Chapter 7

Sector Impacts and Adaptation

This chapter examines key issues related to climate change impacts and adaptation within seven sectors in Canada—forestry, agriculture, fisheries, energy, mining, transportation, and tourism—as well as the status of corporate adaptation.

Recommended citation

Lemmen, D., Lafleur, C., Chabot, D., Hewitt, J., Braun, M., Bussière, B., Kulcsar, I., Scott, D. and Thistlethwaite, J. (2021): Sector Impacts and Adaptation; Chapter 7 in Canada in a Changing Climate: National Issues Report, (ed.) F.J. Warren and N. Lulham; Government of Canada, Ottawa, Ontario.

Coordinating lead authors

  • Donald S. Lemmen, PhD
  • Catherine Lafleur (PhD, Natural Resources Canada)

Lead and contributing authors

Forestry: 

  • Catherine Lafleur, PhD, Natural Resources Canada
  • James MacLellan, PhD, University of Toronto

Fisheries: 

  • Denis Chabot, PhD, Fisheries and Oceans Canada
  • Nancy Shackell, PhD, Fisheries and Oceans Canada
  • Helen Gurney-Smith, PhD, Fisheries and Oceans Canada
  • Jackie King, PhD, Fisheries and Oceans Canada

Agriculture: 

  • Jamie Hewitt, Agriculture and Agri-food Canada

Energy: 

  • Marco Braun, PhD, Ouranos

Mining: 

  • Bruno Bussière, PhD, Université du Québec en Abitibi-Témiscamingue
  • Émilie Bresson, PhD, Polytechnique Montréal

Transportation: 

  • Irene Kulcsar, Transport Canada
  • Elizabeth Smalley, Transport Canada
  • Matt Osler, City of Surrey

Tourism: 

  • Daniel Scott, PhD, University of Waterloo
  • Jackie Dawson, PhD, University of Ottawa

Corporate Adaptation: 

  • Jason Thistlethwaite (PhD, University of Waterloo)
Figure 7.1-a
View details
Figure 7.1-b
View details
Figure 7.2
View details
Figure 7.3
View details
Figure 7.4
View details
Figure 7.5
View details
Figure 7.6
View details
Figure 7.7
View details
Figure 7.8
View details
Figure 7.9
View details
Figure 7.10
View details
Figure 7.11
View details
Figure 7.12
View details
Figure 7.13
View details
Figure 7.14
View details
Figure 7.15
View details
Figre 7.16
View details
Figure 7.17
View details

Key Messages

Climate change affects almost every economic sector in Canada

Virtually every sector of the Canadian economy is affected, directly and/or indirectly, by climate change. Assessments of risks and opportunities that consider connections within and between sectors can help establish priorities for investments in adaptation actions.

Sustainable forest management is challenged by wildfires and pest outbreaks

The forest sector is dealing with a wide range of climate change risks, from pest outbreaks to wildfire and long-term species shifts. The impacts of extreme events, such as wildfire, highlight the need for actions that build more resilient forests and communities, and contribute to climate change mitigation.

Changes to ocean health are affecting fisheries and associated livelihoods

Changes in ocean temperature and chemistry are already affecting fish populations. While some future impacts will be positive, many will present challenges to harvesters’ economic livelihoods and regulators’ responsibility for sustaining ocean health. Effective management depends upon realistic models of the future abundance and distribution of commercial species in response to both climatic and non-climatic stressors.

Climate change brings benefits and threats to the agriculture sector

Climate change brings both opportunities and challenges to Canada’s agricultural sector. Longer growing seasons and the potential to grow crops farther north may benefit agriculture, while changes in water availability, extreme weather events, and pests and diseases will present challenges. Adaptation actions that enhance climate resilience and consider the linkages between agriculture and interconnected sectors, such as water management and transportation, will benefit both local sustainability and global food security.

Climate change brings new environmental challenges for mining

Impacts on the chemical and physical stability of tailings containment and reclamation structures are among the greatest climate-related challenges to the Canadian mining industry. Failure of such structures can lead to severe environmental contamination and present risks for surrounding communities and ecosystems. Considering long-term climate change at the design phase of mining projects is necessary to reduce these risks.

Each link of the energy value chain can be vulnerable to climate change

Changing climate affects energy demand and the full energy value chain, from exploration and production through to transmission and distribution. Climate risks can be integrated into current business planning by considering co-benefits, no-regret options and incremental approaches. Climate resilience needs to be a key consideration in converting to low-carbon energy systems.

Extreme weather events impact transportation, disrupting supply chains 

Road, rail, marine and air transportation in Canada are vulnerable to extreme weather events and slow-onset climate change, with major disruptions having significant economic and social impacts. To fully assess these impacts, linkages between transportation systems, and between transportation modes and a wide range of other economic sectors, need to be accounted for. Coordinating adaptation responses across jurisdictions and sectors will benefit transportation asset owners, operators and those dependent on vulnerable supply chains and corridors.

Climate change is leading to transformational changes in tourism

All tourism destinations need to adapt to climate change impacts on tourism assets and altered competitiveness within the highly interconnected tourism economy. While Canadian tourism competitiveness is expected to increase under climate change, the specific market and regional implications of this change for national competitiveness remain under-researched. Winter and northern tourism and recreation are particularly sensitive to climate variability, and transformational changes in ski, snowmobile and Arctic cruise tourism are expected.

Increased private sector involvement will accelerate adaptation across sectors

Despite growing awareness of climate change impacts, there is no widespread evidence of corporate adaptation in Canada. When adaptation does occur, it tends to focus on short-term actions to address physical risks, such as disruptions in construction and interruptions in supply chains. Increased involvement of the private sector would accelerate adaptation in Canada as a whole.

7.1

Introduction

Virtually every sector of the Canadian economy is affected, directly and/or indirectly, by climate change. Assessments of risks and opportunities that consider connections within and between sectors can help establish priorities for investments in adaptation actions. 

It is increasingly evident that climate change is impacting Canada’s economy as a whole, and that these impacts will increase in future (Canadian Council of Academies, 2019). Adaptation will be necessary across all sectors to limit climate risks and, in some cases, to benefit from new opportunities. Proactive adaptation enables innovation and growth, and can enhance economic competitiveness (Kovacs and Thistlethwaite, 2014). Very few studies have undertaken quantitative economic analysis of the impacts of specific climate scenarios on Canadian business and industry (see Costs and Benefits of Climate Change Impacts and Adaptation chapter; Eyzaguirre, 2016), but research elsewhere demonstrates that costs accelerate with continued warming (IPCC, 2014). Additionally, at higher rates of climate change, adaptation options become increasingly limited, presenting critical risks ranging from local economic viability to global food security (IPCC, 2019, 2018, 2014).

From a Canadian perspective, most sectoral research on impacts and adaptation has focused on food and natural resource sectors where climate change directly affects primary production, such as agriculture, fisheries, forestry and hydroelectricity generation (Warren and Lemmen, 2014). The economic significance of these sectors is amplified at both the local scale, with many Canadian communities deriving 80% or more employment income from these sectors (see Rural and Remote Communities chapter), and on the global scale, where Canada is among the world leaders in agriculture, forestry and mineral exports.

Less attention has been paid to other sectors of Canada’s economy, with the exception of human health (Berry et al., 2014). However, there is growing recognition that climate-related health and social impacts on communities and workers, as well as climate impacts on supply chains and other infrastructure, represent significant material and financial risks throughout the economy (Canadian Council of Academies, 2019). As a result, a growing body of literature on these other sectors is emerging (see Climate Disclosure, Litigation and Finance chapter).

Previous sectoral assessments conducted in Canada, such as Warren and Lemmen (2014)—particularly chapters 4, 5 and 6—highlight:

  • Vulnerabilities to both extreme weather events and to slow-onset climate changes;
  • Amplified impacts in northern and remote communities;
  • Opportunities that climate change presents for many sectors, in addition to the changing nature of climate risks;
  • Increased implementation of climate adaptation measures and expanded engagement of industry, governments and civil society; enhancing both social and economic resilience;
  • Processes that can help advance adaptation actions, including risk disclosure, environmental assessment and sustainable management reporting;
  • Interdependencies between sectors, with transportation systems being particularly important; and
  • A lack of information related to indirect impacts of climate change, including those related to consumer demand, supply chains, real estate or other assets, legal liability and government regulation.

This chapter builds on the findings of Warren and Lemmen (2014) and other relevant Canadian assessments (e.g., Palko and Lemmen, 2017) by examining key climate change impacts and adaptation in seven sectors of Canada’s economy—forestry, agriculture, fisheries, energy, mining, transportation and tourism—, as well as broad perspectives on corporate adaptation. Issues related to human health are not included in this chapter, as they are addressed in a separate assessment report (see Health of Canadians in a Changing Climate Report). The authors focused on a limited number of priority issues identified through assessment of the breadth of available knowledge. As a result, the subsequent sections of this chapter do not provide a comprehensive assessment of climate change impacts and adaptation responses within each sector, but rather focus on topics where knowledge has advanced recently and where the assessed knowledge relates directly to the ongoing decision-making process.

7.2

Sustainable forest management is challenged by wildfires and pest outbreaks

The forest sector is dealing with a wide range of climate change risks, from pest outbreaks to wildfire and long-term species shifts. The impacts of extreme events, such as wildfire, highlight the need for actions that build more resilient forests and communities, and contribute to climate change mitigation.

Climate variability and extreme weather events associated with climate change are challenging forest management by limiting access to forest resources and increasing operational costs. More frequent wildfires and forest pest infestations are constraining local timber supplies and impacting the social and economic well-being of forest communities. In addition, the forest sector is facing longer-term climate impacts, such as changes in tree species composition, stand structure, productivity and health. In response to these and other challenges, such as changing social values and market demand, forest companies are addressing multiple environmental, economic and social needs simultaneously. Adaptive, risk-based management approaches that apply research, monitoring and evaluation will help inform future management policies to promote healthy, resilient forests and enhance carbon storage.

7.2.1

Introduction

The Canadian forest sector is sensitive to the impacts of climate on ecosystem goods and services. In the short term, projected increases in temperature will likely surpass the potential moderating effects of increasing precipitation on fire weather (Zhang et al., 2019), leading to an increased risk of wildland fire and drought (Boucher et al., 2018; Boulanger et al., 2017; Flannigan et al., 2009). Climate change is also a critical driver of progressive disturbances, such as pest infestations, which influence the likelihood of immediate disturbance events, while also affecting long-term forest structure and composition (Sulla-Menashe et al., 2018; van Lierop et al., 2015; Price et al., 2013; Sturrock et al., 2011; Burton, 2010). The cumulative effects of these changes decrease the health and resilience of Canada’s forests, constraining timber supply and increasing risks to the forest sector (Boucher et al., 2018; Taylor et al., 2017; McKenney et al., 2016; Gauthier et al., 2014; Price et al., 2013; Coulombe et al., 2010; Williamson et al., 2009). Better understanding of these projected changes will help the forest sector better prepare for both risks and opportunities.

Past management responses to climate-related impacts in the forestry sector have tended to be reactive, as exemplified by responses to the mountain pine beetle outbreak in Western Canada (Jones and Preston, 2011; Bentz et al., 2010; Williamson et al., 2009). Recently, forest managers, policy specialists and researchers have developed an array of knowledge resources, tools, and protocols to help practitioners and stakeholders take a proactive approach to managing the impacts of a changing climate and related environmental and socioeconomic stressors. The phases of adaptive management are reviewed in a series of past assessments (Gauthier et al., 2014; Lemmen et al., 2014; Price et al., 2013; NRTEE, 2011; Williamson et al., 2009), recognizing that no single “road map” exists to guide the implementation of adaptation responses (Samy et al., 2015). This section builds upon past assessments by focusing on wildland fire in the context of recent events.

7.2.2

Impacts of wildfire

While forest fires are a natural and essential element of forest ecology, there is growing awareness of the dramatic impact that wildfires have on Canadians. As climate changes occur, fire regimes are changing, often with increasing frequency, severity and size (Mori and Johnson, 2013; Flannigan et al., 2009). Non-climate factors, such as forest condition, forest management practices, land cover (Marchal et al., 2017) and cumulative disturbances are also important considerations in explaining these increases. The fire season is becoming longer, starting earlier in the spring and ending later in the fall (Zhang et al., 2019; Hanes et al., 2018; Jolly et al., 2015; Flannigan et al., 2013), with more frequent fires (expressed by a shorter fire return interval) expected throughout this century (see Figure 7.1).

Figure 7.1-a
Baseline (1981–2010) RCP2.6 (2011–2040) RCP2.6 (2041–2070) RCP2.6 (2071–2100)
Figure 7.1-a

Interactive maps of baseline and projected fire likelihood in Canada, presented in terms of fire return interval (years), for a low greenhouse gas (GHG) emissions scenario (RCP2.6). Shorter fire return intervals (in brown and light brown) represent a higher fire likelihood, and longer cycles (in yellow, light green and dark green) represent a low fire likelihood. The maps show local fire likelihood (pixels represent a 250 m2 resolution) using the homogeneous fire regime zones of Boulanger et al. (2014) and display fire selectivity based on vegetation composition and stand age at the pixel level.

Source

Boulanger et al., 2014.

Figure 7.1-b
Baseline (1981–2010) RCP8.5 (2011–2040) RCP8.5 (2041–2070) RCP8.5 (2071–2100)
Figure 7.1-b

Interactive maps of baseline and projected fire likelihood in Canada, presented in terms of fire return interval (years), for a high greenhouse gas (GHG) emissions scenario (RCP8.5). Shorter fire return intervals (in brown and light brown) represent a higher fire likelihood, and longer cycles (in yellow, light green and dark green) represent a low fire likelihood. The maps show local fire likelihood (pixels represent a 250 m2 resolution) using the homogeneous fire regime zones of Boulanger et al. (2014) and display fire selectivity based on vegetation composition and stand age at the pixel level.

Source

Boulanger et al., 2014.

Ecological impacts following wildfires include changes to forest stand structure, such as age class distribution and species composition (Price et al., 2013; Brown and Johnstone, 2012; Lynch, 2004). Shorter return intervals, in combination with growing fire severity, increase the risk of regeneration failure (see Figure 7.2; Whitman et al., 2019), and transitioning of forested areas to non-forested terrain (Boiffin and Munson, 2013; Price et al., 2013; Brown and Johnstone, 2012). Loss of tree cover on the landscape can lead to flooding and mass wasting in some areas (Bladon, 2018; Creed et al., 2016). Changes to fire regimes threaten not only timber supply (quality, quantity and tree species mix), but also the provision of ecosystem services, such as biodiversity, habitat for species at risk, carbon storage, water quality and water quantity (see Ecosystem Services chapter; Gauthier et al., 2014; Price et al., 2013). While it is difficult to predict how such changes will manifest locally, analysis of past mountain pine beetle infestations suggests that creating more resilient forests results in multiple benefits (Dymond et al., 2015).

Figure 7.2

-

Photograph showing a post-fire clearing with scattered charred wood and ash. Trees can be seen in the distance.
Figure 7.2

Photo of a post-fire landscape at risk of regeneration failure: a site in the Northwest Territories, one year after a 2014 wildfire. This site had previously burned in 2004 and the short (10-year) interval between fires has led to a complete lack of tree recruitment. Poor regeneration is attributed to the absence of both seedlings and soil organic matter prior to the fire.

Source

Photo courtesy of Natural Resources Canada.

There is a growing awareness of the social impacts of wildland fire (McGee et al., 2015; Gill et al., 2013), including the numbers of wildfire evacuees (see Figure 7.3). Evacuations cause physical and mental health issues, disrupt the lives of evacuees and create economic stress on individuals and communities (McCaffrey et al., 2015; Beverly and Bothwell, 2011; Marshall et al., 2007; Morton et al., 2003). With more wildfires, health issues due to smoke are increasing well beyond the immediate vicinity of the fire (Liu et al., 2016; Reid et al., 2016; Finlay et al., 2012), and visibility issues are impacting transportation (Goodrick et al., 2013). To date, these impacts have disproportionately affected small and Indigenous communities: one third of all wildfire evacuees are Indigenous and over half of smoke-related evacuations involve Indigenous communities (see Rural and Remote Communities chapter; Sankey, 2018; Scharbach and Waldram, 2016; Christianson, 2015).

Figure 7.3

-

Bar graph showing the number of wildfire evacuees in Canada from 1980 to 2020. The graph shows an increase in evacuees over time, with spikes in 1989, 2003, 2009 and 2011. The largest number of evacuees are recorded in 2016 and 2017.
Figure 7.3

Number of wildfire evacuees in Canada (1980‒2020). Many factors, including fire frequency, size and location relative to population density, influence the number of evacuees. British Columbia, Alberta, and Ontario have had the most evacuations from 1980 to 2020.

Source

Government of Canada, 2020.

Economic impacts associated with wildland fire are far-reaching. Direct costs, which include fire management and suppression activities (Rijal et al., 2018; Wotton et al., 2010), have increased from an average of $290 million per year in the early 1970s to about $1 billion annually in recent years (Natural Resources Canada, 2019, 2017). These costs are projected to further increase by over 100% by the end of the century under a high emissions scenario (Hope et al., 2016). Maintaining current effective levels of fire suppression appears to be unsustainable (Wotton et al., 2017; Hope et al., 2016). Climate change impacts and loss of timber supply have cascading economic impacts on the forest sector that are difficult to quantify (Ochuodho and Lantz, 2014). Other sectors also face direct costs from wildfires, including the following: evacuation expenditures; damage to infrastructure; shutdown of businesses and industries; insurance costs; and loss of forest ecosystem services (McGee et al., 2015; Gauthier et al., 2015; Peter et al., 2006). The increased emphasis on forests in the low-carbon economy highlights the importance of quantifying the costs associated with the release of greenhouse gases. There are also health costs associated with smoke and evacuation stress (Reid et al., 2016; McCaffrey et al., 2015; Beverly and Bothwell, 2011; Morton et al., 2003).

Catastrophic insurable losses due to extreme events such as wildfire are increasing. The Fort McMurray wildfire in 2016 was the largest insurance event in Canada, assessed at almost $4 billion (see Case Story 7.1; Insurance Bureau of Canada, 2019, 2016; Swiss Re Institute, 2018; Statistics Canada, 2017). This assessment vastly underestimates overall costs, given that uninsured costs for repairing and rebuilding for governments and homeowners can be three to four times those of private insurance companies (Dixon et al., 2018). The link between climate change and extreme events is clear (e.g., IPCC, 2012), with recent attribution analysis in Canada indicating that climate change has increased the likelihood of extreme dry conditions, extreme wildfire risk and the length of fire seasons (Kirchmeier-Young et al., 2017; Tett et al., 2017). All of these factors are relevant to the Fort McMurray wildfire (Zhang et al., 2019).

7.2.3

Adaptation

Since the risks associated with wildfire and other climate impacts have effects beyond the forest sector, implementing adaptation measures requires the involvement of stakeholders across multiple sectors (Furness and Nelson, 2016; Nelson et al., 2016). Some adaptation options mentioned below are specific to wildfire, while others contribute to increasing the overall resilience of forests and forest communities (see Edwards et al., 2015 and Gauthier et al., 2014 for compendiums of adaptation measures).

Various adaptation options are available to reduce the risk and impacts of forest disturbances such as wildfire (Leduc et al., 2015; McGee et al., 2015; Blackwell et al., 2008). For example, fire risk can be reduced through active fuel management, involving thinning, debris removal and prescribed burning (Astrup et al., 2018; Schroeder, 2010; Ohlson et al., 2006; Spittlehouse, 2005), and adjusting harvest schedules to favour older and insect-damaged stands (Dymond et al., 2015; Raulier et al., 2014). Regeneration planning could include a greater proportion of more fire-tolerant species and deciduous trees (Bernier et al., 2016).

At the local level, FireSmart activities (Hirsch et al., 2001), such as creating fire breaks around communities, building with fire-resistant materials, and cleaning up debris around properties to reduce fuel load, help increase resilience to wildland fire (FireSmart Canada, 2019a, 2019b, 2018; Spittlehouse, 2005). Communities are actively improving their emergency preparedness by creating plans for culturally-sensitive evacuations and hosting of evacuees (Scharbach and Waldram, 2016; Beardy’s and Okemasis’ Cree Nation, n.d.).

While a number of adaptation responses have been developed to reduce the costs resulting from the effects of wildfire on timber supply (Rijal et al., 2018; Leduc et al., 2015; Raulier et al., 2014; Raulier et al., 2013); addressing the broader economic impacts of wildland fires extends far beyond the forest sector (Orwig, 2016). For example, the insurance sector is examining the impact of wildland fires on mining, forestry, energy, agriculture, transportation and utilities, and is collaborating with organizations such as FireSmart (Hirsch et al., 2001) to incentivize actions to reduce fire risks around communities and infrastructure.

The status of implementation provides a benchmark for measuring adaptation progress in the forest sector. The Canadian Council of Forest Ministers’ Climate Change Task Force produced a series of nine interrelated reports (Canadian Council of Forest Ministers, n.d.), including a practical guidebook (Edwards et al., 2015), to support mainstreaming of climate change into forest management planning. Early adopters are using this guidebook to assess vulnerabilities and rank adaptation options based on current and future ability for implementation (Andrews-Key, 2018; Gatin and Johnston, 2017). Guidance documents have also been produced to help private woodlot owners adapt to a changing climate (Ontario Woodlot Association, 2015). National and regional climate change initiatives (e.g., Gatin and Johnston, 2017) include climate change strategies and action plans, and updated policies and regulations. Supportive research has placed an emphasis on integrated assessment approaches, drawing on expertise from a range of disciplines to complement ecological research with socioeconomic analyses.

7.2.4

Moving forward

Climate change is already affecting the forest sector, especially through extreme events such as wildfire. These effects are expected to continue and intensify, requiring greater efforts to implement adaptation measures. While no single road map exists to guide adaptation implementation, regional risk assessments in Canada have highlighted the need for integration of ecological, social and economic pressures beyond the forest sector. Integration and the capture of synergies across sectors can be greatly facilitated by adopting common data sources and scenarios (Environment and Climate Change Canada, 2018) as well as risk-based methods and frameworks (see Section 7.9; Johnston et al., 2020; ISO, 2018; Daniel et al., 2017; Calkin et al., 2014; Jones and Preston, 2011). Implementation of actions that enhance the climate resilience of forests is important for supporting the ecological, social and economic services that forests provide.

7.3

Changes to ocean health are affecting fisheries and associated livelihoods

Changes in ocean temperature and chemistry are already affecting fish populations. While some future impacts will be positive, many will present challenges to harvesters’ economic livelihoods and regulators’ responsibility for sustaining ocean health. Effective management depends upon realistic models of the future abundance and distribution of commercial species in response to both climatic and non-climatic stressors.

Climate change is increasing water temperatures and acidity, decreasing oxygen content and increasing the salinity of the world’s oceans. Of these environmental variables, changes in temperature, dissolved oxygen and acidification will be responsible for most of the direct impacts on fisheries and aquaculture (farming of fish or shellfish) in Canada over the next few decades. These variables are driving, and will continue to drive, changes in distribution, productivity, reproduction and timing of seasonal events (e.g., moulting, migration, spawning and hatching) for many aquatic species. Extreme events, particularly abrupt warming events lasting many months, are also important considerations in aquatic management. Planning of adaptation measures needs to account for the complex interactions between various climate stressors, as well as non-climatic stressors such as fishing pressures.

7.3.1

Introduction

The critical role and climatic sensitivity of fisheries and aquaculture in food security at the global scale, and in local-scale resilience of coastal communities, are well established (Bindoff et al., 2019; Lemmen et al., 2016; Porter et al., 2014). An overview of climate change impacts on food production in Canada, and related adaptation actions, was presented by Campbell et al. (2014). For the fisheries sector, the researchers concluded that regional impacts from physical habitat changes, invading species, and societal responses will determine future patterns of use and overall economic implications. Aquaculture was noted as having the greatest scope for adaptation measures, making it less vulnerable and better positioned to take advantage of opportunities compared to capture fisheries, while traditional subsistence fisheries were deemed to be particularly vulnerable.

This section builds on the findings of Campbell et al. (2014) by focusing on near-term challenges for the sustainable management of marine ecosystems arising from physical habitat changes (e.g., warming, acidification, and lowering or depletion of dissolved oxygen), and highlights how climate change is being incorporated into scientific advice that informs fisheries and aquaculture management decisions. Greater details on the physical impacts of changing climate are described in Chapter 7 of Canada’s Changing Climate Report (Greenan et al., 2019a). Impacts on freshwater recreational and commercial fishing and aquaculture are not discussed here, but it is clear that increased water temperatures and decreased pH and oxygen are already impacting many freshwater ecosystems. Additionally, it is noted that anadromous species will be subjected to negative impacts during their stay in freshwater, and therefore even marine fisheries will be affected by changes in freshwater ecosystems.

7.3.2

Water temperature

With a few exceptions, the internal temperature of fishes and crustaceans closely matches that of the water in which they live. As a result, temperature exerts a strong influence on their physiology (e.g., metabolism and growth) and aquatic animals are adapted to a species-specific temperature range. When mobile species encounter temperatures approaching their upper tolerance limit, they tend to move towards cooler, more optimal temperatures. Off the west coast of British Columbia, a persistent warming event from 2013 to 2015 had several ecosystem impacts (see Case Story 7.2).

The general warming trends documented for the last century over most of the world’s oceans have already resulted in distribution shifts (Mueter and Litzow, 2008). Changes in distribution of commercial species will cause changes in the location and success of fishing effort. Eventually, new species may replace old ones in the fisheries. Smaller movements or changes in abundance can also be observed at the regional scale, as fishes and crustaceans adjust their distribution to local changes in temperature (see Case Story 7.3).

