RIVER RESEARCH AND APPLICATIONS River Res. Applic. 29: 939–960 (2013) Published online 3 July 2012 in Wiley Online Library (wileyonlinelibrary.com) DOI: 10.1002/rra.2590 RESTORING SALMON HABITAT FOR A CHANGING CLIMATE T. BEECHIEa*, H. IMAKIa, J. GREENEa, A. WADEb, H. WUc,d, G. PESSa, P. RONIa, J. KIMBALLe, J. STANFORDe, P. KIFFNEYa AND N. MANTUAf a b Northwest Fisheries Science Center, National Marine Fisheries Service, Seattle, Washington, USA National Center for Ecological Analysis and Synthesis, University of California Santa Barbara, Santa Barbara, California, USA c Earth System Science Interdisciplinary Center, University of Maryland, College Park, Maryland, USA d NASA Goddard Space Flight Center, Greenbelt, Maryland, USA e Flathead Lake Biological Station, University of Montana, Missoula, Montana, USA f School of Aquatic and Fishery Sciences, University of Washington, Seattle, Washington, USA ABSTRACT An important question for salmon restoration efforts in the western USA is ‘How should habitat restoration plans be altered to accommodate climate change effects on stream flow and temperature?’ We developed a decision support process for adapting salmon recovery plans that incorporates (1) local habitat factors limiting salmon recovery, (2) scenarios of climate change effects on stream flow and temperature, (3) the ability of restoration actions to ameliorate climate change effects, and (4) the ability of restoration actions to increase habitat diversity and salmon population resilience. To facilitate the use of this decision support framework, we mapped scenarios of future stream flow and temperature in the Pacific Northwest region and reviewed literature on habitat restoration actions to determine whether they ameliorate a climate change effect or increase life history diversity and salmon resilience. Under the climate change scenarios considered here, summer low flows decrease by 35–75% west of the Cascade Mountains, maximum monthly flows increase by 10–60% across most of the region, and stream temperatures increase between 2 and 6 C by 2070–2099. On the basis of our literature review, we found that restoring floodplain connectivity, restoring stream flow regimes, and re-aggrading incised channels are most likely to ameliorate stream flow and temperature changes and increase habitat diversity and population resilience. By contrast, most restoration actions focused on in-stream rehabilitation are unlikely to ameliorate climate change effects. Finally, we illustrate how the decision support process can be used to evaluate whether climate change should alter the types or priority of restoration actions in a salmon habitat restoration plan. Copyright © 2012 John Wiley & Sons, Ltd. key words: restoration; climate change; decision support; adaptation; salmon habitat; stream flow; stream temperature Received 24 April 2012; Accepted 30 May 2012 INTRODUCTION Climate change is predicted to have significant effects on Pacific salmon and their ecosystems in western North America, and several reports suggest that restoring habitats for salmon in some places may be pointless because climate change will make their habitats inhospitable (Lackey, 2003; Nelitz et al., 2007). By contrast, recent modelling of the combined effects of climate change and habitat restoration indicates restoration actions are likely to result in a net benefit to salmon populations despite future shifts in temperature and hydrology (Battin et al., 2007). This lack of consensus on how climate change will affect salmon populations inhibits the development of clear guidance on how to modify habitat restoration efforts in response to climate change. With millions of dollars spent each year to restore habitats for threatened and endangered salmon in the western USA, there is increasing concern that climate change *Correspondence to: T. Beechie, Northwest Fisheries Science Center, 2725 Montlake Blvd. E., Seattle, WA 98112, USA. E-mail: tim.beechie@noaa.gov Copyright © 2012 John Wiley & Sons, Ltd. effects on freshwater habitats may limit the future effectiveness of certain salmon recovery efforts (Lackey et al., 2006; Battin et al., 2007; Mantua et al., 2010) and that the priority or design of specific restoration actions should be altered to accommodate future climate change (Mote et al., 2003). Making the decision to adapt a restoration plan for climate change is not straightforward, as predicted climate change effects vary widely throughout the Pacific salmon range, and some species have life histories that will likely allow them to persist throughout most of their range despite shifts in temperature and precipitation (Waples et al., 2009). Stream temperatures are expected to increase in most rivers, and the threat to salmon recovery is high where temperatures are near lethal or sub-lethal thresholds for salmon, but low in many rivers with current temperatures well below those thresholds. Furthermore, some rivers are expected to see large increases in peak flows, whereas other rivers are expected to experience decreased low flows (Arnell, 1999, Mantua et al., 2010). However, past land uses and water abstraction have often degraded habitats to a greater degree than that predicted from climate change, presenting substantial 940 T. BEECHIE ET AL. opportunities to improve salmon habitats more than enough to compensate for expected climate change effects over the next several decades (Battin et al., 2007). For salmonids, variation in life history strategies and habitat requirements—along with their demonstrated capacity to adapt to new environments— further complicates the development of general guidance for modifying restoration strategies to accommodate climate change (Quinn, 2005; Beechie et al., 2006; Bryant, 2009). This complex interplay of climate effects, restoration opportunities, and potential salmon responses poses a considerable challenge for effectively restoring salmon populations in a changing climate. In this paper, we present a simple logic framework and data sets to assist managers in adapting salmon habitat restoration efforts to climate change in the Pacific Northwest, USA (PNW). Our approach consists of four components: (1) a set of guiding questions that serve as a starting point for evaluating the potential effects of climate change on freshwater habitat restoration effectiveness; (2) maps showing future stream flow and temperature scenarios; (3) a review of the ability of specific river restoration actions to ameliorate future effects of climate change or to increase salmon resilience; and (4) a simple decision support structure that integrates these three components to help managers evaluate whether salmon restoration actions should be reprioritized or redesigned for a climate-altered future. Together, these components guide decisions on whether and how to adapt or reprioritize actions in light of expected climate-induced habitat changes. STUDY AREA The study area encompasses the Columbia River basin and coastal drainages of Oregon and Washington (Figure 1), with climatic and ecological conditions ranging from wet forests in the Cascade Mountains to semi-arid and desert regions in the central plateaus (Omernik and Bailey, 1997). The study area is bordered by the Rocky Mountains to the east, and the Cascade Mountains separate coastal drainages from the interior Columbia basin. Mean annual precipitation ranges from <200 mm/year in the central deserts to 3550 mm/year in the Cascade Range (Daly et al., 2002), and elevations range from sea level to over 3700 m in the Rocky Mountains and over 4200 m in the Cascade Mountains. Five anadromous salmon species (Oncorhynchus spp.) and steelhead (O. mykiss) are found in the study area, along with bull trout (Salvelinus confluentus), Dolly Varden char (S. malma), and rainbow and Figure 1. Map of the study area indicating major rivers, mountain ranges, and current and historical ranges of salmon Copyright © 2012 John Wiley & Sons, Ltd. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE cutthroat trout (O. mykiss, O. clarkii). Current ranges of salmon and steelhead are limited to coastal rivers, the Columbia River basin downstream of Chief Joseph dam and the Snake River basin downstream of Hells Canyon Dam. Historical ranges extended into the upper reaches of the Columbia River (Lake Windermere) and into the Snake River basin up to Shoshone Falls. Resident trout species (rainbow, cutthroat, and bull trout) occupy streams throughout the region, with bull trout generally restricted to colder streams at higher elevations. FRAMEWORK AND GUIDING QUESTIONS Determining how a specific change in stream flow or temperature will impact a salmon population depends in part upon species-specific tolerances and life history requirements, and in part upon the expected change in stream flow and temperature relative to those tolerances. The timing of important salmon life history events varies both within and among species (e.g. Groot and Margolis, 1991; Quinn, 2005; Figure 2). For example, salmonids with ocean-type life histories (e.g. pink, chum, and some Chinook) tend to spawn in the fall and winter, do not rear in freshwater during summer, and migrate to sea in the winter or early spring. Salmonids with stream-type life histories (e.g. coho salmon, 941 steelhead, and some Chinook salmon) spawn between fall and late spring, rear in freshwater for 1 or 2 years, and usually migrate to sea in spring or early summer. Hence, each species and life history strategy will encounter a different suite of stream flow and temperature effects because they occupy different habitats and vary in timing of life history events. Climate change effects will also vary among rivers (e.g. Mote et al., 2003; Beechie et al., 2006; Rieman et al., 2007; Crozier et al., 2008), adding additional complexity to understanding how climate change will affect salmonid populations. Because there are many possible combinations of climate change effects and life history responses to evaluate across the study region, we do not attempt a comprehensive review of all possible effects, nor do we use detailed population models to estimate climate change effects on restoration actions (e.g. Battin et al., 2007), mainly because many types of restoration actions cannot be modelled with any certainty and evaluation of hundreds of salmon populations is not feasible (Bartz et al., 2006; Scheuerell et al., 2006). Rather, we summarize key temperature thresholds by species and life history stage, summarize climate change scenarios for stream temperature and flow, and allow local practitioners to relate local climate change scenarios to locally relevant salmonid tolerances. We also review the likely effectiveness of various restoration techniques in a climate-altered future. Figure 2. Timing of climate change effects on stream flow and temperature by life history stages of ocean type Chinook salmon, coho salmon, and steelhead Copyright © 2012 John Wiley & Sons, Ltd. