climate change risk assessment for the biodiversity and...

244
(Defra Project Code GA0204) Climate Change Risk Assessment for the Biodiversity and Ecosystem Services Sector January 2012 1 Brown, I., 2 Ridder, B., 2 Alumbaugh, P., 2 Barnett, C., 2 Brooks, A., 2 Duffy, L., 2 Webbon, C., 3 Nash, E., 3 Townend, I., 1 Black, H. and 1 Hough, R. Contractors: 3 HR Wallingford 1 The James Hutton Institute 2 AMEC Environment & Infrastructure UK Ltd (formerly Entec UK Ltd) The Met Office Collingwood Environmental Planning Alexander Ballard Ltd Paul Watkiss Associates Metroeconomica

Upload: others

Post on 09-Aug-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

(Defra Project Code GA0204)

Climate Change Risk Assessment for the Biodiversity and Ecosystem Services Sector January 2012 1Brown, I., 2Ridder, B.,

2Alumbaugh, P.,

2Barnett, C.,

2Brooks, A.,

2Duffy, L.,

2Webbon, C.,

3Nash, E.,

3Townend, I.,

1Black, H.

and 1Hough, R.

Contractors: 3HR Wallingford

1 The James Hutton Institute 2AMEC Environment & Infrastructure UK Ltd (formerly Entec UK Ltd) The Met Office Collingwood Environmental Planning Alexander Ballard Ltd Paul Watkiss Associates Metroeconomica

Page 2: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

ii Biodiversity Sector Report

Statement of use See full statement of use on Page v Keywords: Climate, risks, biodiversity, ecosystems Research contractor: HR Wallingford Howbery Park, Wallingford, Oxon, OX10 8BA Tel: +44(0)1491 835381 (For contractor quality control purposes this report is also numbered EX 6434) Defra project officer: Dominic Rowland Defra contact details: Adapting to Climate Change Programme, Department for Environment, Food and Rural Affairs (Defra) Area 3A Nobel House 17 Smith Square London SW1P 3JR

Tel: 020 7238 3000

Document History: Date Release Prepared Notes

07/10/10 0.1 Entec UK Ltd Review copy for project team only

16/11/10 0.2 Entec UK Ltd Revised review copy for project team only

22/11/10 1.0 Entec UK Ltd, MLURI, HR Wallingford For peer review

10/02/11 2.0 MLURI Restructuring and response to peer review and Government department review comments

29/03/11 3.0 HR Wallingford Revised in response to Government department 2

nd review comments.

13/05/11 3.0A HR Wallingford Revised in preparation for development of release 4 stand alone reports

14/06/11 3.0A2 HR Wallingford Minor amendments

12/08/11 4.0 MLURI Major revision

21/10/11 4.0A JHI Revised in preparation for development of release 5 stand alone report

05/12/11 5.0 JHI and HR Wallingford Minor amendments and final review steps

13/01/12 6.0 JHI and HR Wallingford Minor edits.

23/04/12 7.0 JHI and HR Wallingford Minor edits.

18/05/12 8.0 HR Wallingford Minor edits.

Page 3: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report iii

Amended 23rd April 2012 from the version published on 25th January 2012. Amendments to the version published on 25th January 2012 The following corrections have been made: Pages 11, 32, 75, 76: Minor typographic and clarification edits. Page 215: Corrections to the magnitude table under the ‘social’ category.

Amended 18th May 2012 from the previous version 7.0 The following amendments have been made: Pages xiii, 70, 75 and 79: Minor typographic edits

© Crown copyright 2012 You may use and re-use the information featured in this document/publication (not including logos) free of charge in any format or medium, under the terms of the Open Government Licence http://www.nationalarchives.gov.uk/doc/open-government-licence/open-government-licence.htm Any email enquiries regarding the use and re-use of this information resource should be sent to: [email protected]. Alternatively write to The Information Policy Team, The National Archives, Kew, Richmond, Surrey, TW9 4DU. Printed on paper containing 75% recycled fibre content minimum.

Page 4: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

iv Biodiversity Sector Report

This report is available online at: http://www.defra.gov.uk/environment/climate/government/

Page 5: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report v

Statement of use This report presents the research completed as part of the UK Climate Change Risk Assessment (CCRA) for a selected group of risks in the Biodiversity and Ecosystem Services sector. Whilst some broader context is provided, it is not intended to be a definitive or comprehensive analysis of the sector.

Before reading this report it is important to understand the process of evidence gathering for the CCRA.

The CCRA methodology is novel in that it has compared over 100 risks (prioritised from an initial list of over 700) from a number of disparate sectors based on the magnitude of the consequences and confidence in the evidence base. A key strength of the analysis is the use of a consistent method and set of climate projections to look at current and future threats and opportunities.

The CCRA methodology has been developed through a number of stages involving expert peer review. The approach developed is a tractable, repeatable methodology that is not dependent on changes in long term plans between the 5 year cycles of the CCRA.

The results, with the exception of population growth where this is relevant, do not include societal change in assessing future risks, either from non-climate related change, for example economic growth, or developments in new technologies; or future responses to climate risks such as future Government policies or private adaptation investment plans.

Excluding these factors from the analysis provides a more robust ‘baseline’ against which the effects of different plans and policies can be more easily assessed. However, when utilising the outputs of the CCRA, it is essential to consider that Government and key organisations are already taking action in many areas to minimise climate change risks and these interventions need to be considered when assessing where further action may be best directed or needed.

Initially, eleven ‘sectors’ were chosen from which to gather evidence: Agriculture; Biodiversity & Ecosystem Services; Built Environment; Business, Industry & Services; Energy; Forestry; Floods & Coastal Erosion; Health; Marine & Fisheries; Transport; and Water.

A review was undertaken to identify the range of climate risks within each sector. The review was followed by a selection process that included sector workshops to identify the most important risks (threats or opportunities) within the sector. Approximately 10% of the total number of risks across all sectors was selected for more detailed consideration and analysis.

The risk assessment used UKCP09 climate projections to assess future changes to sector risks. Impacts were normally analysed using single climate variables, for example temperature.

A final Evidence Report draws together information from the 11 sectors (as well as other evidence streams) to provide an overview of risk from climate change to the UK.

Neither this report nor the Evidence Report aims to provide an in depth, quantitative analysis of risk within any particular ‘sector’. Where detailed analysis is presented using large national or regional datasets, the objective is solely to build a consistent picture of risk for the UK and allow for some comparison between disparate risks and regional/national differences.

Page 6: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

vi Biodiversity Sector Report

This is a UK risk assessment with some national and regional comparisons. The results presented here should not be used by the reader for re-analysis or interpretation at a local or site-specific scale.

In addition, as most impacts were analysed using single climate variables, the analysis may be over-simplified in cases where the consequence of climate change is caused by more than one climate variable (for example, higher summer temperatures combined with reduced summer precipitation).

Page 7: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report vii

Executive summary

Key findings

Many aspects of biodiversity are vulnerable to climate change but this is particularly evident in the coastal zone, uplands and wetlands which are generally high-value areas that are particularly sensitive to change.

Direct effects of climate change are already apparent and would be further exacerbated by increased water stress, changes in species ‘climate space’, the timing of natural seasonal events, migration patterns and sea level rise.