Attribution of impacts must take into account the multiple climatic and non-climatic factors that affect species distributions. For example, fishing pressure, population size and bottom temperature all affect, to varying degrees, changes in the distribution of groundfish (Adams et al., 2018). This illustrates the importance of examining the joint role of fishing and climate in the distribution of fish stocks in order to provide sound scientific advice for management.

7.3.3

Dissolved oxygen

Fishes and invertebrates require oxygen to live, although they differ in their sensitivity to lack of dissolved oxygen (hypoxia). Each species has a minimum level of oxygen required for survival, with severe hypoxia resulting in habitat loss and changes in distribution for mobile species, and increased mortality for immobile (sessile) species (Breitburg et al., 2018). Moderate hypoxia limits the amount of energy that animals can spend (Claireaux and Chabot, 2016), which usually translates into reduced feeding and growth rates for individual animals (Hrycik et al., 2017; Townhill et al., 2016), and ultimately reduced productivity for affected populations.

Climate change will exacerbate hypoxia because warming reduces the solubility of oxygen, increases biological oxygen consumption, and reduces ventilation in the world’s oceans, which increases residence time of deep water. Increased residence time means that fishes, invertebrates and bacteria have more time to remove oxygen through respiration, and deoxygenation becomes more pronounced (Breitburg et al., 2018). Modelling analysis of the impacts of increasing temperature and hypoxia in the Gulf of St. Lawrence indicates significant declines in biomass production (see Case Story 7.4).

7.3.4

Acidification

Ocean acidification is caused by higher concentrations of CO2 in the air, increasing the dissolution of CO2 in the oceans to create carbonic acid and exceeding the ocean’s ability to buffer against the increase in acid concentration. One measure of acidity is pH, with lower values representing a higher level of acidity. Since the Industrial Revolution, ocean pH has declined globally on average by 0.1 pH units, representing an increase of 26% in the number of hydrogen ions, and is predicted to further decrease by 0.4 pH units by 2100 under a high-emissions scenario (RCP8.5) (Ciais et al., 2013). Many studies have shown that the important commercial species of fish, shellfish, crustaceans and gastropods can suffer negative effects when exposed to acidified water (Alin et al., 2019; Parker et al., 2013; Kroeker et al., 2010). Ocean acidification has already had a massive impact on shellfish, where changes in seawater carbonate chemistry alter  the ways in which juvenile (larval) shellfish build their shells, and can cause shell dissolution, deformities, slow growth and even death (Waldbusser et al., 2015; Gazeau et al., 2013). Ocean acidification impacts on shellfish aquaculture are evident through clear linkages between high CO2 seawater concentrations and poor larval development of Pacific oysters in the United States shellfish industry (Barton et al., 2012). Shellfish hatcheries generating juvenile oysters for field outplanting were severely affected, with losses of money and jobs (Ekstrom et al., 2015). In the Atlantic, research is underway to examine the impacts of ocean acidification on the commercially valuable American lobster (see Case Story 7.5).

7.3.5

Moving forward

Fish stocks are assessed using the best available evidence, produced through basic research and using research monitoring programs. The resulting data are used to create physical forecast models that enable a longer-term understanding of climate change. The known associations between species abundance and environmental variables can be used to assess species vulnerability to warming (Stortini et al., 2015), to project future available habitat, and to facilitate planned adaptation to changes in distribution and productivity (Stortini et al., 2017; Marras et al., 2015; Shackell et al., 2014).

Fisheries management combines scientific advice with social and economic information to make decisions about the fishery. Canada has initiated a process to incorporate climate change into science advice on fisheries, with initial emphasis on the stock assessment process as part of ecosystem-based management (Fisheries and Oceans Canada, 2018e). Commercial stocks are assessed and managed within stock area units, with borders separating distinct stocks. Climate change adds another layer of complexity to the assessment and management process, in that we anticipate changes in distribution/productivity within and across borders. Increased understanding of the impact of multiple stressors on fishes and aquatic invertebrates will make it possible to develop more realistic models of the future abundance and distribution of commercial species, providing a stronger scientific foundation for resource management strategies.

7.4

Climate change brings benefits and threats to the agriculture sector

Climate change brings both opportunities and challenges to Canada’s agricultural sector. Longer growing seasons and the potential to grow crops farther north may benefit agriculture, while changes in water availability, extreme weather events, and pests and diseases will present challenges. Adaptation actions that enhance climate resilience and consider the linkages between agriculture and interconnected sectors, such as water management and transportation, will benefit both local sustainability and global food security. 

Agriculture is inherently sensitive to climate. Increasing temperatures, shifting precipitation patterns, and changes in the frequency and intensity of some extreme climate events will affect crops and livestock operations, by both amplifying existing risks, and bringing new risks and opportunities. The type and degree of impacts will vary across agricultural regions and production systems. For example, increasing abundance of insect and disease species, and greater risk of new pests and diseases will impact crop and livestock health and could increase the risk of trade barriers. Additional risks arise from the sector’s strong dependence on transportation systems to maintain access to markets—systems that are themselves vulnerable to climate change impacts. Improvements to on-farm management practices―including improved fertilizer management, adoption of no-till practices to minimize soil disturbance, and improved water-use efficiency―have enhanced the sector’s climate resilience and illustrate a high capacity to adapt.

7.4.1

Introduction

While Canada’s agriculture sector is intrinsically adaptive (Campbell et al., 2014), its ability to adapt is being challenged by the impacts of climate change, including a northward shift in distribution of many insect and disease species and new invasive species. Changes in temperature and precipitation patterns, discussed comprehensively in Zhang et al. (2019), will bring shifts in the distribution of crops. Despite the fact that precipitation is projected to increase for Canada as a whole, areas such as the southern Prairies and the interior of British Columbia are expected to see increased moisture deficits during the growing season (Zhang et al., 2019), increasing the importance of irrigation and good water management.

Canadian agriculture is also highly dependent on reliable transportation networks to ensure that inputs (e.g., seed, fuel, fertilizer, pest control products and equipment) are delivered in an efficient, cost-effective and timely manner, and to maintain access to domestic and international markets. Flooding and other extreme weather events will challenge Canada’s grain handling and transportation system (Phillips and Towns, 2017).

Improvements over the last twenty-five years have made Canada’s agriculture sector more climate-resilient (see Prairie Provinces chapter; Campbell et al., 2014). There has been an increased focus on collaboration between producers, researchers, and decision makers within interconnected sectors to assess impacts and implement new adaptation measures, and this will contribute to increasing the sector’s sustainability. This section examines recent developments in understanding climate change impacts on agriculture in Canada and the critical inter-connectivity between agriculture and transportation. A more in-depth discussion of agriculture at the regional level can be found in the regional chapters (see Atlantic Provinces, Ontario, Quebec, Prairie Provinces and British Columbia chapters).

7.4.2

Climate risks and regional adaptation planning

Understanding climate risks to agriculture is informed by inputting observed and projected climate data into agricultural models (e.g., for crops, livestock, pests and diseases) to identify potential physical and economic impacts. Planning and implementation are required at various scales to target and identify actions within whole agricultural value chains, to inform on-farm impacts and to identify appropriate adaptation actions.

Analysis of climate change impacts on agricultural production in Canada, summarized previously by Campbell et al. (2014), continues to improve. Under medium-emissions (RCP4.5) and high-emissions (RCP8.5) scenarios, production of wheat (Canada’s largest crop and greatest value export) and other spring-seeded grains is projected to increase by between 8% and 11% by the 2050s (2041–2070), relative to a 2006–2015 baseline, across most of Canada’s agricultural regions (Qian et al., 2019; He et al., 2018). Increased heat stress could reduce canola yields across the southern Prairies, which will likely lead to a northward shift in crop production areas (Qian et al., 2018). Climate impacts on corn and soybean production in Canada remain less conclusive, with some studies showing increased suitability in Atlantic and central Canada, and opportunity for expansion in the Prairies, provided moisture limitations are not too restrictive (see Prairie Provinces chapter; Agriculture and Agri-Food Canada, 2018; He et al., 2018; Schauberger et al., 2017; Sauchyn and Kulshreshtha, 2008).

Severe moisture deficits by the 2050s are projected for certain areas of Canada under a high-emissions scenario (see Figure 7.8), which will particularly affect spring-seeded grain crops, such as wheat and barley. Water scarcity would be amplified by increased demand for water by other sectors, especially where irrigation and access to water are already important constraints on the sustainability of agriculture. Climate change is projected to increase variability in water supply, and to strain irrigation, drainage and flood control systems. For example, glacial meltwater, which contributes approximately 10% of summer flow in the Columbia River and Bow River, will be reduced to almost zero over the next 50 to 60 years as glaciers disappear (Derksen et al., 2019; Fyfe et al., 2017). Glacial water melt is very important for maintaining summer flows in the rivers of western Canada, during the key growing period. Increasing irrigation and on-farm water-use efficiency is a key adaptation mechanism. Currently, irrigation is practiced on 1.1 million hectares of Canada’s 44 million hectares of arable land (2.27%). While there is irrigation in every province, over 90% occurs in Saskatchewan, Alberta and British Columbia.

Figure 7.8
Figure 7.8

Growing season moisture deficit (precipitation minus potential evapotranspiration) across Canada for the period 1981–‍2010 and for the period 2041–‍2070 as projected under a high emissions (RCP8.5) scenario. In most areas, increased precipitation is more than offset by increased temperature and associated evaporation, leading to greater moisture deficits.

Source

Agriculture and Agri-Food Canada, 2018

The risks posed by animal diseases and plant pests are increasing as a result of climate change, with the likelihood of new invasive alien species (IAS) being introduced to Canada having significant consequences for Canada’s agriculture sector and economy. It is estimated that invasive species currently cause $4.2 billion in annual measurable losses to Canada’s agriculture industry (Agriculture and Agri-Food Canada and Canadian Food Inspection Agency, 2008). These relate to reduced crop yields and revenue, increased IAS control costs, loss of access to export or domestic markets, and decreased native biodiversity resulting from competition for resources.

Modelling of pest and disease expansion under a changing climate indicates that the ranges of the striped flea beetle, Phyllotreta striolata (Fabricius), and the crucifer flea beetle, P. cruciferae, both IAS to North America, are projected to shift and potentially cause economic losses over an expanded area in the future (Olfert et al., 2017). Increased winter temperatures have contributed to higher overwinter survival of insect pests (Olfert et al., 2016) and some pathogens, including cereal rusts in the Prairie region (Xi et al., 2015; Kumar et al., 2013). The linkage between climate change and the rapid north and westward range expansion of the deer tick (Ixodes scapularis), which transmits the Lyme disease agent that can infect cattle and other livestock, as well as humans, is well documented (Ogden et al., 2014). Bluetongue, a viral disease of ruminants transmitted by the biting midge Culicoides sonorensis, extends from the southwestern United States into southern British Columbia and Alberta (Lysyk and Dergousoff, 2014), and has recently expanded into southern Ontario (Jewiss-Gaines et al. 2017). Species distribution models based on different climate change scenarios indicate that the range of C. sonorensis is likely to expand northward in Alberta (Zuliani et al., 2015).

Barriers to adaptation across the agriculture sector broadly (see Figure 1 of Campbell et al., 2014) include a lack of knowledge of climate impacts, limited technical capacity to assess agricultural risks at an appropriate scale, and uncertainties associated with projected climate impacts and adaptation actions (Agriculture Adaptation Working Group, 2016). Recent efforts by the agriculture sector to assess climate risks and consider adaptation actions, both regionally and at the farm level, have started to address these barriers. Examples include developments in climate risk assessment methodology and agricultural adaptation planning in Quebec (Ouranos, 2015), Manitoba (Goertzen, 2018), Nova Scotia (Nova Scotia Federation of Agriculture, 2018) and British Columbia (British Columbia Ministry of Environment and Climate Change Strategy, 2019). Continued work could further improve understanding of risks, the agricultural sector’s awareness of climate impacts, and mechanisms to support adaptation action. New knowledge will continue to inform initiatives such as the agriculture sector’s Emergency Management Framework (Federal, Provincial and Territorial Emergency Management Framework Task Team, 2016), which includes elements to help proactively manage the increasing risk of pests and diseases and potential associated trade risk issues.

7.4.3

Inter-connectivity

The success of agricultural production in Canada is highly dependent upon connections with transportation networks. A reliable transportation network allows for efficient delivery of necessary input goods (such as seeds, fuel, fertilizers and farm equipment), as well as access to export markets. Climate change presents a wide range of risks to transportation networks in Canada (see Section 7.7; Palko and Lemmen, 2017), including damage to infrastructure from extreme precipitation and flooding, impacts associated with sea-level rise and storm surge flooding at terminal ports, increased risk of rail buckling due to extreme heat, cracking of roadways as a result of desiccation of clay sub-soils during drought conditions and washouts during flooding. All of these impacts can cause transportation delays that negatively impact the agriculture sector value chain (e.g., farmers, processors, suppliers, etc.) from local to international scales.

Canada’s grain handling and transportation system is frequently impacted by extreme weather events. For example, excessive rainfall in Vancouver between January and March 2011, and severe weather events in the railway mountain corridors of British Columbia contributed to long delays in loading grain onto vessels in Vancouver ports (Quorum Corporation, 2014). Again in 2013, record totals of harvested crops in the Prairie provinces, followed by extreme cold winter conditions, contributed to transportation delays in accessing markets. Such delays impact Canada’s reputation as a reliable supplier of agricultural products and result in added costs for both grain companies and producers (Gray, 2015; Quorum Corporation, 2015).

The importance of export markets for the Canadian agricultural sector cannot be overstated, with approximately 42% of its annual production being shipped out of the country. In 2013, Canada was the world’s fifth-largest exporter of agriculture and agri-food products (Agriculture and Agri-Food Canada, 2016), with exports representing 85% of canola and canola products, 85% of flax, 66% of soybeans and soybean products, 65% of hogs and pork products, and 48% of cattle and beef products. Over the past 10 years, half of all Canadian grain production has been exported, averaging 41 million tonnes per year. Agri-food export sales in 2017 totalled $57.7 billion (Canadian Agri-Food Trade Alliance, 2020). While regional climate change impacts on agriculture production in Canada will be significant, impacts are expected to be much greater in most other major agriculture producing areas of the world (Food and Agriculture Organization, 2017; Ignaciuk and Mason-D’Croz, 2014). This represents both an opportunity for Canadian producers (Ignaciuk and Mason-D’Croz, 2014) and a critical need with respect to global food security (Mbow, 2019).

7.4.4

Moving forward

Mechanisms that support regional capacity to utilize and interpret climate change impacts and adaptation information are improving the ability of the sector to proactively plan and implement adaptation actions. It is also increasingly important for the sector to collaborate with other appropriate sectors, including transportation and water management, on shared adaptation issues. Initiatives such as Drought Watch (Agriculture and Agri-Food Canada, 2019), which provides information on current climate risks and provides a mechanism for producers to submit details on impacts experienced at the farm level, are an important step forward in enhancing engagement and collaboration.

7.5

Climate change brings new environmental challenges for mining

Impacts on the chemical and physical stability of tailings containment and reclamation structures are among the greatest climate-related challenges to the Canadian mining industry. Failure of such structures can lead to severe environmental contamination and present risks for surrounding communities and ecosystems. Considering long-term climate change at the design phase of mining projects is necessary to reduce these risks. 

Climate change can affect every phase of a mine’s life cycle. While short-term challenges, such as impacts on daily operations and disruption of critical supply chains, require adaptive actions, impacts on tailings containment and reclamation structures represent key vulnerabilities for the Canadian mining industry. Tailings containment and reclamation structures must remain in place for many decades and even centuries, which increases their vulnerability to climate change. While current designs of such structures take into account both average and extreme historical conditions, this does not adequately capture the full range of likely future conditions. Adaptation priorities include incorporating climate change projections into future designs, enhancing monitoring of existing structures and developing new methods and tools to improve climate resilience.

7.5.1

Introduction

Mining is a major component of the Canadian economy, with activities occurring in all provinces and territories. The effects of climate change on the mining sector are both direct and indirect, with the potential to affect every phase of a mine’s life cycle (Bussière et al., 2017; Pearce et al., 2011; Stratos, 2009). In many cases, climate change will exacerbate existing climate risks and create new risks, but it will also afford some new opportunities (Stratos, 2017, 2009; Bussière et al., 2017; Pearce et al., 2011). For example, during the exploration phase, the shorter period of frozen ground will make it difficult to access some exploration sites. On the other hand, a longer warm season will allow more time for mapping and the delivery of raw material supplies by ship or floatplane.

An overview of climate change impacts and adaptation in the Canadian mining sector by Lemmen et al. (2014) highlighted risks to built infrastructure, transportation, extraction and processing, and daily operations. The project phases most affected are the management of mine tailings during operations and the reclamation of waste storage areas. These two phases have a long life span, and climate considerations are incorporated into their design (Bussière et al., 2017). Lemmen et al. (2014) also noted a lack of proactive adaptation planning for climate change in the sector, a conclusion supported by subsequent analysis that found little evidence of government or corporate policies having advanced in terms of addressing climate change impacts and adaptation (Stratos, 2017).

This section builds on previous Canadian assessments by focusing on what has emerged as a key climate vulnerability for the Canadian mining industry—risks to the physical and chemical stability of structures designed to contain tailings during and following mining activities, and risk to the long-term effectiveness of reclamation methods (Bussière et al., 2017; Stratos, 2017).

7.5.2

Tailings containment structures

Mining waste, composed of finely crushed rock and water, is stored in tailings ponds that are generally located at surface-level close to the mine (Blight, 2010; Bussière, 2007; Bussière et al., 2005; Aubertin et al., 2002). The purpose of these ponds is to contain the greatest possible volume of solid waste and permit the transfer of any overflow into one or more secondary ponds (called sedimentation and/or polishing ponds). These structures are usually created by closing off a natural valley or cordoning off an existing area with one or more dykes. They are equipped with drainage systems to allow the water, once solids settle, to be reused or processed, with spillways to control effluence during periods of extreme precipitation (Blight, 2010; Aubertin et al., 2002). Given the high water content and low density of most mine tailings, the physical stability of these structures is difficult to maintain over both the short and long term. Structures are frequently modified throughout the extraction stage of a mine (generally a few decades) to adapt to evolving operating conditions.

Current approaches to designing tailings containment structures consider both average and extreme climate parameters, such as probable maximum precipitation (PMP) and probable maximum flood (PMF) (Aubertin et al., 2015). PMP and PMF are based on frequency analyses of historical climate data, and have a low annual probability of recurrence (Canadian Dam Association, 2013; Aubertin et al., 2011; Koutsoyiannis, 1999). The use of PMF is recommended in the conversion to permanent structures when a mine is closed, in order to avoid breaches and potential environmental, human and material consequences (Centre d’expertise hydrique du Québec, 2019; Canadian Dam Association, 2013; Ministère du Développement durable, de l’Environnement et des Parcs, 2012; Aubertin et al., 2002, 1997; Vick, 2001). Importantly, there is currently no uniform methodological approach to factor climate change into the PMP and PMF calculations.

The frequency and intensity of extreme precipitation events, as well as average annual precipitation levels, are projected to increase in most regions across Canada (Zhang et al., 2019), increasing the risk of physical instability of waste containment structures (Bussière et al., 2017; Guthrie et al., 2010; Jakob and Lambert, 2009). Recovering spilled waste resulting from a dyke breach is difficult and expensive. For example, the tailings pond failure at the Aznalcollar mine in Spain in 1998 led to site reclamation costs estimated at about US$230 million (Eriksson and Adamek, 2000). In northern Canada, increasing temperatures are leading to permafrost degradation and increased depth of the annual thaw (active layer) (Derksen et al., 2019; Arzhanov and Mokhov, 2013; Zhang et al., 2008a, b). As a result, the integrity and stability of waste-retaining structures may be compromised. Increased thickness of the active layer can create problems in locating suitable construction sites for mining infrastructure; it should be noted that technologies for building on permafrost do exist and that relevant guidelines are well documented (Mine Environment Neutral Drainage, 2012; Doré and Zubeck, 2009; Andersland and Ladanyi, 2003).

7.5.3

Reclamation of mining sites

Reclamation of mining sites includes preventing contaminants such as acid mine drainage (AMD) from escaping into the environment (Blowes et al., 2003). This is possible by maintaining the long-term chemical stability of mine waste, as AMD results from the oxidation of sulphide minerals when they come into contact with water and atmospheric oxygen. Specific reclamation methods for preventing AMD are strongly influenced by climate conditions (see Box 7.1 and Case Story 7.6).

The long-term effectiveness of reclamation systems is affected by many factors (Aubertin et al., 2015, 2002) related to the properties of the covering materials, the configuration and location of the containment area to be reclaimed, and climate. As the required life of reclamation structures is greater than 100 years, this is the stage of the mining cycle most likely to be severely impacted by climate change.

The main climate change risks during the reclamation of mining sites are reduced effectiveness of insulation covers, water infiltration barriers and oxygen barriers, and failure of retaining structures on sites undergoing reclamation (see Figure 7.11; Bussière et al., 2017; Lemmen et al., 2014). Direct climate change impacts of concern include: 1) higher temperatures that increase the depth of the active layer in sub-Arctic and Arctic regions, potentially resulting in the thaw of waste that could generate AMD; 2) an increase in extreme precipitation that could lead to the physical failure of dykes or other structures when employing reclamation methods such as water cover and elevated water table techniques; and 3) longer or more frequent droughts that could compromise the effectiveness of oxygen barriers that need high levels of saturation in the tailings and/or covers. Indirect climate change impacts, such as vegetation changes on revegetated sites affecting the water balance and the properties of the covering materials, can also positively or negatively impact the effectiveness of reclamation in both the short and long terms (Guittonny et al., 2018; Reinecke and Brodie, 2012).

Figure 7.11

-

Flow chart shows how changes in climate will lead to parameter changes and potential environmental impacts. Extreme precipitation (more intense and more frequent) will lead to increased water pressure on retaining infrastructures (such as dams and dykes), therefore posing an increased risk to the stability of retaining infrastructure, which could release solid and liquid tailings and mine waste. An increase in annual mean precipitation could lead to increased saturation level of cover materials and tailings, decreasing oxygen flux, and resulting in an increase in long-term performance for oxygen barrier methods. An increase in the number of consecutive days without precipitation in the summer leads to a decreased saturation level of cover materials and tailings, leading to an increase in oxygen flux, and a decrease in long-term performance for oxygen barrier methods. Changes in precipitation and temperature distribution lead to changes in spatial distribution of vegetation and therefore changes in water infiltration, evapotranspiration and material properties, which leads to changes in long-term performance of reclamation methods. Increases in annual mean temperatures leads to a disappearance of permafrost and increased thickness of active layer, which leads to a great risk of tailings and mine waste thawing and an increased risk to the physical stability of mining infrastructure. Ultimately, this leads to a decrease in long-term performance of reclamation methods.
Figure 7.11

Summary of expected major climate changes and their implications for mine waste containment and reclamation structures.

7.5.4

Adaptation

Maintaining the stability of tailing containment structures currently represents a major challenge for the mining industry when following conventional containment methods (Aubertin et al., 2011; Azam and Li, 2010). Climate change will amplify these risks. Adaptation will involve considering approaches that reduce reliance on dykes to retain large volumes of water and tailings pulp. Thickened or filter-pressed tailings may be one viable option (Bussière, 2007). Co-disposal approaches of waste rock and tailings could also improve the mechanical resistance of containment structures for mine tailings (Aubertin et al., 2016, 2011; Wilson et al., 2003). The reuse of mine openings (open pits or underground voids) for the storage of waste rock and tailings can also help to reduce the problems of dyke instability (Aubertin et al., 2016, 2015). If mining companies choose to use conventional methods for mine waste storage, the effects of climate change must be integrated into risk management analysis during the design stage. Approaches with this aim are being developed in conjunction with the Mining Association of Canada and the Canadian Dam Association.

While most governments now require proof that climate change has been considered in the design of any reclamation project (e.g., Ministère du Développement durable, de l’Environnement et des Parcs, 2012 for Quebec), the lack of specific guidance on how to modify reclamation methods in the context of climate change remains a barrier to effective adaptation.

Specific tools and analysis are needed: for example, to inform the design of oxygen barriers, where an extended drought could lead to desaturation of the cover, increasing oxygen ingress (Bresson et al., 2018; Hotton et al., 2018). While studies in Quebec have simulated a period of approximately two months with no water intake (Éthier et al., 2018; Bussière et al., 2003), the choice of this period of time was not based on systematic analysis of meteorological data, and the probability of such an event occurring is unknown. Tools are needed that quantify the changing risk of extreme drought to be considered in the design of oxygen barriers. For sites located in cold climates, where increasing temperatures and precipitation will have a major impact on the performance of reclamation methods, the need for adaptation planning tools is particularly urgent.

Enhanced understanding of local impacts is also needed to inform the planting of perennial vegetation at reclamation sites. Over time, ecosystem changes driven by changing climate will result in species changes at the sites, but the impacts on the long-term effectiveness of reclamation systems are largely unknown (Guittonny et al., 2018). Monitoring the performance of such sites, to document whether such changes lead to deterioration or enhancement of the reclamation system, would provide important insights.

7.5.5

Moving forward

New approaches and tools that factor climate change into the planning, design, and risk management processes of mining activities provide a sound basis for the mining industry and governments to reduce the sector’s vulnerability to climate change. While numerous studies have recommended that climate change be taken into account in mine waste management (e.g., Rousseau et al., 2014; Stratz and Hossain, 2014; Pearce et al., 2009; Aubertin et al., 2002), uncertainty in projecting changes, particularly extreme precipitation events (Mailhot et al., 2014), remains a barrier to action. Solutions require collaboration between mining engineers and climate services in developing updated or new methods and guidelines for assessing climate impacts that can be factored into the design of tailings containment structures and reclamation methods.