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 942 T. BEECHIE ET AL. For example, actions that create new summer rearing habitats in an area expected to exceed upper lethal temperature limits will not likely improve salmonid populations over the long term. By contrast, actions that significantly reduce stream temperature or create thermal refugia for the same species are more likely to retain their effectiveness in a future with increasing temperatures. Therefore, we reviewed recent literature to evaluate whether specific restoration action types will likely ameliorate climate change effects on flood flows, low flows, or stream temperature. We also developed four guiding questions to evaluate the potential impacts of climate change on project prioritization and design: (1) What habitat restoration actions are necessary for recovery of local salmon populations? (2) Do future stream flow and temperature scenarios alter the types of habitat restoration actions that are necessary for recovery? (3) Does the restoration plan or action ameliorate a predicted climate change effect on stream flow or temperature? (4) Will the restoration plan or action increase habitat diversity and salmon population resilience? Systematic consideration of these questions will help determine whether restoration objectives or priorities should be altered to accommodate future climate change. Answering the first question requires local information about restoration plans and objectives, which restoration planners and practitioners can acquire from salmon recovery plans developed under the Endangered Species Act. However, information required to answer the last three questions is rarely readily available to restoration groups. Therefore, in the following sections, we summarize information needed to address questions 2–4 for salmon restoration actions in the PNW. SCENARIOS FOR CLIMATE CHANGE EFFECTS ON STREAM FLOW AND TEMPERATURE IN THE PACIFIC NORTHWEST Recent climate change scenarios modelled for the PNW suggest a clear warming trend, but the magnitude of estimated temperature increase varies with choice of climate model and emissions scenario (Elsner et al., 2010). By contrast, even the sign of future precipitation changes is not consistent among different scenarios, with some predicting precipitation increases and others predicting decreases (e.g. Elsner et al., 2010). A multi-model averaged climate change scenario under A1B emissions indicates an average temperature increase in the PNW of 3.5 C by 2080, with wetter winters, drier summers, and an increase in average annual precipitation of 5% (Elsner et al., 2010; Mote and Salathé, 2010). Because all the future Copyright © 2012 John Wiley & Sons, Ltd. climate scenarios evaluated by Elsner et al. (2010) predict warming trends, the models predict that more precipitation will fall as rain and less as snow and that this effect will be most pronounced in mid-elevation areas (Hamlet and Lettenmaier, 1999). In this paper, we use this multi-model average from Elsner et al. (2010) to drive a coupled stream flow and temperature model (Whited et al., in press) to produce scenarios of stream temperature and flow regimes that may have significant impacts on salmon populations and food webs that support them. Although the A1B multi-model average is commonly considered to be an informative future climate scenario (i.e. it is closest to most model estimates and the weighting scheme discounts extreme values; Mote and Salathé, 2010), there remains considerable uncertainty around any estimate of future precipitation or air temperature. Uncertainties around future temperature and precipitation predictions have three main sources: (1) the factors that force climate change (including future greenhouse gas and aerosol emissions), (2) global climate model (GCM) errors, and (3) ‘natural’ variability in the climate system (Deser et al., 2010; Hawkins and Sutton, 2011). In general, the variation in temperature or precipitation predictions among different emissions scenarios is smaller than the variation among different GCMs (Mote and Salathé, 2010). For example, the multi-model average predictions of PNW climate for three emissions scenarios in the 2040s indicate a temperature increase of 1.7–2.4 C and a precipitation increase of 1–2% (Figure 3); the 2080s simulations indicate a 2.7–4.7 C increase in temperature and 3–6% increase in precipitation. However, the variation among GCMs for the 2040s is roughly 2 C for each emissions scenario (compared with a range of <1 C among emission scenarios) and as high as 3.5 C among GCMs in the 2080s (compared with ~2 C among emissions scenarios). Variation among GCMs is even greater for precipitation predictions, with a range as high as À8% to +23% for precipitation by the 2080s (compared with 3–6% between emissions scenarios). The combined emissions scenario, model uncertainties, and natural variability for air temperature in the PNW suggest an increase of 1–3 C by the 2040s and 2–6 C by the 2080s (multi-model averages of 2 C and 3.5 C, respectively). The combined uncertainties for precipitation suggest a À6% to +14% change in precipitation by the 2040s and a À8% to +23% change by the 2080s (multi-model averages of +2% and +5%, respectively). Finally, both air temperature and precipitation are expected to continue increasing through the end of the 21st century regardless of the GCM used, and our use of climate scenarios for the 2080s is not intended to suggest a stable future climate. Stream flow and temperature methods Changes in stream flow and temperature were simulated using a two-step modelling process that (1) predicted daily River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE 943 model was based on a hierarchical dominant river tracing algorithm that defines the underlying hydrography for stream flow and temperature calculations (Wu et al., 2011). The coupled model, called the dominant river tracing-based stream flow and temperature model, produces gridded daily stream flow and stream temperature data on the basis of water and heat transport in river networks, thermal dynamics of stream water and the surrounding environment, and the coupling of hydrologic routing processes and associated thermal dynamics (Whited et al., in press). Stream flow and temperature scenarios were based on the multi-model average future climate scenario described previously that provided daily gridded precipitation and air temperature data at one-sixteenth degree resolution (Elsner et al., 2010). Only cells with flow accumulation areas greater than six upstream cells (equivalent to a drainage area of approximately 200 km2) were included in the regional mapping because smaller drainage areas were not considered reliable for future scenarios (Wu and Kimball, unpublished data). Stream flow and temperature were calibrated to measured stream flow and temperatures at seven US Geological Survey sites in the Columbia River basin and then validated against an independent 10-year record of daily stream flows at 12 gauges and a 7-year record of daily stream temperatures at 11 gauges. Model validation indicated strong correlation between measured and simulated stream flow and temperature at the majority of gauges (Figure 4). Although there was no consistent positive or negative bias in either stream flow or temperature, some deviations between modelled and measured stream flow or temperature are likely due to differences between the way VIC models runoff, groundwater, and stream flow compared with local physical processes. For example, in the Willamette basin, groundwater discharge Figure 3. Variation in global climate model (GCM) predictions of precipitation change (%) and temperature increase ( C) in the Pacific Northwest. Circles indicate ensemble model averages for each of three emissions scenarios, and lines indicate range of predictions from 20 different GCMs for each emissions scenario. Based on data from Mote and Salathé (2010) runoff using the macroscale variable infiltration capacity (VIC) model and (2) dynamic runoff routing, stream flow and stream temperature simulations based on VIC that estimate water balance, energy balance, and runoff outputs. The VIC model produces daily runoff and soil moisture, as well as associated forcing variables including incoming shortwave and long wave radiation that are used later in the stream temperature model. The coupled stream flow and temperature Copyright © 2012 John Wiley & Sons, Ltd. Figure 4. Summary of dominant river tracing-based stream flow and temperature model validation results, illustrating distribution of correlation coefficients (R) for modelled versus measured stream flow and temperature. Heavy horizontal line indicates median R, box indicates 25th and 75th percentiles, lines indicate 10th and 90th percentiles, and circles indicate range of values. (Wu and Kimball, unpublished data) River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 944 T. BEECHIE ET AL. from the fractured basalts of the Cascade Mountains leads to higher summer stream flow and lower summer temperature than predicted by the dominant river tracing-based stream flow and temperature model, primarily because the VIC model does not effectively model deep groundwater storage and its influence on flow. Therefore, the VIC model underestimates summer stream flow, which in turn leads to an overestimate of summer stream temperature. Other studies have found regional biases in mean annual runoff associated with arid regions or snowmelt systems and showed overall underestimation of mean annual stream flow in the PNW from VIC modelling (Gangopadhyay and Pruitt, 2011). However, validation of stream flows for this study did not suggest a clear spatial pattern of positive or negative biases. We focused our flow analysis on mean monthly flows for four periods: 1970–1999, 2000–2029, 2030–2069, and 2070–2099. We calculated change in magnitude of the maximum and minimum monthly flows between periods for each stream cell, as well as the change in timing of maximum and minimum monthly flows between periods. We focused on predicted change in flow relative to the modelled historical baseline rather than absolute stream flows to minimize impacts of errors in the flow model on our results. That is, we assume that biases in the stream flow model will be in the same direction for all periods within a grid cell and that using the change in flow as our primary metric will reduce the impact of those biases on our analysis. Finally, we used cluster analysis to map three flow regimes (rainfall-dominated, snowmelt-dominated, and transitional) during each period (Beechie et al., 2006). We chose to map temperature predictions directly to allow the greatest flexibility in biological interpretations. That is, we avoided selection of specific thermal limits because thermal tolerances vary considerably among species for each life stage. To aid in biological interpretation of these temperature maps, we provide both species-specific and generalized salmonid thermal limits (Table I), as well as published temperature criteria that are recommended for protection of Pacific salmon from negative temperature effects (Table II). Most upper lethal limits are between 20 and 24 C and recommended temperature thresholds for the 7-day average daily maximum range from 13 to 18 C. In the absence of local data on thermal tolerances of salmonids (which vary among species and environments), these data can be used to gauge the likelihood that stream temperature changes will be significant for local species. Results: stream flow change scenario Areas of the PNW with a snowmelt-dominated hydrologic regime (in which the maximum monthly flows are during the spring snowmelt) shrink considerably under the ensemble climate change scenario as snow level rises across the region (Figure 5). By 2070–2099, the snowmelt hydrologic regime no longer exists in the north Cascades and upper Snake River basin, and the only remaining snowmelt-dominated area is in the Canadian Rockies. The transitional regime, which has Table I. Temperature thresholds ( C) for critical parts of the salmonid life cycle, including general and species specific information Life stage Chinook (O. tshawytscha) Adult migration Optimal threshold Lethal threshold 22 Thermal blockage Adult holding and spawning Optimal threshold 14.5 Detrimental to internally held gametes Incubation and early fry development Upper threshold 14.5 Juvenile rearing Optimal threshold 14.8a Lethal threshold 24 UZNGb Smoltification Impairment threshold 12–17 Chum (O. keta) Coho (O. kisutch) 21 15.6 21 Pink (O. gorbuscha) Sockeye (O. nerka) Steelhead (O. mykiss) 22 12.8 15.6 20 10 12 15 21 19.8 17 23 23.4 15 12.8 12 12.5 12 19 20 21 24 13 Temperatures cited are for constant exposure, unless otherwise noted. Data compiled from Bjornn and Reiser (1991), Eaton and Scheller (1996), McCullough et al. (2001), and Richter and Kolmes (2005). a Natural rations level. b Upper zero net growth (UZNG) temperature: maximum weekly temperature at which fish can live for several days but at which they do not ingest enough food to gain weight. Copyright © 2012 John Wiley & Sons, Ltd. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 945 RESTORING SALMON HABITAT FOR A CHANGING CLIMATE Table II. Recommended temperature criteria (upper thresholds; C) for Pacific salmon and steelhead Life stage Reference Adult migration Spawning Incubation Juvenile rearing Smoltification Steelhead (non-steelhead) smoltification 7-DADM 18 13 13 16 16 14 Weekly mean 16 10 10 15 15 12 Richter and Kolmes, 2005; US EPA, 2003 Richter and Kolmes, 2005 Figure 5. Modelled hydrologic regime through time, based on cluster analyses of mean monthly flows for each period both spring snowmelt and fall–winter high flows, shifts inland and northwards, leaving only small areas of the transitional regime in the north Cascades and the Rocky Mountains. The rainfall-dominated regime, which has the highest flows during fall–winter floods, has historically been limited to the maritime climate west of the Cascades and small low-elevation portion Copyright © 2012 John Wiley & Sons, Ltd. of the interior Columbia basin. By 2070–2099 however, the rainfall-dominated regime expands to nearly the entire interior Columbia basin in this climate change scenario. The ensemble climate scenario suggests the largest decreases in summer low flows will be west of the Cascade Mountains, where minimum monthly flows decrease by 10–70% over the River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 946 T. BEECHIE ET AL. course of the 21st century (Figure 6). The largest changes are in basins that currently have a transitional hydrologic regime (with both a fall/winter storm peak and a late spring snowmelt peak), and large decreases in minimum monthly flows result from a nearly complete loss of the spring snowmelt peak and a concomitant decline in late summer flows. More modest decreases in low flow (10–35%) are predicted in the Rocky Mountains south of the US–Canada border and in the Blue Mountains of northeastern Oregon. In these areas, the hydrologic regime shifts from snowmelt to transitional, and the decline in snowpack results in earlier spring melt and a decrease in late summer stream flows (Figure 6). Minimum monthly flows in the Canadian Rockies increase by 10% or more, largely as a result of a predicted increase in precipitation and snowpack. Increased minimum flows are also predicted in the upper Snake River basin, caused by a shift in minimum monthly flows from the cold January–February period to minimum flows in summer following spring snowmelt. In this region, there is little predicted change in summer low flows, but the minimum winter flows increase significantly. Simulated maximum monthly flows increase by 10–50% across most of the region as a result of an increasing fraction of precipitation falling as rain rather than snow (Figure 7). The few areas where maximum monthly flows are expected to increase by more than 50% are located in the Cascade Mountains and in the middle and lower Snake River basin. The large increases in the Cascade Mountains are predominantly a result of a shift from transitional to rainfalldominated hydrographs, with future flood flows in the fall and winter being considerably larger than at present. Large predicted increases in the monthly average peak stream flow in the Snake River tributaries result mainly from large increases in spring precipitation. Overall, the ensemble scenario suggests increasing volume of winter runoff and increased flooding in transitional basins and increasing spring flows in snow-dominant basins. Reductions in summer low flows are projected to be largest in the transitional basins in the Cascade and Olympic Mountain ranges. Results: stream temperature change scenario Increased air temperatures will lead to increased water temperatures on both the west and east sides of the Cascade Mountains, and the scenario indicates a 1–4 C increase in stream temperatures (maximum weekly mean temperature) across the region by the 2030–2069 period and a 2–6 C increase by the 2070–2099 period (Figure 8). Highest mean weekly water temperatures vary significantly across the region in all periods, with highest temperatures in reaches of the Snake and Willamette River basins (Figure 9). Because these areas are close to or exceed published thermal tolerances of most salmon species even during the historical period (1970–1999), they are most Copyright © 2012 John Wiley & Sons, Ltd. likely to shift to stressful or lethal thermal conditions in the future. Notably, many rivers within the current salmon range have modelled temperatures above published lethal or protective thresholds (Tables I and II), yet salmon currently occupy the majority of these rivers. Stream temperatures in the northern part of the Columbia basin are currently within thermal tolerances of most rearing juvenile salmonids, and under this climate change scenario, projected temperature increases remain the lowest in the region and within thermal tolerances. This area is outside the current salmon range because of blocked migration by dams, but within their historical range. Most coastal river systems and rivers originating on the west slope of the Cascade Mountains are likely to remain within published thermal tolerances even in the 2070–2099 period. These river systems have the smallest projected temperature increases, whereas the largest increases are expected in the main stem Columbia River and tributaries in the middle and lower Columbia basin. REVIEW OF RESTORATION ACTIONS AND CLIMATE CHANGE We grouped restoration actions on the basis of the watershed processes or functions they attempt to restore (Beechie et al., 2010) and then classified them as either likely or not likely to ameliorate a climate change effect on high stream flows, low stream flows, and stream temperatures (Table III). We classified actions on the basis of a literature review of restoration action effectiveness and watershed processes to develop a comprehensive summary of each action’s likelihood of ameliorating climate change effects. Our basic rules were to (1) classify an action as likely to ameliorate an effect if we could find literature support for that response and (2) avoid including effects that were theoretically possible but not supported by data. In a few cases, the literature was sparse and suggested mixed effects depending on the context. In those cases, we classified the action as having a contextdependent effect on stream flow or temperature to indicate that the ability of the action to ameliorate a climate effect depends on the situation in which the action is employed. Although these rules may omit a few effects, we felt that it was more important to provide clear guidance on the dominant effects and avoid including actions that only rarely would ameliorate the climate change effect. Nevertheless, this review is not intended to imply that less robust actions should be avoided in all circumstances. For instance, where summer rearing habitats constrain population recovery and summer stream flow and temperature are not expected to change significantly, any action that addresses causal factors for habitat and population declines should be implemented even if it does not ameliorate a climate change effect. Only in cases where River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE 947 Figure 6. Modelled change in the minimum mean monthly flow and shift in timing of minimum mean monthly flow. Ratio is minimum mean monthly flow from the future period divided by the minimum mean monthly flow from the period 1970–1999. For shift in low flow timing, areas mapped in red indicate a shift from minimum flow in winter (usually February) to minimum flow in August. Areas in yellow predominantly indicate a shift from minimum monthly flow in September to August Copyright © 2012 John Wiley & Sons, Ltd. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 948 T. BEECHIE ET AL. Figure 7. Modelled change in the maximum mean monthly flow and shift in timing of maximum monthly flows. Ratio is maximum mean monthly flow from the future period 2070–2099 divided by the maximum mean monthly flow from the period 1970–1999. For timing of maximum monthly flows, areas mapped in red indicate a shift from maximum monthly flow during spring snowmelt (usually April or May) to maximum flows during December and January winter storms. Areas in yellow indicate areas with maximum monthly flow remaining in late winter or spring, but shifted 1–2 months earlier (generally April–May to February–March) Copyright © 2012 John Wiley & Sons, Ltd. