rise and future flows. Northwest Science 90(1):94-118. Koenigstein, S., F. C. Mark, S. Gossling-Reisemann, H. Reuter, and H. O. Poertner. 2016. Modelling climate change impacts on marine fish populations: process- based integration of ocean warming, acidification and other environmental drivers. Fish and Fisheries 17(4):972-1004. Lam, V. W. Y., W. W. L. Cheung, and U. R. Sumaila. 2016. Marine capture fisheries in the Arctic: Winners or losers under climate change and ocean acidification? Fish and Fisheries 17(2):335-357. Lee, S. Y., A. F. Hamlet, and E. E. Grossman. 2016. Impacts of climate change on regulated streamflow, hydrologic extremes, hydropower production, and sediment discharge in the Skagit River Basin. Northwest Science 90(1):23-43. Lefevre, S. 2016. Are global warming and ocean acidification conspiring against marine ectotherms? A meta-analysis of the respiratory effects of elevated temperature, high CO2 and their interaction. Conservation Physiology 4:31. Lopes, A. F., and coauthors. 2016. Behavioural lateralization and shoaling cohesion of fish larvae altered under ocean acidification. Marine Biology 163(12):10. Marshall, K. N., and coauthors. 2017. Risks of ocean acidification in the California Current food web and fisheries: ecosystem model projections. Global Change Biology:1-15. McKenzie, D., and J. Littell. 2017. Climate change and the eco-hydrology of fire: Will area burned increase in a warming western USA? Ecological Applications 27(1):26-36. Miller, S. H., D. L. Breitburg, R. B. Burrell, and A. G. Keppel. 2016. Acidification increases sensitivity to hypoxia in important forage fishes. Marine Ecology Progress Series 549:1-8. Morley, S. A., and coauthors. 2016. Rates of warming and the global sensitivity of shallow water marine invertebrates to elevated temperature. Journal of the Marine Biological Association of the United Kingdom 96(1):159-165. Mostofa, K. M. G., and coauthors. 2016. Reviews and Syntheses: Ocean acidification and its potential impacts on marine ecosystems. Biogeosciences 13(6):1767-1786. Mote, P. W., J. Abatzoglou, and K. E. Kunkel. 2013. Variability and change in the past and the future. Pages 25-40 in M. M. Dalton, P. W. Mote, and A. K. Snover, editors. Climate Change in the Northwest: Implications for Our Landscapes, Waters, and Communities. Island Press, Washington, DC. Nagelkerken, I., B. D. Russell, B. M. Gillanders, and S. D. Connell. 2016. Ocean acidification alters fish populations indirectly through habitat modification. Nature Climate Change 6:89-95. NOAA, National Centers for Environmental Information. 2017. State of the climate: Global climate report for annual 2016. retrieved on July 5, 2017 from https://www.ncdc.noaa.gov/sotc/global/201613. Partain Jr., J. L., Alden, Sharon, Bhatt, Uma S., Bieniek, Peter A., Brettschneider, Brian R.,, R. T. Lader, Olsson, Peter Q., Rupp, T. Scott, Strader, Heidi, Thoman Jr., Richard L.,, and J. E. Walsh, York, Alison D., and Ziel, Robert H. . 2016. An assessment of the role of anthropogenic climate change in the Alaska fire season of 2015. Bulletin of the American Meteorological Society 97(12). Praskievicz, S. 2016. Impacts of projected climate changes on streamflow and sediment transport for three snowmelt-dominated rivers in the interior Pacific Northwest. River Research and Applications 32(1):4-17.