The longevity of tailings containment and reclamation structures increases their vulnerability to climate change. The physical and chemical stability of such structures is essential to avoid release of harmful contaminants into the environment. Tools and guidance are required to better enable designers to accommodate both changes in extreme weather caused by climate change, and slow-onset changes such as permafrost degradation, and to develop climate-resilient reclamation methods. In some cases, it is possible that, for a given site, a “best” solution for current climate conditions may not be the optimum solution for future conditions. Better designed tailings containment and reclamation structures at active and closed mine sites can significantly reduce the risk of environmental contamination over the long term. In the case of older infrastructure, it would be advisable to conduct analyses to ensure that these facilities will be able to withstand future conditions. If not, remedial measures will have to be implemented to maintain the integrity of the structures over the long term.

7.6

Each link of the energy value chain can be vulnerable to climate change

Changing climate affects energy demand and the full energy value chain, from exploration and production through to transmission and distribution. Climate risks can be integrated into current business planning by considering co-benefits, no-regret options and incremental approaches. Climate resilience needs to be a key consideration in converting to low-carbon energy systems.

Warming climate increases the demand for cooling in summer and decreases the demand for heating in winter. Energy supply is sensitive to a wide range of climate impacts, including changes in permafrost, ice cover, sea level, wave regimes, precipitation patterns, river flows and extreme weather events, such as hurricanes and ice storms. These impacts can all disrupt the energy value chain, with significant economic and social consequences. The use of climate risk screening tools can help to embed adaptation within current business planning practices and to identify opportunities for introducing adaptation measures at cost-effective times (e.g., during maintenance, upgrades and retrofits). As significant investments are made to convert to low-carbon energy systems over the next few decades, it will be important to ensure that climate resilience is considered in infrastructure design. 

7.6.1

Introduction

Energy is an important component of the Canadian economy, accounting for almost 10% of gross domestic product (GDP). Energy systems are the backbone of critical services, such as transportation, communication, health systems, drinking water and wastewater, as well as businesses and households. As such, their reliability is of paramount importance to economic activity and human well-being in Canada. Energy assets and operations have always been exposed to highly variable and extreme climate conditions, and were built to perform well under these conditions. Nonetheless, they were designed based on historical climate norms, which raises concern about their resilience in the face of ongoing changes in climate. Observed and projected climate changes that significantly affect the energy sector are described in Bush and Lemmen (2019), including changes in temperature and precipitation extremes (Zhang et al., 2019), permafrost, snow and sea ice (Derksen et al., 2019), streamflow (Bonsal et al., 2019) and ocean climate (Greenan et al., 2019a). Energy sector vulnerabilities to climate are highlighted by recent extreme events, including the 2016 Fort McMurray (Horse River) wildfire that devastated the community and caused substantial reductions in oil sands production, with an estimated loss of approximately 47 million barrels, costing producers 1.4 billion dollars in lost revenues (see Case Study 7.1; National Energy Board, 2016; Antunes et al., 2016).

Climate change also has direct and indirect impacts on energy demand. Warmer winters reduce fossil fuel and electricity demand for heating (Mantle314, 2019), while the increasing number of hot days in summer increase electricity demand for cooling (Ortiz et al., 2018; Jaglom et al., 2014). Indirect impacts are associated with societal and government responses to reduce greenhouse gas emissions, and these responses have differing impacts on renewable and non-renewable energy sources (International Energy Agency, 2016a).

Previous assessments of climate change impacts and adaptation highlight the diversity of the Canadian energy sector, noting that most available research focuses on hydroelectricity generation and electricity distribution (Lemmen et al., 2014). Key risks identified include impacts of extreme weather on critical infrastructure, and a potential mismatch between reduced hydroelectricity production and increased electricity demand during heat waves. Smart grid technologies and urban design measures that reduce heat island effects were noted as examples of adaptation measures (Lemmen et al., 2014). Knowledge gaps identified included the limited amount of research on the impacts of climate change on renewable energy supplies other than hydroelectricity. Environmental assessment and risk disclosure were identified as emerging drivers of adaptation action in the sector (see Climate Disclosure, Litigation and Finance chapter; Lemmen et al., 2014).

7.6.2

Risks to energy production and transmission

The impacts of a changing climate will vary between energy type and region, and may occur at different time horizons. The Canadian oil and gas sector fuels most transportation and accounts for 10% of Canada’s electricity generation (Natural Resources Canada, 2018). Oil and gas exploration, extraction, production and delivery will be affected by climate change (Mantle314, 2019; Cruz and Krausmann, 2013). Pipelines, roads and buildings are affected by permafrost degradation, slope failures and flooding. In northern regions, thawing permafrost and reduced availability of stable ice roads affect transportation and require changes in schedules. River ice breakup will happen earlier in the year and ice jam flooding may occur more frequently. Sea-level rise will change flood risks and increase coastal erosion. A large water demand for energy production, including for oil sands, may pose a risk as overall water availability declines in some areas and droughts may intensify (Bonsal et al., 2019). Offshore activities may benefit from increased access to resources in the Arctic due to reduced sea ice, but will still face significant risks from ice in northern regions and increased impacts from waves elsewhere (Mantle314, 2019; Stantec Consulting Ltd., 2012). While not directly related to production, carbon capture and storage (CCS) technology is critical to the oil and gas sector’s transformation to low-carbon systems. Indeed, the World Energy Council stated that “global warming is unlikely to be kept under 2°C without introduction and widespread adoption of CCS, and the cost of mitigation would be higher in the absence of CCS” (World Energy Council et al., 2014, p. 12). However, CCS technologies are water-intensive, using almost double the volume of water compared to non-carbon capture and storage plants, making them vulnerable to climate-related changes in fresh water availability (International Energy Agency, 2015).

Nuclear and fossil fuel thermal generation facilities depend on sufficient cooling water and may face reduced cooling efficiency due to higher water temperatures or contamination of cooling water intake (e.g., from algae, zebra mussels or ice particles during freeze/thaw cycles) (Braun and Fournier, 2016; Canadian Electricity Association, 2016). Maintenance schedules for nuclear facilities may be impacted by increased ambient temperatures, as temperature affects the number of workers allowed in vaults. Extreme impacts are exemplified by the shutdown of nuclear power plants in France due to lack of cooling water during the 2003 heat wave (Kopytko and Perkins, 2011).

Hydropower accounts for 59% of Canada’s electricity generation (Natural Resources Canada, 2018). As this is a renewable energy source depending directly on climate conditions, hydropower production is impacted by changes in temperature, precipitation and snow cover. Shorter winters will result in earlier spring floods with a smaller contribution from melting snow. Winter flows are expected to be higher, and late summer flows may be reduced in many parts of Canada (Bonsal et al., 2019). At higher latitudes, where higher precipitation is not offset by increased evapotranspiration, generation potential may increase. However, with changes in rainfall frequency and intensity, risks to infrastructure may also increase and changes to water management would be needed to minimize losses. The high flexibility provided by the storage capacity of hydropower operations means that they can play a key role in transitioning to a low-carbon energy system, providing stability for electricity grids that integrate more volatile renewables such as wind and solar (International Energy Agency, 2018). A review of 200 projects in the energy sector suggests that hydropower is the most advanced subsector with respect to climate change adaptation (Braun and Fournier, 2016).

Non-hydro renewable energy, which includes wind, solar and biomass, is becoming increasingly important in Canada’s energy mix, a trend that will continue in the future (see Figure 7.12; National Energy Board, 2018, 2017). Wind generation will generally be reduced from lower density air at higher temperatures. A 5°C increase in air temperature leads to a decrease in air density of 1–‍2%, with a proportionate decline in energy density affecting power production (Pryor and Barthelmie, 2010). However, wind generation may benefit from more sustained and higher winds. Fast market growth of renewable wind power (see Figure 7.12) reflects the influence of a larger share of renewable energy sources being integrated into the economy in response to climate change. Solar capacity is increasing even more rapidly than wind, and is expected to triple by 2040 (see Figure 7.12; National Energy Board, 2018, 2017). While there are concerns that an increase in cloudiness due to changing weather patterns and extreme weather events could affect solar energy generation (World Energy Council et al., 2014), the significance of these impacts is regionally dependent. More research is needed in Canada.

Figure 7.12

-

Bar chart shows the growth of renewable energy capacity (Gigawatts) in Canada between 2005 and 2040. In 2005, total renewable energy capacity is approximately 2.5 Gigawatts. By 2040, it is expected to reach approximately 33 Gigawatts. Energy capacity from biomass, wind, and solar is expected to increase, with the largest growth coming from wind.
Figure 7.12

Observed and projected changes in non-hydro renewable energy capacity in Canada between 2005 and 2040 under the National Energy Board reference case scenario. Higher rates of growth are projected under a technology scenario (NEB, 2018).

Source

Adapted from National Energy Board, 2018.

The electricity transmission and distribution sector will be affected by many factors, including the following: 1) increased temperatures, which can increase line resistance by approximately 0.4% per 1°C rise and decrease line load capacity by 0.5 to 1% per 1°C rise; 2) thermal expansion of power lines, which affects safety line clearances (a 1°C rise can result in 4.5 cm of sag); 3) extreme precipitation events that increase the risk of flooding of “underground” assets; and 4) high winds and gusts that can cause line damage (Asian Development Bank, 2013, 2012).

7.6.3

Adaptation

There are several examples of the energy industry adapting to changing climate conditions in Canada and around the world. Western Power Distribution in the UK studied the impact of higher ambient temperatures on its grid and developed a “low-regret” approach to adapt to an increase of their line rating by 5°C. To compensate for projected increased sag of the lines, they have started to gradually replace the poles in the course of normal maintenance (Western Power Distribution, 2011). Hydro‑Québec has adopted a method to improve its energy demand forecasting by applying a climate-model derived adjustment factor to the historical temperature record prior to running the forecast model (Braun, 2016). A similar approach to improve flow forecasts has been adopted by Iceland’s National Power provider Landsvirkjun, which has adjusted its streamflow record based on climate projections (see Case Story 7.7; Fournier, 2016).

New technologies are a key tool for climate change adaptation in the energy sector. The International Energy Agency’s Sixth Forum on the Climate-Energy Security Nexus identified “technology innovation” as a priority area to advance energy sector resilience to climate change in North America (International Energy Agency, 2016b). While decentralized renewable energy technologies have been driven largely by demand for low-emitting energy sources, they also provide redundancy that can create sub-systems for generating and distributing energy in the event that other parts of the grid are disrupted (American Council on Renewable Energy, 2018; IISD, 2017; U.S. Department of Energy, 2016; International Energy Agency, 2015). Smart grid technologies, including renewable energy generation, smart meters, smart appliances, and automatic power production have been proven to improve recovery from extreme weather, such as in the US with Hurricane Irene (2011) and Hurricane Sandy (2012) (IISD, 2017; Executive Office of the President, 2013). For thermal generation, new dry-cooling technologies are being implemented to reduce the vulnerability to changing temperatures and diminish dependency on cooling water sources (Braun and Fournier, 2016).

7.6.4

Moving forward

Increased awareness is a critical first step in developing a climate-resilient energy sector that can ensure future energy security and reliable service for Canadians. Detailed case studies of projected impacts and actual or proposed business responses are a valuable tool for building such awareness (Braun and Fournier, 2016). Responses should be feasible, economically sound, and able to be implemented in adequate time frames. Cost-effective and beneficial responses can be built into existing business operations―for example, by addressing climate change as an integral part of environmental impact assessments, as is required for projects undergoing a federal impact assessment (Environment and Climate Change Canada, 2020).

Proactive energy providers in Canada and abroad have established in-house climate expertise and climate change committees. These committees collaborate with researchers and climate change centres to understand, produce and use climate data for establishing operational thresholds that are sensitive to climate. Engaging asset managers can be an important component of a holistic, integrated risk-based approach for energy planning, as is strong monitoring and evaluation. Sharing adaptation case studies between companies is another way to advance climate resilience in the energy sector.

7.7

Extreme weather events impact transportation, disrupting supply chains

Road, rail, marine and air transportation in Canada are vulnerable to extreme weather events and slow-onset climate change, with major disruptions having significant economic and social impacts. To fully assess these impacts, linkages between transportation systems, and between transportation modes and a wide range of other economic sectors, need to be accounted for. Coordinating adaptation responses across jurisdictions and sectors will benefit transportation asset owners, operators and those dependent on vulnerable supply chains and corridors.

Canada relies on efficient, safe and reliable transportation infrastructure and operations that enable the movement of goods, services and people across the country. Current climate risk assessment and adaptation initiatives in the transportation sector tend to focus on individual, direct climate impacts associated with air, road, rail or marine infrastructure and other assets. This approach underestimates the potential cascading risks and impacts from surrounding assets or networks, and vice versa. Examining how various elements of transportation are dependent upon each other, and how transportation is linked to a wide range of other sectors, can help to identify opportunities for collaboration and cost efficiencies, and can prevent situations in which actions taken to reduce individual or specific risks inadvertently increase the climate vulnerability of others. 

7.7.1

Introduction

A well-functioning transportation sector connects people and communities, products, resources and services to domestic and international markets (Palko, 2017). In Canada, responsibilities for transportation are dispersed across various levels of government, with private-sector stakeholders playing an important role as owners, operators and managers of infrastructure and assets, including railway infrastructure, vehicles, ships and aircraft (Andrey and Palko, 2017). Canada’s transportation infrastructure is concentrated mainly in the southern portion of the country, where most of the trade and transportation movements occur (see Figure 7.14). However, northern transportation systems are particularly sensitive to climate change, and are already impacted by thawing permafrost, reduced river, lake and sea ice cover, and increases in coastal erosion and storm surge flooding (Palko, 2017; Hori et al., 2017).

Figure 7.14
National Airport System
National Highway System
Canada Port Authorities
National Railway Network
Figure 7.14

Maps of Canada’s national transportation systems, showing the National Airport System, the National Highway System, Canada Port Authorities and the National Railway Network.

Source

Adapted from Transport Canada, 2017

A comprehensive, regional-based report of climate change impacts and adaptation actions in Canada was released in 2017 (Palko and Lemmen, 2017). Key findings include:

  • Canada’s transportation infrastructure is vulnerable to damage and disruptions from a changing climate and extreme weather, and this can pose risks to other sectors of the economy;
  • Northern transportation systems are experiencing some of the greatest impacts from warming, and temperatures in the North will continue to increase at a faster rate than in any other region of Canada;
  • While reactive approaches to managing climate risks remain common in Canada’s transportation sector, there are examples from all regions and transportation modes of adaptation actions being taken in anticipation of future climate conditions.

This section supplements the report, Climate Risks and Adaptation Practices for the Canadian Transportation Sector (2017), by examining the interconnections that exist within intermodal transportation systems and across a wide range of sectors. It draws upon Canadian and international literature, and focuses primarily on physical infrastructure.

7.7.2

Climate impacts on transportation systems

Impacts of extreme weather events and climate change are already affecting transportation infrastructure, operations, systems and services across all modes in all regions of Canada (Palko, 2017); it should be noted that climate change has already increased the likelihood of some types of extreme weather events (Zhang et al., 2019). In some cases, these impacts have resulted in travel disruptions and unsafe conditions, affecting the movement of both freight and people, and leading to increased operating costs, reduced revenues or compensation for service disruptions.

Extreme weather events can incur large costs and accelerate the deterioration of transportation infrastructure, shortening its lifespan (Boyle et al., 2013). For example, a torrential downpour in Toronto in 2013 resulted in flooding that caused road closures, flight delays and cancellations, and stranded 1,400 train commuters. Damages were estimated to be $999.5 million (Amec Foster Wheeler Environment Infrastructure, 2017). That same year, flooding in southern Alberta caused $6 billion in damages, washing out 1,000 km of roads and destroying several hundred bridges and culverts (Andrey and Palko, 2017). In 2017 and 2018, the Churchill rail line was closed for 18 months (CTV, 2018) due to the cumulative effects of flooding and permafrost degradation beneath the railway line. Without a viable alternative mode of transportation to bring supplies in and ship goods out, living expenses escalated and access to essential services, including medical services located in southern communities, were compromised (Globe and Mail, 2018).

7.7.3

Understanding interdependencies

Infrastructure sectors and systems, including transportation, energy, telecommunications, water supply, wastewater treatment, solid waste management, and buildings, are “highly interdependent on each other, containing multiple connections, feedback and feed-forward paths and intricate branching” (Sudhalkar et al., 2017, p. 3).

Examples of critical interdependencies in transportation systems include situations where two sectors are essential to one another (e.g., transportation and energy), and where technology serves to strengthen connections between infrastructure, such as railway signals and traffic control systems being controlled by computer systems that depend on electrical power grids (Sudhalkar et al., 2017). Infrastructure networks containing one or more of these features “are at a higher risk of failure from external shocks or stresses, including climate hazards” (Sudhalkar et al., 2017, p. 3). Climate change risk assessments that do not address such interconnections could lead to the miscalculation of risks (Dawson, 2015)

Within the Canadian transportation sector, transportation modes (e.g., rail, marine, aviation and trucking) are often interconnected (intermodal) within supply chains and across transportation gateways and corridors. Port infrastructure and facilities, for example, form a central hub for transportation, logistics and supply chains, and are key convergence points for marine, rail and road infrastructure to facilitate movements of domestic and foreign goods (Becker et al., 2018).

Disruptions and delays in domestic and international supply chains, including those caused by weather events, can spread across networks with negative social and economic impacts (Becker et al., 2018; Allen et al., 2016; Zorn et al., 2016). In this way, vulnerabilities in one mode of transportation or sector can cascade to others.

7.7.4

Adaptation

Transportation asset owners and operators across Canada are undertaking actions to consider their climate risks and strengthen resiliency, often independently (Kwiatkowski, 2017). The distinct mandates, priorities, commercial sensitivities and resources of the many players within the sector create governance challenges and barriers to action (Sudhalkar et al., 2017). In the absence of holistic adaptation strategies, individual investments to address adaptation will be limited in effectiveness and could have unexpected and unintended negative consequences on neighbouring or interdependent assets and systems (Kwiatkowski, 2017). For example, the City of Toronto undertook a high-level climate risk assessment and interdependencies exercise for critical infrastructure and found that stakeholders had insufficient knowledge about the impacts of their activities on other interdependent systems, and that there was a need to implement roles and responsibilities to tackle joint risks (Sudhalkar et al., 2017).

Challenges for adaptation arise from many sources, including the fact that many transportation assets have long life-cycles, requiring lengthy planning and design, whereas technological development and replacement in areas such as information and communication technologies occur very quickly, leading to mismatched planning horizons and potential timing challenges (Man, 2013; Dewar and Wachs, 2008; Finley and Schuchard, nd). Opportunities arise when transportation planning is able to overcome such challenges and facilitate broad collaboration and engagement across sectors and jurisdictions. These include the following: early identification of multiple benefits and potential solutions (Man, 2013), which can be particularly important in areas of shared land use (Department of Environment, Food and Rural Affairs, 2011); improved understanding of interdependencies amongst multiple infrastructure owners; and increased innovation and efficiency across the supply chain (Dubois et al., 2011).

Tools have been developed to help planners consider risks associated with interdependencies. An example developed in Australia enables a systems-wide analysis of climate risks to organizations, as well as upstream and downstream interdependencies (Cross Dependency Initiative, 2019). Outcomes can demonstrate which ‘third-party risks” can affect an organization’s system, as well as the consequences of their own failure risks to other critical infrastructure, and can facilitate collaborative adaptation measures (Cross Dependency Initiative, 2019).

Case Story 7.8 provides an example of how stakeholders have come together to collectively explore interdependencies with transportation, as well as climate change risks to infrastructure and operations, in Surrey, BC.

7.7.5

Moving forward

Changing climate and extreme weather continue to pose challenges to Canada’s transportation sector. A range of interdependencies exists within the sector (between transportation modes) and across sectors. Without consideration of interdependencies and interrelated risks, individual investments to improve climate resilience will be limited in effectiveness or will result in unexpected or unintended consequences. Collaborative approaches can help asset owners and operators better understand the range of potential risks and identify adaptation solutions that respond to and cut across multiple risk areas. There are practical examples, including the City of Surrey (see Case Story 7.8), that demonstrate the multiple benefits that can be produced through such partnerships, trust and more informed planning.

7.8

Climate change is leading to transformational changes in tourism

All tourism destinations need to adapt to climate change impacts on tourism assets and altered competitiveness within the highly interconnected tourism economy. While Canadian tourism competitiveness is expected to increase under climate change, the specific market and regional implications of this change for national competitiveness remain under-researched. Winter and northern tourism and recreation are particularly sensitive to climate variability, and transformational changes in ski, snowmobile and Arctic cruise tourism are expected. 

Tourism is Canada’s largest service sector. Climate change is influencing Canadian tourism operations, investment and travel patterns, requiring adaptation by all tourism destinations. Many risks and some opportunities will progressively increase over this century, causing geographical shifts to enable seasonal and nature-based tourism markets to thrive. Earliest impacts have been visible in winter tourism markets, where declining snow and ice conditions are affecting the viability of multi-billion dollar tourism industries across Canada, much of which is concentrated in small and rural communities. The ski industry has invested hundreds of millions of dollars in snowmaking, providing substantial capacity to adapt to future warming and reduced natural snow fall in some locations. Higher warming scenarios will exceed the technical limits of snowmaking adaptation in many locations. The snowmobile industry is highly vulnerable to the same changes because snowmaking is not technically or economically viable for thousands of kilometres of trails, and ice crossings will become increasingly unsafe. Melting glaciers and reduced sea ice are altering tourist attractions from the Rocky Mountain national parks to polar bear and Arctic cruise tourism in the Canadian Arctic Archipelago. Tourists, tourism operators and destination communities are adapting to these diverse climate risks and opportunities, sometimes with unintended consequences for non-tourism government agencies (e.g., search and rescue), communities (e.g., overtourism) and visitor experience. 

7.8.1

Introduction

Climate change poses diverse risks and opportunities to domestic and international tourism across Canada (Scott et al., 2020; Hewer and Gough, 2018; Scott et al., 2012). Climate change is already influencing tourism operations, investment, and travel patterns so that all tourism destinations will need to adapt to impacts on local tourism assets, as well as to altered competitiveness within the highly interconnected international tourism economy (Scott et al., 2020, 2016). Globally, high emission pathways are widely considered to be incompatible with projected tourism growth (Scott et al., 2019; IPCC, 2018). Based on the various analyses of the multiple impacts of climate change on the tourism sector, Canada’s international tourism competitiveness is expected to improve (Scott et al., 2019; Roson and Sartori, 2016; OECD, 2015). The impacts of changing seasonality, landscapes (e.g., beaches, water levels), biodiversity, extreme events (e.g., heat waves, forest fires) and transnational markets have far-reaching, yet under-researched, implications for tourism development and competitiveness, travel patterns and livelihoods.

An overview of broad climate impacts and adaptation actions in the Canadian tourism industry was provided in Kovacs and Thistlethwaite (2014), including a discussion of implications for the parks system, warm- and cold-weather recreation, and nature-based tourism. While noting examples of advanced adaptation planning efforts, the overview concluded that the level of preparedness across the tourism sector to deal with climate change was low, which is consistent with assessments of the sector’s preparedness globally (Becken et al., 2020; Scott et al., 2016). Climate change also presents emerging opportunities for some tourism markets, who would also need to adapt in order to realize potential economic benefits, sustain tourism assets and maintain visitor experience.

This section builds on the findings of Kovacs and Thistlethwaite (2014) by focusing on snow- and ice-dependent markets like ski, snowmobile and Arctic cruise tourism, because these are where the impacts of changing climate are being experienced first and where adaptation by the tourism industry and investors is ongoing. Details of observed and projected changes in snow and ice cover (including sea, lake and river ice, glaciers and permafrost) are presented in Chapter 5 of Canada’s Changing Climate Report (Derksen et al., 2019).

7.8.2

Winter sports tourism

The ski industry of North America has invested hundreds of millions of dollars in snowmaking over the last 30 years to reduce its climate sensitivity. While average winter temperatures have continued to rise, the length of ski seasons increased throughout the 1980s, 1990s and 2000s in all five regional ski markets in the United States (Scott and Steiger, 2013). Only in the 2010s has this trend reversed, suggesting that advanced snowmaking capacity may no longer be able to offset winter warming. Recent record warm winters that are analogues for normal future winter conditions provide important insight into how the ski industry and ski tourists adapt. During the record warm winter of 2011–2012, the Ontario ski market experienced an average decrease in the ski season length (-17%) and in skiable terrain (-9%), reduced snow quality (-46% days with packed powder), fewer snowmaking days (-18%) and increased early season snowmaking (+300% in December), as well as a 10% decrease in overall skier visits, when compared to a climatically normal winter for the 1981–2010 period (Rutty et al., 2017). Similar impacts on season length and visitation have been observed in the Quebec market, with skier visits reduced by 12.5% in the record warm winter of 2015‒2016 (Association des Stations de Ski du Québec, 2016).

Differences in exposure to climate change risk among ski destinations have important implications for intra- and inter-regional market competitiveness and geographical shifts in ski tourism. A comparison of climate change impacts on the ski seasons in ski areas in Ontario, Quebec and the Northeastern United States revealed different futures for these regional markets for the 2050s and 2080s (see Figure 7.16). Ski areas in Quebec and high-elevation locations in Vermont and New Hampshire are more climate-resilient than those in Ontario and lower-elevation/lower-latitude locations in the US Northeast (Scott et al., 2020). Analysis of three ski areas in Quebec projected almost identical season losses of 10 to 20 days in the 2050s, and an expected 10% reduction in skier visits (Da Silva et al., 2019). Important information gaps remain, such as for Canada’s largest ski tourism markets (British Columbia and Alberta) and regarding the implications of differential impacts on market dynamics, tourism and community-level employment, development pressures and vacation real estate values (Scott et al., 2017; Rutty et al., 2015).