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE 949 climate effects are expected to impact a project’s or plan’s effectiveness within a few decades do we suggest adjustments to project priorities or designs. We also review restoration actions in the context of their ability to maintain or increase resilience of river ecosystems and salmon populations (Waples et al., 2009). We define resilience as the ability of a system to absorb change and still maintain its basic ecosystem functions and relationships, even though the balance of habitat types or species may shift slowly through time (Holling, 1973; Waples et al., 2009). Pacific salmon are adapted to wide array of natural disturbance regimes by virtue of their life history diversity, and restoration actions designed to reduce constraints on life history diversity allow Pacific salmon a broader range of options by which to respond to climate change (Waples et al., 2008, 2009)—conferring resilience to both populations and meta-populations (Greene et al., 2010; Schindler et al., 2010). That is, restoration actions that increase habitat diversity to the point that salmon have the ability to express alternative life history strategies are considered to potentially increase population resilience. For example, restoring diverse floodplain habitats or reconnecting cold-water tributaries to main stem habitats by barrier removals offers salmon a variety of physical and thermal conditions, allowing multiple species to persist and to express varied life history strategies within species (Poole et al., 2008; Waples et al., 2009). In contrast, creation of pools by adding wood to a small stream generates a small increase in habitat diversity but does not offer an array of habitats that allow expression of alternative life histories. Hence, we consider the former action to potentially increase resilience but not the latter. Restoring connectivity Restoring connectivity (longitudinal, lateral, and vertical) typically improves both physical and biological functions of river systems. Restoring longitudinal connectivity for salmon is primarily intended to reestablish salmon migration to diverse habitats that have been lost through construction of artificial barriers such as dams or culverts, but it often also restores downstream transport of essential flows, sediment, and wood or organic matter. Restoring lateral connectivity generally refers to reconnection of rivers to their floodplains by removal of levees or bank armouring. These actions restore the ability of the river system to create and sustain diverse habitats and to allow migration of salmon into those habitats. Actions that aim to restore vertical connectivity seek to aggrade incised or scoured channels, which increases the connection between surface and subsurface flows and increases floodplain connectivity over time. Figure 8. Modelled increase in maximum average weekly temperature through time in the Columbia River basin Copyright © 2012 John Wiley & Sons, Ltd. Longitudinal connectivity (barrier removal). The primary aims of restoring longitudinal connectivity by removal of dams or other blocking structures are to (1) reestablish upstream and River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 950 T. BEECHIE ET AL. Figure 9. Modelled maximum weekly temperature through time in the Columbia River basin downstream fish migration pathways and (2) restore natural stream flow, sediment, and organic matter transport (Pess et al., 2005). Removal of dams or providing fish passage on dams that cannot be removed will allow fish to access important upstream cool water habitats when downstream areas become too warm (McClure et al., 2008; Pess et al., 2008), thereby increasing habitat and life history diversity at the population and meta-population scales (Beechie et al., 2006; Waples et al., 2009). Where dams or other structures contribute to reduced low flows or increased stream temperature, dam removal can also ameliorate low base flow and high temperature problems by restoring downstream movement of sediment and water (Burroughs et al., 2009). Lateral connectivity (floodplain reconnection). The aims of re-establishing lateral connectivity between river channels Copyright © 2012 John Wiley & Sons, Ltd. and floodplains are often twofold: to restore river floodplain dynamics that create diverse habitats and to restore fish access to floodplain habitats (Pess et al., 2005; Waples et al., 2009). These actions, which typically include reconnection or creation of side channels and sloughs, removal or set back of levees and dikes, and re-meandering of dredged or straightened channels, can ameliorate peak flow increases by storing flood water and reducing flood peaks (Sparks et al., 1998; McAlister et al., 2000) or by increasing the availability of velocity and thermal refugia (Sommer et al., 2001; Morley et al., 2005; Jeffres et al., 2008; Poole et al., 2008). Similarly, removing levees or re-meandering channels can ameliorate temperature increases by increasing length of hyporheic flow paths beneath the floodplain, which can cool water during the summer (Arrigoni et al., 2008; Konrad et al., 2008; Poole et al., 2008; Opperman et al., River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 951 RESTORING SALMON HABITAT FOR A CHANGING CLIMATE Table III. Summary of restoration action types and their ability to ameliorate climate change effects on peak flow, low flow, stream temperature, or to increase salmon population resilience Category Common techniques Longitudinal connectivity (barrier removal) Removal or breaching of dam Barrier or culvert replacement/removal Lateral connectivity (floodplain reconnection) Levee removal Reconnection of floodplain features (e.g. channels, ponds) Creation of new floodplain habitats Vertical connectivity (incised channel restoration) Reintroduce beaver (dams increase sediment storage) Remove cattle (restored vegetation stores sediment) Install grade controls Stream flow regimes Restoration of natural flood regime Reduce water withdrawals, restore summer baseflow Reduce upland grazing Disconnect road drainage from streams Natural drainage systems, retention ponds, other urban stormwater techniques Erosion and sediment delivery Road resurfacing Landslide hazard reduction (sidecast removal, fill removal) Reduced cropland erosion (e.g. no-till seeding) Reduced grazing (e.g. fencing livestock away from streams) Riparian functions Grazing removal, fencing, controlled grazing Planting (trees, other vegetation) Thinning or removal of understory Remove non-native plants Instream rehabilitation Re-meandering of straightened stream, channel realignment Addition of log structures, log jams Boulder weirs and boulders Brush bundles, cover structures Gravel addition Nutrient enrichment Addition of organic and inorganic nutrients Ameliorates temperature increase ● Ameliorates base flow decrease Ameliorates peak flow increase Increases salmon resilience ● ● ● ● ● ● ○ ○ ● ○ ○ ● ● ● ● ● ● ● ● ○ ○ ● ● ● ● ● ● ○ ● ● ● ● ● ○ ○ ○ ○ ◒ ○ ○ ○ ○ ○ ◒ ◒ ● ● ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ● ● ○ ○ ○ ○ ◒ ○ ○ ○ ○ ○ ○ ○ ○ ◒ ○ ○ ◒ ◒ ◒ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ○ ◒ ◒ ◒ Actions are grouped by major processes or functions they attempt to restore: connectivity (longitudinal, lateral and vertical), watershed-scale processes (stream flow and erosion regimes), riparian processes, instream rehabilitation, and nutrient enrichment. Filled circles indicate positive effect, empty circles indicate no effect, and partially filled circles indicate context-dependent effects. See text for supporting citations. 2010). Increasing habitat diversity by restoring lateral connectivity generally allows for increased life history diversity within a population (Waples et al., 2009), which has been linked to increased population resilience (Greene et al., 2010; Schindler et al., 2010). Floodplain reconnection actions generally do not ameliorate base flow decreases. Copyright © 2012 John Wiley & Sons, Ltd. Vertical connectivity (restoring incised channels). Channel incision (or downcutting) has degraded stream and riparian habitats in many rivers of the PNW by lowering water tables, reducing exchange between surface and subsurface flows, and decreasing late summer stream flows. Associated losses in riparian vegetation lead to reduced shading River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 952 T. BEECHIE ET AL. and organic matter inputs to streams and increased stream temperatures (Elmore and Beschta, 1987; Poole and Berman, 2001). Moreover, incised streams rarely access their floodplains, high flows are concentrated within the incised channel, and fish have no access to slow-water refugia during floods (Harvey and Watson, 1986; Elmore and Beschta, 1987; Shields et al., 1995). Efforts to restore incised streams by increasing sediment storage and aggrading the channel aim to restore floodplain aquifer storage, increase summer base flow and decrease summer stream temperature, and increase availability of flood refugia (Pollock et al., 2007; Beechie et al., 2008b). Some restoration techniques, such as use of beaver dams to increase sediment storage, have the added effects of increasing summer base flows, locally decreasing or buffering stream temperature and increasing habitat diversity and productivity (Ponce and Lindquist, 1990; McRae and Edwards, 1994; Pollock et al., 2003; Pollock et al., 2007). Hence, restoration of incised channels has the potential to ameliorate climate-induced increases in stream temperature, effects on peak flows and low flows, and also to increase life history diversity through creation of off-channel and pond habitats. We emphasize that ameliorating climate change effects through aggradation techniques is specific to incised channels and that the same techniques may have different responses in other settings. For example, water may be warmer in a beaver pond within an otherwise closed canopy system, although stream cooling may still occur downstream of the pond (Robison et al., 1999). Restoring stream flow regimes Flood flows are increased to a moderate degree by logging and forest roads (Jones and Grant, 1996), grazing effects (Belsky et al., 1999), and to a much greater degree by impervious surfaces in urban areas (Booth et al., 2002). The primary mechanism by which logging roads increase peak flows is interception of subsurface flow through soils (which moves relatively slowly) and rapid routing of water to streams through ditches (Furniss et al., 1991; Jones and Grant, 1996). Road rehabilitation actions to decrease peak flow effects generally focus on addition of cross-drains to reduce routing of water directly from road ditches to the stream (Furniss et al., 1991). In an urban environment, the primary focus is to reduce the impacts of impervious surfaces by creating additional stormwater retention structures or modifying impervious surface areas so that rapid runoff is routed into groundwater storage rather than storm drains (e.g. Booth and Leavitt, 1999). In many cases, increased runoff and flood flows cause summer baseflows to decrease due to loss of infiltration and water storage in soils (e.g. Belsky et al., 1999). Hence, reductions of grazing or logging effects on flood flows may also increase low flows in summer. Copyright © 2012 John Wiley & Sons, Ltd. Low stream flows are often reduced by withdrawal of water from streams for irrigation or consumptive uses (Poff et al., 1997; Myers et al., 1998), and both peak and low flows may be dramatically reduced by water storage behind dams (Stanford et al., 1996; Poff et al., 1997). Restoring some or all of abstracted water to streams through purchase of water rights or increased irrigation efficiency can dramatically increase low flows to streams (Poff et al., 2010) and directly ameliorate climate-induced decreases in low stream flow or increased stream temperature. In some cases, flow regulation has decreased peak flows to the point that many geomorphological and ecological functions of streams are lost (Olden and Poff, 2004). Moreover, low flows may be reduced in summer, which can also lead to increased stream temperature. In such cases, restoring ‘environmental flow regimes’ can ameliorate not only low stream flows, but can also increase habitat diversity by restoring channel-forming flows that maintain habitat diversity and other ecological functions (Stanford et al., 1996; Poff et al., 2010). Hence, where water storage or withdrawal has decreased low flows, purchase of water rights or use of water conservation measures that leave more water in the stream can ameliorate predicted decreases in low flows. Some dams can release cool water from deep in the reservoir, allowing dam operations to ameliorate stream temperature increases. Where water storage has decreased peak flows, restoration of channel forming flows can increase