Indirect effects are expected to further compound direct effects including increases in invasive non-native species, impacts on soil and water quality, and increases in wildfire risk.

Indirect effects from actions in other sectors, such as climate mitigation strategies, have the potential if poorly implemented to endanger biodiversity or if properly implemented to enhance biodiversity.

Ecosystems’ processes can naturally adapt to change but anthropogenic pressures (habitat fragmentation, land use change, pollution, water abstraction and fixed flood or erosion defences), are currently acting to restrict this adaptation.

The complexity of ecosystems means it is difficult to project future risks with certainty. The evidence suggests that further changes seem inevitable, but rates of change are less certain. If the natural resilience of ecosystems is degraded and the rate of change is high, an abrupt step change in ecosystem processes can occur. Integrated responses across sectors are therefore required to enhance the resilience and adaptive capacity of the ‘ecosystem services’ that sustain human wellbeing.

Overall

Why the risks were chosen

Biodiversity is crucial to human wellbeing and sustainable development. Biodiversity and ecosystems have large cultural value for recreation, tourism, education and as inspiration for the arts and religion. Biodiversity loss can have negative effects on our health and material wealth. Our food and energy security strongly depend on biodiversity and so does our vulnerability to natural hazards such as fires and flooding. Biodiversity also regulates and maintains the quality of soil, air and water.

Changes in soil moisture deficits and drying (BD11).

Most UK habitats and species are adapted to a temperate wet climate with water availability only a problem in extreme dry years. They are therefore sensitive to a change in climate towards a state where present-day ‘extreme’ dry conditions become the norm. Some habitats and species are particularly sensitive to even small changes in moisture. This can lead to loss of ecosystem function and key services provided by wetlands and other habitats which could impact upon provisioning, regulating and 1 “BD” numbers are unique identifiers that identify each risk and are not related to the relative

importance of the impacts.

Page 8: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

viii Biodiversity Sector Report

cultural ecosystem services (food provision, water supply, water quality and landscape amenity).

Coastal evolution, extreme flooding events or coastline reconfiguration (BD2 and BD7).

Coasts are dynamic environments with habitats occupying particular zones relative to salinity levels, sediments, and hydrodynamic forces. These habitats have high biodiversity value because of the range of priority species that they support, particularly overwintering birds. Changes in sea level therefore lead to shifts in habitat extent and the niche space for species. This can provide both risks and opportunities for habitat change, depending on whether each habitat can migrate to a new position or has capacity to withstand increased erosion and inundation.

The position of any coast defence structures can restrict the movement of intertidal habitats on the seaward side (‘‘coastal squeeze”) but they can also protect habitats on the landward side. Some habitats, notably saltmarsh and dunes, have an important role in coastal defence and therefore can provide protection to humans from rising sea levels and extreme water levels. Impacts for habitats and species become particularly pronounced when change occurs in a short time period over a large area. Such change can be induced by extreme events caused by storm surges and breaches of low-lying coasts and their hinterland. They can also be the consequence of a deliberate decision to realign the coast by moving coastal defences inland.

Increased risk from invasive non-native species, pests and diseases (BD3 and BD4).

Biodiversity is vulnerable to the spread of organisms that damage natural ecosystems. Of particular concern is the threat from invasive non-native species introduced by human agency (deliberate or accidental), which can overwhelm native species and habitats, and potentially disrupt key ecosystem functions. Milder winters could increase the survival rates and spread of these nuisance organisms. Complex interactions and pathways may also develop due to the ‘natural’ spread of species in response to the changing climate. The presence of a wide diversity of species in ecosystems may have an important role in regulating the spread of pests and diseases.

Species unable to track changing climate space (BD5).

Species’ distributions are typically associated with an array of climate parameters that define a notional ‘climate space’. A changing climate implies that this climate space could be shifted geographically, or may contract or expand. Many species would then shift their range towards the new climate space if it defines the most favourable ecological conditions for that species to maintain a viable population. However, if species cannot track their changing climate space due to limits on dispersal or habitat availability then their populations are vulnerable, possibly to the point of extinction. These changes have the potential to cause disaggregation of ecological communities with consequent impacts upon ecosystem functioning and services.

Impacts of climate mitigation schemes (BD6).

Actions to reduce greenhouse gas emissions, such as via renewable energy schemes, are fundamental to a healthy future planet. As a result of ambitious targets, large-scale expansion of these schemes is occurring, especially through wind energy and bioenergy. However, if these schemes are not properly planned to account for current and future change in the natural environment, then they could cause damage to biodiversity. Conversely, well-planned schemes could increase opportunities for biodiversity.

Page 9: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report ix

Changes in soil organic carbon (BD8).

The organic content of soils is a key regulator of plant nutrient cycling and water availability. Changes in the relative rates of biomass production and decomposition due to climate change (temperature, precipitation and CO2 levels) can therefore impact on above- and below-ground biota and ecosystem functioning. These changes can also lead to modification of the carbon stocks held by the soil, manifest as either increased CO2 emissions or sequestration, with major implications for climate change policy.

Changes in migration patterns (BD9).

Many animals, particularly birds, migrate to seek favourable habitat and a reliable food supply during the annual cycle of seasons. A change in climate can therefore have many implications for these migration patterns, either at the source or destination site, or on the actual migration route. For some species this may provide new opportunities whereas for others it might introduce new risks to the viability of the species. The UK has particular international obligations for its populations of overwintering migratory birds in coastal areas.

Increased water temperature and changes in stratification (BD10).

Many aquatic species have life cycles that are based upon particular thermal requirements and are, therefore, sensitive to changes in water temperatures. Temperature changes may also change the thermal or salinity stratification of water bodies, such that the mixing between surface and subsurface water layers is altered, affecting the supply of oxygen and nutrients. Aquatic ecosystems feature intricate food webs therefore changes in individual species can modify the structure and functioning of the whole system. Aquatic and wetland ecosystems provide a key service to humans by regulating water quality and flow levels.

Generalist species benefiting at the expense of specialists (BD11).

A changing environment can modify ecological niches, meaning species will need to adapt to change. This typically means that those species that have less specific habitat preferences have advantages over those that have developed specialist requirements. As a consequence, change could lead to more ‘generalist’ species benefiting at the expense of ‘specialist’ species, with a reduction in biological diversity. Environmental change could be a product of changing climate or land use, or both. Impacts on ecosystem structure could also be as a consequence of invasive generalist species (BD3). Changes in ecosystem structure from species loss could disrupt the delivery of key services such as pollination.

Increased risk of wildfire (BD12).

An increased prevalence of hotter, drier conditions as indicated by increased soil moisture deficits (BD1) also implies a greater risk of fire. Some ecosystems, such as woodlands, semi-natural grasslands, heathlands, and those on peat soils (e.g. bogs) are particularly sensitive to fire. Therefore an increased prevalence of large fires on these important habitats could lead to significant loss of biodiversity and ecosystem function for example carbon sequestration in woodlands and peatlands.

Impacts on water quality (BD13).

Pollution of water bodies occurs through runoff of fertiliser from agricultural land, influx of human wastes, and deposition from the atmosphere. The impact of these pollutants depends also on the temperature and quantity of water for dilution, but is already a major source of damage to aquatic habitats.