Figre 7.16

-

Bar graph showing baseline and projected length of the ski season, measured by number of days, in Quebec, Ontario, and the US Northeastern region. In Quebec the baseline season length is 137 days, under RCP4.5 the season is projected to be 121 days in the 2050s and 119 in the 2080s. Under RCP8.5 it is projected to be 116 in the 2050s and 106 in the 2080s. In Ontario the baseline season length is 117 days, under RCP4.5 the season is projected to be 102 days in the 2050s and 96 in the 2080s. Under RCP8.5 it is projected to be 93 in the 2050s and 46 in the 2080s. In the Northeastern US region the baseline season length is 121 days, under RCP4.5 the season is projected to be 104 days in the 2050s and 98 in the 2080s. Under RCP8.5 it is projected to be 94 in the 2050s and 67 in the 2080s.
Figre 7.16

Inter-comparison of changes to ski season length in Ontario, Quebec and New England markets with advanced snowmaking capacity under medium emissions (RCP4.5) and high emissions (RCP8.5) scenarios for the 2050s and 2080s. Baseline refers to the period 1981‒2010.

Source

Adapted from Scott et al., 2020.

Studies of the impact of ongoing and projected reductions in winter season snow on the snowmobile industry of North America show that while riders are adapting to changing conditions, continued deterioration of the season length will result in loss of this tourism market in many regions of the US and Canada (Perry et al., 2018; Hatchett and Eisen, 2018; Tercek and Rodman, 2016; McBoyle et al., 2007). Under the high-emissions scenario projected for the 2050s, a reliable snowmobiling season will be largely eliminated in regions of Ontario and Quebec with the densest network of snowmobile trails (McBoyle et al., 2007). A 20% decline in registered snowmobiles in Canada from 1995 to 2015 (International Snowmobile Manufacturers Association data) may signify a climate adaptation by individuals that will induce changes in this tourism marketplace. Bombardier Recreational Products (2017) advised investors that global climate change might impact future snowmobile sales to a greater extent than previously anticipated. A shift to all-terrain vehicles is not thought to be a feasible adaptation strategy, as these vehicles tend to be more impactful on trail surfaces, preventing shared trail networks and access by other users to parks and farmlands (Perry et al., 2018; McBoyle et al., 2007).

7.8.3

Arctic cruise tourism

Declining summer sea ice cover has allowed greater marine access to the majestic land, seascapes and cultures of the Canadian Arctic (see Northern Canada chapter). This has opened up access to areas and communities that were previously inaccessible to tourist vessels. Tourism operators and tourists have been quick to adapt to these new opportunities, with strong growth in commercial (cruise ships) and non-commercial (private yachts) tourism vessel traffic since the late 2000s (Dawson et al., 2018; Johnston et al., 2017). Although tourism development is considered highly strategic by pan-Arctic governments (Dawson et al., 2017), there are concerns related to infrastructure needed for hosting the growing tourist numbers, environmental impacts, uneven economic opportunities, negative local social and cultural impacts, and limited search and rescue capacity (Dawson et al., 2018; Stewart et al., 2011). Limited hydrographic charts and changing ice conditions increase the risk of a high-impact cruise ship incident (Dawson et al., 2016). Even relatively minor incidents can be associated with high costs (see Case Story 7.9). Integrated multi-sector adaptation is essential to respond to these challenges and support sustainable tourism in the Canadian Arctic.

7.8.4

“Last chance” tourism

Some evidence supports the emergence of a travel trend of “last chance” tourism. This refers to travel by tourists to visit sites before their attractions vanish or are irrevocably degraded, or to witness the impact of climate-induced landscape changes, such as rapidly melting glaciers or biodiversity changes (Lemelin et al., 2010). Tourist market surveys about Canada’s Rocky Mountain Parks and Churchill, Manitoba, the self-declared polar bear capital of the world, reveal last-chance motivations among a segment of travellers, suggesting short- to medium-term opportunities for increased visitation, as well as visitor education and interpretive activities (Weber et al., 2019; Lemieux et al., 2017; Groulx et al., 2017; Groulx et al., 2016; Dawson et al., 2010). There remain important uncertainties associated with longer-term tourist responses to degraded tourism assets. For example, while tourists have indicated their motivation and intent to visit the Rocky Mountain Parks or Churchill, the number of tourists would decline if glaciers and polar bear populations were significantly impacted, and it is uncertain whether future generations of tourists with no experience or expectations of current ecotourism attractions would respond in the same way (Scott et al., 2007).

7.8.5

Moving forward

Tourism is Canada’s largest service sector and is projected to become more competitive in the global market as a result of climate change (Scott et al., 2019; Roson and Sartori, 2016; OECD, 2015). The current literature is insufficient to determine the net economic impact of low- and high-emissions scenarios on sub-national tourism markets. Most adaptation in the sector is focused on addressing ongoing operational risks and emerging market opportunities. A key barrier to longer-term strategic adaptation planning remains the lack of integrated sectoral assessments that consider the full range of potentially compounding impacts at national and destination scales, and their interactions with other major drivers of tourism. Tourism is not part of any major national climate policy documents in Canada, nor is there a climate change response in national and most sub-national tourism strategies (Becken et al., 2020). Adaptation that is occurring is often at the company scale, with limited examples of coordinated adaptation planning at destination scale (e.g., Resort Municipality of Whistler, 2016).

7.9

Increased private sector involvement will accelerate adaptation across sectors

Despite growing awareness of climate change impacts, there is no widespread evidence of corporate adaptation in Canada. When adaptation does occur, it tends to focus on short-term actions to address physical risks, such as disruptions in construction and interruptions in supply chains. Increased involvement of the private sector would accelerate adaptation in Canada as a whole.

There is considerable uncertainty about the role of the private sector in climate change adaptation in Canada. While there are several case studies of Canadian businesses adapting to changing climate, particularly in the insurance and natural resources sectors, there is no evidence that such actions are broadly representative of the business community response to climate change. Actions that have been taken are frequently “low-hanging fruit” focused on site-specific vulnerabilities to the current climate that would have been implemented regardless of increasing risk in the future. Drivers for corporate adaptation include strategic incentives associated with physical risk, growing awareness among stakeholders of the need to adapt, and government regulation. The lack of available data on corporate adaptation may be influenced by the proprietary protection of internal information; however, there are also clear barriers to adaptation related to capacity and the short-term time horizons of most business operations. Additional effort, including research, could help reduce these barriers by identifying the appropriate roles for the private sector in supporting adaptation. 

7.9.1

Introduction

The private sector, which is a key component of all the sectors discussed previously in this chapter, is both a source of risk and opportunity in Canada’s approach to climate change adaptation. Research highlights the exposure of firms and industries to climate change, particularly how changing environmental conditions (e.g., more frequent extreme weather) could limit growth, disrupt operations and devalue investments. At the same time, firms and industries are critical actors in supporting the development of expertise, tools and knowledge related to adaptation.

Private sector engagement on climate change has historically focused on mitigation rather than adaptation. However, governments, scholars and other organizations have started to explore the potential role of corporate adaptation, given the scale of the investment and resources required to manage climate change risks (Dougherty-Choux et al., 2015). The private sector could help to address demands for new technologies, expertise in risk management and modelling, capacity to scale up solutions beyond individual communities, and financial resources needed to achieve domestic and international adaptation objectives (UNFCCC, 2012; UNEP, 2012).

An overview of climate impacts and adaptation actions in Canadian businesses was included in multiple chapters of Warren and Lemmen (2014), with Kovacs and Thistlethwaite (2014) noting that corporate adaptation had been largely reactive, responding to variations in weather or extreme events, rather than involving analysis of long-term projected changes in climate. They further noted that successful adaptation can create new opportunities through expanded markets and products. They emphasized the scarcity of published research detailing climate impacts and adaptation on Canadian businesses, noting that in some cases adaptation actions may be under-reported for strategic reasons.

This section builds on the findings of Kovacs and Thistlethwaite (2014) by focusing on the status of corporate adaptation in Canada broadly. It complements the discussion of specific adaptation actions in other sections of this chapter, as well as a number of emerging issues of importance to the private sector, which are discussed in the Climate Disclosure, Litigation and Finance chapter of this report.

7.9.2

Corporate adaptation in Canada

Corporate climate change adaptation is defined as a “process of adjustment by companies to actual or expected climate and its effects through changes in business strategies, operations, practices and/or investment decisions” (Averchenkova et al., 2016, p. 520). Cases of corporate adaptation in Canada, as well as internationally, tend to be limited and difficult to generalize beyond sector and site-specific actions. Comprehensive literature reviews on corporate adaptation (Linnenluecke and Smith, 2018; Averchenkova et al., 2016) did not identify any recent (post-2011) Canadian studies on corporate adaptation.

Research findings from other industrialized countries indicate that firms may have extensive experience in making management or operational adjustments in response to changing economic or competitive environments, but they have generally been unable to translate that experience into making adjustments to changing climate (Linnenlueke et al., 2013). As a result, firms often separate climate change adaptation from the main value-driving processes, and instead treat it as a social responsibility issue (Thistlethwaite and Wood, 2018; Furrer et al., 2009).

It is often difficult, and perhaps not particularly useful, to distinguish actions supporting climate change adaptation from actions arising from existing risk management processes, given that anticipating and identifying external instability that could interrupt services is effective business continuity planning (Agrawala et al., 2011). Experts have argued that adaptation can be incorporated into standard enterprise risk management strategies by expanding the scope of risk assessment, prioritization and response actions to include climate change (Berkhout, 2012).

7.9.3

Adaptation actions

Risk management represents the main strategy for firms to reduce their exposure to climate change. In analyzing private sector adaptation, it is possible to differentiate between firm-level management actions to limit operational exposure to climate change, and the production of adaptation tools and services that help manage climate vulnerability (Schaer and Kuruppu, 2018). In terms of managing operational exposure, adaptation actions can be categorized as either “soft” or “hard” measures. “Soft” adaptations constitute “low-regret” measures, as they require limited investment and will yield benefits, although not always direct financial returns. They include activities such as climate change risk and opportunity assessment, adjusting operational practices, education and awareness programs, stakeholder and political engagement, and initiating partnerships with external actors. In contrast, “hard” adaptation measures require a significant investment or change in practice and operations, such as building or renovating structural defenses, relocating infrastructure or offices, or divesting from climate-exposed property and sectors (Averchenkova et al., 2016). Adaptation services represent a response to demand for climate change risk and disaster risk management, and include climate risk and opportunity assessments, mapping and communication technologies, climate change-resilient agricultural products, and new insurance products (e.g., parametric insurance).

Adaptation is growing in the corporate sector in response to many drivers, including increasing awareness of climate change among firms and sectors, strategic incentives associated with addressing physical risks, (Williams and Schaefer, 2013) and government regulations (Revell et al., 2009). In Canada, attention has focused primarily on the understanding and disclosure of physical risks. These include disruptions in construction, interruptions in supply chains, volatile energy costs associated with climate-related changes in demand, and closures or relocation in situations where employees or customers are unable to access a firm as a result of extreme weather (see Climate Disclosure, Litigation and Finance chapter; Certified Professional Accountants Canada, 2016a; Linnenluecke et al., 2011). Despite this emphasis, most firms struggle to measure physical risk that involves a lot of uncertainty and can be perceived as a more long-term concern (see Climate Disclosure, Litigation and Finance chapter; Mazzacurati, 2018).

Overall, there is limited evidence of clear responses to these drivers in Canada. A 2018 survey focused on business responses to climate change found that a majority of businesses are not taking action and one quarter of respondents suggest that they plan to engage in adaptation (Earnscliffe Strategy Group, 2018). One example is the insurance industry implementing strategies to support climate change adaptation in response to significant increases in water-related property damage (see Case Story 7.10; McBean, 2012). Climate change, combined with ageing infrastructure and development in areas where climate risks are high, have increased the number and costs of claims for flood damage (Henstra and Thistlethwaite, 2017).

7.9.4

Knowledge gaps

Apart from a few examples in the insurance industry and natural resource sectors, there is a paucity of available research and data on corporate adaptation in Canada. Knowledge gaps start with the fact that corporate adaptation remains poorly defined in practice, which limits our understanding of which management, strategic or investment behaviours constitute adaptations. As a consequence, firms might be under-reporting adaptation since they are unsure if an action can be defined as such.

There also remains uncertainty over the motivation for and barriers to corporate adaptation. In Canada, most attention has focused on physical risks to business operations, but there is almost no research on the topics of firm awareness or regulation. There is also little research on the outcomes of corporate adaptation and its impacts on firms or the communities where they are located. Monitoring and evaluation of corporate adaptation could improve understanding of these outcomes (Surminski and Hankinson, 2018; Averchenkova et al., 2016). Finally, additional research would help to clarify the role of businesses in broader adaptation policies. The division of responsibility for adaptation between different stakeholders remains a source of ambiguity.

These knowledge gaps are particularly concerning for small and medium-sized enterprises (SMEs), which are more vulnerable to climate risk (Linnenluecke and Smith, 2018). SMEs lack the resources of their larger counterparts and may have greater difficulty prioritizing perceived long-term issues like climate change, given short-term concerns around sustaining operations. The 2018 survey of Canadian business responses to climate change found that SMEs were far less likely to be engaging in actions supporting adaptation (Earnscliffe Strategy Group, 2018). This gap is especially concerning since the recovery of a local community in the aftermath of a disaster is often contingent on the resiliency of local SMEs. There is evidence that 40% of SMEs fail to reopen after a disaster and that many are unable to sustain their business even if they do reopen (McKay, 2018). Without these SMEs, the local community may never support the economic growth and quality of life enjoyed before the disaster, since many provide critical services such as access to food and medicine.

7.9.5

Moving forward

While the lack of available data precludes strong conclusions about the status of corporate adaptation in Canada, it suggests that firms may lack sufficient adaptive capacity. Most firms face limitations in the human and financial resources required to interpret climate change data, assess the costs and benefits of actions, and incorporate the flexibility required to adjust strategies as new information emerges (Wedawatta and Ingirige, 2016; Downing, 2012). Research also suggests that firms are organizationally biased in favour of short-term and local scales, and resistant to acting when faced with temporal and spatial uncertainty associated with climate change (Bansal et al., 2017; Slawinski et al., 2017). The lack of adaptive capacity and absence of organizational interest in addressing climate change call into question the current ability of businesses to have a role in supporting climate change adaptation.

7.10

Moving forward

7.10.1

Knowledge gaps and emerging issues

It has been noted for more than a decade that, in most situations, existing knowledge is sufficient to start taking adaptation action in Canada (Lemmen et al., 2008). Nonetheless, the need to accelerate the implementation of adaptation measures has been acknowledged in both scientific and policy analyses at the global and national scale (e.g., Canadian Council of the Academies, 2019; IPCC, 2018; Government of Canada, 2016; UNFCCC, 2015). Risk and opportunity assessments are frequently a prerequisite to enhancing investment in adaptation. A wide range of methodologies are available, with the most appropriate being determined by many factors, including the scope and goal of the assessment and available resources (e.g., British Columbia Ministry of Environment and Climate Change Strategy, 2019). The Canadian Council of Academies undertook an assessment to prioritize climate change risks for Canada and the Canadian federal government, based largely upon collective expert judgement (Canadian Council of Academies, 2019). Major climate change risk areas included some of the sectors discussed in this chapter as stand-alone sections (agriculture, fisheries, forestry) as well as themes woven throughout (e.g., physical infrastructure and governance and capacity). Importantly, analysis considered the potential to reduce damages through adaptation, in addition to the likelihood and potential consequences associated with each risk. Of the topics addressed in this chapter, infrastructure (including transportation) and agriculture were identified as having the highest adaptation potential, and fisheries the lowest (Canadian Council of Academies, 2019).

While many factors contribute to the lack of progress on adaptation, uncertainties and gaps in knowledge are frequently highlighted as a barrier to action (Eyzaguirre and Warren, 2014). Knowledge gaps specific to individual sectors are identified in the preceding sections of this chapter and the publications cited in those sections. There are also a number of emerging cross-cutting issues that represent important knowledge gaps with respect to economic sectors in Canada.

The state of adaptation in the private sector

While climate disclosure is emerging as a key instrument for understanding how the private sector is assessing and responding to physical climate risk (see Climate Disclosure, Litigation and Finance chapter; TCFD, 2017), such information is largely restricted to large, publicly-traded companies. Very little information is available on adaptation actions undertaken by small and medium-sized businesses, even though many are highly exposed to climate risks. Benchmark surveys (e.g., Earnscliffe Strategy Group, 2018) provide a foundation for future work.

Transnational climate impacts

Also referred to as transboundary or indirect impacts, these refer to climate impacts that occur in one country and affect the adaptation measures taken within other countries (Hedlund et al., 2018). These could relate to impacts on global supply chains, international competitiveness, financial flows and trade (see International Dimensions chapter). For example, flooding in Thailand in 2011 affected global electronics and automotive supply chains, with economic implications for many countries and corporations (Shughrue and Seto, 2018). The magnitude of such vulnerabilities in Canada is essentially unknown, although research elsewhere indicates that open and export-intensive economies are particularly exposed (Hedlund et al., 2018).

Interdependencies

Much of the existing sectoral research examining climate impacts in Canada, and elsewhere, has been focused on individual sectors. While the importance of understanding interdependencies between sectors is growing (see Section 7.7.3), quantitative analysis of these connections remains limited. Without such analysis, there is potential to significantly underestimate the risks associated with climate change (Canadian Council of Academies, 2019).

Potential for stranded assets

Understanding of the risk of assets losing significant value as a result of policy changes to address climate change is well developed, particularly with respect to the energy sector (e.g., IPCC, 2018; IRENA, 2018). Far less attention, particularly in Canada, has been given to stranded assets that could result from the physical impacts of climate change (see Circle of Blue, 2018 for examples related to water resources). In Canada, closure of the rail link to the Port of Churchill for 18 months as a result of flooding represents temporary stranding of the port assets (see Section 7.7.3).

Strengthened economic analysis

A commonly cited barrier to adaptation action is the absence of a strong business case (Eyzaguirre and Warren, 2014). While examples of detailed economic analysis, including cost-benefit, cost-effectiveness and multi-criteria analyses, exist (e.g., UNFCCC, 2011) there is limited application of these techniques in Canada (see Costs and Benefits of Climate Change Impacts and Adaptation chapter). At a national scale, quantitative analysis of economic impacts under a range of future climate scenarios is lacking, potentially hindering action to enhance climate resilience and reduce greenhouse gas emissions.

7.11

Conclusion

Examination of the sector-specific key messages and associated discussions in this chapter reveals several integrative conclusions:

  1. A tremendous breadth of sector activities are impacted by climate change, which is evident despite authors focusing on a limited number of key issues per sector. While research has traditionally focused on direct climate impacts on production (e.g., agriculture, forestry, fisheries and hydroelectricity) and to a lesser degree on consumer demand (e.g., energy and tourism), it is clear that cumulative and cascading impacts of climate change ultimately affect virtually all elements of sectoral systems. This breadth emphasizes the value of comprehensive assessments of both risks and opportunities to inform adaptation.
  2. It is important to understand the interconnections within and between multiple sectors (see Section 7.7.3). For some sectors, such as transportation, this is particularly evident given its critical role in supply chains (see Section 7.7.3). The reliability and resilience of Canada’s freight transportation system regarding access to domestic and international markets is a critical issue for the sustainability of agriculture in Canada (see Section 7.4.3). The response of the forest sector to wildfire has implications for many other sectors, including costs related to evacuations, damage to buildings, roads, pipelines and other physical infrastructure, shutdown of businesses and industries, and insurance costs (see Section 7.2.2). Emerging tourism in the Arctic is having unforeseen social and cultural impacts, and is placing stress on search and rescue capacities (see Section 7.8.3). Further modelling studies could help elucidate the nature of these interconnections (see Section 7.7.4).
  3. Adaptation is occurring within all sectors, but needs to be accelerated. This applies to actions that reduce risks and those that take advantage of new opportunities. It is noteworthy that successful adaptation not only reduces the vulnerability of sectors and communities within Canada, but also enhances global resilience (e.g., food security) (see Sections 7.3.1, 7.4.3). Most examples of implemented adaptation action relate to those sectors where the direct impacts of climate change are already evident. Evidence of widespread adaptation within the private sector is particularly limited, despite the key role that it plays in all the sectors discussed here (see Section 7.9.2). The need for collaboration using a systems approach that includes producers, asset managers, regulators, researchers and relevant stakeholders is a commonly identified need (see Sections 7.2.3, 7.4.1, 7.5.5, 7.6.4, 7.7.4). Finally, it is extremely important to monitor and report on adaptation measures that have been implemented in order to inform future plans (see Sections 7.5.4, 7.6.4, 7.10).
  4. There is urgency associated with accelerating adaptation action. Urgency is most clear where current climate risks are not adequately managed and where investment decisions made today have implications extending for many decades to come: for example, decisions regarding infrastructure (see Sections 7.6, 7.7, 7.8), forest management (see Section 7.2) and mining reclamation (see Section 7.5). While proactive adaptation is generally recognized as being more effective and cost-efficient than reactive approaches (i.e., responding to impacts as they happen), and provides opportunities for innovation and competitive advantage (e.g., Eyzaguirre and Warren, 2014), it is recognized that investment decisions in both the public and private sectors take place within a context of competing priorities and associated opportunity costs. Comprehensive assessments of risks and opportunities can be critical in identifying action priorities, particularly those that include consideration of adaptation potential (see Sections 7.7.3, 7.8.5, 7.10). There is also an urgent need to reduce greenhouse gas emissions as the range of viable adaptation options decreases under higher rates of climate change, and limits to adaptation can be exceeded (IPCC, 2018, 2014).
7.12

References

Adams, C.F., Alade, L.A., Legault, C.M., O’Brien, L., Palmer, M.C., Sosebee, K.A. and Traver, M.L. (2018). Relative importance of population size, fishing pressure and temperature on the spatial distribution of nine Northwest Atlantic groundfish stocks; PLoS ONE, 13(4), e0196583. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0196583>

Agrawala, S., Carraro, M., Kingsmill, N., Lanzi, E., Mullan, M. and Prudent-Richard, G. (2011). Private Sector Engagement in Adaptation to Climate Change. OECD Environment Working Papers 39, OECD Publishing. Retrieved June 2020, from <https://doi.org/10.1787/5kg221jkf1g7-en>

Agriculture Adaptation Working Group (2016). Barriers to Adaptation. Analytic Paper Series Paper #2, Developed under Canada’s Climate Change Adaptation Platform. 

Agriculture and Agri-Food Canada (2016). An Overview of the Canadian Agriculture and Agri-Food System 2016. Retrieved June 2020, from <http://www.agr.gc.ca/eng/about-us/publications/economic-publications/an-overview-of-the-canadian-agriculture-and-agri-food-system-2016/?id=1462288050282>

Agriculture and Agri-Food Canada (2018). Climate change scenario maps drawn from CMIP5 data -26 GCMs as supplied by PCIC. Retrieved June 2020, from <https://www.pacificclimate.org/data/statistically-downscaled-climate-scenarios>

Agriculture and Agri-Food Canada (2019). Drought Watch. Retrieved June 2020, from <http://www.agr.gc.ca/eng/programs-and-services/drought-watch/?id=1461263317515>

Agriculture and Agri-Food Canada and Canadian Food Inspection Agency (2008). Invasive Alien Species (IAS) Backgrounder; Agriculture and Agri-Food Portfolio, Environmental Health Research Branch, AAFC and Plant Health Division, Policy and Programs Branch, CFIA.

Alin, S., Evans, W., Gao, Z., Gurney-Smith, H., Lee, K. and Tischenko, P. (2019). Ocean acidification and deoxygenation in the North Pacific Ocean, (Eds.) J. Christian and T. Ono, PICES Special Publication 6, 116 p. Retrieved June 2020, from <https://meetings.pices.int/publications/special-publications/Special-Report%205-Ocean_Acidification.pdf>

Allen, R., Whelen, M. and Khan, A. (2016). Building Resilience to Counter the Impact of International Supply Chain Vulnerabilities. Canadian Transportation Research Forum, 8 p. Retrieved June 2020, from <https://ctrf.ca/wp-content/uploads/2016/05/CTRF2016AllenWhelenKhanLogisticsTrade.pdf>

Amec Foster Wheeler Environment Infrastructure (2017). Canada’s Climate Change Adaptation Platform: Infrastructure and Buildings Working Group Adaptation State of Play Report, 162 p.

American Council on Renewable Energy (2018). The role of Renewable Energy in National Security. Retrieved June 2020, from <https://acore.org/wp-content/uploads/2018/10/ACORE_Issue-Brief_-The-Role-of-Renewable-Energy-in-National-Security.pdf>

Andersland, O.B. and Ladanyi, B. (2003). Frozen ground engineering. John Wiley and Sons, 384 p.