habitat diversity through restoration of physical functions that create diverse habitat features in streams and across the floodplain and also maintain riparian functions (Poff et al., 1997)—thereby increasing resilience of river ecosystems to climate change. Reducing erosion and sediment delivery In forested environments of the PNW, sediment supply to stream channels is typically increased through surface erosion on unpaved roads or by increased landsliding from roads or clearcuts (Reid and Dunne, 1984; Sidle et al., 1985). Therefore, sediment reduction efforts in forest environments commonly focus on road rehabilitation to decrease landslide hazards and surface erosion (Beechie et al., 2005; Roni et al., 2008). Landslide hazard reduction is typically achieved by removing or abandoning roads, or rebuilding stream crossings to avoid fill failures when culverts become blocked (Madej, 2001; McCaffery et al., 2007). Despite these efforts, future increases in storm intensity and a shift from snow to rainfall may drive more frequent mass wasting in forest environments, especially where road management has not yet achieved reductions in landslide hazard. Effects of increased surface erosion on roads can be abated by resurfacing the road or adding cross drains or water bars to prevent delivery of eroded sediments to streams (Furniss et al., 1991). River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE In croplands, surface erosion is often increased by erosion of exposed soil in fallow fields (Wendt and Burwell, 1985; Ebbert and Roe, 1998). An increasingly common strategy to manage surface erosion in agricultural lands is no-till seeding, which preserves vegetative cover on croplands and dramatically reduces erosion and sediment delivery to streams (Wendt and Burwell, 1985; Ebbert and Roe, 1998). Grazing effects on sediment supply include removal of hillslope vegetation and erosion of exposed soils, as well as trampling of banks and increased bank erosion (Medina et al., 2005). Grazing impacts can be controlled either by removal of livestock from key areas (especially stream banks and riparian areas) or by grazing rotations that retain sufficient vegetative cover to reduce surface erosion (Medina et al., 2005). Although each of these actions can improve stream habitat by decreasing fines in the stream bed, increasing pool depth, or narrowing widened channels—none of these actions ameliorate decreased low flows, increased flood magnitude, or increased stream temperature (although increased pool depth may create thermal refugia in rare cases). Moreover, these actions do little to increase habitat or life history diversity except in cases where extremely high sediment supply has filled pools and reduced the diversity of habitat types (see Beechie et al., 2005 for examples). Restoring riparian functions Riparian rehabilitation actions aim to restore riparian functions such as stream shading, root reinforcement of banks, supply of large wood and organic matter, and trapping sediment or filtering nutrients (Kauffman et al., 1997; Pollock et al., 2005). In forested environments, restoration of riparian functions commonly focuses on thinning or replanting of riparian forests to restore wood recruitment and shade functions and, secondarily, to restore other functions (Beechie et al., 2000; Welty et al., 2002; Meleason et al., 2003). Restored riparian functions do not directly ameliorate the stream flow changes predicted by climate change models, but may mitigate stream temperature increases via increased shading (Johnson, 2004), or via increased wood recruitment and sediment storage in headwater channels that have been scoured to bedrock (Pollock et al., 2009). However, removal of certain non-native species that use more water than native species and provide less shade can ameliorate increased stream temperatures or decreased flows. In non-forest environments, replanting of denuded or managed riparian zones and removal or reduction of livestock grazing typically results in regrowth of riparian vegetation and should also ameliorate increases in stream temperatures through increased shade, bank stability, and narrowing of stream channels (Medina et al., 2005). Riparian restoration can lead to modest increases in habitat diversity over the long term via formation of pools or Copyright © 2012 John Wiley & Sons, Ltd. 953 hiding cover (Beechie et al., 2000), whereas actions that seek to thin riparian zones are unlikely to affect either stream flow or temperature (Pollock et al., 2009). Finally, restoration of normative flow regimes on regulated rivers should help recovery of riparian areas on larger rivers, as seedling establishment for key riparian species is often dependent on flood magnitudes and duration (Stanford et al., 1996; Mahoney and Rood, 1998). Riparian restoration can be expected to increase ecosystem resilience in the sense that rivers with intact riparian buffers can buffer ecological functions against changes in stream flow, but it is unlikely to increase life history diversity and salmon resilience beyond the buffering of temperature effects. Instream rehabilitation Instream rehabilitation includes restoration actions that seek to improve habitat conditions by actively altering channel habitat structure (e.g. adding wood debris, spawning gravel), reconstructing channel characteristics (re-meandering), or by providing cover for fish (Roni et al., 2008). Such fixed structures are susceptible to failure or require maintenance, especially in the face of increased magnitude and frequency of peak flow events as predicted by climate change models. Although instream rehabilitation actions such as wood and boulder placement have been documented to provide quick improvements in both physical habitat and fish production (Cederholm et al., 1997; Solazzi et al., 2000; Roni and Quinn, 2001), they do not restore the underlying disrupted process (typically large wood delivery) and are unlikely to last more than one or two decades without additional intervention or maintenance (Roni et al., 2002). Moreover, instream rehabilitation actions generally do not ameliorate changes in temperature, base flow, or peak flows. For example, some studies have shown that creation of poolriffle sequences can lead to increased hyporheic exchange and increased temperature variability, but none has shown a significant net decrease in stream temperature (Crispell and Endreny, 2009; Hester et al., 2009). This is most likely because the subsurface flow path is too short to significantly affect stream temperature (Poole et al., 2008). By contrast, restoring sediment storage to channels that are incised to bedrock may reduce stream temperatures if the loss of sediment has completely eliminated hyporheic exchange and increased stream temperature (Pollock et al., 2009). In such cases, use of wood or boulder structures to store sediment may decrease stream temperature. Finally, instream habitat actions can increase local habitat complexity (particularly if a large portion of the stream is treated), but such actions are unlikely to increase life history diversity or resilience of salmon populations. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 954 T. BEECHIE ET AL. Nutrient enrichment The purpose of nutrient enrichment is to compensate for lack of marine-derived nutrients from reduced salmon returns by adding nutrients and carbon to boost stream or lake productivity, and ultimately fish production (Bilby et al., 1998; Gresh et al., 2000; Kiffney et al., 2005). These exogenous sources of nutrients are important to the productivity of naturally oligotrophic rivers of the PNW where many salmonid populations are food-limited (e.g. Boss and Richardson, 2002). As with instream rehabilitation actions, nutrient additions do not address the ultimate cause of low nutrient levels as a result of reduced salmon runs and, in the absence of increased salmon returns, are dependent on continually adding nutrients to maintain any benefits (Roni et al., 2008). Nutrient additions do not ameliorate climate change effects on stream flow, stream temperature, or habitat diversity. However, an important secondary effect of increased stream temperature is increased metabolism in juvenile fishes, which increases food requirements to maintain positive growth (McCullough et al., 2001; Boughton et al., 2007). Where reduced nutrients and food resources have already compromised growth of juvenile salmonids, rehabilitation actions to increase nutrient supply—thereby increasing invertebrate abundance and prey availability for juvenile salmonids—may indirectly ameliorate temperature effects on salmonid growth rates (Wipfli and Baxter, 2010). However, this would require a consistent, long-term nutrient supplementation programme and would not lead to self-sustaining nutrient levels without continual intervention. DECISION SUPPORT STRUCTURE TO EVALUATE CLIMATE CHANGE EFFECTS ON SALMON RESTORATION We proposed four guiding questions to help determine whether restoration plans or actions should be altered to accommodate climate change (first column of Figure 10). We have also provided key maps and information to answer these questions (Figures 3–7, Tables 1–3). However, translation of this information into adaptation of restoration plans or actions can be ambiguous, so we offer two simple decision support tools to assist in evaluating restoration plans or actions in the context of climate change. Both are simple flow charts that illustrate how answers to the guiding questions might lead to logical adaptations of restoration plans or actions. We do not intend these to be rigid protocols with predefined outcomes because there are many possible combinations of future climates, restoration strategies, and species responses, and it is difficult to arrive at a set of rules that will apply to all possible cases. Rather, we intend these tools to illustrate how answers to the guiding questions can be integrated to Copyright © 2012 John Wiley & Sons, Ltd. arrive at management decisions in the context of local goals and objectives, as well as in the context of local climate change scenarios. Evaluating a salmon recovery plan In most cases, the first of the guiding questions has been answered in the process of developing local salmon recovery plans (e.g. Shared Strategy Development Committee, 2007). That is, an important component of salmon recovery plans is the identification of habitat impairments that constrain salmon population growth, and which, if addressed, will increase abundance or productivity (population growth rate) of the population (McElhany et al., 2000; Beechie et al., 2003). From this analysis, a list of important habitat recovery actions can be developed and prioritized (Beechie et al., 2008a). However, these lists are commonly developed without consideration of future climate change effects on habitats and therefore have not considered how climate change might alter the suite of restoration actions identified as necessary to achieve salmon recovery. Evaluating whether potential climate change effects on stream flow or temperature will change the list of restoration actions necessary for salmon recovery (question 2) begins a decision tree that helps restoration planners determine whether a salmon recovery plan should be revised to accommodate future climate change effects (Figure 10). Answering