24 Rana, A., and H. Moradkhani. 2016. Spatial, temporal and frequency based climate change assessment in Columbia River Basin using multi downscaled-scenarios. Climate Dynamics 47(1-2):579-600. Renault, L., and coauthors. 2016. Partial decoupling of primary productivity from upwelling in the California Current system. Nature Geoscience 9(7):505-510. Robinson, C. J. 2016. Evolution of the 2014-2015 sea surface temperature warming in the central west coast of Baja California, Mexico, recorded by remote sensing. Geophysical Research Letters 43(13):7066-7071. Rossi, T., I. Nagelkerken, J. C. A. Pistevos, and S. D. Connell. 2016. Lost at sea: Ocean acidification undermines larval fish orientation via altered hearing and marine soundscape modification. Biology Letters 12(1):4. Rupp, D. E., J. T. Abatzoglou, and P. Mote. 2016. Projections of 21st century climate of the Columbia River Basin. Climate Dynamics:doi: 10.1007/s00382-016-3418-7. Ruzicka, J. J., K. H. Brink, D. J. Gifford, and F. Bahr. 2016a. A physically coupled end- to-end model platform for coastal ecosystems: Simulating the effects of climate change and changing upwelling characteristics on the Northern California Current ecosystem. Ecological Modelling 331:86-99. Ruzicka, J. J., E. A. Daly, and R. D. Brodeur. 2016b. Evidence that summer jellyfish blooms impact Pacific Northwest salmon production. Ecosphere 7(4):22. Sabal, M. C., and coauthors. 2016. Contrasting patterns in growth and survival of Central Valley fall run Chinook salmon related to hatchery and ocean conditions. Environmental Biology of Fishes 99(12):949-967. Scannell, H. A., A. J. Pershing, M. A. Alexander, A. C. Thomas, and K. E. Mills. 2016. Frequency of marine heatwaves in the North Atlantic and North Pacific since 1950. Geophysical Research Letters 43(5):2015GL067308. Sharma, R., D. Graves, A. Farrell, and N. Mantua. 2016. Investigating freshwater and ocean effects on Pacific lamprey and Pacific eulachon of the Columbia River Basin: Projections within the context of climate change. Columbia River Inter- Tribal Fish Commission, Portland, OR. Technical Report 16-05, 139p. Shields, C. A., and J. T. Kiehl. 2016. Atmospheric river landfall-latitude changes in future climate simulations. Geophysical Research Letters 43(16):8775-8782. Silva, C. S. E., and coauthors. 2016. Effects of ocean acidification on the swimming ability, development and biochemical responses of sand smelt larvae. Science of the Total Environment 563:89-98. Somero, G. N., and coauthors. 2016. What changes in the carbonate system, oxygen, and temperature portend for the northeastern Pacific Ocean: A physiological perspective. Bioscience 66(1):14-26. Stern, M., L. Flint, J. Minear, A. Flint, and S. Wright. 2016. Characterizing changes in streamflow and sediment supply in the Sacramento River basin, California, using hydrological simulation program-fortran (hspf). Water 8(10):21. Stiasny, M. H., and coauthors. 2016. Ocean acidification effects on Atlantic cod larval survival and recruitment to the fished population. Plos One 11(8):11. Stumbaugh, M., and A. F. Hamlet. 2016. Effects of climate change on extreme low-flows in small lowland tributaries in the Skagit River Basin. Northwest Science 90(1):44-56. Sweet, W. V., and coauthors. 2017. Global and regional sea level rise scenarios for the United States. National Oceanic and Atmospheric Administration NOAA Technical Report NOS CO-OPS 083. Szeto, K., Zhang, Xuebin, White, Robert Edward, and Brimelow, Julian 2016. The 2015 extreme drought in Western Canada. Bulletin of the American Meteorological Society 97(12). Thompson, J. N., and D. A. Beauchamp. 2016. Growth of juvenile steelhead Oncorhynchus mykiss under size-selective pressure limited by seasonal bioenergetic and environmental constraints. Journal of Fish Biology 89(3):1720-

25 1739. Thompson, N. F., M. R. Christie, M. L. Marine, L. D. Curtis, and M. S. Blouin. 2016. Spawn date explains variation in growth rate among families of hatchery reared Hood River steelhead (Oncorhynchus mykiss). Environmental Biology of Fishes 99(6-7):581-591. Tim, N., E. Zorita, B. Hunicke, X. Yi, and K. C. Emeis. 2016. The importance of external climate forcing for the variability and trends of coastal upwelling in past and future climate. Ocean Science 12(3):807-823. Turi, G., Z. Lachkar, N. Gruber, and M. Nunnich. 2016. Climatic modulation of recent trends in ocean acidification in the California Current System. Environmental Research Letters 11(1):17. van der Wiel, K., S. B. Kapnick, and G. A. Vecchi. 2016. Shifting patterns of mild weather in response to projected . Climatic Change:649-658. Vano, J. A., J. B. Kim, D. E. Rupp, and P. W. Mote. 2015a. Selecting climate change scenarios using impact-relevant sensitivities. Geophysical Research Letters 42(13):5516-5525. Vano, J. A., B. Nijssen, and D. P. Lettenmaier. 2015b. Seasonal hydrologic responses to climate change in the Pacific Northwest. Water Resources Research 51(4):1959- 1976. von Biela, V. R., and coauthors. 2016. Influence of basin- and local-scale environmental conditions on nearshore production in the northeast Pacific Ocean. Marine and Coastal Fisheries 8(1):502-521. Wade, A. A., and coauthors. 2016. Accounting for adaptive capacity and uncertainty in assessments of species' climate-change vulnerability. Conservation Biology 31(1):136-149. Wahl, M., V. Saderne, and Y. Sawall. 2016. How good are we at assessing the impact of ocean acidification in coastal systems? Limitations, omissions and strengths of commonly used experimental approaches with special emphasis on the neglected role of fluctuations. Marine and Freshwater Research 67(1):25-36. Waples, R. S., and A. Audzijonyte. 2016. Fishery-induced evolution provides insights into adaptive responses of marine species to climate change. Frontiers in Ecology and the Environment 14(4):217-224. Wells, B. K., and coauthors. 2016. Marine ecosystem perspectives on Chinook salmon recruitment: A synthesis of empirical and modeling studies from a California upwelling system. Marine Ecology Progress Series 552:271-284. Wiens, J. J. 2016. Climate-related local extinctions are already widespread among plant and animal species. PLOS Biology 14(12):e2001104. Woltemade, C. J., and T. W. Hawkins. 2016. Stream temperature impacts because of changes in air temperature, land cover and stream discharge: Navarro River Watershed, California, USA. River Research and Applications 32(10):2020-2031. Zickfeld, K., S. Solomon, and D. M. Gilford. 2017. Centuries of thermal sea-level rise due to anthropogenic emissions of short-lived greenhouse gases. Proceedings of the National Academy of Sciences of the United States of America 114(4):657- 662.

26