Andrews-Key, S.A. (2018). Vulnerability and Adaptation to Climate Change in Sustainable Forest Management and the Forest Industry in Saskatchewan (PhD Thesis). University of Saskatchewan. Retrieved June 2020, from <http://hdl.handle.net/10388/11671>

Andrey, J. and Palko, K. (2017). Introduction, in Climate risks and adaptation practices for the Canadian transportation sector 2016, (Eds.) K. Palko and D.S. Lemmen. Government of Canada, Ottawa, 2–10. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2016/Chapter-1e.pdf>

Antunes, P., Bernard, M.-C. and Owusu, P. (2016). The Economic Impacts of the 2016 Alberta Wildfires. The Conference Board of Canada, Ottawa. Retrieved June 2020, from <https://www.conferenceboard.ca/e-library/abstract.aspx?did=8420&AspxAutoDetectCookieSupport=1>

Arzhanov, M.M., and Mokhov, I.I. (2013). Temperature trends in the permafrost of the Northern     Hemisphere: Comparison of model calculations with observations. Doklady Earth Sciences, 449(1), 319–323. Retrieved June 2020, from <10.1134/S1028334X1303001X>

Astrup, R., Bernier, P.Y., Genet, H., Lutz, D.A. and Bright, R.M. (2018). A sensible climate solution for the boreal forest. Nature Climate Change, 8, 11–12. Retrieved June 2020, from <https://doi.org/10.1038/s41558-017-0043-3>

Asian Development Bank (2012). Climate Risk and Adaptation in the Electric Power Sector. Retrieved June 2020, from <https://www.adb.org/sites/default/files/publication/29889/climate-risks-adaptation-power-sector.pdf>

Asian Development Bank (2013). Guidelines for Climate Proofing Investment in the Energy Sector. Retrieved June 2020, from <https://www.adb.org/documents/guidelines-climate-proofing-investment-energy-sector>

Associated Engineering Limited (2018). Final Report: Improving Coastal Flood Adaptation Approaches. FCM MCIP 15274. Retrieved June 2020, from <https://www.surrey.ca/files/CFAS-ICFAA-FinalReport-29032018.pdf>

Association des Stations de Ski du Québec (2016). Congrès 2016: Le Québec a atteint le 5,125 millions de jours/ski pour la saison 2015-2016. Tourismexpress. Retrieved June 2020, from <https://tourismexpress.com/nouvelles/congres-assq-2016-le-quebec-a-atteint-le-5-125-millions-de-jours-ski-pour-la-saison-2015-16>

Aubertin, M., Bussière, B. and Bernier, L. (2002). Environnement et gestion des résidus miniers. Presses Internationales de Polytechnique, Corporation de l’École Polytechnique de Montréal, Montréal, Canada.

Aubertin, M., Bussière, B., Barbera, J.M., Chapuis, R.P., Monzon, M. and Aachib, M. (1997). Construction and instrumentation of in situ test plots to evaluate covers built with clean tailings. 4th International Conference on Acid Rock Drainage, Vancouver, British Columbia, Canada, vol. 2, 715–730.

Aubertin, M., Bussière, B., James, M., Jaouhar, E.M., Martin, V., Pépin, N., Mbonimpa, M. and Chapuis, R.P. (2011). Vers une mise à jour des critères de stabilité géotechnique pour la conception des ouvrages de retenue de résidus miniers. Symposium 2011 sur l’environnement et les mines, Rouyn-Noranda, Québec, Canada.

Aubertin, M., Bussière, B., Pabst, T., James, M. and Mbonimpa, M. (2016). Review of the reclamation techniques for acid-generating mine wastes upon closure of disposal sites. Geo-Chicago 2016, 343–358. Retrieved June 2020, from <https://ascelibrary.org/doi/10.1061/9780784480137.034>

Aubertin, M., Pabst, T., Bussière, B., James, M., Mbonimpa, M. Benzaazoua and Maqsoud, A. (2015). Revue des meilleures pratiques de restauration des sites d’entreposage de rejets miniers générateurs de DMA. Symposium 2015 sur l’environnement et les mines, Rouyn-Noranda, Québec, Canada.

Averchenkova, A., Crick, F., Kocornik-Mina, A., Leck, H. and Surminski, S. (2016). Multinational and large national corporations and climate adaptation: Are we asking the right questions? A review of current knowledge and a new research perspective. Wiley Interdisciplinary Reviews: Climate Change, 7(4), 517–536. Retrieved June 2020, from <https://doi.org/10.1002/wcc.402>

Azam, S. and Li, Q. (2010). Tailings dam failures: a review of the last one hundred years. Geotechnical News, 28(4), 50–54. Retrieved June 2020, from <https://www.semanticscholar.org/paper/Tailings-Dam-Failures%3A-A-Review-of-the-Last-One-Azam-Li/e57ebdac0a801b412cefd42017c2dded29cafd41>

Bansal, P., Kim, A. and Wood, M.O (2017). Hidden in Plain Sight: The Importance of Scale on Organizational Attention to Issues. Academy of Management Review, 43(2). January, amr.2014.0238. Retrieved June 2020, from <https://doi.org/10.5465/amr.2014.0238>

Barton, A., Hales, B., Waldbusser, G., Langdon, C. and Feely, R.A. (2012). The Pacific oyster, Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels: Implications for near-term ocean acidification effects. Limnology and Oceanography, 57(3), 698–710. Retrieved June 2020, from <https://doi.org/10.4319/lo.2012.57.3.0698>

Beardy’s and Okemasis’ Cree Nation (n.d.). Fire. Retrieved January 2019, from <https://bofn9697.com/tag/fire>

Becken, S., Whittlesea, E., Schliephack, J. and Scott, D. (2020). Tourism and climate change: evaluating the extent of policy integration. Journal of Sustainable Tourism, 25(10), 1–18. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2020.1745217>

Becker, A., Ng, A., McEvoy, D. and Mullett, J. (2018). Implications of climate change for shipping: Ports and supply chains. Wiley Interdisciplinary Reviews: Climate Change, 9(2). Retrieved June 2020, from <https://doi.org/10.1002/wcc.508>

Bentz, B.J., Régnière, J., Fettig, C.J., Hansen, E.M., Hayes, J.L., Hicke, J.A., Kelsey, R.G., Negron, J.F. and Seybold, S.J. (2010). Climate change and bark beetles of the western United States and Canada: direct and indirect effects. BioScience, 60(8), 602–613. Retrieved June 2020, from <https://doi.org/10.1525/bio.2010.60.8.6>

Berkhout, F. (2012). Adaptation to Climate Change by Organizations. Wiley Interdisciplinary Reviews: Climate Change, 3(1), 91–106. Retrieved June 2020, from <https://doi.org/10.1002/wcc.154>

Bernier, P.Y., Gauthier, S., Jean, P.-O., Manka, F., Boulanger, Y., Beaudoin, A. and Guindon, L. (2016). Mapping local effects of forest properties on fire risk across Canada. Forests, 7(8), 157. Retrieved June 2020, from <https://doi.org/10.3390/f7080157>

Berry, P., Clarke, K., Fleury, M.D. and Parker, S. (2014). Human health, Chapter 7 in Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation, (Eds.) F.J. Warren and D.S. Lemmen. Government of Canada, Ottawa, Ontario, 191–232. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2014/pdf/Chapter7-Human-Health_Eng.pdf>

Beverly, J.L. and Bothwell, P. (2011). Wildfire evacuations in Canada 1980–2007. Natural Hazards, 59(1), 571–596. Retrieved June 2020, from <https://doi.org/10.1007/s11069-011-9777-9>

Bindoff, N.L., Cheung, W.W.L., Kairo, J.G., Arístegui, J., Guinder, V.A., Hallberg, R., Hilmi, N., Jiao, N., Karim, M.S., Levin, L., O’Donoghue, S., Purca Cuicapusa, S.R., Rinkevich, B., Suga, T., Tagliabue, A. and Williamson, P. ( 2019). Changing Ocean, Marine Ecosystems, and Dependent Communities, in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate, (Eds.) H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M. Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer. Retrieved June 2020, from <https://www.ipcc.ch/srocc/chapter/chapter-5/>

Blackwell, B.A., Shrimpton, G., Steele, F., Ohlson, D.W. and Needoba, A. (2008). Development of a Wildfire Risk Management System for British Columbia Transmission Corporation’s Rights-of-Way. Environment Concerns in Rights-of-Way Management 8th International Symposium, 793–801. Retrieved June 2020, from <https://doi.org/10.1016/B978-044453223-7.50085-X>

Bladon, K.D. (2018). Rethinking wildfires and forest watersheds. Science, 359(6379), 1001–1002. Retrieved June 2020, from <https://doi.org/10.1126/science.aar8120>

Blight, G.H. (2010). Geotechnical Engineering for Mine Waste Storage Facilities. Taylor and Francis Group, London, United Kingdom, CRC Press, 652 p.

Blowes, D.W., Ptacek, C.J., Jambor, J.L. and Weisener, C.G. (2003). The geochemistry of acid mine drainage. Treatise on Geochemistry, 9, 149–204. Retrieved June 2020, from <https://doi.org/10.1016/B0-08-043751-6/09137-4>

Boiffin, J. and Munson, A.D. (2013). Three large fire years threaten resilience of closed crown black spruce forests in eastern Canada. Ecosphere, 4(5), 1–20. Retrieved June 2020, from <https://doi.org/10.1890/ES13-00038.1>

Bombardier Recreational Products (2017). Annual Information Form – 23 March 2017 (Fiscal Year Ended 31 January 2017). Retrieved June 2020, from <http://ir.brp.com/phoenix.zhtml?c=251742&p=irol-reportsannual>

Bonsal, B.R., Peters, D.L., Seglenieks, F., Rivera, A. and Berg, A. (2019). Changes in freshwater availability across Canada, Chapter 6 in Canada’s Changing Climate Report, (Eds.) E. Bush and D.S. Lemmen. Government of Canada, Ottawa, Ontario, 261–342. Retrieved June 2020, from <https://changingclimate.ca/CCCR2019/chapter/6-0/>

Boucher, D., Boulanger, Y., Aubin, I., Bernier, P.Y., Beaudoin, A., Guindon, L. and Gauthier, S. (2018). Current and projected cumulative impacts of fire, drought, and insects on timber volumes across Canada. Ecological Applications, 285, 1245–1259. Retrieved June 2020, from <https://doi.org/10.1002/eap.1724>

Boulanger, Y., Gauthier, S. and Burton, P.J. (2014). A refinement of models projecting future Canadian fire regimes using homogeneous fire regime zones. Canadian Journal of Forest Research, 44(4), 365–376. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2013-0372>

Boulanger, Y., Girardin, M., Bernier, P.Y., Gauthier, S., Beaudoin, A. and Guindon, L. (2017). Changes in mean forest age in Canada’s forests could limit future increases in area burned but compromise potential harvestable conifer volumes. Canadian Journal of Forest Research, 47(6), 755–764. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2016-0445>

Boyle, J., Cunningham, M. and Dekens, J. (2013). Climate Change Adaptation and Canadian Infrastructure: A review of the literature. International Institute of Sustainable Development Report, 35 p. Retrieved June 2020, from <https://www.iisd.org/pdf/2013/adaptation_can_infrastructure.pdf>

Braun, M. (2016). New Climate Normals for Electricity Demand Forecasting, Case Study prepared for Climate Change Impacts and Adaptation Division, Natural Resources Canada, 10. Retrieved June 2020, from <https://www.ouranos.ca/publication-cientifique/CaseStudy7_EN_final.pdf>

Braun, M. and Fournier, E. (2016): Adaptation Case Studies in the Energy Sector – Overcoming Barriers to Adaptation, Report presented to Climate Change Impacts and Adaptation Division, Natural Resources Canada, 114 p. Retrieved June 2020, from <https://www.ouranos.ca/publication-scientifique/ReportCaseStudies-EN.pdf>

Breitburg, D., Levin, L.A., Oschlies, A., Grégoire, M., Chavez, F.P., Conley, D.J., Garçon, V., Gilbert, D., Gutiérrez, D., Isensee, K., Jacinto, G.S., Limburg, K.E., Montes, I., Naqvi, S.W.A., Pitcher, G.C., Rabalais, N.N., Roman, M.R., Rose, K.A., Seibel, B.A., Telszewski, M., Yasuhara, M. and Zhang, J. (2018). Declining oxygen in the global ocean and coastal waters. Science, 359(6371), eaam7240. Retrieved June 2020, from <http://dx.doi.org/10.1126/science.aam7240>

Bresson, É., Demers, I., Chavaillaz, Y., Roy, P. and Pabst, T. (2018). Effectiveness of oxygen barriers as reclamation methods under climate change: definition of a drought index. Proceedings of the Tailings and Mine Waste 2018 Conference, Keystone, Colorado, USA. Retrieved June 2020, from <ftp://ftp.mern.gouv.qc.ca/Public/Dc/Conferences_Quebec-Mines-2018/Formation_Nouvelles%20exigences_changements%20climatiques/Bresson%20et%20al%202018.pdf>

Brickman, D., Hebert, D. and Wang, Z. (2018). Mechanism for the recent ocean warming events on the Scotian Shelf of Eastern Canada. Continental Shelf Research, 156(15), 11–22. Retrieved June 2020, from <https://doi.org/10.1016/j.csr.2018.01.001>

British Columbia Ministry of Environment and Climate Change Strategy (2019). Preliminary Strategic Climate Risk Assessment for British Columbia, Report prepared for the Government of British Columbia, Victoria, British Columbia. Retrieved June 2020, from <https://www2.gov.bc.ca/gov/content/environment/climate-change/adaptation/risk-assessment>

Brodeur, R., Perry, I., Boldt, J., Flostrand, L., Galbraith, M., King, J., Murphy, J., Sakuma, K. and Thompson, A. (2018). An unusual gelatinous plankton event in the NE Pacific: The Great Pyrosome Bloom of 2017. PICES Press, 26(1), 22–27. Retrieved June 2020, from <https://www.researchgate.net/publication/322665037_An_unusual_gelatinous_plankton_event_in_the_NE_Pacific_The_Great_Pyrosome_Bloom_of_2017>

Brown, C.D. and Johnstone, J.F. (2012). Once burned, twice shy: Repeat fires reduce seed availability and alter substrate constraints on Picea mariana regeneration. Forest Ecology and Management, 266, 34–41. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2011.11.006>

Bureau of Transportation Statistics (2018). Border Crossing/Entry Data. Retrieved June 2020, from <https://www.bts.gov/content/border-crossingentry-data>

Burton, P.J. (2010). Striving for sustainability and resilience in the face of unprecedented change: the case of the mountain pine beetle outbreak in British Columbia. Sustainability, 2(8), 2403–2423. Retrieved June 2020, from <https://doi.org/10.3390/su2082403>

Bush, E. and Lemmen, D.S. (Eds.) (2019). Canada’s Changing Climate Report. Government of Canada, Ottawa, Ontario, 444. Retrieved June 2020, from <https://changingclimate.ca/CCCR2019/>

Bussière, B. (2007). Colloquium 2004: Hydrogeotechnical properties of hard rock tailings from metal mines and emerging geoenvironmental disposal approaches. Canadian Geotechnical Journal, 44(9), 1019–1052. Retrieved June 2020, from <https://doi.org/10.1139/T07-040>

Bussière, B., Aubertin, M. and Chapuis, R.P. (2003). The behavior of inclined covers used as oxygen barriers. Canadian Geotechnical Journal, 40(3), 512–535. Retrieved June 2020, from <https://doi.org/10.1139/t03-001>

Bussière, B., Aubertin, M., Zagury, G. J., Potvin, R. and Benzaazoua, M. (2005). Principaux défis et pistes de solution pour la restauration des aires d’entreposage de rejets miniers abandonnées. Symposium 2005 sur l’environnement et les mines. Retrieved June 2020, from <https://www.researchgate.net/publication/267855816_PRINCIPAUX_DEFIS_ET_PISTES_DE_SOLUTION_POUR_LA_RESTAURATION_DES_AIRES_D’ENTREPOSAGE_DE_REJETS_MINIERS_ABANDONNEES>

Bussière, B., Potvin, R., Dagenais, A., Aubertin, M., Maqsoud, A. and Cyr, J. (2009). Restauration du site minier Lorraine, Latulipe, Québec: Résultats de 10 ans de suivi. Déchets – Revue francophone d’écologie industrielle, 1–16. Retrieved June 2020, from <https://doi.org/10.4267/dechets-sciences-techniques.1254>

Bussière, B., Demers, I., Charron, P., Bossé, B., Roy, P., Jébrak, M. and Trépanier, S. (2017). Analyse de risques et de vulnérabilités liés aux changements climatiques pour le secteur minier québécois, 331 p. Retrieved June 2020, from <https://mern.gouv.qc.ca/wp-content/uploads/analyse-changements-climatiques-secteur-minier.pdf>

Calkin, D.E., Cohen, J.D., Finney, M.A. and Thompson, M.P. (2014). How risk management can prevent future wildfire disasters in the wildland-urban interface. Proceedings of the National Academy of Sciences, 111(2), 746–751. Retrieved June 2020, from <https://doi.org/10.1073/pnas.1315088111>

Campbell, I.D., Durant D.G., Hunter, K.L. and Hyatt, K.D. (2014). Food production, Chapter 4 in Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation, (Eds.) F.J. Warren and D.S. Lemmen. Government of Canada, Ottawa, Ontario, 99–134. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2014/pdf/Chapter4-Food-Production_Eng.pdf>

Canadian Agri-Food Trade Alliance (2020). Agri-Food Exports. Retrieved June 2020, from <http://cafta.org/agri-food-exports/>

Canadian Council of Forest Ministers (n.d.). Climate Change. Retrieved June 2020, from <http://www.ccfm.org/english/coreproducts-cc.asp>

Canadian Electricity Association (2016). Adapting to Climate Change: State of Play and Recommendations for the Electricity Sector in Canada. Retrieved June 2020, from <https://electricity.ca/wp-content/uploads/2016/02/The_Canadian_Electricity_Association_Releases_a_Roadmap_to_Climate_Change_Adaptation_final.pdf>

Canadian Dam Association (2013). Dam Safety Guidelines. Library and Archives Canada Cataloguing in Publication Data, Ottawa, Ontario.

Centre d’expertise hydrique du Québec (2019). Ministère de l’Environnement et de la Lutte contre les changements climatiques. Retrieved May 2019, from <www.cehq.gouv.qc.ca>

Certified Professional Accountants Canada (2016a). How Organizations Can Adapt to Climate Change. Chartered Professional Accountants Canada. Retrieved June 2020, from <https://www.cpacanada.ca/-/media/site/business-and-accounting-resources/docs/climate-change-mitigation-adaptation-brief-2-how-organizations-can-adapt-to-climate-change-february-2016.pdf?la=en&hash=53D86B64BA875656F49B457BD7CCFB5D4B94859F>

Certified Professional Accountants Canada (2016b). Adaptation Case Study #4: The Co-Operators. Chartered Professional Accountants Canada. <https://www.cpacanada.ca/-/media/site/business-and-accounting-resources/docs/climate-change-adaptation-case-study-4-the-co-operators-october-2015.pdf?la=en&hash=6A57F3569EF436E13C7E2BB0DA6184D7BE4EDF6B>

Chandler, P.C., King, S.A. and Perry, R.I. (Eds.) (2016). State of the physical, biological and selected fishery resources of Pacific Canadian marine ecosystems in 2015. Canadian Technical Report on Fisheries and Aquatic Sciences, 3179, viii + 230 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/365564.pdf>

Christianson, A. (2015). Social science research on Indigenous wildfire management in the 21st century and future research needs. International Journal of Wildland Fire, 24(2), 190–200. Retrieved June 2020, from <https://doi.org/10.1071/WF13048>

Ciais, P., Sabine, C., Bala, G., Bopp, L., Brovkin, V., Canadell, J., Chhabra, A., DeFries, R., Galloway, J., Heimann, M., Jones, C., Le Quéré, C., Myneni, R.B., Piao, S. and Thornton, P. (2013). Carbon and Other Biogeochemical Cycles, in Climate Change 2013: The Physical Science Basis (Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change), (Eds.) T.F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley. Cambridge University Press, Cambridge, United Kingdom and New York, New York, USA. Retrieved June 2020, from <https://www.ipcc.ch/site/assets/uploads/2018/02/WG1AR5_Chapter06_FINAL.pdf>

Circle of Blue (2018). Stranded Assets. Retrieved June 2020, from <https://www.circleofblue.org/stranded-assets>

City of Surrey (2018). CFAS Primer Part 1: Coastal Flooding in Surrey. Retrieved June 2020, from <http://www.surrey.ca/files/CFAS-Primerpart1.pdf>

Claireaux, G. and Chabot, D. (2016). Responses by fishes to environmental hypoxia: integration through Fry’s concept of aerobic metabolic scope. Journal of Fish Biology, 88, 232–251. Retrieved June 2020, from <https://doi.org/10.1111/jfb.12833>

Coulombe, S., Bernier, P.Y. and Raulier, F. (2010). Uncertainty in detecting climate change impact on the projected yield of black spruce (Picea mariana). Forest Ecology and Management, 259(4), 730–738. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2009.06.028>

Council of Canadian Academies (2019). Canada’s Top Climate Change Risks. The Expert Panel on Climate Change Risks and Adaptation Potential, Council of Canadian Academies, Ottawa, Ontario, 70 p. Retrieved June 2020, from <https://cca-reports.ca/wp-content/uploads/2019/07/Report-Canada-top-climate-change-risks.pdf>

Creed, I., Weber, M., Accatino, F. and Kreutzweiser, D. (2016). Managing forests for water in the Anthropocene—the best kept secret services of forest ecosystems. Forests, 7(3), 60. Retrieved June 2020, from <https://doi.org/10.3390/f7030060>

Cross Dependency Initiative (2019). Building climate resilience across critical infrastructure, 4 p. Retrieved June 2020, from <https://xdi.systems/wp-content/uploads/2019/05/XDI_Sydney-A3toA4__singlepages.pdf>

Crossin, G.T., Al-Ayoub, S.A., Jury, S.H., Howell, W.H. and Watson III, W.H. (1998) Behavioral thermoregulation in the American lobster Homarus americanus. Journal of Experimental Biology, 201, 365–374. Retrieved June 2020, from <https://jeb.biologists.org/content/jexbio/201/3/365.full.pdf>

Cruz, A.M. and Krausmann, E. (2013). Vulnerability of the oil and gas sector to climate change and extreme weather events. Climatic Change, 121(1), 41–53. Retrieved June 2020, from <https://link.springer.com/article/10.1007/s10584-013-0891-4>

CTV (2018). All aboard: First passenger train in 18 months departs for Churchill, Man. Retrieved June 2020, from <https://www.ctvnews.ca/mobile/canada/all-aboard-first-passenger-train-in-18-months-departs-for-churchill-man-1.4201440>

Da Silva, L., Desrochers, F.-A., Pineault, K., Gosselin, C.-A., Grenier, P. and Larose, G. (2019). Analyse économique des mesures d’adaptation aux changements climatiques appliquée au secteur du ski alpin au Québec. Ouranos, Montréal, 119 p. Retrieved June 2020, from <https://www.ouranos.ca/publication-scientifique/RapportDaSilva2019.pdf>

Dagenais, A.-M., Aubertin, M., Bussière, B., Bernier, L. and Cyr, J. (2001). Monitoring at the Lorraine Mine Site: A Follow-Up on the Remediation Plan. National Association of Abandoned Mine Land Programs Annual Conference, August 19–22, Athens, Ohio, 19 p. Retrieved June 2020, from <http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.492.4689&rep=rep1&type=pdf>

Daniel, C.J., Ter-Mikaelian, M.T., Wotton, B.M., Rayfield, B. and Fortin, M.-J. (2017). Incorporating uncertainty into forest management planning: Timber harvest, wildfire and climate change in the boreal forest. Forest Ecology and Management, 400, 542–554. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2017.06.039>

Dawson, J., Stewart, E.J., Lemelin, R.H. and Scott, D. (2010). The carbon cost of polar bear viewing tourism in Churchill, Canada. Journal of Sustainable Tourism, 18(3), 319–336. Retrieved June 2020, from <https://doi.org/10.1080/09669580903215147>

Dawson, J., Stewart, E.J., Johnston, M.E. and Lemieux, C.J. (2016). Identifying and evaluating adaptation strategies for cruise tourism in Arctic Canada. Journal of Sustainable Tourism, 24(10), 1425–1441. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2015.1125358>

Dawson, J., Copland, L., Johnston, M.E., Pizzolato, L., Howell, S., Pelot, R., Etienne, L., Matthews, L. and Parsons, J. (2017). Climate change adaptation strategies and policy options for Arctic shipping, Report prepared for Transport Canada. Retrieved June 2020, from <https://ruor.uottawa.ca/handle/10393/36016>

Dawson, J., Pizzolato. L., Howell, S., Copland, L. and Johnston, M. (2018). Temporal and spatial patterns of ship traffic in the Canadian Arctic from 1990 to 2015. Arctic, 71(1), 15–26. Retrieved June 2020, from <https://doi.org/10.14430/arctic4698>

Dawson, R. (2015). Handling interdependencies in climate change risk assessment. Climate, 3(4), 1079–1096. Retrieved June 2020, from <https://doi.org/10.3390/cli3041079>

Department of Environment, Food and Rural Affairs (2011). Climate Resilient Infrastructure: Preparing for a Changing Climate, 71 p. Retrieved June 2020, from <https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/69269/climate-resilient-infrastructure-full.pdf>

Derksen, C., Burgess, D., Duguay, C., Howell, S., Mudryk, L., Smith, S., Thackeray, C. and Kirchmeier-Young, M. (2019). Changes in snow, ice, and permafrost across Canada, Chapter 5 in Canada’s Changing Climate Report, (Eds.) E. Bush and D. S. Lemmen. Government of Canada, Ottawa, Ontario, 194–260. Retrieved June 2020, from <https://changingclimate.ca/CCCR2019/chapter/5-0/>

Dewar, J.A. and Wachs, M. (2008). Transportation planning, climate change, and decision making under uncertainty. Transportation Research Board, 26 p. Retrieved June 2020, from <http://onlinepubs.trb.org/onlinepubs/sr/sr290dewarwachs.pdf>

Dixon, L., Tsang, F. and Fitts, G. (2018). The Impact of Changing Wildfire Risk on California’s Residential Insurance Market. The RAND Corporation, California, USA, 105 p. Retrieved June 2020, from <http://climateassessment.ca.gov/techreports/docs/20180827-Forests_CCCA4-CNRA-2018-008.pdf>

Doré, G. and Zubeck, H. K. (2009). Cold regions pavement engineering. ASCE Press, Reston, Virginia, 403 p.