this question requires examination of potential climate change effects on stream flow and temperature (Figures 3–6) and a qualitative assessment of whether future stream flows or temperatures are likely to alter conclusions about which habitat restoration actions are necessary for salmon recovery. If climate change effects on stream flow or temperature are not expected to change the types or priority of restoration actions, then restoration actions may proceed according to the current plan. An important caution is that any assessment using mapped scenarios of flow or temperature changes should recognize that the resolution of the maps is quite low (in our analysis, the smallest grid cell representing a stream is 6 Â 6 km) and that there is considerable variation in predicted stream flow and temperature changes among emissions scenarios and GCMs (Elsner et al., 2010). Therefore, any evaluation of these stream flow and temperature scenarios should acknowledge that uncertainty in the climate scenarios is high. If the evaluation of climate change scenarios indicates a change in the types of actions needed for recovery, then planned actions should be evaluated to determine whether they ameliorate the local climate change effects (question 3). For example, a coho salmon population may currently be constrained by winter rearing habitat availability, but potential increases in summer stream temperatures and decreases in summer stream flows (Figure 9) may reduce summer rearing habitat availability to the point that it River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE 955 Figure 10. Schematic diagram of the decision-making process for adapting salmon recovery plans or individual restoration actions to climate change, and relationship to the four guiding questions becomes the primary constraint on a population’s recovery. If the planned restoration actions are likely to ameliorate climate change effects on stream flow or temperature (from Table III) and preserve the anticipated effectiveness of the restoration plan, then restoration can also proceed without modification. However, if actions will not ameliorate the climate change effects, revisions to the restoration plan should be considered. We acknowledge that determining whether the plan will ameliorate a temperature effect enough to prevent stream temperatures from exceeding critical thresholds is extremely difficult. In most cases, available data and models are not sufficient to answer this question quantitatively and with confidence. However, most restoration actions are not expected to reduce stream temperatures by more than 1–2 C (e.g. Medina et al., 2005; Arrigoni et al., 2008), so we suggest as a rule of thumb that restoration actions in areas that are less than 1–2 C below a critical threshold be considered unlikely to ameliorate a climate change effect. Nevertheless, it is possible that some combinations of actions might reduce stream temperature by more than 2 C, and local experience with restoration actions and changes in stream flow or temperature should be considered in management decisions. Moreover, salmon are often adapted to higher temperatures than are typically reported in Copyright © 2012 John Wiley & Sons, Ltd. the literature, and data on local thermal tolerances of salmon should be considered. When revision of a plan is warranted, the degree to which the plan should be revised depends on whether the proposed actions contribute to increasing resilience of the population (question 4). Where the main habitat restoration actions in a plan contribute significantly to increasing resilience of populations (e.g. by increasing habitat diversity; Waples et al., 2009), the plan may be followed with the understanding that climate change may reduce effectiveness of the habitat restoration plan over the long term. There will always be considerable uncertainty about whether the actions can increase resilience enough to allow population recovery. But in any case, restoring diverse habitats will increase resilience of the riverine ecosystem—thereby increasing the likelihood that a salmon population can recover under a warming climate. Where the main habitat restoration actions do not contribute significantly to increasing habitat diversity or resilience of a population, then the restoration plan should be revised to increase the likelihood that actions either ameliorate climate change effects or increase habitat diversity and ecosystem resilience. The re-evaluation should focus on identifying actions that will help the population recover under both existing and future River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 956 T. BEECHIE ET AL. limiting habitats, so that climate change effects on habitat conditions and population performance do not hinder recovery of the population. Restoration of key physical and biological processes will allow a river ecosystem to adjust naturally to changes in key ecosystem drivers such as stream flow and temperature and will be more robust to variation in future climate patterns than actions that attempt to control river behaviour or build specific habitat features (Waples et al., 2009; Beechie et al., 2010). We stress that even where detailed models are used to assess how climate change will affect habitats that limit population recovery (e.g. Battin et al., 2007), there will always be considerable uncertainty in both model structure and climate change scenarios. Therefore, we encourage adjustments to recovery plans that broaden the portfolio of actions to accommodate a wide range of potential future climate scenarios. Although we did not model future stream flow and temperature for multiple emissions scenarios and climate models, previous studies give us some indication of the range of potential outcomes for air temperature and precipitation (e.g. Elsner et al., 2010, Mote and Salathé, 2010). On the basis of those studies, it seems prudent to consider a range of potential stream temperature increases at least 2 C higher than those predicted from our A1B ensemble climate scenario (although we recognize that there is not always a strong correlation between air temperature and stream temperature). We cannot suggest a similar range of values for stream flows because no studies have modelled variation in stream flow among GCMs in the PNW (although Elsner et al., 2010 modelled stream flow differences among two emissions scenarios). Nevertheless, uncertainty in precipitation predictions is very high, suggesting that a conservative approach might anticipate that changes in stream flow (either high or low) might be considerably larger than our map illustrates. Finally, population status may also influence management choices. Populations at very low abundance may require an emphasis on near-term habitat recovery actions to stabilize abundance (i.e. longevity of actions may be only 10–20 years), whereas more stable populations may benefit more from restoration of processes that persist for much longer periods. For near-term actions, climate change will likely produce relatively small effects on habitat conditions, and plans that emphasize near-term actions may need little revision regardless of climate change threats. However, this is a relatively rare case (most populations are threatened rather than endangered), and emphasizing restoration of habitat-forming processes is more likely to succeed over the long term (Beechie et al., 2010). Evaluating an individual habitat restoration action Individual restoration actions are perhaps simpler to evaluate because their future effectiveness is primarily a function of Copyright © 2012 John Wiley & Sons, Ltd. their intended purpose, which can be very narrowly defined. Answering the four guiding questions leads to one of three potential decisions: proceed with the action as planned, redesign the project to accommodate climate change, or abandon the project in favour of other projects that maintain their effectiveness in the face of climate change. The first outcome—proceeding with the action as planned—is appropriate when climate change does not alter effectiveness or longevity of the action, or when the action ameliorates the climate change effect sufficiently to maintain project effectiveness in the future (Figure 10). Redesigning a project to accommodate climate change is most appropriate where the action effectiveness is not reduced but longevity may be reduced as a function of a climate change effect (e.g. if peak flows will increase then the project should be designed to accommodate larger peak flows than observed at present). Finally, it is prudent to consider abandoning a specific project if climate change will likely negate its intended purpose, the action does not ameliorate the effect, and the action does not increase diversity and resilience. In these cases, the action may be abandoned in favour of other actions that will maintain their effectiveness in the face of climate change. In evaluating the potential effects of climate change on individual restoration projects, it is first necessary to know which species and life stage the restoration action targets in order to evaluate whether climate change alters the effectiveness of the action. For example, if an action is intended to restore a winter rearing habitat, then changes to winter stream flows will be an important evaluation criterion whereas summer stream temperatures may not. Once the purpose of an action is identified, one can ask whether climate change will alter action effectiveness. If the effectiveness is not altered, then the action can proceed pending evaluation of climate change effects on project longevity (Figure 10). If the effectiveness is likely to be reduced, then one should consider whether the action type significantly ameliorates climate-related changes in stream flow and temperature. Evaluating the ecological impacts of different temperature change scenarios is relatively straightforward, as each species and life stage has a relatively specific range of thermal tolerances (Tables I and II), and temperature change magnitudes (Figures 2 and 3) can be compared with those tolerances to judge whether climate change will likely reduce project effectiveness. Finally, it is also important to consider how long the restoration action will last when determining whether it will ameliorate the impacts of climate change. Actions such as restoring floodplain connectivity or removing migration barriers restore underlying watershed processes, can last many decades, and will likely be the most effective longterm strategies for climate change because they both ameliorate climate change effects and increase habitat diversity and resilience. Other actions that may last only two or three decades without continued intervention will only provide River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 957 RESTORING SALMON HABITAT FOR A CHANGING CLIMATE short-term amelioration for climate change impacts and are unlikely to appreciably increase resilience over the long term. SUMMARY AND CONCLUSIONS We developed a set of guiding questions and data to inform adaptation of habitat restoration plans for salmonids in northwestern USA. These same questions are applicable to any salmon restoration effort, and—moreover—generally applicable to restoration of many species or ecosystems. Key elements of adapting any restoration strategy