Page 10: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

x Biodiversity Sector Report

Impacts on water quantity: low flows, societal demand, drought (BD14-16).

During low flow periods, reduced oxygen is available for biodiversity and effluent from sewage and other sources can become more concentrated. These can cause ecological stress but if the event is temporary and adequate flows return, then eventually the ecosystem may recover. However, levels of water demand and abstraction rates are increasing in many catchments which will act in combination with any direct impact on changes in water supply from climate change to modify flow regimes. Recovery of the ecosystem may therefore be further jeopardised, particularly during extended drought periods.

Other risks to terrestrial and freshwater habitats

A series of other risks were also evaluated during the assessment process but not in detail, including:

change in species interactions due to differing growth and survival rates, with the potential to alter the food web.

asynchronous timing of annual life cycle events between species, such as between predators and prey species or plant growth (‘phenological mismatch’).

impacts of agricultural land use intensification or land abandonment, particularly in marginal areas that have habitats sustained by a limited amount of human disturbance.

modification to soil nutrient availability and cycling, through enhanced biomass production and decomposition.

an increase in flood defence structures and consequent further isolation of rivers from floodplain and wetland habitats.

Marine Biodiversity

Marine ecosystems are already experiencing major change. A synthesis is also provided of key risks that have the potential to induce major consequences to the marine environment including: harmful algal blooms, ocean acidification, species range shifts, invasive non-native species, disease hosts and pathogens.

Current pressures and vulnerability

Climate is a key factor influencing the distribution of ecosystems, habitats and species. Therefore many aspects of biodiversity are sensitive to change, and adaptation to this change is a natural process that has occurred for millions of years.

Biodiversity is sensitive to anthropogenic climate change because it is already under pressure from a range of other stresses and as the rate of climate change increases it is happening at a level that exceeds the usual adaptive responses. Other pressures include land-use change and pollution, and reflect a trend over recent decades for increased intensification of land use, particularly in agriculture. This has tended to result in large areas of land optimised for crop or livestock production but with limited biodiversity and simplified ecosystems. Areas of good quality habitat that do remain, such as in protected sites, have been left as fragments that are often too small or lack the coherence to functionally

Page 11: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xi

adapt to change. The larger areas of habitat that remain are mainly in the uplands but these are also under pressure from atmospheric pollution.

An additional pressure is the presence of man-made fixed defences on coastal and river floodplains, and zones of erosion. These separate active and non-active areas of the coast or fluvial zone and impose a static structure on dynamically responsive landforms and habitats. This acts to the detriment of many habitats, resulting in both coastal ‘squeeze’ of intertidal habitats, and the detachment of wetlands from the river or sea.

These additional co-stressors and their consequences for reduced resilience mean that current vulnerability to change is generally ‘high’. It has been exacerbated because the natural capacity to adapt to change has been severely hindered by human activities, particularly habitat degradation and fragmentation. This also means that it is difficult for species to track their changing ‘climate space’; recombination of different species may occur, thus changing communities and ecosystems. By reducing the ability of ecosystems to respond to change, there is an increased likelihood of non-linear step changes that can cause both a major decline in biodiversity but also a reduction in the delivery of key ecosystem services that support human well-being.

Current policy on biodiversity is strongly influenced by the requirements of the EU Habitats Directive and Birds Directives that aim to protect species and habitats through the ecological site network (Natura 2000), and other site designations (e.g. SSSI, Ramsar). The Habitats Directive does provide scope to modify the site network to maintain favourable conservation status for priority habitats and species. However, the implementation of the Directive has to-date tended to follow a static rather than a dynamic approach, which acts to increase current vulnerability. Positive examples do exist: for example, notification of parts of the Humber estuary SSSI includes 50years of projected future coastal change.

Ecosystem services are not explicitly covered by current policy (but recognised by emerging frameworks in the UK and devolved countries). In addition, some policies in other sectors (e.g. the Water Framework Directive) include measures that recognise the key role of biodiversity and the natural environment in maintaining water quality and resources for humans. These measures may also need to be further refined with regard to the additional pressures from climate change. For other sectors, the role of ecosystem services is less well recognised (e.g. agriculture) which means that these services can be vulnerable to both autonomous (unplanned) adaptation and inadvertently through planned adaptation responses that have only a narrow sectoral focus. Vulnerability may also be increased through climate mitigation policies that do not consider the effects of future climate change on the local landscape.

Ecosystems typically contain a multitude of complex interactions and natural responses to climate change which, combined with the human influences, makes prediction of change extremely difficult. Habitats and species can exhibit rather different responses depending on their local context and antecedent conditions. This complexity can increase vulnerability because although good monitoring data is available from some sites, there has been limited systematic collation and analysis of data on a large scale as is required to understand the relative influences of different factors.

Page 12: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xii Biodiversity Sector Report

Vulnerability is generally greatest for uplands, coasts and wetlands. Habitats and species at either the southern end or lower altitudinal limit of their climate range are particularly vulnerable in the UK as they will have nowhere to retreat to as the climate changes. The inherent vulnerability of a species or habitat is a key component in an assessment of the potential impacts of climate or other change. The national nature conservation agencies of the UK are currently developing procedures for assessing this vulnerability (see Appendix 4: example for Countryside Council of Wales).

Risk assessment

Increased soil moisture deficits and drying (BD1).

The current risk is primarily manifest during dry years but the projected future trend to drier summers would increase the exposure to harmful events. In addition to the loss of priority habitats and species that need wetter conditions, this is likely to lead to consequences for ecosystem services. Analysis of this risk has highlighted two exemplar priority habitats, beech woodland and blanket bog, from lowland and upland environments respectively, but a wide range of habitats are vulnerable, including grassland, heathland, wetlands and coastal habitats. Key services provided by these habitats include the regulation of water quantity and quality, storage of carbon, and landscape amenity. Increased summer drying would have impacts on invertebrates, such as worms and crane flies, which are a key component of ecosystem food webs. Variations in local conditions will determine the actual change in risk based upon factors such as soils and topography. Land management (e.g. drainage systems) will also combine with climate change to exacerbate or alleviate changes in soil moisture levels.

Coastal evolution, extreme flooding events or coastline reconfiguration (BD2 and BD7).

Rising sea levels mean that ‘coastal squeeze’ of intertidal habitats on the seaward side of defences is currently a problem, particularly in East and South England. This is highly likely to be further exacerbated by an accelerated rate of sea level rise. The consequence would be a major loss of intertidal habitat unless sea defences are shifted to allow the habitat to ‘migrate’ inland. If sediment supply is available, saltmarsh habitats have the potential to maintain their elevation with sea level rise and provide additional natural coastal protection from inundation. On the landward side of the sea defences, freshwater and other habitats may be lost in some locations if the defence line is realigned. Hence, coastal evolution is projected to produce both habitat gains and losses in all circumstances, with major implications for the rich variety of species they support.

Extreme events are by definition, low frequency and high magnitude risks. When they do occur (e.g. the storm surge of 1953) they can cause major damage to both the human and natural environment. Extreme high water levels from rising sea levels combined with storm surge and large waves could lead to an increased risk of a breach in low-lying areas of the coast and reorganisation of coastal ecosystems. This risk is particularly apparent in East and South England, although other areas of the UK have also been identified as risk zones. It also possible that society in the future will decide that investment in sea defences for some sections of the coast is unsustainable and over time a large-scale realignment of the coast could occur which would have implications for the habitats and species in that area.