Dougherty-Choux, L., Terpstra, P., Kamila, S. and Kurukulasuriya, P. (2015). Adapting from the Ground Up. World Resources Institute, Washington, DC. Retrieved June 2020, from <https://www.preventionweb.net/go/47027>

Downing, T.E. (2012). Views of the frontiers in climate change Adaptation economics. Wiley Interdisciplinary Reviews: Climate Change, 3(2), 161–70. Retrieved June 2020, from <https://doi.org/10.1002/wcc.157>

Dubois, A., Hulthen, K. and Pedersen, A. (2011). Interdependence Within and Among “Supply Chains”. 21 p. Retrieved June 2020, from <https://www.researchgate.net/profile/Kajsa_Hulthen/publication/242489338_INTERDEPENDENCE_WITHIN_AND_AMONG_aposSUPPLY_CHAINSapos/links/54006e0c0cf24c81027dec7e/INTERDEPENDENCE-WITHIN-AND-AMONG-aposSUPPLY-CHAINS-apos.pdf>

Dupont-Prinet, A., Vagner, M., Chabot, D. and Audet, C. (2013). Impact of hypoxia on the metabolism of Greenland halibut (Reinhardtius hippoglossoides). Canadian Journal of Fisheries and Aquatic Sciences, 70, 461–469. Retrieved June 2020, from <https://doi.org/10.1139/cjfas-2012-0327>

Dymond, C.C., Spittlehouse, D.L., Tedder, S., Hopkins, K., McCallion, K. and Sandland, J. (2015). Applying resilience concepts in forest management: a retrospective simulation approach. Forests, 6(12), 4421–4438. Retrieved June 2020, from <https://doi.org/10.3390/f6124377>

Earnscliffe Strategy Group (2018). National Climate Change Adaptation Survey, Report prepared for Climate Change Impacts and Adaptation Division, Natural Resources Canada, Ottawa.

Edwards, J.E., Pearce, C., Ogden, A.E. and Williamson, T.B. (2015). Climate change and sustainable forest management in Canada: a guidebook for assessing vulnerability and mainstreaming adaptation into decision making. Canadian Council of Forest Ministers, Climate Change Task Force, 172 p. Retrieved June 2020, from <https://www.ccfm.org/pdf/Vulnerability%20Guidebook_June2_EN.pdf>

Ekstrom, J.A., Suatoni, L., Cooley, S.R., Pendelton, L.H., Waldbusser, G.G., Cinner, J.E., Ritter, J., Langdon, C., van Hooidonk, R., Gledhill, D., Wellman, K., Beck, M.W., Brander, L.M., Rittschof, D., Doherty, C., Edwards, P.E.T. and Portela, R. (2015). Vulnerability and adaptation of US shellfisheries to ocean acidification. Nature Climate Change, 5, 207–214. Retrieved June 2020, from <https://doi.org/10.1038/nclimate2508>

Environment and Climate Change Canada (2018). Canadian Centre for Climate Services. Retrieved December 2019, from <https://www.canada.ca/en/environment-climate-change/services/climate-change/canadian-centre-climate-services.html>

Environment and Climate Change Canada (2020). Strategic assessment of climate change (revised October 2020), 20 p.  Retrieved October 2020, from <https://www.canada.ca/en/services/environment/conservation/assessments/strategic-assessments/climate-change.html>

Eriksson, N. and Adamek, P. (2000). The tailings pond failure at the Aznalcóllar mine, Spain. British Columbia MEND ML/ARD Annual Workshop, 8 p. Retrieved June 2020, from <http://bc-mlard.ca/files/presentations/2016-19-ERIKSSON-ADAMEK-tailings-pond-failure-aznalcollar.pdf>

Ernst, B., Orensanz, J.M. and Armstrong, D.A. (2005). Spatial dynamics of female snow crab (Chionoecetes opilio) in the eastern Bering Sea. Canadian Journal of Fisheries and Aquatic Sciences, 62, 250–268. Retrieved June 2020, from <https://doi.org/10.1139/f04-201>

Ethier, M.P., Bussière, B., Broda, S. and Aubertin, M. (2018). Three-dimensional hydrogeological modeling to assess the elevated-water-table technique for controlling acid generation from an abandoned tailings site in Quebec, Canada. Hydrogeology Journal, 26(4), 1201–1219. Retrieved June 2020, from <https://doi.org/10.1007/s10040-017-1713-y>

Executive Office of the President (2013). Economic benefits of increasing electric grid resilience to weather outages. Retrieved June 2020, from <http://energy.gov/ sites/prod/files/2013/08/f2/Grid%20Resiliency%20 Report_FINAL.pdf>

Eyzaguirre, J. and Warren, F.J. (2014). Adaptation: Linking research and practice, Chapter 9 in Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation, (Eds.) F.J. Warren and D. S. Lemmen. Government of Canada, Ottawa, Ontario, 253–286. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2014/pdf/Chapter9-Adaptation_Eng.pdf>

Eyzaguirre, J. (2016). FAQ 11: How do the costs of climate change impacts compare to the costs of adaptation?, in Canada’s Marine Coasts in a Changing Climate, (Eds.) D.S. Lemmen, F.J. Warren, T.S. James and C.S.L. Mercer Clarke. Government of Canada, Ottawa, Ontario, 270‒272. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2016/Coastal_Assessment_Chapter7_FrequentlyAskedQuestions.pdf>

Federal, Provincial and Territorial Emergency Management Framework Task Team. (2016). Emergency Management Framework for Agriculture in Canada; Agriculture and Agri-Food Canada, 22 p. Retrieved June 2020, from <http://www.agr.gc.ca/resources/prod/doc/pdf/emf-gu_2016-eng.pdf>

Finlay, S.E., Moffat, A., Gazzard, R., Baker, D. and Murray, V. (2012). Health impacts of wildfires. PLoS Currents Disasters, Edition 1. Retrieved June 2020, from <https://doi.org/10.1371/4f959951cce2c>

Finley, T. and Schuchard, R. (n.d.). Adapting to climate change: a guide for the transportation industry. BSR Industry Series, 7 p. Retrieved June 2020, from <https://www.bsr.org/reports/BSR_Climate_Adaptation_Issue_Brief_Transportation.pdf>

FireSmart Canada (2018). FireSmart Guidebook for the Oil and Gas Industry. Retrieved June 2020, from <https://www.firesmartcanada.ca/mdocs-posts/firesmart-guidebook-for-the-oil-and-gas-industry>

FireSmart Canada (2019a). FireSmart Videos. Retrieved June 2020, from <https://www.firesmartcanada.ca/resources-library/firesmart-videos>

FireSmart Canada (2019b). Vegetation Management. Retrieved June 2020, from <https://www.firesmartcanada.ca/what-is-firesmart/vegetation-management>

Fisheries and Oceans Canada (2018a). Seafisheries: 2016 value of Atlantic Coast commercial landings, by region. Retrieved June 2020, from <http://www.dfo-mpo.gc.ca/stats/commercial/land-debarq/sea-maritimes/s2016av-eng.htm>

Fisheries and Oceans Canada (2018b). Assessment of the Estuary and northern Gulf of St. Lawrence (areas 13 to 17, 12A, 12B, 12C and 16A) snow crab stocks in 2017. DFO Canadian Science Advisory Secretariat Science Advisory Report 2018/047, 30 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/40738565.pdf>

Fisheries and Oceans Canada (2018c). Assessment of northern shrimp stocks in the Estuary and Gulf of St. Lawrence in 2017. DFO Canadian Advisory Secretariat Science Advisory Report 2018/015, 17 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/40688495.pdf>

Fisheries and Oceans Canada (2018d). Assessment of the Greenland halibut stock in the Gulf of St. Lawrence (4rst) in 2017. DFO Canadian Advisory Secretariat Science Advisory Report 2018/035, 20 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/40714007.pdf>

Fisheries and Oceans Canada (2018e). Proceedings on the National Peer Review on Incorporating environmental and climate change considerations into population assessments in DFO’s Science advisory process, May 8–9, 2018. DFO Canadian Scientific Advisory Secretariat.

Flannigan, M.D., Krawchuk, M.A., De Groot, W.J., Wotton, B.M. and Gowman, L.M. (2009). Implications of changing climate for global wildland fire. International Journal of Wildland Fire, 18(5), 483–507. Retrieved June 2020, from <https://doi.org/10.1071/WF08187>

Flannigan, M., Cantin, A.S., De Groot, W.J., Wotton, M., Newbery, A. and Gowman, L.M. (2013). Global wildland fire season severity in the 21st century. Forest Ecology and Management, 294, 54–61. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2012.10.022>

Food and Agriculture Organization (2017). The future of food and agriculture: Trends and challenges; Food and Agriculture Organization of the United Nations Rome, 180 p. Retrieved June 2020, from <http://www.fao.org/3/a-i6583e.pdf>

Fournier, É. (2016). Fine-Tuning Observations to Better Manage and Design Hydroelectricity Assets, Case Study prepared for Climate Change Impacts and Adaptation Division, Natural Resources Canada, 9 p. Retrieved June 2020, from <https://www.ouranos.ca/publication-scientifique/CaseStudy2_EN_final.pdf>

Furness, E. and Nelson, H. (2016). Are human values and community participation key to climate adaptation? The case of community forest organisations in British Columbia. Climatic Change, 135(2), 243–259. Retrieved June 2020, from <https://doi.org/10.1007/s10584-015-1564-2>

Furrer, B., Hoffmann, V. and Marion S. (2009). Banking and Climate Change: Opportunities and Risks. SAM (Sustainable Asset Management), Zurich, 51 p. Retrieved June 2020, from <http://www.sustainabilitycompany.it/img/text/SAM_ETH_Study_Banking.pdf>

Fyfe, J., Derksen, C., Mudryk, L., Flate, G.M., Santer, B.D., Swart, N.C., Molotech, N.P., Zhang, X., Wan, H., Arora, V.K., Scinocca, J. and Jiao, Y. (2017). Large near-term projected snowpack loss over the western United States. Nature Communications, 8. Retrieved June 2020, from <https://doi.org/10.1038/ncomms14996>

Galbraith, P.S., Chassé, J., Caverhill, C., Nicot, P., Gilbert, D., Pettigrew, B., Lefaivre, D., Brickman, D., Devine, L., Lafleur, C. (2018). Physical oceanographic conditions in the Gulf of St. Lawrence during 2017. DFO Canadian Science Advisory Secretariat, Research Document 2018/050, v + 79 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/40759702.pdf>

Gatin, J. and Johnston, M. (2017). State of Play Report: Forestry Adaptation Initiatives across Canada. Canadian Forestry Service, Natural Resources Canada, Ottawa.

Gauthier, S., Bernier, P. Y., Boulanger, Y., Guo, J., Guindon, L., Beaudoin, A. and Boucher, D. (2015). Vulnerability of timber supply to projected changes in fire regime in Canada’s managed forests. Canadian Journal of Forest Research, 45(11), 1439–1447. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2015-0079>

Gauthier, S., Bernier, P., Burton, P.J., Edwards, J., Isaac, K., Isabel, N., Jayen, K., Le Goff, H. and Nelson, E.A. (2014). Climate change vulnerability and adaptation in the managed Canadian boreal forest. Environmental Reviews, 22(3), 256–285. Retrieved June 2020, from <https://doi.org/10.1139/er-2013-0064>

Gazeau, F., Parker, L.M., Comeau, S., Gattuso, J.-P., O’Connor, W.A., Martin, S., Pörtner, H.-O. and Ross, P. (2013). Impacts of ocean acidification on marine shelled molluscs. Marine Biology, 160 (8), 2207–2245. Retrieved June 2020, from <https://doi.org/10.1007/s00227-013-2219-3>

Gill, A.M., Stephens, S.L. and Cary, G.J. (2013). The worldwide “wildfire” problem. Ecological Applications, 23(2), 438–454. Retrieved June 2020, from <https://doi.org/10.1890/10-2213.1>

Gledhill, D.K., White, M.M., Salisbury, J., Thomas, H., Mlsna, I., Liebman, M., Mook, B., Grear, J., Candelmo, A.C., Chambers, R.C., Gobler, C.J., Hunt, C.W. King, A.L., Price, N.N., Signorini, S.R., Stancioff, E., Stymiest, C., Wahle R.A., Waller, J.D., Rebuck, N.D., Wang, Z.A., Capson, T.L., Morrison, J.R., Cooley, S.R. and Doney, S.C. (2015). Ocean and Coastal Acidification off New England and Nova Scotia. Oceanography, 28(2), 182–197. Retrieved June 2020, from <https://doi.org/10.5670/oceanog.2015.41>

Globe and Mail (2018). “It’s devastating to see”: How a rail line crisis is threatening Churchill, Manitoba. Retrieved June 2020, from <https://www.theglobeandmail.com/politics/article-churchill-man-residents-angry-and-frustrated-as-solutions-to-rail>

Goertzen, S. (2018). Agricultural Solutions to Climate Change: Findings from the Manitoba Agriculture Climate Initiative. Keystone Agricultural Producers, 30 p. Retrieved June 2020, from <https://486d1ae5-bba6-4d54-8d48-301060955887.filesusr.com/ugd/b43985_b680623ee3504c099fc30c3ccf48d960.pdf>

Goodrick, S.L., Achtemeier, G.L., Larkin, N.K., Liu, Y. and Strand, T.M. (2013). Modelling smoke transport from wildland fires: a review. International Journal of Wildland Fire, 22(1), 83–94. Retrieved June 2020, from <https://doi.org/10.1071/WF11116>

Government of Canada (2016). Pan-Canadian framework on clean growth and climate change: Canada’s plan to address climate change and grow the economy. 77 p. Retrieved June 2020, from <https://www.canada.ca/content/dam/themes/environment/documents/weather1/20170125-en.pdf>

Government of Canada (2020) Wildland fire evacuations, Natural Resources Canada. Retrieved May 2020 from <https://www.nrcan.gc.ca/climate-change/impacts-adaptations/climate-change-impacts-forests/forest-change-indicators/wildland-fire-evacuations/17787>

Gray, R. (2015). The Economic Impacts of Elevated Export Basis Levels on Western Canadian Grain Producers 2012/13, 2013/14, and 2014/15, Report submitted to the Saskatchewan Wheat Development Commission. Retrieved June 2020, from <https://static1.squarespace.com/static/5c40f31a620b85cf0d073e7b/t/5cdd901fc9048b0001f4599d/1558024227090/The-Economic-Impacts-Of-Elevated-Export-Basis-Levels-On-Western-Canadian-Grain-Producers-2012-2015.pdf>

Greenan, B.J.W., James, T.S., Loder, J.W., Pepin, P., Azetsu-Scott, K., Ianson, D., Hamme, R.C., Gilbert, D., Tremblay, J-E., Wang, X.L. and Perrie, W. (2019a). Changes in oceans surrounding Canada, Chapter 7 in Canada’s Changing Climate Report, (Eds.) Bush and Lemmen. Government of Canada, Ottawa, Ontario, 343–423. Retrieved June 2020, from <https://changingclimate.ca/CCCR2019/chapter/7-0/>

Greenan, B.J.W., Shackell, N.L., Ferguson, K., Greyson, P., Cogswell, A., Brickman, D., Wang, Z., Cook, A., Brennan, C.E. and Saba, V.S. (2019b). Climate Change Vulnerability of American Lobster Fishing Communities in Atlantic Canada. Frontiers in Marine Science, 6, 579 p. Retrieved June 2020, from <https://doi.org/10.3389/fmars.2019.00579>

Groulx, M., Lemieux, C., Dawson, J., Stewart, E. and Yudina, O. (2016). Motivations to engage in last chance tourism in the Churchill wildlife management area and Wapusk National Park: The role of place identity and nature relatedness. Journal of Sustainable Tourism, 24(11), 1523–1540. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2015.1134556>

Groulx. M., Lemieux, C.J., Lewis, J.L. and Brown, S. (2017). Understanding consumer behaviour and adaptation planning responses to climate-driven environmental change in Canada’s parks and protected areas: a climate futurescapes approach. Journal of Environmental Planning and Management, 60(6), 1016–1035. Retrieved June 2020, from <https://doi.org/10.1080/09640568.2016.1192024>

Guittonny, M., Bussière, B., Maqsoud, A., Proteau, A., Ben Khouya, T. and Botula, Y.D. (2018). Colonisation racinaire dans les recouvrements miniers et impact sur leur fonctionnement. Proceedings of the 6th Symposium on Mines and the Environment, June 2018, Rouyn-Noranda, Canada.

Guthrie, R.H., Mitchell, S.J., Lanquaye-Opoku, N. and Evans, S. G. (2010). Extreme weather and landslide initiation in coastal British Columbia. Quarterly Journal of Engineering Geology and Hydrogeology, 43(4), 417–428. Retrieved June 2020, from <https://doi.org/10.1144/1470-9236/08-119>

Hanes, C., Wang, X., Jain, P., Parisien, M.-A., Little, J. and Flannigan, M. (2018). Fire regime changes in Canada over the last half century. Canadian Journal of Forest Research, 49(3), 256-269. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2018-0293>

Hatchett, B. and Eisen, H. (2018). Early season snowpack loss and implications for over-snow vehicle recreation travel planning. The Cryosphere. Retrieved June 2020, from <https://doi.org/10.5194/tc-2018-181>

He, W., Yang, J.Y., Qian, B., Drury, C.F., Hoogenboom G., He, P., Lapen, D. and Zhou, W. (2018). Climate change impacts on crop yield, soil water balance and nitrate leaching in the semiarid and humid regions of Canada. PLoS ONE, 13(11), e0207370. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0207370>

Hedlund, J., Fick, S., Carlsen, H. and Benzie, M. (2018). Quantifying transnational climate impact exposure: New perspectives on the global distribution of climate risk. Global Environmental Change, 52, 75-85. Retrieved June 2020, from <https://doi.org/10.1016/j.gloenvcha.2018.04.006>

Henstra, D. and Thistlethwaite, J. (2017). Climate Change, Floods, and Municipal Risk Sharing in Canada. IMFG Papers on Municipal Finance and Governance, No. 30, Institute on Municipal Finance and Governance, Toronto, 48 p. Retrieved June 2020, from <https://munkschool.utoronto.ca/imfg/uploads/373/1917_imfg_no_30_online_final.pdf>

Hewer, M. and Gough, W. (2018). Thirty years of assessing the impacts of climate change on outdoor recreation and tourism in Canada. Tourism Management Perspectives, 26, 179–192. Retrieved June 2020, from <https://doi.org/10.1016/j.tmp.2017.07.003>

Hirsch, K., Kafka, V., Tymstra, C., McAlpine, R., Hawkes, B., Stegehuis, H., Quintilio, S., Gauthier, S. and Peck, K. (2001). Fire-smart forest management: a pragmatic approach to sustainable forest management in fire-dominated ecosystems. The Forestry Chronicle, 77(2), 357–363. Retrieved June 2020, from <https://doi.org/10.5558/tfc77357-2>

Holubec, I. (2004). Covers for Reactive Tailings Located in Permafrost Regions Review. MEND Report 1.61. 6., Minerals and Metals Sector, Natural Resources Canada, Ottawa. Retrieved June 2020, from <http://mend-nedem.org/wp-content/uploads/1.61.4CoversPermafrostRegions.pdf>

Hope, E.S., McKenney, D.W., Pedlar, J.H., Stocks, B.J. and Gauthier, S. (2016). Wildfire suppression costs for Canada under a changing climate. PLoS ONE, 11(8), e0157425. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0157425>

Hotton, G., Bresson, É., Bussière, B., Pabst, T., Demers, I. and Roy, P. (2018). Assessment of CCBE performance with climate change: case study of the Lorraine mine site. Proceedings of the Tailings and Mine Waste 2018 Conference, Keystone, Colorado, USA.

Hotton, G., Bussière, B., Pabst, T., Bresson, E. and Roy, P. (2019). Influence of climate change on the ability of a cover with capillary barrier effects to control acid generation. Hydrogeology Journal. Retrieved June 2020, from <https://doi.org/10.1007/s10040-019-02084-y>

Hori, Y., Gough, W.A., Butler, K. and Tsuji, L.J.S. (2017). Trends in the seasonal length and opening dates of a winter road in the western James Bay region, Ontario, Canada. Theoretical and Applied Climatology, 129(3), 1309–1320. Retrieved June 2020, from <https://doi.org/10/gbq6gv>

Hrycik, A.R, Almeida, L.Z. and Höök, T.O. (2017). Sub‐lethal effects on fish provide insight into a biologically‐relevant threshold of hypoxia. Oikos, 126, 307–317. Retrieved June 2020, from <https://doi.org/10.1111/oik.03678>

Ignaciuk, A. and Mason-D’Croz, D. (2014). Modelling Adaptation to Climate Change in Agriculture. OECD Food, Agriculture and Fisheries Papers, No. 70, OECD Publishing, Paris. Retrieved June 2020, from <https://doi.org/10.1787/5jxrclljnbxq-en>

IISD [International Institute for Sustainable Development] (2017). Building a Climate-Resilient City: Electricity and information and communication technology infrastructure. Retrieved June 2020, from <https://www.iisd.org/library/building-climate-resilient-city-electricity-and-information-and-communication-technology>

Insurance Bureau of Canada (2016). Northern Alberta Wildfire Costliest Insured Natural Disaster in Canadian History. Retrieved January 2020, from <http://www.ibc.ca/bc/resources/media-centre/media-releases/northern-alberta-wildfire-costliest-insured-natural-disaster-in-canadian-history>

Insurance Bureau of Canada (2019). Facts of the Property and Casualty Insurance Industry in Canada, 41st edition. Retrieved June 2020, from <http://assets.ibc.ca/Documents/Facts%20Book/Facts_Book/2019/IBC-2019-Facts.pdf>

International Energy Agency (2015). Making the energy sector more resilient to climate change. COP 21 Resilience Brochure. Retrieved June 2020, from <https://www.iea.org/publications/freepublications/publication/COP21_Resilience_Brochure.pdf>

International Energy Agency (2016a). Key World Energy Statistics 2016. Retrieved June 2020, from <https://www.iea.org/publications/freepublications/publication/key-world-energy-statistics.html>

International Energy Agency (2016b). International Energy Agency Workshop Report Sixth Forum on the Climate-Energy Security Nexus. Retrieved June 2020, from <https://www.iea.org/media/workshops/2016/6thnexusforum/6thNexusForumSummary.pdf>

International Energy Agency (2018). Getting Wind and Sun onto the Grid: A Manual for Policy Makers. Retrieved June 2020, from <https://www.iea.org/publications/insights/insightpublications/Getting_Wind_and_Sun.pdf>

IPCC [Intergovernmental Panel on Climate Change] (2012). Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. (A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change), (Eds.) C.B. Field, V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,M. Tignor, and P.M. Midgley. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 582 p. Retrieved June 2020, from <https://www.ipcc.ch/report/managing-the-risks-of-extreme-events-and-disasters-to-advance-climate-change-adaptation/>

IPCC [Intergovernmental Panel on Climate Change] (2014). Summary for policymakers, in Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects (Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change), (Eds.) C.B. Field, V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White. Cambridge University Press, Cambridge, United Kingdom and New York, New York, USA, 1–32. Retrieved June 2020, from <https://doi.org/10.1017/CBO9781107415379.003>

IPCC [Intergovernmental Panel on Climate Change] (2018). Summary for Policymakers, in Global Warming of 1.5°C (An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty), (Eds.) V. Masson-Delmotte, P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T. Maycock, M. Tignor, and T. Waterfield. World Meteorological Organization, Geneva, Switzerland, 24 p. Retrieved June 2020, from <https://www.ipcc.ch/sr15/chapter/spm/>

IPCC [Intergovernmental Panel on Climate Change] (2019). Summary for Policymakers, in Climate Change and Land (An IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems), (Eds.) P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pörtner, D. C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal Pereira, P. Vyas, E. Huntley, K. Kissick, M. Belkacemi and J. Malley. Retrieved June 2020, from <https://www.ipcc.ch/srccl/chapter/summary-for-policymakers/>

IRENA [International Renewable Energy Agency] (2018). Global Energy Transformation: A Roadmap to 2050. 76 p. Retrieved June 2020, from <https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2018/Apr/IRENA_Report_GET_2018.pdf>

ISO [International Standards Organization] (2018). ISO 31000:2018. Risk Management. Retrieved June 2020, from <http://www.iso.org/cms/render/live/en/sites/isoorg/contents/data/standard/06/56/65694.html>

Jaglom, W.S., McFarland, J.R., Colley, M.F., Mack, C.B., Venkatesh, B., Miller, R.L., Haydel, J., Schultz, P.A., Perkins, B., Casola, J.H., Martinich, J.A., Cross, P., Kolian, M.J. and Kayin, S. (2014). Assessment of projected temperature impacts from climate change on the U.S. electric power sector using the Integrated Planning Model®. Energy Policy, 73, 524–539. Retrieved June 2020, from <https://doi.org/10.1016/j.enpol.2014.04.032>