to climate change include (1) understanding the current recovery needs, (2) evaluating whether climate change effects will likely alter those needs, (3) determining whether restoration actions can ameliorate climate change effects, and (4) determining whether restoration actions can increase ecosystem resilience. These components are not specific to salmon, nor are they specific to aquatic species. These same questions can be used for any ecosystem in which restoration actions might need to be adapted to accommodate environmental effects of climate change. The key questions that must be answered for any adaptation strategy are as follows: Does climate change alter restoration needs in the future? And can restoration actions increase ecosystem resilience by reducing climate change effects or increasing habitat diversity? When these questions are applied to other species or environments, data needs include an understanding of current restoration plans, an assessment of how climate change might alter restoration needs, data on likely environmental effects of climate change, and a review of potential restoration actions to evaluate their likely effectiveness under future climate scenarios. Although habitat restoration can contribute to increasing resilience of salmon populations to climate change, restoration of freshwater habitats alone may not be enough for their recovery. Climate change effects are imposed on top of a long history of insults, including harvest and hatchery effects on population status, a broad array of habitat losses that have dramatically reduced salmon abundance in the western USA, and continuing changes in ocean conditions that are at least partly a result of climate change. These combined constraints have reduced wild salmon populations to the point that many have been listed under the US Endangered Species Act. Hence, recovery of these populations may also require adjustments to hatchery production and harvest levels that impact wild populations, which we did not address here. In combination, such actions will likely increase abundance and diversity in wild populations, allowing them to adapt to a changing climate. ACKNOWLEDGEMENTS We thank the Moore Foundation and National Marine Fisheries Service for providing funding for this research. We also thank Sarah Morley for her insightful review of the manuscript. Copyright © 2012 John Wiley & Sons, Ltd. REFERENCES Arnell NW. 1999. Climate change and global water resources. Global Environmental Change 9: S31–S49. Arrigoni AS, Poole GC, Mertes LAK, O’Daniel SJ, Woessner WW, Thomas SA. 2008. Buffered, lagged, or cooled? Disentangling hyporheic influences on temperature cycles in stream channels. Water Resources Research 44: W09418. doi: 10.1029/2007WR006480 Bartz KL, Lagueux K, Scheuerell MD, Beechie TJ, Haas A, Ruckelshaus MH. 2006. Translating restoration scenarios into habitat conditions: an initial step in evaluating recovery strategies for Chinook salmon (Oncorhynchus tshawytscha). Canadian Journal of Fisheries and Aquatic Sciences 63: 1578–1595. Battin J, Wiley MW, Ruckelshaus MH, Palmer RN, Korb E, Bartz KK, Imaki H. 2007. Projected impacts of climate change on salmon habitat restoration. Proceedings of the National Academy of Sciences 104: 6720–6725. Beechie T, Pess G, Roni P, Giannico G. 2008a. Setting river restoration priorities: a review of approaches and a general protocol for identifying and prioritizing actions. North American Journal of Fisheries Management 28: 891–905. Beechie TJ, Pollock MM, Baker S. 2008b. Channel incision, evolution and potential recovery in the Walla Walla and Tucannon River basins, northwestern USA. Earth Surface Processes and Landforms 33: 784–800. Beechie TJ, Pess G, Kennard P, Bilby RE, Bolton S. 2000. Modeling recovery rates and pathways for woody debris recruitment in northwestern Washington streams. North American Journal of Fisheries Management 20: 436–452. Beechie TJ, Ruckelshaus M, Buhle E, Fullerton A, Holsinger L. 2006. Hydrologic regime and the conservation of salmon life history diversity. Biological Conservation 130: 560–572. Beechie TJ, Sear D, Olden J, Pess GR, Buffington J, Moir H, Roni P, Pollock MM. 2010. Process-based principles for restoring river ecosystems. BioScience 60: 209–222. Beechie TJ, Steel EA, Roni PR, Quimby E (eds). 2003. Ecosystem Recovery Planning for Listed Salmon: An Integrated Assessment Approach for Salmon Habitat. U.S. Dept. Commerce, NOAA Tech. Memo. NMFSNWFSC-58: Seattle, WA; 183. Beechie TJ, Veldhuisen CN, Schuett-Hames DE, DeVries P, Conrad RH, Beamer EM. 2005. Monitoring treatments to reduce sediment and hydrologic effects from roads. In Monitoring Stream and Watershed Restoration, Roni P (ed). American Fisheries Society: Bethesda; 35–65. Belsky AJ, Matske A, Uselman S. 1999. Survey of livestock influences on stream and riparian ecosystems in the western United States. Journal of Soil and Water Conservation 54: 419–431. Bilby RE, Fransen BR, Bisson PA, Walter JK. 1998. Response of juvenile coho salmon (Oncorhynchus kisutch) and steelhead (O. mykiss) to the addition of salmon carcasses to two streams in southwestern Washington, U.S.A. Canadian Journal of Fisheries and Aquatic Sciences 55: 1909–1918. Bjornn TC, Reiser DW. 1991. Habitat requirements of salmonids in streams. In Influences of Forest and Rangeland Management on Salmonid Fishes and Their Habitats, Meehan WR (ed.). American Fisheries Society Special Publication 19: Bethesda; 83–138. Booth DB, Leavitt JL. 1999. Field evaluation of permeable pavement systems for improved stormwater management. Journal of the American Planning Association 65: 314–325. Booth DB, Hartley D, Jackson R. 2002. Forest cover, impervious-surface area, and the mitigation of stormwater impacts. Journal of the American Water Resources Association 38: 835–845. Boss SM, Richardson JS. 2002. Effects of food and cover on the growth, survival, and movement of cutthroat trout (Oncorhynchus clarki) in coastal streams. Canadian Journal of Fisheries and Aquatic Sciences 59: 1044–1053. Boughton DA, Gibson M, Yedor R, Kelley E. 2007. Stream temperature and the potential growth and survival of juvenile Oncorhynchus mykiss in a southern California Creek. Freshwater Biology 52: 1353–1364. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 958 T. BEECHIE ET AL. Bryant MD. 2009. Global climate change and potential effects on Pacific salmonids in freshwater ecosystems of southeast Alaska. Climatic Change 95: 169–193. Burroughs BA, Hayes DB, Klomp KD, Hansen JF, Mistak J. 2009. Effects of Stronach Dam removal on fluvial geomorphology in the Pine River, Michigan, United States. Geomorphology 110: 96–107. Cederholm CJ, Bilby RE, Bisson PA, Bumstead, TW, Fransen, BR, Scarlett WJ, Ward JW. 1997. Response of juvenile coho salmon and steelhead to placement of large woody debris in a coastal Washington Stream. North American Journal of Fisheries Management 17: 947–963. Crispell JK, Endreny TA. 2009. Hyporheic exchange flow around constructed in-channel structures and implications for restoration design. Hydrological Processes 23: 1158–1168. Crozier LG, Hendry AP, Lawson PW, Quinn TP, Mantua NJ, Battin J, Shaw RG, Huey RB. 2008. Potential responses to climate in organisms with complex life histories: evolution and plasticity in Pacific salmon. Evolutionary Applications 1: 252–270. Daly C, Gibson WP, Taylor GH, Johnson GL, Pasteris P. 2002. A knowledgebased approach to the statistical mapping of climate. Climate Research 22: 99–113. Deser C, Phillips A, Bourdette V, Teng H. 2010. Uncertainty in climate change projections: the role of internal variability. Climate Dynamics. DOI: 10.1007/s00382–010–0977–x Eaton JG, Scheller RM. 1996. Effects of climate warming on thermal habitat in streams of the United States. Limnology and Oceanography 41: 1109–1115. Ebbert JC, Roe RD. 1998. Soil erosion in the Palouse River basin: indications of improvement. US Geological Survey, USGS Fact Sheet FS-069–98. Elmore W, Beschta RL. 1987. Riparian areas: perceptions in management. Rangelands 9: 260–265. Elsner MM, Cuo L, Voisin N, Deems J, Hamlet AF, Vano JA, Mickelson KEB, Lee SY, Lettenmaier DP. 2010. Implications of 21st century climate change for the hydrology of Washington State. Climatic Change. DOI: 10.1007/s10584–010–9855–0 Furniss MJ, Roelofs TD, Yee CS. 1991. Road construction and maintenance. In Influences of Forest and Rangeland Management on Salmonid Fishes and Their Habitats, Meehan WR (ed.). American Fisheries Society Special Publication 19: Bethesda; 297–324. Gangopadhyay S, Pruitt T. 2011. West-wide climate risk assessments: biascorrected and spatially downscaled surface water projections. Bureau of Reclamation Technical Memorandum No. 86-68210-2011-01. U.S. Department of the Interior, Bureau of Reclamation, Denver, Colorado. Greene CM, Hall JE, Guilbault KR, Quinn TP. 2010. Improved viability of populations with diverse life history portfolios. Biology Letters 6: 382–386. Gresh T, Lichatowich J, Schoonmaker P. 2000. An estimation of historic and current levels of salmon production in the northeast Pacific ecosystem: evidence of a nutrient deficit in the freshwater systems of the Pacific Northwest. Fisheries 25: 15–21. Groot C, Margolis L (eds). 1991. Pacific Salmon Life Histories. University of British Columbia Press: Vancouver, Canada. Hamlet AF, Lettenmaier DP. 1999. Effects of climate change on hydrology and water resources in the Columbia River basin. Journal of the American Water Resources Association 35: 1597–1623. Harvey M, Watson C. 1986. Fluvial processes and morphological thresholds in incised channel restoration. Water Resources Bulletin 22: 359–36. Hawkins E, Sutton R. 2011. The potential to narrow uncertainty in projections of regional precipitation change. Climate Dynamics 37: 407–418. doi: 10.1007/s00382–010–0810–6 Hester ET, Doyle MW, Poole GC. 2009. The influence of in-stream structures on summer water temperatures via induced hyporheic exchange. Limnology and Oceanography 54: 355–367. Holling CS. 