Page 13: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xiii

Increased risk from pests, diseases and invasive non-native species (BD3 and BD4).

Pests and diseases such as Parrot’s-feather, Phytothphora ramorum and Chytridiomycosis currently present a severe risk to some priority habitats and species, particularly in aquatic ecosystems. As low winter temperatures currently act as a climatic control on many pests and diseases, future projections of milder winters suggest that the risk from pests and diseases would therefore increase, and be particularly apparent in the spread of new invasive non-native species during milder temperatures in winter. A trend towards wetter winters would also increase problems from fungi and related organisms such as Phytothphora that thrive on high moisture levels. However, each pest and disease has its own characteristics that preclude generalisations, with the spread of diseases also determined by the population dynamics of the host organism. It is therefore important that climate change is factored into risk assessments for each organism. Pests and diseases do not only affect the natural environment but can be transferred to and from agriculture and forestry, particularly where natural controls are limited as with invasive non-native species. There is some evidence that the complexity of natural ecosystems can regulate the spread of pests and diseases, possibly by increasing the number of species interactions that the organisms must overcome. Climate change may also challenge our current concept of ‘invasive non-natives’ (i.e. those introduced by human agency as defined by the Convention on Biological Diversity) by inducing complex pathways that combine human agency with natural spread mediated by climate.

Species unable to track changing climate space (BD5).

There is already evidence that the ‘climate space’ of many species has been modified by climatic warming. In some cases, particular species have been unable to adapt to this change, primarily due to habitat loss and fragmentation. The greater magnitude of projected future climate change suggests that many species will have to adapt to survive by tracking a general northerly movement of their climate space. In upland areas, the diverse topography means that the same effects would also be apparent with altitude. For some species, namely those at the southerly limit of their range in the UK, or montane species at their altitudinal limit, this may not be possible as they have nowhere to retreat to as the climate changes. For other species, habitat fragmentation due to land use pressures may hinder their potential to disperse to new sites. This can be addressed through planned adaptation measures that improve the size and connectivity of sites within the ecological site network, and measures to reduce the other pressures on species in the wider countryside and to improve general landscape ‘permeability’ to species movement. General measures to improve species’ resilience in situ include enhancing habitat diversity and local microclimatic variation, and improving population abundance of vulnerable species. In some cases, assisted translocation to new sites may be required for the viability of some priority species. The changing climate within the UK may also provide an opportunity for new species, and if these are recognised as good for biodiversity objectives then designated site management objectives may need to be modified in order to accommodate them.

Impacts of climate mitigation schemes (BD6).

Climate mitigation schemes in the UK, such as an expansion of renewable energy sources, are a recent policy development and therefore the evidence base for their impacts on biodiversity remains limited. However, it is clear that some habitats and species are sensitive to particular schemes. There are ambitious future plans to meet GHG emissions targets for 2020 and 2050. The future risk is not possible to identify with certainty as it depends on many factors including the actual type of scheme and its spatial distribution in the landscape. However, the impacts of some of the larger-scale

Page 14: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xiv Biodiversity Sector Report

schemes such as for bioenergy, tidal barrages, wind power and hydro-power could have both positive and negative implications. Environmental impact assessments for these schemes need to factor in both present and future change in biodiversity otherwise they could hinder adaptation. The development of new schemes and technology in suitable locations could also be maximised to enhance benefits for biodiversity and the wider landscape.

Changes in soil organic carbon (BD8).

Recent measurements of changes in soil organic carbon (SOC) suggest that climate change has had a rather less significant impact than land use change. However, the complexity of interactions involved in the cycling of SOC mean that the risk factors are poorly understood. The much larger changes in climate projected for the future (including increased CO2 concentrations) suggests that these interactions will be further modified. It is not currently possible to state with certainty whether this will result in an increased carbon store or sink in the UK but it seems likely that the pattern will be geographically and temporally variable and related to soil moisture deficits (BD1). In addition to the implications for GHG emissions, changes in soil organic matter has implications for soil biota. This in turn is likely to impact on all plants and animals that rely on these biota, on ecosystem functioning (e.g. nutrient cycling), and on ecosystem services such as water retention.

Changes in migration patterns (BD9).

There is abundant observational evidence of present-day changes in migration patterns, particularly with bird species, including both long-distance and short-distance migrants. Arrival and departure dates are often celebrated in folklore and culture as symbols of the changing seasons, particularly the advent of spring, and provide an important link between the natural and human landscape. In some cases, the changes in migration patterns are a signal of behavioural change as the species adapts to changing habitat and food availability during the seasons. It is very likely that these patterns will change increasingly in the future, particularly due to increases in seasonal temperatures. However, it is very difficult to predict these changing patterns as they depend on climate change (and other factors) at both the source and destination site, and along the migration route.

Changes in migration patterns have important implications for the designated site network, particularly with regard to the distribution of overwintering birds on the coast which the UK has international obligations to protect. The site network may therefore need to be modified or expand, as population distributions change. In addition, if migrants arrive earlier in the year, the local habitat may become less able to support the population, for example if plants and invertebrates that provide food have not synchronously advanced their life cycle at the same rate as the arriving migrants. Evidence for this mismatch in the timing of seasonal events (phenology) between species remains limited at present, and is perhaps compensated by other factors, but is potentially a severe risk with the large changes projected for the future.

Increased water temperature and changes in stratification (BD10).

An increase in water temperatures has been recorded at many sites across the UK but ecological consequences remain uncertain because of the complexity of aquatic ecosystems and confounding factors such as water pollution. Nevertheless, there is good evidence of changes in fish and invertebrate populations, and relative abundance of phytoplankton, linked to changes in temperature. These are manifest through changes in food supply, fish growth, and invertebrate life cycles. Future projections indicate increased impacts, and that they may be further compounded in lakes by changes in nutrients and oxygen from increased thermal stratification. The

Page 15: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xv

consequences will depend on the interaction of these factors with other influences such as the level of pollution and the presence of invasive non-native species (BD3).

Generalist species benefiting at the expense of specialists (BD11).

The current evidence base from some well-recorded species (e.g. birds, butterflies) suggests that this impact is already occurring but that land use change and habitat fragmentation is the dominant factor. Implications for the wider functioning of the ecosystem remain uncertain. Future risks will depend on the rate and magnitude of climate change in combination with land use. If the current state of habitat fragmentation is not reversed, the loss of niche space may mean that many specialist species will become extinct, particularly if the rate of climate change becomes more pronounced. The key planned adaptation measure required to address this risk is to make the landscape more diverse to ensure that a wide variety of habitats are available and to encourage dispersal by specialist species (whilst also developing measures to counter the spread of invasive non-native or pest species – BD3/4). This risk may ultimately affect species that provide important ecosystem services to humans, such as pollinators that benefit agriculture and silviculture.

Increased risk of wildfire (BD12).