Jakob, M. and Lambert, S. (2009). Climate change effects on landslides along the southwest coast of British Columbia. Geomorphology, 107(3–4), 275–284. Retrieved June 2020, from <https://doi.org/10.1016/j.geomorph.2008.12.009>

Jewiss-Gaines A, Barelli, L. and Hunter, F.F. (2017). First Records of Culicoides sonorensis (Diptera: Ceratopogonidae), a Known Vector of Bluetongue Virus, in Southern Ontario. Journal of Medical Entomology, 54(3), 757–762. Retrieved June 2020, from <https://doi.org/10.1093/jme/tjw215>

Johnston, L.M., Wang, X., Erni, S., Taylor, S.W., McFayden, C.B., Oliver, J.A., Stockdale, C., Christianson, A., Boulanger, Y., Gauthier, S., Arseneault, D., Wotton, B.M., Parisien, M.-A. and Flannigan, M.D. (2020). Wildland fire risk research in Canada. Environmental Reviews, 28(2), 164–186. Retrieved June 2020, from <https://doi.org/10.1139/er-2019-0046>

Johnston, M., Dawson, J., De Souza, E. and Stewart, E.J. (2017). Management challenges for the fastest growing marine shipping sector in Arctic Canada: Pleasure crafts. Polar Record, 53(1), 67–78. Retrieved June 2020, from <https://doi.org/10.1017/S0032247416000565>

Jolly, W.M., Cochrane, M.A., Freeborn, P.H., Holden, Z.A., Brown, T.J., Williamson, G.J. and Bowman, D.M.J.S. (2015). Climate-induced variations in global wildfire danger from 1979 to 2013. Nature Communications, 6, 7537. Retrieved June 2020, from <https://doi.org/10.1038/ncomms8537>

Jones, R.N. and Preston, B.L. (2011). Adaptation and risk management. Wiley Interdisciplinary Reviews: Climate Change, 2(2), 296–308. Retrieved June 2020, from <https://doi.org/10.1002/wcc.97>

Keppel, E.A., Scrosati, R.A. and Courtenay, S.C. (2012). Ocean acidification decreases growth and development in American lobster (Homarus americanus) larvae. Journal of Northwest Atlantic Fishery Science, 44, 61–66. Retrieved June 2020, from <https://doi.org/10.2960/J.v44.m683>

Kirchmeier-Young, M.C., Zwiers, F.W., Gillett, N.P. and Cannon, A.J. (2017). Attributing extreme fire risk in Western Canada to human emissions. Climatic Change, 144, 365–379. Retrieved June 2020, from <https://doi.org/10.1007/s10584-017-2030-0>

Kopytko, N. and Perkins, J. (2011). Climate change, nuclear power, and the adaptation–mitigation dilemma. Energy Policy, 39(1), 318–333. Retrieved June 2020, from <https://doi.org/10.1016/j.enpol.2010.09.046>

Koutsoyiannis, D. (1999). A probabilistic view of Hershfield’s method for estimating probable maximum precipitation. Water Resources Research, 35(4), 1313–1322. Retrieved June 2020, from <https://doi.org/10.1029/1999WR900002>

Kovacs, P. and Thistlethwaite, J. (2014). Industry, Chapter 5 in Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation, (Eds.) F.J. Warren and D. S. Lemmen. Government of Canada, Ottawa, Ontario, 135–158. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2014/pdf/Chapter5-Industry_Eng.pdf>

Kroeker, K.J., Kordas, R., Crim, R. and Singh, G. (2010). Meta-analysis reveals negative yet variable effects of ocean acidification on marine organisms. Ecology Letters, 13(11), 1419–1434. Retrieved June 2020, from <https://doi.org/10.1111/j.1461-0248.2010.01518.x>

Kumar, K., Holtz, M. D., Xi, K. and Turkington, T.K. (2013). Overwintering potential of the stripe rust pathogen (Puccinia striiformis) in central Alberta. Canadian Journal of Plant Pathology, 35(3), 304–314. Retrieved June 2020, from <https://doi.org/10.1080/07060661.2013.809385>

Kwiatkowski, K. (2017). Modeling Climate Change Adaptation in Transportation Infrastructure Organizations. Civil Engineering Graduate Theses and Dissertations, University of Colorado, 184 p. Retrieved June 2020, from <https://pdfs.semanticscholar.org/67fa/017e57f7e6009dccbdfabca788a6f2e13c90.pdf>

Le Bris, A., Mills, K.E., Wahle, R.A., Chen, Y., Alexander, M.A., Allyn, A.J., Schuetz, J.G., Scott, J.D. and Pershing, A.J. (2018). Climate vulnerability and resilience in the most valuable North American fishery. Proceedings of the National Acadamies of Sciences USA, 115, 1831–1836. Retrieved June 2020, from <https://doi.org/10.1073/pnas.1711122115>

Leduc, A., Bernier, P.Y., Mansuy, N., Raulier, F., Gauthier, S. and Bergeron, Y. (2015). Using salvage logging and tolerance to risk to reduce the impact of forest fires on timber supply calculations. Canadian Journal of Forest Research, 45(4), 480–486. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2014-0434>

Lemelin, R. H., Dawson, J., Stewart, E.J., Maher, P. and Luck, M. (2010). Last-chance tourism: The boom, doom, and gloom of visiting vanishing destinations. Current Issues in Tourism, 13(5), 477–493. Retrieved June 2020, from <https://doi.org/10.1080/13683500903406367>

Lemieux C. J., Groulx, M., Halpenny, E., Stager, H., Dawson, J., Stewart, E.J. and Hvenegaard, G.T. (2017). The End of the Ice Age?: Disappearing World Heritage and the Climate Change Communication Imperative. Environmental Communication, 12(5), 653-671. Retrieved June 2020, from <https://doi.org/10.1080/17524032.2017.1400454>

Lemmen, D.S., Johnston, M., Ste-Marie, C. and Pearce, T. (2014). Natural resources, Chapter 3 in Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation, (Eds.) F.J. Warren and D.S. Lemmen. Government of Canada, Ottawa, Ontario, 65–98. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2014/pdf/Chapter3-Natural-Resources_Eng.pdf>

Lemmen, D.S., Warren, F.J., James, T.S. and Mercer Clarke, C.S.L. (Eds.) (2016). Canada’s Marine Coasts in a Changing Climate. Government of Canada, Ottawa, Ontario, 274 p. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/files/pdf/NRCAN_fullBook%20%20accessible.pdf>

Lemmen, D.S., Warren, F.J., Lacroix, J. and Bush E. (Eds.) (2008). From Impacts to Adaption: Canada in a Changing Climate 2007. Government of Canada, Ottawa, Ontario, 448 p. Retrieved June 2020, from <https://www.nrcan.gc.ca/impacts-adaptation-canada-changing-climate/10253>

Linnenluecke, M.K., Stathakis, A. and Griffiths, A. (2011). Firm relocation as adaptive response to climate change and weather extremes. Global Environmental Change, 21(1), 123–33. Retrieved June 2020, from <https://doi.org/10.1016/j.gloenvcha.2010.09.010>

Linnenluecke, M. and Smith, T. (2018). Adaptation of MSMEs to Climate Change: A Review of the Existing Literature, in Private-Sector Action in Adaptation: Perspectives on the Role of Micro, Small and Medium Size Enterprises, (Eds.) C. Schaer and N. Kuruppu. UNEP DTU Partnership, Copenhagen, 19‒27. Retrieved June 2020, from <https://backend.orbit.dtu.dk/ws/files/162053774/MSME_Adaptation_updated_WEB.pdf>

Linnenlueke, M., Griffiths, A. and Winn, M.I. (2013). Firm and Industry Adaptation to Climate Change: A Review of Climate Adaptation Studies in the Business and Management Field. Wiley Interdisciplinary Reviews: Climate Change, 4, 397–416. Retrieved June 2020, from <https://doi.org/10.1002/wcc.214>

Liu, J.C., Mickley, L.J., Sulprizio, M.P., Dominici, F., Yue, X., Ebisu, K., Anderson, G.B., Khan, R.F.A, Bravo, M.A. and Bell, M.L. (2016). Particulate air pollution from wildfires in the Western US under climate change. Climatic Change, 138(3–4), 655–666. Retrieved June 2020, from <https://doi.org/10.1007/s10584-016-1762-6>

Lynch, D.L. (2004). What Do Forest Fires Really Cost? Journal of Forestry, 102(6), 42–49. Retrieved June 2020, from <https://doi.org/10.1093/jof/102.6.42>

Lysyk, T.J. and Dergousoff, S.J. (2014). Distribution of Culicoides sonorensis (Diptera: Ceratopogonidae) in Alberta, Canada. Journal of Medical Entomology, 51, 560–571. Retrieved June 2020, from <https://doi.org/10.1603/ME13239>

Mailhot, A., Panthou, G. and Talbot, G. (2014) Recommandations sur les majorations à considérer dans les courbes Intensité-durée-Fréquence (IDF) aux horizons 2040-2070 et 2070-2100 pour l’ensemble du Québec, Phase II. Ministère du Transport du Québec, 36 p. Retrieved June 2020, from <espace.inrs.ca/id/eprint/2421/1/R001515.pdf> 

Man, T. (2013). Infrastructure Interdependencies Timelines. The Royal Academy of Engineering, United Kingdom, 17 p. Retrieved June 2020, from <https://www.raeng.org.uk/publications/responses/etf-infrastructure-interdependencies-report>

Mantle314 (2019). Climate Adaptation in Canada’s Oil and Gas Sector: Planning to Thrive in a Changing World, Report submitted to the Climate Change Impacts and Adaptation Division, Natural Resources Canada.

Marchal, J., Cumming, S.G. and McIntire, E.J. (2017). Land cover, more than monthly fire weather, drives fire-size distribution in Southern Quebec forests: Implications for fire risk management. PLoS ONE, 12(9), e0185515. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0179294>

Marras, S, Cucco, A., Antognarelli, F., Azzurro, E., Milazzo, M., Bariche, M., Butenschön, M., Kay, S., Di Bitetto, M. and Quattrocchi, G. (2015). Predicting future thermal habitat suitability of competing native and invasive fish species: from metabolic scope to oceanographic modelling. Conservation Physiology, 3, 1–14. Retrieved June 2020, from <https://doi.org/10.1093/conphys/cou059>

Marshall, G.N., Schell, T.L., Elliott, M.N., Rayburn, N.R. and Jaycox, L.H. (2007). Psychiatric Disorders Among Adults Seeking Emergency Disaster Assistance After a Wildland-Urban Interface Fire. Psychiatric Services, 58(4), 509–514. Retrieved June 2020, from <https://doi.org/10.1176/ps.2007.58.4.509>

Mazzacurati, E. (2018). Advancing TCFD Guidance on Physical Climate Risks and Opportunities. Four Twenty Seven, Paris, 56 p. Retrieved June 2020, from <http://427mt.com/2018/05/31/report-advancing-tcfd-guidance-physical-climate-risk-opportunities/>

Mbow, C., Rosenzweig, C., Barioni, L.G., Benton, T.G., Herrero, M., Krishnapillai, M., Liwenga, E., Pradhan, P., Rivera-Ferre, M.G., Sapkota, T., Tubiello, F.N., and Xu, Y. (2019). Food Security, in Climate Change and Land (An IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems), (Eds.) P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pörtner, D.C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal Pereira, P. Vyas, E. Huntley, K. Kissick, M. Belkacemi, J. Malley. Retrieved June 2020, from <https://www.ipcc.ch/site/assets/uploads/sites/4/2019/11/08_Chapter-5.pdf>

McBean, G. (2012). Telling the Weather Story. Institute for Catastrophic Loss Reduction and Insurance Bureau of Canada. Retrieved June 2020, from <assets.ibc.ca/Documents/Studies/McBean_Report.pdf>

McBoyle, G., Scott, D. and Jones, B. (2007). Climate change and the future of snowmobiling in non-mountainous regions of Canada. Managing Leisure, 12(4), 237–250. Retrieved June 2020, from <https://doi.org/10.1080/13606710701546868>

McCaffrey, S., Rhodes, A. and Stidham, M. (2015). Wildfire evacuation and its alternatives: perspectives from four United States’ communities. International Journal of Wildland Fire, 24(2), 170–178. Retrieved June 2020, from <https://doi.org/10.1071/WF13050>

McGee, T., McFarlane, B. and Tymstra, C. (2015). Wildfire: A Canadian perspective, in Wildfire Hazards, Risks and Disasters, (Ed.) D. Paton. Elsevier, 35–68.

McKay, J. (2018). Small Businesses Are a Vital Part of Community Resiliency but Often Overlook Vulnerabilities. Government Technology. Retrieved July 2018, from <https://www.govtech.com/em/preparedness/Small-Businesses-Are-a-Vital-Part-of-Community-Resiliency-but-Often-Overlook-Vulnerabilities.html>

McKenney, D.W., Yemshanov, D., Pedlar, J., Allen, D., Lawrence, K., Hope, E., Lu, B. and Eddy, B. (2016). Canada’s Timber Supply: Current Status and Future Prospects Under a Changing Climate. Great Lakes Forestry Centre, Canadian Forest Service, Natural Resources Canada.

McLean, E.L., Katenka, N.V. and Seibel, B.A. (2018). Decreased growth and increased shell disease in early benthic phase Homarus americanus in response to elevated CO2. Marine Ecology Progress Series, 596, 113–126. Retrieved June 2020, from <https://doi.org/10.3354/meps12586>

Mine Environment Neutral Drainage (2012). Cold Regions Cover System Design Technical Guidance Document. Report 1.61.5c, Canadian Centre for Mineral and Energy Technology. Retrieved June 2020, from <http://mend-nedem.org/wp-content/uploads/2013/01/1.61.5c.pdf>

Ministère du Développement durable, de l’Environnement et des Parcs (2012). Directive 019 sur l’industrie minière. 105 p. Retrieved June 2020, from <http://www.environnement.gouv.qc.ca/milieu_ind/directive019/directive019.pdf>

MNP (2017). A Review of the 2016 Horse River Wildfire: Alberta Agriculture and Forestry Preparedness and Response, Report prepared for Forestry Division, Alberta Agriculture and Forestry, Edmonton, Alberta. Retrieved June 2020, from <https://www.alberta.ca/assets/documents/Wildfire-MNP-Report.pdf>

Mori, A.S. and Johnson, E.A. (2013). Assessing possible shifts in wildfire regimes under a changing climate in mountainous landscapes. Forest Ecology and Management, 310, 875–886. Retrieved June 2020, form <https://doi.org/10.1016/j.foreco.2013.09.036>

Morton, D.C., Roessing, M.E., Camp, A.E. and Tyrrell, M.L. (2003). Assessing the environmental, social, and economic impacts of wildfire. Global Institute of Sustainable Forestry Research Paper 001, Yale University, 59 p. Retrieved June 2020, from <https://gisf.yale.edu/sites/default/files/files/wildfire_report(1).pdf>

Mueter, F.J. and Litzow, M.A. (2008). Sea ice retreat alters the biogeography of the Bering Sea continental shelf. Ecological Applications, 18, 309–320. Retrieved June 2020, from <https://doi.org/10.1890/07-0564.1>

Nastev, M. and Aubertin, M. (2000). Hydrogeological modelling for the reclamation work at the Lorraine mine site Québec. 53rd Canadian Geotechnical Conference and 1st joint IAHCNC and CGS Groundwater Specialty Conference, Montréal, QC, Canada, 311–318.

National Energy Board (2016). Market Snapshot: Impacts of the Fort McMurray wildfires on Canadian crude oil production. Retrieved June 2020, from <https://www.neb-one.gc.ca/nrg/ntgrtd/mrkt/snpsht/2016/05-02frtmcmrr-eng.html>

National Energy Board (2017). Canada’s Adoption of Renewable Power Sources – Energy Market Analysis. Retrieved June 2020, from <https://www.neb-one.gc.ca/nrg/sttstc/lctrct/rprt/2017cnddptnrnwblpwr/index-eng.html>

National Energy Board (2018). Canada’s Energy Future 2018. Retrieved June 2020, from <https://www.neb-one.gc.ca/nrg/ntgrtd/ftr/2018/index-eng.html>

Natural Resources Canada (2017). The State of Canada’s Forests. Annual Report 2017. Retrieved June 2020, from <https://cfs.nrcan.gc.ca/publications?id=38871>

Natural Resources Canada (2018). The Energy Factbook, 2018. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/energy/pdf/energy-factbook-oct2-2018%20(1).pdf>

Natural Resources Canada (2019). Cost of fire protection. Retrieved June 2020, from <https://www.nrcan.gc.ca/forests/climate-change/forest-change/17783>

Nelson, H.W., Williamson, T.B., Macaulay, C. and Mahony, C. (2016). Assessing the potential for forest management practitioner participation in climate change adaptation. Forest Ecology and Management, 360, 388–399. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2015.09.038>

Nova Scotia Federation of Agriculture (2018). Risk Proofing Nova Scotia’s Agriculture: A Risk Assessment System Pilot (AgriRisk). Final Report, Bible Hill, Nova Scotia, 3-14. Retrieved June 2020, from <https://nsfa-fane.ca/wp-content/uploads/2018/08/Final-Report-for-AgriRisk-project-English.pdf>

NRTEE [National Roundtable on the Environment and the Economy] (2011). Paying the price: the economic impacts of climate change for Canada (Vol. 4). National Round Table on the Environment and the Economy, Ottawa, Canada.

Ochuodho, T.O. and Lantz, V.A. (2014). Economic impacts of climate change in the forest sector: a comparison of single-region and multiregional CGE modeling frameworks. Canadian Journal of Forest Research, 44(5), 449-464. Retrieved June 2020, from <https://doi.org/10.1139/cjfr-2013-0317>

OECD [Organization for Economic Co-operation and Development] (2015). The Economic Consequences of Climate Change. OECD Publishing, Paris. Retrieved June 2020, from <http://dx.doi.org/10.1787/9789264235410-en>

Ogden, N.H., Radojevic, M., Wu, X., Duvvuri, V.R., Leighton, P.A. and Wu, J. (2014). Estimated Effects of Projected Climate Change on the Basic Reproductive Number of the Lyme Disease Vector Ixodes scapularis. Environmental Health Perspectives, 122, 631–638. Retrieved June 2020, from <https://doi.org/10.1289/ehp.1307799>

Ohlson, D.W., Berry, T.M., Gray, R.W., Blackwell, B.A. and Hawkes, B.C. (2006). Multi-attribute evaluation of landscape-level fuel management to reduce wildfire risk. Forest Policy and Economics, 8(8), 824–837. Retrieved June 2020, from <https://doi.org/10.1016/j.forpol.2005.01.001>

Olfert, O., Weiss, R.M. and Elliott, R.H. (2016). Bioclimatic approach to assessing the potential impact of climate change on wheat midge (Diptera: Cecidomyiidae) in North America. The Canadian Entomologist, 148, 52–67. Retrieved June 2020, from <https://doi.org/10.4039/tce.2015.40>

Olfert, O., Weiss, R.M., Elliott, R.H. and Soroka, J.J. (2017). Bioclimatic approach to assessing the potential impact of climate change on two flea beetle (Coleoptera: Chrysomelidae) species in Canada. The Canadian Entomologist, 149, 616-627. Retrieved June 2020, from <https://doi.org/10.4039/tce.2017.39>

Ontario Woodlot Association. (2015). Managing Your Forest in a Changing Climate. Retrieved June 2020, from <https://www.ontariowoodlot.com/publications-and-links/owa-publications/landowner-guides/managing-your-forest-in-a-changing-climate-new>

Ortiz, L., González, J.E. and Lin, W. (2018). Climate change impacts on peak building cooling energy demand in a coastal megacity. Environmental Research Letters, 13(9). Retrieved June 2020, from <https://doi.org/10.1088/1748-9326/aad8d0 >

Orwig, K. (2016). Strategies for funding wildfire mitigation. Risk Management, 14+ p. Retrieved June 2020, from <https://go.gale.com/ps/anonymous?id=GALE%7CA454786513&sid=googleScholar&v=2.1&it=r&linkaccess=abs&issn=00355593&p=AONE&sw=w>

Ouranos (2015). Summary of the Synthesis on Climate Change Knowledge in Quebec, 2015 Edition. 13 p.

Palko, K. (2017). Synthesis, in Climate risks and adaptation practices for the Canadian transportation sector 2016, (Eds.) K. Palko and D.S. Lemmen. Government of Canada, Ottawa, 12–25. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2016/Chapter-2e.pdf>

Palko, K. and Lemmen, D.S. (Eds.) (2017). Climate risks and adaptation practices for the Canadian transportation sector 2016. Government of Canada, Ottawa, 309 p. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2016/ClimatRisk-E-ACCESSIBLE.pdf>

Parker, L.M., Ross, P.M., O’Connor, W.A., Pörtner, H.-O., Scanes, E. and Wright, J.M. (2013). Predicting the response of molluscs to the impact of ocean acidification. Biology, 2, 651–692. Retrieved June 2020, from <https://doi.org/10.3390/biology2020651>

Pearce, T., Ford, J.D., Prno, J. and Duerden, F. (2009). Climate change and Canadian mining: opportunities for adaptation. David Suzuki Foundation, 50 p. Retrieved June 2020, from <https://davidsuzuki.org/science-learning-centre-article/climate-change-canadian-mining-opportunities-adaptation/>

Pearce, T.D., Ford, J.D., Prno, J., Duerden, F., Pittman, J., Beaumier, M. and Smit, B. (2011). Climate change and mining in Canada. Mitigation and Adaptation Strategies for Global Change, 16(3), 347 –368. Retrieved June 2020, from <https://doi.org/10.1007/s11027-010-9269-3>

Perry, E., Manning, R., Xiao, X., Valliere, W. and Reigner, N. (2018). Social climate change: The advancing extirpation of snowmobilers in Vermont. Journal of Park and Recreation Administration, 36, 31–51. Retrieved June 2020, from <https://doi.org/10.18666/JPRA-2018-V36-I2-8307>

Peter, B., Wang, S., Mogus, T. and Wilson, B. (2006). Fire risk and population trends in Canada’s wildland-urban interface, in Canadian Wildland Fire Strategy: Background Synthesis, Analysis, and Perspectives, (Eds.) K.G. Hirsch and P. Fuglem, 37–48. Retrieved June 2020, from <https://www.ccfm.org/pdf/cwfs_Analysis_EN_web.pdf>

Phillips, A. and Towns, W. (2017). The Prairies, in Climate risks and adaptation practices for the Canadian transportation sector 2016, (Eds.) K. Palko and D.S. Lemmen. Government of Canada, Ottawa, 105–137. Retrieved June 2020, from <https://www.nrcan.gc.ca/sites/www.nrcan.gc.ca/files/earthsciences/pdf/assess/2016/Chapter-5e.pdf>

PIEVC [Public Infrastructure Engineering Vulnerability Committee] (n.d.). The Protocol. Engineers Canada. Retrieved June 2020, from <https://pievc.ca/protocol>

Porter, J.R., Xie, L., Challinor, A.J., Cochrane, K., Howden, S.M., Iqbal, M.M., Lobell, D.B. and Travasso, M.I. (2014). Food security and food production systems, in Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects (Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change), (Eds.) C.B. Field, V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea and L.L. White. Cambridge University Press, Cambridge, United Kingdom and New York, New York, USA, 485–533. Retrieved June 2020, from <https://www.ipcc.ch/site/assets/uploads/2018/02/WGIIAR5-Chap7_FINAL.pdf>

Price, D.T., Alfaro, R.I., Brown, K.J., Flannigan, M.D., Fleming, R.A., Hogg, E.H., Girardin, M.P., Lakusta, T., Johnston, M., McKenney, D.W., Pedlar, J.H., Stratton, T., Sturrock, R.N., Thompson, I.D., Trofymow, J.A. and Venier, L.A. (2013). Anticipating the consequences of climate change for Canada’s boreal forest ecosystems. Environmental Reviews, 21(4), 322–365. Retrieved June 2020, from <https://doi.org/10.1139/er-2013-0042>

Pryor, S.C. and Barthelmie, R.J. (2010). Climate change impacts on wind energy: A review. Renewable and Sustainable Energy Reviews, 14(1), 430–437. Retrieved June 2020, from <https://www.sciencedirect.com/science/article/pii/S1364032109001713>

Qian, B., Jing, Q., Bélanger, G., Shang, J., Huffman, T., Liu, J. and Hoogenboom, G. (2018). Simulated canola yield responses to climate change and adaptation in Canada. Agronomy Journal, 110(1), 133–146. Retrieved June 2020, from <https://doi.org/10.2134/agronj2017.02.0076>

Qian, B., Zhang, X., Smith, W., Grant, B., Jing, Q., Cannon, A.J., Neilsen, D., McConkey, B., Li, G., Bonsal, B., Wan, H., Xue, L. and Zhao, J. (2019). Climate change impacts on Canadian yields of spring wheat, canola and maize for global warming levels of 1.5°C, 2.0°C, 2.5°C and 3.0°C. Environmental Research Letters, 14. Retrieved June 2020, from <https://doi.org/10.1088/1748-9326/ab17fb>

Quorum Corporation (2014). Grain Supply Chain Study. Grain Monitoring Program Supplemental Study, 107 p. Retrieved June 2020, from <http://grainmonitor.ca/Downloads/SupplementalReports/Grain%20Supply%20Chain%20Study%20-%20Report.pdf>

Quorum Corporation (2015). Annual Report: 2013–2014 Crop Year. Retrieved June 2020, from <http://grainmonitor.ca/Downloads/AnnualReports/AnnualReport201314.pdf>

Raulier, F., Dhital, N., Racine, P., Tittler, R. and Fall, A. (2014). Increasing resilience of timber supply: How a variable buffer stock of timber can efficiently reduce exposure to shortfalls caused by wildfires. Forest Policy and Economics, 46, 47–55. Retrieved June 2020, from <https://doi.org/10.1016/j.forpol.2014.06.007>

Raulier, F., Le Goff, H., Gauthier, S., Rapanoela, R. and Bergeron, Y. (2013). Introducing two indicators for fire risk consideration in the management of boreal forests. Ecological Indicators, 24, 451–461. Retrieved June 2020, from <https://doi.org/10.1016/j.ecolind.2012.07.023>

Reid, C. E., Brauer, M., Johnston, F.H., Jerrett, M., Balmes, J.R. and Elliott, C.T. (2016). Critical review of health impacts of wildfire smoke exposure. Environmental Health Perspectives, 124(9), 1334–1343. Retrieved June 2020, from <https://doi.org/10.1289/ehp.1409277>

Reinecke, S. and Brodie, J. (2012). Climate Change Impacts on Metal Leaching and Acid Rock Drainage at Canadian Mines ‒ Risks for Operation and Closure, in Proceedings of the 9th International Conference on Acid Rock Drainage (ICARD), Ottawa, 1–12.