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics 4: 1–23. Copyright © 2012 John Wiley & Sons, Ltd. Jeffres CA, Opperman JJ, Moyle PB. 2008. Ephemeral floodplain habitats provide best growth conditions for juvenile Chinook salmon in a California river. Environmental Biology of Fishes 83: 449–458. Johnson SL. 2004. Factors influencing stream temperature in small streams: substrate effects and a shading experiment. Canadian Journal of Fisheries and Aquatic Sciences 61: 913–923. Jones JA, Grant GE. 1996. Peakflow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon. Water Resources Research 32: 959–974. Kauffman JB, Beschta RL, Otting N, Lytjen D. 1997. An ecological perspective of riparian and stream restoration in the western United States. Fisheries 22: 12–24. Kiffney PM, Bilby RE, Sanderson BL. 2005. Monitoring the effects of nutrient enrichment on freshwater ecosystems. In Monitoring Stream and Watershed Restoration, Roni P (ed.). American Fisheries Society: Bethesda; 237–266. Konrad CP, Black RW, Voss F, Neale CMU. 2008. Integrating remotely acquired and field data to assess effects of setback levees on riparian and aquatic habitats in glacial-meltwater rivers. River Research and Applications 24: 355–72. Lackey RT. 2003. Pacific Northwest salmon: forecasting their status in 2100. Reviews in Fisheries Science 11: 35–88. Lackey RT, Lach D, Duncan S. 2006. Salmon 2100: The Future of Wild Pacific Salmon. American Fisheries Society: Bethesda MD. Madej MA. 2001. Erosion and sediment delivery following removal of forest roads. Earth Surface Processes and Landforms 26: 175–190. Mahoney JM, Rood SB. 1998. Streamflow requirements for cottonwood seedling recruitment—an integrative model. Wetlands 18: 634–645. Mantua N, Tolver I, Hamlet A. 2010. Climate change impacts on streamflow extremes and summertime stream temperature and their possible consequences for freshwater salmon habitat in Washington State. Climatic Change, [online], 1–37. DOI: 10.1007/s10584–010–9845–2 McAlister LS, Peniston BE, Leibowitz SG, Abbruzzese B, Hyman JB. 2000. A synoptic assessment for prioritizing wetland restoration efforts to optimize flood attenuation. Wetlands 20: 70–83. McCaffery M, Switalski TA, Eby L. 2007. Effects of road decommissioning on stream habitat characteristics in the South Fork Flathead River, Montana. Transactions of the American Fisheries Society 136: 553–561. McClure MM, Carlson SM, Beechie TJ, Pess GR, Jorgensen JC, Sogard SM, Sultan SE, Holzer DM, Travis J, Sanderson BL, Power ME, Carmichael RW. 2008. Evolutionary consequences of habitat loss for Pacific anadromous Salmonids. Evolutionary Applications 1: 300–318. McCullough DA, Spalding S, Sturdevant D, Hicks M. 2001. Issue Paper 5. Summary of technical literature examining the physiological effects of temperature on salmonids. EPA-910-D-01-005, U.S. EPA Region 10; Portland OR. McElhany P, Ruckelshaus M, Ford M, Wainwright T, Bjorkstedt E. 2000. Viable salmonid populations and the recovery of evolutionarily significant units. U.S. Dept. Commerce, NOAA Tech, Memo. NMFS-NWFSC-42, Seattle. McRae G, Edwards CJ. 1994. Thermal characteristics of Wisconsin streams occupied by beaver: implications for brook trout habitat. Transactions of the American Fisheries Society 123: 641–656. Medina AL, Rinne JN, Roni PR. 2005. Riparian restoration through grazing management: considerations for monitoring project effectiveness. In Monitoring Stream and Watershed Restoration, Roni P (ed.). American Fisheries Society: Bethesda; 97–126. Meleason MA, Gregory SV, Bolte JP. 2003. Implications of riparian management strategies on wood in streams of the Pacific Northwest. Ecological Applications 13: 1212–1221. doi: 10.1890/02–5004 Morley SA, Garcia PS, Bennett TR, Roni P. 2005. Juvenile salmonid (Oncorhynchus spp.) use of constructed and natural side channels in River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra RESTORING SALMON HABITAT FOR A CHANGING CLIMATE Pacific Northwest rivers. Canadian Journal of Fisheries and Aquatic Sciences 62: 2811–2821. Mote PW, Salathé EP. 2010. Future climate in the Pacific Northwest. Climatic Change. DOI: 10.1007/s10584–010–9848–z Mote PW, Parson EA, Hamlet AF, Keeton WS, Lettenmaier D, Mantua N, Miles EL, Peterson DW, Peterson DL, Slaughter R, Snover AK. 2003. Preparing for climatic change: the water, salmon, and forests of the Pacific Northwest. Climatic Change 61: 45–88. Myers JM, Kope RG, Bryant GJ, Teel D, Lierheimer LJ, Wainwright TC, Grant WS, Waknitz FW, Neely K, Lindley ST, Waples RS. 1998. Status review of Chinook salmon from Washington, Idaho, Oregon, and California. U.S. Department of Commerce, NOAA Technical Memorandum NMFS-NWFSC-35. Nelitz M, Wieckowski K, Prckard D, Pawley K, Marmorek DR. 2007. Helping Pacific Salmon Survive the Impact of Climate Change on Freshwater Habitats. Pacific Fisheries Resource Conservation Council: Vancouver, B.C. Olden JD, Poff NL. 2004. Ecological processes driving biotic homogenization: testing a mechanistic model using fish faunas. Ecology 85: 1867–1875. Omernik JM, Bailey RG. 1997. Distinguishing between watersheds and ecoregions, Journal of the American Water Resources Association 33: 935–949. Opperman JJ, Luster R, McKenney BA, Roberts M, Meadows AW. 2010. Ecologically functional floodplains: connectivity, flow regime, and scale. Journal of the American Water Resources Association 46: 211–226. Pess GR, McHenry ML, Beechie TJ, Davies J. 2008. Biological impacts of the Elwha River dams and potential salmonid responses to dam removal. Northwest Science 82: 72–90. Pess GR, Morley SA, Hall JL, Timm RK. 2005. Monitoring floodplain restoration. In Monitoring Stream and Watershed Restoration, Roni P (ed.). American Fisheries Society: Bethesda; 127–166. Poff NL, Allen JD, Bain MB, Karr JR, Prestegaard KL, Richter BD, Sparks RE, Stromberg JC. 1997. The natural flow regime. BioScience 47: 769–784. Poff NL, Richter BD, Arthington AH, Bunn SE, Naiman RJ, Kendy E, Acrement M, Apse C, Bledsoe BP, Freeman MC, Henriksen J, Jacobsen RB, Kennen JG, Merritt DM, O’Keeffe JH, Olden JD, Rovers K, Tharme RE, Warner A. 2010. The ecological limits of hydrologic alteration (ELOHA): a new framework for developing regional environmental flow standards. Freshwater Biology 55: 147–170. Pollock MM, Beechie TJ, Chan S, Bigley R. 2005. Monitoring and evaluating riparian restoration efforts. In Monitoring Stream and Watershed Restoration, Roni P (ed.). American Fisheries Society: Bethesda; 67–96. Pollock MM, Beechie TJ, Jordan CE. 2007. Geomorphic changes upstream of beaver dams in Bridge Creek, in incised stream channel in the interior Columbia River basin, eastern Oregon. Earth Surface Processes and Landforms 32: 1174–1185. Pollock MM, Beechie TJ, Liermann M, Bigley RE. 2009. Stream temperature relationships to forest harvest in the Olympic Peninsula, Washington. Journal of the American Water Resources Association 45: 141–156. Pollock MM, Heim M, Werner D. 2003. Hydrologic and geomorphic effects of beaver dams and their influence on fishes. In The Ecology and Management of Wood in World Rivers. Gregory SV, Boyer K, Gurnell A (eds). American Fisheries Society: Bethesda, MD; 213–233. Ponce VM, Lindquist DS. 1990. Management of baseflow augmentation: a review. Water Resources Bulletin 26: 259–268. Poole GC, Berman CH. 2001. An ecological perspective on in-stream temperature: natural heat dynamics and mechanisms of human-caused thermal degradation. Environmental Management 27: 787–802. Poole GC, O’Daniel SJ, Jones KL, Woessner WW, Bernhardt ES, Helton AM, Stanford JA, Boer BR, Beechie TJ. 2008. Hydrologic spiraling: the role of multiple interactive flow paths in stream ecosystems. River Research and Applications 24: 1018–1031. Copyright © 2012 John Wiley & Sons, Ltd. 959 Quinn TP. 2005. The Behavior and Ecology of Pacific Salmon and Trout. University of Washington Press: Seattle. Reid LM, Dunne T. 1984. Sediment production from forest road surfaces. Water Resources Research 20: 1753–1761. Richter A, Kolmes SA. 2005. Maximum temperature limits for Chinook, coho, and chum salmon, and steelhead trout in the Pacific Northwest. Reviews in Fisheries Science 13: 23–49. Rieman BE, Isaak D, Adams S, Horan D, Nagel D, Luce C, Myers D. 2007. Anticipated climate warming effects on bull trout habitats and populations across the interior Columbia River basin. Transactions of the American Fisheries Society 136: 1552–1565. doi: 10.1577/T07–028.1 Robison, GE, Runyon J, Andrus C. 1999. Oregon Department of Forestry cooperative stream temperature monitoring: project completion report for 1994–1995. Forest Practices Technical Report 2, Oregon Department of Forestry, Forest Practices, Salem, OR. 47. Roni P, Quinn TP. 2001. Density and size of juvenile salmonids in response to placement of large woody debris in western Oregon and Washington streams. Canadian Journal of Fisheries and Aquatic Sciences 58: 282–29. Roni P, Beechie TJ, Bilby RE, Leonetti FE, Pollock MM, Pess GR. 2002. A review of stream restoration techniques and a hierarchical strategy for prioritizing restoration in Pacific Northwest watersheds. North American Journal of Fisheries Management 22: 1–20. Roni P, Hanson K, Beechie T. 2008. Global review of the physical and biological effectiveness of stream habitat rehabilitation techniques. North American Journal of Fisheries Management 28: 856–890. Scheuerell MD, Hilborn R, Ruckelshaus MH, Bartz KK, Lagueux KM, Haas AD, Rawson K. 2006. The Shiraz model: a tool for incorporating anthropogenic effects and fish–habitat relationships in conservation planning. Canadian Journal of Fisheries and Aquatic Sciences 63: 1596–1607. Schindler DE, Hilborn R, Chasco B, Boatright CP, Quinn TP, Rogers LA, Webster MS. 2010. Population diversity and the portfolio effect in an exploited species. Nature 465: 609–612. Shared Strategy Development Committee. 2007. Puget Sound Salmon Recovery Plan. National Marine Fisheries Service, Portland, Oregon. (http://www.nwr.noaa.gov/Salmon-Recovery-Planning/Recovery-Domains/ Puget-Sound/PS-Recovery-Plan.cfm). Shields FD Jr, Knight SS, Cooper CM. 1995. Rehabilitation of watersheds with incising channels. Water Resources Bulletin 31:971–982. Sidle RC, Pierce AJ, O’Loughlin CL. 1985. Hillslope Stability and Land Use. Water Resources Monograph Series Volume 11. American Geophysical Union: Washington, D.C. Solazzi MF, Nickelson TE, Johnson SL, Rodgers JD. 2000. Effects of increasing winter rearing habitat on abundance of salmonids in two coastal Oregon streams. Canadian Journal of Fisheries and Aquatic Sciences, 57: 906–91. Sommer TR, Nobriga ML, Harrell WC, Batham W, Kimmerer WJ. 2001. Floodplain rearing of juvenile Chinook salmon: evidence of enhanced growth and survival. Canadian Journal of Fisheries and Aquatic Sciences 58: 325–333. Sparks RE, Nelson JC, Yin Y. 1998. Naturalization of the flood regime in regulated rivers. BioScience 48: 706–720. Stanford JA, Ward JV, Liss WJ, Frissell CA, Williams RN, Lichatowich JA, Coutant CC. 1996. A general protocol for restoration of regulated rivers. Regulated Rivers: Research & Management, 12: 391–413. U.S. Environmental Protection Agency. 2003. EPA Region 10 guidance for Pacific Northwest state and tribal temperature water quality standards. EPA-910-B-03-002. Region 10 Office of Water, Seattle, Washington. Waples RS, Beechie TJ, Pess GR. 2009. Evolutionary history, habitat disturbance regimes, and anthropogenic changes: what do these mean for resilience of Pacific salmon populations? Ecology and Society 14: 3. [online] http://www.ecologyandsociety.org/vol14/iss1/art3/. River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra 960 T. BEECHIE ET AL. Waples RS, Pess GR, Beechie T. 2008. Evolutionary history of Pacific salmon in dynamic environments. Evolutionary Applications 1: 189–206.h Welty JT, Beechie TJ, Sullivan K, Hyink DM, Bilby RE, Andrus C, Pess G. 2002. Riparian aquatic interaction simulator (RAIS): a model of riparian forest dynamics for the generation of large woody debris and shade. Forest Ecology and Management 162: 299–318. Wendt RC, Burwell RE. 1985. Runoff and soil losses for conventional, reduced, and no-till corn. Journal of Soil and Water Conservation 40: 450–454. Copyright © 2012 John Wiley & Sons, Ltd. Whited D, Kimball J, Lucotch J, Maumenee N, Wu H, Chilcote S, Standford J. in press A riverscape analysis tool developed to assist wild salmon conservation. Fisheries. Wipfli MS, Baxter CV. 2010. Linking ecosystems, food webs, and fish production: subsidies in salmonid watershed. Fisheries 35: 373–387. Wu H, Kimball JS, Mantua N, Stanford JA. 2011. Automated upscaling of river networks for macroscale hydrological modeling. Water Resources Research 47: W03517. doi: 10.1029/WR008871 River Res. Applic. 29: 939–960 (2013) DOI: 10.1002/rra