At present, fire risk is particularly apparent during hot dry years or unusually dry seasons. At these times a large-scale fire can cause severe damage to woodlands, heathland, or grassland habitats and the species they support. The risk can also spread to or from land use activities such as agriculture, forestry and grouse moor management. An increased prevalence of these hot dry conditions as suggested by projections based upon UKCP09 data is implied to increase the background risk factors. This has important implications not just for biodiversity but also for these other sectors and emphasises the requirement for integrated land use and emergency planning to develop risk reduction strategies that take full account of important ecosystem services. Further work is required to ascertain the full extent of this risk.

Impacts on water quality (BD13).

Impacts of climate change will depend on the magnitude of pollution combined with changes in temperature and quantity of water for dilution. These may be further compounded by increased runoff during intense rainfall events. One particularly severe consequence is likely to be an increase in harmful algal blooms. Impacts are likely to extend from aquatic ecosystems to key services such as water purification and fisheries.

Impacts on water quantity (BD14-16).

A change in rainfall patterns, particularly a trend to warmer drier summers, is likely to result in lower river flows. This seems also likely to be accompanied by continuing increases in demand for water. Reduction in flow levels combined with higher temperatures would decrease oxygen supplies available for aquatic habitats. It would also likely to be accompanied by less dilution of harmful pollutants. These impacts would be particularly exacerbated during drought events when the risk of irreversible ecological change would be much higher. The consequences for fisheries, amenity value of wetlands, and the supply of clean water, will depend not just on climate change but on the implementation of appropriate regulatory limits to abstraction.

Marine biodiversity

Harmful algal blooms. As with terrestrial freshwater habitats, pollution combined with high temperatures increases the risk of harmful algal blooms due to the high nutrient

Page 16: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xvi Biodiversity Sector Report

status. This can deoxygenate the water and produce harmful toxins to both humans and biodiversity, with knock-on effects for fisheries and tourism. This risk is expected to increase due to future temperature rises.

Ocean acidification. Higher dissolved CO2 concentrations in seawater are leading to an acidification of the marine environment which will be further exacerbated as CO2

levels continue to increase in the future. This has particular consequences for shellfish and corals because it retards the process of shell production from calcium carbonate. In addition, they are likely to be many secondary impacts across marine ecosystems related to growth and life cycles but as yet these consequences remain to be fully understood.

Species range shifts. A continued rise in sea temperatures may cause many species to shift their geographic range in order to maintain a similar relationship with their environment. These impacts are already apparent and are having important consequences for species interactions, in combination with other pressures such as acidification and overfishing.

Invasive non-native species. As defined by the Convention on Biological Diversity, these harmful species are introduced by human agency and are often without natural predators in their new environment. A change to a warmer climate is likely to lead to an increased prevalence and persistence of invasive non-native species. In addition to the severe detrimental impacts on native species and ecosystems, this also implies important economic and social consequences.

Disease hosts and pathogens. Modifications to species range and changes in species interactions due to climate change also increase the risk from pathogens. Diseases therefore have greater potential to move to new hosts and produce new outbreaks in previously-unaffected areas.

Priority habitats and species. In addition to impacts on areas of high biodiversity value in the coastal zone, changes in the marine environment are highly likely to impact other priority habitats and species. In the subtidal region, eelgrass beds may be affected by increased temperature and any potential increase in storminess and erosion. The growth rate of corals and other reef organisms would be particularly affected by changing CO2 concentrations through ocean acidification.

Emerging challenges

Biodiversity and ecosystem services are clearly sensitive to change. The key challenge for this sector is dealing with the complexity of the natural environment and ensuring that these issues are effectively incorporated within other sectors. Risks are usually interconnected and there are key interdependencies with other sectors.

The evidence base for natural responses to climate change is sometimes contradictory because of this complexity, and the presence of other drivers, notably land use change. The importance of natural (autonomous) adaptation in this sector means that the development of risk metrics through response functions, as employed throughout the CCRA, is a less robust method to assess future risks, although still constructive. In addition many cause-effect relationships are non-linear and have important thresholds beyond which a step change can occur. Recently, the concept of ecosystem services has emerged as a potential framework through which to better understand systemic interdependencies between biodiversity and other sectors, but there are many gaps in our knowledge of these services.

Change also presents a range of conservation challenges. What do we mean by ‘native’ species? How do we make the ecological site network more dynamic and adaptable to change whilst still protecting priority species and habitats? How can we

Page 17: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xvii

better build resilience to undesirable change whilst accepting some change seems inevitable? In some cases, these challenge existing management practice.

Institutional adaptive capacity/awareness in sector

Awareness of climate change is typically high in this sector but institutional adaptive capacity is also defined by the need to co-ordinate with and influence actions in other sectors. This is particularly manifest through decision making in the wider countryside beyond the designated site network. This can act to reduce adaptive capacity because risks and opportunities from climate change are framed by different policy or institutional contexts in other sectors. As a consequence, establishing a consensus for intervention can be difficult. However, exemplars are available through partnership schemes that have demonstrated success. Even within the designated site network, adaptation options are sometimes constrained because the land is in private ownership.

In addition, although there is recognition of dynamic ecological processes within legislation (e.g. EU Habitats Directive or for SSSIs) and in guidance for spatial planning, systematic use is not currently being made of planning procedures that could deliver this in practice. Furthermore, some of the moral and ethical issues relating to biodiversity conservation (e.g. translocation of species, definitions of “invasive species”) present weighty dilemmas and are the subject of ongoing debate. For all these reasons, adaptation action ‘on the ground’ has sometimes been constrained within this sector. More broadly, development of an ecosystem approach to integrate the benefits from the natural environment within a cross-sectoral framework that explicitly identifies its economic and social value remains in its early stages, and is currently at the capacity-building stage.

Interdependencies

Key links to other CCRA risks/reports

This sector has many interdependencies with other sectors. These include the following:

Floods and coastal erosion: the presence and fixed location of defence structures can constrain the capacity of ecosystems to adapt to change. The presence of some ecosystems such as wetlands and woodlands can help regulate water flows and reduce flood peaks therefore providing natural flood management both in fluvial and coastal environments.

Water: the regulation of water abstraction and protection of water quality are key controls that can maintain good ecological status of wetlands and water bodies, as recognised by current legislation. In return, the maintenance of a healthy ecosystem can help regulate water quality and buffer low and high lows.

Built environment: the presence of ‘greenspace’ in urban environments can offer significant opportunities for biodiversity and help regulate the impacts of climatic change due to higher temperatures (e.g. heat stress) and heavier rainfall (e.g. pluvial flooding), particularly during extreme events.

Agriculture: increased intensification of production systems can reduce landscape and biological diversity and potentially degrade water quality. This can damage the natural benefits that are obtained from ecosystems such as pollination, soil nutrients, and control of pests and diseases. Agri-environment schemes can be instrumental in delivering wider

Page 18: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xviii Biodiversity Sector Report

environmental benefits. Reform of the Common Agricultural Policy is likely to be an important policy driver for changes in land use and land management.

Forestry: similarly to agriculture, increased emphasis on production systems can damage the natural environment with potential loss of ecosystem services such as pollination, soil nutrients, and control of pests and diseases. However, the forestry sector has implemented several initiatives in recent decades that have enhanced woodland ecosystems, including long-term planning for woodland habitat networks. The recently-published UK Forestry Standard and its Climate Change Guidelines identify the importance of maintaining biodiversity benefits. With good planning and management, opportunities provided by the climate change mitigation agenda may provide for expansion of woodland habitats.