Resort Municipality of Whistler (2016). Community Energy and Climate Action Plan. Retrieved June 2020, from <https://www.whistler.ca/sites/default/files/related/cecap.0_final.pdf>

Revell, A., Stokes, D. and Chen, H. (2009). Small businesses and the environment: turning over a new leaf? Business Strategy and the Environment, 19(5), 273–288. Retrieved June 2020, from <https://doi.org/10.1002/bse.628>

Rijal, B., Raulier, F., Martell, D. L. and Gauthier, S. (2018). The economic impact of fire management on timber production in the boreal forest region of Quebec, Canada. International Journal of Wildland Fire, 27(12), 831–844. Retrieved June 2020, from <https://doi.org/10.1071/WF18041>

Roson, R. and Sartori, M. (2016). Estimation of Climate Change Damage Functions for 140 Regions in the GTAP 9 Database. Journal of Global Economic Analysis, 1(2), 78-115. Retrieved June 2020, from <http://dx.doi.org/10.21642/JGEA.010202AF>

Ross, T. (2016). La Niña, the Blob and another warmest year, in State of the Physical, Biological and Selected Fishery Resources of Pacific Canadian Marine Ecosystems in 2015, (Eds.) P.C. Chandler, S.A. King and R.I. Perry. Canadian Technical Reports of Fisheries and Aquatic Sciences, 3179, viii + 230 p. Retrieved June 2020, from <https://waves-vagues.dfo-mpo.gc.ca/Library/40617944.pdf>

Rousseau, A.N., Klein, I.M., Freudiger, D., Gagnon, P., Frigon, A. and Ratté-Fortin, C. (2014). Development of a methodology to evaluate probable maximum precipitation (PMP) under changing climate conditions: Application to southern Quebec, Canada. Journal of Hydrology, 519, 3094–3109. Retrieved June 2020, from <https://doi.org/10.1016/j.jhydrol.2014.10.053>

Rutty, M., Scott, D., Steiger, R., Johnson, P. and Pons, M. (2015). The geography of skier adaptation to adverse conditions in the Ontario ski market. The Canadian Geographer, 59(4). Retrieved June 2020, from <https://doi.org/10.1111/cag.12220>

Rutty, M., Scott, D., Johnson, P., Steiger, R., Pons, M. and Vilella, M. (2017). Using ski industry response to climatic variability to assess climate change risk: An analogue study in Eastern Canada. Tourism Management, 58, 196–205. Retrieved June 2020, from <https://doi.org/10.1016/j.tourman.2016.10.020>

Samy, G.M., Samy, C.P. and Ammasaiappan, M. (2015). Integrated management systems for better environmental performance and sustainable development – a review. Environmental Engineering and Management Journal, 14(5), 985–1000. Retrieved June 2020, from <https://doi.org/10.13140/rg.2.2.35566.51528>

Sankey, S. (2018). Blueprint for wildland fire science in Canada (2019–2029). Northern Forestry Centre, Canadian Forest Service, Natural Resources Canada, Edmonton, 1–27. Retrieved June 2020, from <http://cfs.nrcan.gc.ca/pubwarehouse/pdfs/39429.pdf>

Sauchyn, D. and Kulshreshtha, S. (2008). Prairies, in From Impacts to Adaptation: Canada in a Changing Climate 2007, (Eds.) D.S. Lemmen, F.J. Warren, J. Lacroix and E. Bush. Government of Canada, Ottawa, 275‒328. Retrieved June 2020, from <https://www.nrcan.gc.ca/impacts-adaptation-canada-changing-climate/10253>

Schaer, C. and Kuruppu, N. (Eds.) (2018). Private-sector action in adaptation: Perspectives on the role of micro, small and medium size enterprises. UNEP DTU Partnership, Copenhagen, 203 p. Retrieved June 2020, from <https://backend.orbit.dtu.dk/ws/files/162053774/MSME_Adaptation_updated_WEB.pdf>

Scharbach, J. and Waldram, J.B. (2016). Asking for a disaster: being “at risk” in the emergency evacuation of a northern Canadian Aboriginal community. Human Organization, 75(1), 59–70. Retrieved June 2020, from <https://doi.org/10.17730/0018-7259-75.1.59>

Schauberger, B., Archontoulis, S., Arneth, A., Balkovic, J., Ciais, P., Deryng, D., Elliott, J., Folberth, C., Khabarov, N., Müller, C., Pugh, T.A.M., Rolinski, S., Schaphoff, S., Schmid, E., Wang, X., Schlenker, W. and Frieler, K. (2017). Consistent negative response of US crops to high temperatures in observations and crop models. Nature Communications, 8(13931). Retrieved June 2020, from <https://doi.org/10.1038/ncomms13931>

Schroeder, D. (2010). Fire behaviour in thinned jack pine: two case studies of FireSmart treatments in Canada’s Northwest Territories. FPInnovations, Eastern Region, Pointe-Claire, Quebec and Western Region, Vancouver, British Columbia. Advantage Report, 12(7), 12.

Scott, D. and Steiger, R. (2013). Vulnerability of the Ski Industry, in Climate Vulnerability: Understanding and Addressing Threats to Essential Resources, (Ed.) R.A. Pielke. Elsevier Inc., Academic Press, 305–313.

Scott, D., Hall, C.M. and Gössling, S. (2016). A report on the Paris Climate Change Agreement and its implications for tourism: why we will always have Paris. Journal of Sustainable Tourism, 2(7), 933–948. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2016.1187623>

Scott, D., Jones, B. and Konopek, J. (2007). Implications of climate and environmental change for nature-based tourism in the Canadian Rocky Mountains: A case study of Waterton Lakes National Park. Tourism Management, 28(2), 570–579. Retrieved June 2020, from <https://doi.org/10.1016/j.tourman.2006.04.020>

Scott, D., Gössling, S. and Hall, C.M. (2012). International Tourism and Climate Change. Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213–232. Retrieved June 2020, from <https://doi.org/10.1002/wcc.165>

Scott, D., Steiger, R., Rutty, M., Pons, M., Johnson, P. and Vilella, M. (2017). The differential futures of ski tourism in Ontario (Canada) under climate change: the limits of snowmaking adaptation. Current Issues in Tourism, 22(11), 1327‒1342. Retrieved June 2020, from <https://doi.org/10.1080/13683500.2017.1401984>

Scott, D., Hall, C.M. and Gössling, S. (2019). Global Tourism Vulnerability to Climate Change. Annals of Tourism Research, 77, 49–61. Retrieved June 2020, from <https://doi.org/10.1016/j.annals.2019.05.007>

Scott, D., Steiger, R. Knowles, N. and Fang, Y. (2020). Regional ski tourism risk to climate change: An inter-comparison of Eastern Canada and US Northeast markets. Journal of Sustainable Tourism, 28(4), 568–586. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2019.1684932>

Shackell, N.L., Ricard, D. and Stortini, C. (2014). Thermal Habitat Index of Many Northwest Atlantic Temperate Species Stays Neutral under Warming Projected for 2030 but Changes Radically by 2060. PLoS ONE, 9(3), e90662. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0090662>

Shughrue, C. and Seto, K.C. (2018). Systematic vulnerabilities of the global urban-industrial network to hazards. Climatic Change, 151, 173-187. Retrieved June 2020, from <https://doi.org/10.1007/s10584-018-2293-0>

Slawinski, N., Pinkse, J., Busch, T. and Banerjee, S.B. (2017). The role of short-termism and uncertainty avoidance in organizational inaction on climate change: A multi-level framework. Business and Society, 56(2), 253–82. Retrieved June 2020, from <https://doi.org/10.1177/0007650315576136>

Spittlehouse, D.L. (2005). Integrating climate change adaptation into forest management. The Forestry Chronicle, 81(5), 691–695. Retrieved June 2020, from <https://doi.org/10.5558/tfc81691-5>

Stantec Consulting Limited (2012). Preliminary assessment report, Canadian Beaufort Sea Region, Report prepared for Aboriginal Affairs and Northern Development Canada, Ottawa, Ontario, 1–171.

Statistics Canada (2017). Infographic: Fort McMurray 2016 wildfire – economic impact. Retrieved June 2020, from <https://www.statcan.gc.ca/pub/11-627-m/11-627-m2017007-eng.htm>

Stewart, E.J., Dawson, J. and Draper, D. (2011). Cruise tourism and residents in Arctic Canada: Development of a resident attitude typology. Journal of Hospitality and Tourism Management, 18(1), 95–106. Retrieved June 2020, from <https://doi.org/10.1375/jhtm.18.1.95>

Stortini, C.H, Shackell, N.L., Tyedmers, P. and Beazley, K. (2015). Assessing marine species vulnerability to projected warming on the Scotian Shelf, Canada. ICES Journal of Marine Science, 72, 1731–1743. Retrieved June 2020, from <https://doi.org/10.1093/icesjms/fst022>

Stortini, C.H., Chabot, D. and Shackell, N.L. (2017). Marine species in ambient low-oxygen regions subject to double jeopardy impacts of climate change. Global Change Biology, 23, 2284–2296. Retrieved June 2020, from <https://doi.org/10.1111/gcb.13534>

Stratos (2009). Climate change impacts on mining operations and infrastructure, Report prepared for the Centre for Excellence in Mining Innovation, 2009. Retrieved June 2020, from <http://www.miningexcellence.ca>

Stratos (2017). Mining State of Play: Climate Change Adaptation in the Canadian Mining Sector, Report submitted to Canada’s Climate Change Adaptation Platform Mining Working Group, 57 p.

Stratz, S.A. and Hossain, F. (2014). Probable maximum precipitation in a changing climate: Implications for dam design. Journal of Hydrologic Engineering, 19(12). Retrieved June 2020, from <https://doi.org/10.1061/(ASCE)HE.1943-5584.0001021>

Sturrock, R.N., Frankel, S.J., Brown, A.V., Hennon, P.E., Kliejunas, J.T., Lewis, K.J., Worrall, J.J. and Woods, A.J. (2011). Climate change and forest diseases. Plant Pathology, 60(1), 133–149. Retrieved June 2020, from <https://doi.org/10.1111/j.1365-3059.2010.02406.x>

Subedi, N., Taylor, S. and Peter, B. (2016). Increasing threats of mega-fires on the WUI communities of Canada. Retrieved June 2020, from <http://www.wildlandfire2016.ca/wp-content/uploads/2017/05/Mega-fire-presentation-Nirmal-Subedi.pdf>

Sudhalkar, A., Chan, C., Bonham-Carter, C. and Smith, M. (2017). C40 Infrastructure Interdependencies and Climate Risks Report, 1–26.

Sulla-Menashe, D., Woodcock, C.E. and Friedl, M.A. (2018). Canadian boreal forest greening and browning trends: An analysis of biogeographic patterns and the relative roles of disturbance versus climate drivers. Environmental Research Letters, 13(1). Retrieved June 2020, from <https://doi.org/10.1088/1748-9326/aa9b88>

Surminski, S. and Hankinson, J. (2018). MSMEs, climate change risks and insurance: reflections on the use of insurance for climate adaptation, in Private-sector action in adaptation: Perspectives on the role of micro, small and medium size enterprises, (Eds.) C. Schaer and N. Kuruppu. UNEP DTU Partnership, Copenhagen, 65–78. Retrieved June 2020, from <https://backend.orbit.dtu.dk/ws/files/162053774/MSME_Adaptation_updated_WEB.pdf>

Sveinsson, O.G.B. (2015). Case study: Influence of climate change on hydropower in Iceland, in Resilience of hydropower and dams to climate change and natural disasters. International Hydropower Association, World Bank Group. Retrieved June 2020, from <https://www.hydropower.org/sites/default/files/OliSveinssonLandsvirkjun.pdf>

Swiss Re Institute (2018). Natural catastrophes and man-made disasters in 2017: a year of record-breaking losses. SIGMA No 1, Zurich. Retrieved June 2020, from <https://reliefweb.int/sites/reliefweb.int/files/resources/sigma1_2018_en.pdf>

Taylor, A.R., Boulanger, Y., Price, D.T., Cyr, D., McGarrigle, E., Rammer, W. and Kershaw, J.A. (2017). Rapid 21st century climate change projected to shift composition and growth of Canada’s Acadian Forest Region. Forest Ecology and Management, 405, 284–294. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2017.07.033>

Tercek, M. and Rodman, A. (2016). Forecasts of 21st century snowpack and implications for snowmobile and snowcoach use in Yellowstone National Park. PLoS ONE, 11(7). Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0159218>

Tett, S.F.B., Falk, A., Rogers, M., Spuler, F., Turner, C., Wainwright, J., Dimdore-Miles, O., Knight, S., Freychet, N., Mineter, M.J. and Lehmann, C.E.R. (2017). Anthropogenic forcings and associated changes in fire risk in western North America and Australia during 2015/16, in Explaining Extreme Events of 2016 from a Climate Perspective. Bulletin of the American Meteorological Society, 98, S60–S64. Retrieved June 2020, from <https://doi.org/10.1175/BAMS-D-17-0096.1>

TCFD [Task Force on Climate-Related Financial Disclosures] (2017). Final Report: Recommendations of the Task Force on Climate-Related Financial Disclosures. Retrieved June 2020, from <https://www.fsb-tcfd.org/publications/final-recommendations-report/>

Thistlethwaite, J. (2016). The Emergence of Flood Insurance in Canada: Navigating Institutional Uncertainty. Risk Analysis, 37(4), 744–755. Retrieved June 2020, from <https://doi.org/10.1111/risa.12659>

Thistlethwaite, J. and Wood, M. (2018). Insurance and Climate Change Risk Management: Rescaling to Look Beyond the Horizon. British Journal of Management, 29(2), 279–298. Retrieved June 2020, from <https://doi.org/10.1111/1467-8551.12302>

Thistlethwaite, J., Henstra, D., Peddle, S. and Scott, D. (2017). Canadian Voices on Changing Flood Risk: Findings from a National Survey. University of Waterloo, Waterloo, Ontario. Retrieved June 2020, from <https://uwaterloo.ca/climate-centre/sites/ca.climate-centre/files/uploads/files/canadian_voices_on_changing_flood_risk_fnl.pdf>

Toth, K. (2018). Grounded cruise ship rescue in Nunavut cost Canada’s Armed Forces $513K. CBC News. Retrieved June 2020, from <https://www.cbc.ca/news/canada/north/rcaf-akademik-ioffe-expensive-1.4810589>

Townhill, B.L., Pinnegar, J.K., Righton, D.A. and Metcalfe, J.D. (2016). Fisheries, low oxygen and climate change: how much do we really know? Journal of Fish Biology, 90, 723–750. Retrieved June 2020, from <https://doi.org/10.1111/jfb.13203>

Transport Canada (2017). Transportation in Canada: Overview report 2017. Transport Canada, Ottawa, 58 p. Retrieved June 2020, from <https://www.tc.gc.ca/media/documents/policy/Transportation_in_Canada_2017nwe.pdf>

Tremblay, M.J. (1997). Snow crab (Chionoecetes opilio) distribution limits and abundance trends on the Scotian Shelf. Journal of Northwest Atlantic Fishery Science, 21, 7–22. Retrieved June 2020, from <https://doi.org/10.2960/J.v21.a1>

UNEP [United Nations Environment Programme] (2012). Business and Climate Change Adaptation: Toward Resilient Companies and Communities. United Nations Global Compact Office, Geneva, 4‒50. Retrieved June 2020, from <https://www.unglobalcompact.org/docs/issues_doc/Environment/climate/Business_and_Climate_Change_Adaptation.pdf>

UNFCCC [United Nations Framework Convention on Climate Change] (2011). Assessing the costs and benefits of adaptation options: An overview of approaches. Retrieved June 2020, from <https://unfccc.int/resource/docs/publications/pub_nwp_costs_benefits_adaptation.pdf>

UNFCCC [United Nations Framework Convention on Climate Change] (2012). Nairobi Work Programme on Impacts, Vulnerability and Adaptation to Climate Change. Retrieved June 2020, from <http:// unfccc.int/adaptation/workstreams/nairobi_work_programme/ items/6547.php>

UNFCCC [United Nations Framework Convention on Climate Change] (2015). Paris Agreement. Retrieved June 2020, from <https://unfccc.int/sites/default/files/english_paris_agreement.pdf>

U.S. Department of Energy (2016). Climate Change and the Electricity Sector: Guide for Climate Change Resilience Planning. Retrieved June 2020, from <https://www.energy.gov/sites/prod/files/2016/10/f33/Climate%20Change%20and%20the%20Electricity%20Sector%20Guide%20for%20Climate%20Change%20Resilience%20Planning%20September%202016_0.pdf>

van Lierop, P., Lindquist, E., Sathyapala, S. and Franceschini, G. (2015). Global forest area disturbance from fire, insect pests, diseases and severe weather events. Forest Ecology and Management, 352, 78–88. Retrieved June 2020, from <https://doi.org/10.1016/j.foreco.2015.06.010>

Vick, S.G. (2001). Stability aspects of long-term closure for sulphide tailings, in Seminar on Safe Tailings Dam Constructions, Gallivare, 20–21.

Waldbusser, G.G., Hales, B., Langdon, C.J., Haley, B.A., Schrader, P., Brunner, E.L., Gray, M.W., Miller, C.A. and Gimenez, I. (2015). Saturation-state sensitivity of marine bivalve larvae to ocean acidification. Nature Climate Change, 5, 273–280. Retrieved June 2020, from <https://doi.org/10.1038/nclimate2479>

Waller, J.D., Wahle, R.A., McVeigh, H. and Fields, D.M. (2017). Linking rising pCO2 and temperature to the larval development and physiology of the American lobster (Homarus americanus). ICES Journal of Marine Science, 74, 1210–1219. Retrieved June 2020, from <https://doi.org/10.1093/icesjms/fsw154>

Warren, F.J. and Lemmen, D.S. (Eds.) (2014). Canada in a Changing Climate: Sector Perspectives on Impacts and Adaptation. Government of Canada, Ottawa, 286 p. Retrieved June 2020, from <https://www.nrcan.gc.ca/climate-change/impacts-adaptations/what-adaptation/canada-changing-climate-sector-perspectives-impacts-and-adaptation/16309>

Weber, M., Groulx, M., Lemieux, C., Scott, D. and Dawson, J. (2019). Balancing the dual mandate of conservation and visitor use at a Canadian world heritage site in an era of rapid climate change. Journal of Sustainable Tourism, 27(9), 1318‒1337. Retrieved June 2020, from <https://doi.org/10.1080/09669582.2019.1620754>

World Energy Council, the Cambridge Institute for Sustainability Leadership (CISL) and the Cambridge Judge Business School (2014). Climate Change: Implications for the Energy Sector, IPCC Climate Science Business Briefings. Retrieved June 2020, from <https://www.cisl.cam.ac.uk/business-action/low-carbon-transformation/ipcc-climate-science-business-briefings/energy>

Wedawatta, G. and Ingirige, B. (2016). A conceptual framework for understanding resilience of construction SMEs to extreme weather events. Built Environment Project and Asset Management, 6(4), 428–43. Retrieved June 2020, from <https://doi.org/10.1108/BEPAM-06-2015-0023>

Western Power Distribution (2011). Western Power Distribution Adaptation to Climate Change Report. 23 p. Retrieved June 2020, from <https://www.westernpower.co.uk/downloads/4117>

Whitman, E., Parisien, M., Thompson, D.K. and Flannigan, M.D. (2019). Short-interval wildfire and drought overwhelm boreal forest resilience. Scientific Reports, 9, 18796. Retrieved June 2020, from <https://doi.org/10.1038/s41598-019-55036-7>

Williams, S. and Schaefer, A. (2013). Small and Medium-Sized Enterprises and Sustainability: Managers’ Values and Engagement with Environmental and Climate Change Issues. Business Strategy and the Environment, 22(3), 173–186. Retrieved June 2020, from <https://doi.org/10.1002/bse.1740>

Williamson, T.B., Colombo, S.J., Duinker, P.N., Gray, P.A., Hennessey, R.J., Houle, D., Johnston, M., Ogden, A. and Spittlehouse, D.L. (2009). Climate change and Canada’s forests: from impacts to adaptation. Sustainable Forest Management Network and Northern Forest Center, Canadian Forest Service, Natural Resources Canada, Edmonton, 58 p.

Wilson, R.E. and Swanson, R.L. (2005). A perspective on bottom water temperature anomalies in Long Island Sound during the 1999 lobster mortality event. Journal of Shellfish Research, 24, 825–830. Retrieved June 2020, from <https://doi.org/10.2983/0730-8000(2005)24[825:APOBWT]2.0.CO;2>

Wilson, G.W., Plewes, H.D, Williams, D. and Robertson, J. (2003). Concepts for co-mixing of tailings and waste rock. In Proceedings of 6th International Conference on Acid Rock Drainage (ICARD), in Application and Sustainability of Technologies, Cairns, Australia, (Eds.) T. Farrell and G. Taylor. Australian Institute of Mining and Metallurgy, Carlton South, Australia, 437–443.

Wotton, B.M., Flannigan, M.D. and Marshall, G.A. (2017). Potential climate change impacts on fire intensity and key wildfire suppression thresholds in Canada. Environmental Research Letters, 12(9). Retrieved June 2020, from <https://doi.org/10.1088/1748-9326/aa7e6e>

Wotton, B.M., Nock, C.A. and Flannigan, M.D. (2010). Forest fire occurrence and climate change in Canada. International Journal of Wildland Fire, 19(3), 253–271. Retrieved June 2020, from <https://doi.org/10.1071/WF09002>

Xi, K., Kumar, K., Holtz, M.D., Turkington, T.K. and Chapman, B.P. (2015). Understanding the development and management of stripe rust in central Alberta. Canadian Journal of Plant Pathology, 37, 21–39. Retrieved June 2020, from <https://doi.org/10.1080/07060661.2014.981215>

Youcef, W.A., Lambert, Y. and Audet, C. (2015). Variations in length and growth of Greenland Halibut juveniles in relation to environmental conditions. Fisheries Research, 167, 38–47. Retrieved June 2020, from <https://doi.org/10.1016/j.fishres.2015.01.007>

Zhang, X., Flato, G., Kirchmeier-Young, M., Vincent, L., Wan, H., Wang, X., Rong, R., Fyfe, J., Li, G. and Kharin, V.V. (2019). Changes in Temperature and Precipitation Across Canada, Chapter 4 in Canada’s Changing Climate Report, (Eds.) E. Bush and D.S. Lemmen. Government of Canada, Ottawa, 112–193. Retrieved June 2020, from <https://changingclimate.ca/CCCR2019/chapter/4-0/>

Zhang, Y., Chen, W.J. and Riseborough, D.W. (2008a). Disequilibrium response of permafrost thaw to climate warming in Canada over 1850–2100. Geophysical Research Letters, 35, L02502. Retrieved June 2020, from <https://doi.org/10.1029/2007GL032117>

Zhang, Y., Chen, W.J. and Riseborough, D.W. (2008b). Transient projections of permafrost distribution in Canada during the 21st century under scenarios of climate change. Global and Planetary Change, 60(3), 443–456. Retrieved June 2020, from <https://doi.org/10.1016/j.gloplacha.2007.05.003>

Zheng, J. and Kruse, G.H. (2000). Recruitment patterns of Alaskan crabs in relation to decadal shifts in climate and physical oceanography. ICES Journal of Marine Science, 57(2), 438–451. Retrieved June 2020, from <https://doi.org/10.1006/jmsc.1999.0521>

Zisserson, B. and Cook, A. (2017). Impact of bottom water temperature change on the southernmost snow crab fishery in the Atlantic Ocean. Fisheries Research, 195, 12–18. Retrieved June 2020, from <https://doi.org/10.1016/j.fishres.2017.06.009>

Zorn, C., Thacker, S., Pant, R. and Shamseldin, A. (2016). The Dependence of National Transportation Infrastructure on Electricity. University of Canterbury, Christchurch, New Zealand. Retrieved June 2020, from <https://ir.canterbury.ac.nz/bitstream/handle/10092/12780/POSTER%20-%20Zorn_DependenceTrnsptInfstructureElect_QCAM_2016.pdf?sequence=1>

Zuliani, A., Massolo, A., Lysyk, T., Johnson, G., Marshall, S., Berger, K. and Cork, S.C. (2015). Modelling the Northward Expansion of Culicoides sonorensis (Diptera: Ceratopogonidae) under Future Climate Scenarios. PLoS ONE, 10(8), e0130294. Retrieved June 2020, from <https://doi.org/10.1371/journal.pone.0130294>

Next

Climate Disclosure, Litigation and Finance