Health: many diseases affecting humans also have interactions with the natural environment; therefore risks may have a similar profile. ‘Green’ infrastructure and networks are increasingly recognised for their benefits to human well-being in addition to their value for biodiversity.

Energy: policy targets for climate mitigation measures and renewable energy schemes are likely to have an increasing land footprint, and these will interact with changes in the climate and natural ecosystems. Well-planned schemes could have benefits for enhancing ecosystem resilience to climate change.

Marine and Fisheries: the coastal zone represents the key interface zone between terrestrial and marine processes, and between land use and marine policy.

Transport: ecosystems can buffer transport infrastructure from flooding, erosion, and other hazards such as slope instability. However this infrastructure can often provide fixed barriers which hinder natural adaptation through species dispersal.

Business, Industry and Services: the natural heritage value of many landscapes in the UK is a key asset that boosts tourism and other businesses, particularly in National Parks and other areas of high landscape quality.

In addition, some risks such as fire, pests and diseases, or the impacts of extreme events (e.g. floods or droughts) are likely to involve several sectors together. The following non-climate factors have been identified to have a major influence on future risks and opportunities for this sector:

Land use change, which in some cases may also be attributable to climate change hence causing indirect secondary effects on the natural environment. Changes in agri-environment schemes or forestry grants could also act as positive or negative incentives.

Pollution, especially of water bodies, from existing and new sources.

Invasive non-native species, introduced either deliberately or accidentally.

Attitudes to the natural environment and ecosystem services, including the ability to value non-market goods in economic assessments and to include irreversible change.

Governance and regulation: as most ecosystems already exist in a partly managed state, they require human intervention to adapt to change, but

Page 19: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xix

this requires governance at appropriate scales (especially landscape scale) rather than isolated piecemeal interventions at site level.

Each of these factors is also likely to vary into the future due to changes in socioeconomic drivers such as technological, demographic and behavioural changes.

About the analysis

Data quality and modelling issues

Significant uncertainties exist in the risk assessment for the sector. In some cases, due to the complexity of interactions and other confounding factors, even the current level of sensitivity to climate change remains to be established with reasonably high confidence. Furthermore, high quality data is often available only for specific sites and due to the high spatial variability of factors influencing biodiversity and ecosystem functions (e.g. climate, soils, land use), and inherent time lags in cause-effect relationships, extrapolating from these sites to the wider landscape can often only be indicative of the magnitude of future risks. For these reasons developing risk metrics based upon climate response functions is only a partial approach for this sector, and a broader contextualisation of risk is also required. Models of ecosystem dynamics and the consequences for ecosystem functions and services are at an early stage of development, although this is a very active area of research at present.

What is certain and what is uncertain (better framing)

We have reasonably good knowledge of the response of some species to current climate change but less on species interactions and of habitat change. At a higher level there are fundamental uncertainties in our knowledge of ecosystem responses and of changes in soil functions, both of which should be a major source of concern because of the vital services provided by ecosystems to human well-being.

The natural sensitivity of many ecosystems to climate means that some form of change is inevitable. However, this will be further compounded by it acting in conjunction with other stressors such as land use change and pollution. If ecosystems are not able to adapt to change because of constraints imposed by human use of the land then the evidence suggests significant loss of biodiversity. This may be also manifest in major or irreversible loss of ecosystem functioning, although our knowledge of these interlinked processes is rather incomplete. This will be particularly pronounced in environments where the rate of change is expected to be highest and other pressures are already pronounced, including coasts, the uplands and aquatic habitats.

The complexity of ecosystems means it has not been practical to use simple climate response functions to the same degree as in other CCRA sectors because of inherently high spatial variability and the importance of dynamic natural (autonomous) adaptation. A naive implementation could mislead as to the magnitude of future risk. Where response functions have been used then they are best considered as indicators of current change rather than a universal metric, and it is quite possible that adaptive responses will change the future relationship. Nevertheless, there is enough evidence from high-quality empirical observation data, and from theory and modelling, to indicate the need to develop planned adaptation responses now that operate in tandem with natural adaptation. The risk of irreversible change to biodiversity and ecosystem services is substantially increased if no further actions are implemented beyond the status quo.

Page 20: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xx Biodiversity Sector Report

Page 21: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xxi

Key term glossary The key terms are defined below.

Adaptation (IPCC, 2007)

Autonomous adaptation – Adaptation that does not constitute a conscious2 response to climatic stimuli but is triggered by ecological changes in natural systems and by market or welfare changes in human systems. Also referred to as spontaneous adaptation.

Planned adaptation – Adaptation that is the result of a deliberate policy decision, based on an awareness that conditions have changed or are about to change and that action is required to return to, maintain, or achieve a desired state.

Adaptive capacity -The ability of a system to design or implement effective adaptation strategies to adjust to information about potential climate change (including climate variability and extremes), to moderate potential damages, to take advantage of opportunities, or to cope with the consequences (modified from the IPCC to support project focus on management of future risks) (Ballard, 2009). This sector report differs from the generic CCRA approach as herein adaptive capacity also includes, where it is present, the natural resilience of ecosystems to respond to climate change and maintain their key functions.

Adaptation costs and benefits

The costs of planning, preparing for, facilitating, and implementing adaptation measures, including transition costs.

The avoided damage costs or the accrued benefits following the adoption and implementation of adaptation measures.

BAP species/habits or priority species/habitats - The UK Biodiversity Action Plan (UK BAP), published in 1994 sets out a programme for conserving biodiversity in the UK. The UK BAP has published lists of species and habitats that are conservation priorities which are under threat because of their rarity and rate of decline. A review of the UK BAP priority list in 2007 led to the identification of 1,150 species and 65 habitats that meet the BAP criteria at UK level (for more information, see http://www.naturalengland.org.uk/ourwork/conservation/biodiversity/protectandmanage/prioritylist.aspx).

Consequence - The end result or effect on society, the economy or environment caused by some event or action (e.g. economic losses, loss of life). Consequences may be beneficial or detrimental. This may be expressed descriptively and/or semi-quantitatively (high, medium, low) or quantitatively (monetary value, number of people affected etc).

Impact - An effect of climate change on the socio-bio-physical system (e.g. flooding, rails buckling).

Response function - Defines how climate impacts or consequences vary with key climate variables; can be based on observations, sensitivity analysis, impacts modelling and/or expert elicitation.

2 The inclusion of the word ‘conscious’ in this IPCC definition is a problem for the CCRA and we

treat this as anticipated adaptation that is not part of a planned adaptation programme. It may include behavioural changes by people who are fully aware of climate change issues.

Page 22: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xxii Biodiversity Sector Report

Risk - Combines the likelihood an event will occur with the magnitude of its outcome.

Sensitivity - The degree to which a system is affected, either adversely or beneficially, by climate variability or change.

Uncertainty - A characteristic of a system or decision where the probabilities that certain states or outcomes have occurred or may occur is not precisely known.

Vulnerability - Climate vulnerability defines the extent to which a system is susceptible to, or unable to cope with, adverse effects of climate change including climate variability and extremes. It depends not only on a system’s sensitivity but also on its adaptive capacity.

Pests – These species represent either native or non-native (alien) organisms that cause nuisance value through damage to native species or ecosystems.

Diseases – These species are micro-organisms (pathogens such as bacteria, fungi or viruses) that cause harm when transmitted to a particular host, such as Dutch elm disease in trees and sea distemper virus in mammals.

Invasive – These are species that cause major disruption to ecosystems and can cause severe environmental, economic or social damage. Invasive non-native species can be particularly problematic if they have no native natural predator to act as a control. The Convention of Biological Diversity defines invasive non-native species as those that are introduced by human agency into new biogeographical regions in order to distinguish them from species movements by natural mechanisms.

Page 23: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xxiii

Acknowledgements The authors of this report wish to acknowledge Alistair Hunt of Metroeconomica for his contribution to the costs assessment in Chapter 6.

Page 24: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xxiv Biodiversity Sector Report

Page 25: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xxv

Contents

Statement of use v

Executive summary vii

Key term glossary xxi

Acknowledgements xxiii

Contents xxv

1 Introduction 1

1.1 Background 1

1.2 Biodiversity and ecosystem services sector 3

1.3 Overview of the biodiversity and ecosystem services sector 5

1.4 Policy context 8

1.5 Structure of this report 16

2 Methods 18

2.1 Introduction: CCRA framework 18

2.2 Outline of the method used to assess impacts, consequences and risks 19

2.3 Identify and characterise the impacts 21

2.4 Assess vulnerability 22

2.5 Identify the main risks 22

2.6 Assess current and future risk 23

2.7 Report on risks 24

2.8 Confidence levels 25

3 Impacts and risk characterisation 26

3.1 High-level identification of climate impacts (Tier 1) 26

3.2 Identification of priority (Tier 2) risks 34

3.3 Synthesis of other key risks (terrestrial biodiversity) 37

3.4 Synthesis of other risks (marine biodiversity) 41

3.5 Cross-sectoral risks 45

3.6 Ecosystem services 47

4 Sector risk analysis (Tier 2) 50

4.1 Introduction 50

4.2 Increased soil moisture deficits and drying (BD1) 51

4.3 Coastal evolution impacts on intertidal, grazing marsh etc (BD2) and Major coastal flood / reconfiguration (BD7) 60

4.4 Increased risks from pests, diseases and invasive non-native species (BD3 and BD4) 70

Page 26: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xxvi Biodiversity Sector Report

4.5 Species unable to track changing climate space (BD5) 80

4.6 Climate mitigation measures (positive/negative) (BD6) 87

4.7 Changes in soil organic carbon (BD8) 92

4.8 Changes in species migration patterns (BD9) 97

4.9 Increased water temperature and stratification of water bodies (BD10) 102

4.10 Water quality and pollution risk and eutrophication (BD13) 106

4.11 Generalists favoured over specialists (BD11) 109

4.12 Increased risk of wildfires (BD12) 113

4.13 Impact on water quantity: low flows (BD14); increased societal water demand (BD15); major drought (BD16) 118

5 Socio-economic change 121

5.1 Introduction 121

5.2 Increased soil moisture deficits and drying (BD1) 123

5.3 Coastal evolution impacts and major coastal reconfiguration (BD2 and BD7) 124

5.4 Increased risk from pests, diseases and invasive non-native species (BD3 and BD4) 125

5.5 Species unable to track changing climate space (BD5) 125

5.6 Climate mitigation measures (positive/negative) (BD6) 126

5.7 Changes in soil organic carbon (SOC) (BD8) 127

5.8 Changes in species migration patterns (BD9) 127

5.9 Increased water temperature and stratification of water bodies (BD10) 128

5.10 Water quality and pollution risk and eutrophication (BD13) 128

5.11 Generalists favoured over specialists (BD11) 128

5.12 Increased risk of wildfire (BD12) 129

5.13 Impact on water quality: low flows, increased societal water demand, major drought (BD14/15/16) 129

5.14 Social vulnerability Issues 129

6 Monetisation 132

6.1 Summary 132

6.2 Introduction 133

6.3 Specific risks 141

7 Adaptive capacity 153

7.1 Overview 153

7.2 Assessing structural and organisational adaptive capacity 153

7.3 Natural adaptive capacity 155

7.4 Interactions between natural and organisational adaptive capacity 156

7.5 Enhancing adaptive capacity 157

8 Discussion: implications for adaptation 159

8.1 Risks 159

8.2 Costs and benefits 164

Page 27: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report xxvii

8.3 Risk in wider context 165

8.4 Limitations of analysis and knowledge gaps 167

9 Conclusion 170

References 171

Appendices 193

Appendix 1 Acknowledgements 195

Appendix 2 Tier 1 impacts scores 197

Appendix 3 Application of climate change projections 207

Appendix 4 Climate vulnerability assessment of designated sites in Wales – a summary 209

Appendix 5 Magnitude, confidence and presentation of results 213

Tables Table 1.1 Ecosystem Services 4 Table 3.1 High-level impacts identified in Tier 1 32 Table 3.2 Tier 2 primary risks 35 Table 3.3 Key ecosystem services provided to other sectors 48 Table 4.1 Estimated golden plover population declines at Snake Summit and probabilities of extinction for a

range of temperature scenarios by 2100 59 Table 4.2 Coastal erosion and coastal defences in the UK 64 Table 4.3 Current coastal erosion and flood risk for National Trust sites in England and Wales 64 Table 4.4 Species currently identified as ‘High Risk’ as defined by the standard risk methodology of the GB

non-native species secretariat 74 Table 4.5 Overview of general results from MONARCH3 based upon projected changes in UKCIP02 83 Table 4.6 The likelihood of European species gaining over 50% new potential suitable climate space from

2020s to the 2080s UKCIP02 HIGH scenarios 84 Table 4.7 The likelihood of extinction of species based upon loss of climate space from 2020s to the 2080s

UKCIP02 HIGH scenarios 85 Table 4.8 Onshore wind energy generation in the UK (MW) in 2010 90 Table 4.9 Current anticipated responses of soil carbon to direct and indirect effects of climate change 95 Table 4.10 McArthur Forest Fire Danger Index applied to UK national parks 116 Table 5.1 Influence of socio-economic drivers on biodiversity risks 123 Table 5.2 Summary and ranking of six future (2050) scenarios for GB based upon four ecosystem services

(‘biodiversity’ values were not monetised) 131 Table 6.1 Summary of economic impacts based upon informed judgement of existing evidence base

(climate change only – central estimate [50% value]; socio-economic factors following current trajectories; current levels of adaptation) 132

Table 6.2 Indicative habitat restoration and creation costs when established through management agreements 141

Table 6.3 Ecosystem service valuations for coastal wetlands 143 Table 6.4 Ecosystem service valuations for inland wetlands 148 Table 6.6 Mid-point estimates of the direct costs of ocean acidification on commercial species (£million per

annum) (Pinnegar et al., 2012) 151 Table 7.1 Physiological and life history traits that may make a species more or less vulnerable or resilient to

climate-related disturbances 156

Figures Figure 1.1 Changes in the status (‘extent’ and ‘quality’) of UK BAP priority habitats: 1999 to 2008 6 Figure 1.2 Influence and trend of different drivers on ecosystem services 7 Figure 1.3 Relative influences of drivers of change on UK broad habitats 8 Figure 2.1 Stages of the CCRA (yellow) and other actions for Government (grey) 18 Figure 2.2 Steps of the CCRA method (that cover stage 3 of the CCRA framework: Assess risks) 21 Figure 2.3 Framework for the confidence levels 25 Figure 3.1 Preliminary vulnerability assessment of UK BAP habitats 36 Figure 3.2 Cross-sectoral linkages with other sectors in the CCRA 49 Figure 4.1 Distribution of presence of lowland beech/yew woodland and blanket bog per 10km grid-square.

For reference annual maximum soil moisture deficits are shown (1971-2000 mean) 52 Figure 4.2 Correlation between beech foliage cover and previous July monthly rainfall 54 Figure 4.3 Correlation between August temperature of 2 years earlier (mean daily maximum temperature)

and golden plover population (log-ratio of change between consecutive years). Negative changes in population are associated with higher temperatures 55

Page 28: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

xxviii Biodiversity Sector Report

Figure 4.4 Indicative maps of suitability for beech under UKCIP02 low and high emissions scenarios for 2020s, 2050s and 2080s 57

Figure 4.5 Future projections of blanket peat distribution using a range of bioclimate envelope models and the UKCIP02 low and high emissions scenario 58

Figure 4.6 Likelihood matrix for grazing marsh grassland 66 Figure 4.7 Change in widely-established invasive non-native species for Great Britain: 1960-2008 73 Figure 4.8 Risk of P.ramorum derived from three different models 77 Figure 4.9 Bird sensitivity map for Scotland at 1km scale 89 Figure 4.10 Current and planned wind farms in the UK 91 Figure 4.11 Conceptual model of changes in soil organic carbon linked to climate 93 Figure 4.12 UK wintering waterbird indicator 98 Figure 4.13 Correlation of swallow arrival time (from 1959 – 2002) with mean temperature for February to

April 99 Figure 4.14 Average minimum temperatures in eastern England and proportions of UK populations of

sanderling and ringed plover overwintering in south west of Britain from 1974-1998 100 Figure 4.15 Populations of butterflies in the UK: 1976-2007 110 Figure 4.16 Farmland bird populations in England 111 Figure 4.17 Outdoor fires in the UK (1995 – 2008) and heatwave years 115 Figure 4.18 FFDI Index for 1980s and 2080s using HadRM3 climate model 116 Figure 5.1 Current and emerging threats faced by priority habitats and species as identified by UK BAP

plans 121 Figure 5.2 Projected trends in ecosystem services for the six scenarios explored by the UK NEA 131 Figure 6.1 Components of total economic value and associated techniques used to elicit values for

biodiversity and ecosystems 136 Figure 6.2 Schematic illustration of the role of monetary valuation within the wider context of quantitative and

qualitative knowledge 139 Figure 6.3 The limits to monetary valuation for complex systems and where multiple values prevail 139 Figure 7.1 Key principles for biodiversity to adapt to climate change 157 Figure 8.1 Cumulative habitat area created in the UK for all completed managed realignment schemes

(orange bars) and net habitat area created (blue bars – accounting for habitat compensation schemes) 160

Page 29: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 1

1 Introduction

1.1 Background

It is widely accepted that the world’s climate is being affected by the increasing anthropogenic emissions of greenhouse gases into the atmosphere. Even if efforts to mitigate these emissions are successful, the Earth is already committed to significant climatic change (IPCC, 2007).

Over the past century, the Earth has warmed by approximately 0.7°C3. Since the mid-1970s, global average temperature increased at an average of around 0.17°C per decade4. UK average temperature increased by 1°C since the mid-1970s (Jenkins et al., 2009), however recent years have been below the long-term trend highlighting the significant year-to-year variability. Due to the time lag between emissions and temperature rise, past emissions are expected to contribute an estimated further 0.2°C increase per decade in global temperatures for the next 2-3 decades (IPCC, 2007), irrespective of mitigation efforts during that time period.

The sorts of impacts expected later in the Century are already being felt in some cases, for example:

Global sea levels rose by 3.3 mm per year (± 0.4 mm) between 1993 and 2007; approximately 30% was due to ocean thermal expansion due to ocean warming and 55% due to melting of land ice. The rise in sea level is slightly faster since the early 1990s than previous decades (Cazenave and Llovel, 2010).

Acidification of the oceans caused by increasing atmospheric Carbon dioxide (CO2) concentrations is likely to have a negative impact on the many marine organisms and there are already signs that this is occurring, e.g. reported loss of shell weight of Antarctic plankton, and a decrease in growth of Great Barrier coral reefs (ISCCC, 2009).

Sea ice is already reducing in extent and coverage. Annual average Arctic sea ice extent has decreased by 3.7% per decade since 1978 (Comiso et al., 2008).

There is evidence that human activity has doubled the risk of a very hot summer occurring in Europe, akin to the 2003 heatwave (Stott et al., 2004).

The main greenhouse gas responsible for recent climate change is CO2 and CO2 emissions from burning fossil fuels have increased by 41% between 1990 and 2008. The rate of increase in emissions has increased between 2000 and 2007 (3.4% per year) compared to the 1990s (1.0% per year) (Le Quéré et al., 2009). At the end of 2009 the global atmospheric concentration of CO2 was 387.2 ppm (Friedlingstein et al., 2010); this high level has not been experienced on earth for at least 650,000 years (IPCC 2007).

3 Global temperature trends 1911-2010 were: HadCRUT3 0.8˚C/century, NCDC 0.7˚C/century,

GISS 0.7˚C/century. Similar values are obtained if we difference the decadal averages 2000-2009 and 1910-1919, or 2000-2009 and 1920-1929. 4 Global temperature trends 1975-2010 were: HadCRUT3 0.16˚C/decade, NCDC

0.17˚C/decade, GISS 0.18˚C/decade.

Page 30: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

2 Biodiversity Sector Report

The UK Government is committed to action to both mitigate and adapt to climate change5 and the Climate Change Act 20086 makes the UK the first country in the world to have a legally binding long-term framework to cut carbon emissions, as well as setting a framework for building the nation’s adaptive capacity.

The Act sets a clear and credible long term framework for the UK to reduce its greenhouse gas (GHG) emissions including:

A legal requirement to reduce emissions by at least 80% below 1990 levels by 2050 and by at least 34% by 2020.

Compliance with a system of five-year carbon budgets, set up to 15 years in advance, to deliver the emissions reductions required to achieve the 2020 and 2050 targets.

In addition it requires the Government to create a framework for building the UK's ability to adapt to climate change and requires Government to:

Carry out a UK wide Climate Change Risk Assessment (CCRA) every five years.

Put in place a National Adaptation Programme for England and reserved matters to address the most pressing climate change risks as soon as possible after every CCRA.

The purpose of this CCRA is to provide underpinning new evidence, assessing the key risks and opportunities to the UK from climate change, and so enable the UK and Devolved Governments to prioritise adaptation options and inform current and future policy development. The CCRA will also inform devolved Governments’ policy on climate change mitigation and adaptation.

Climate Change Act: First 5 year Cycle

The scope of the CCRA covers an assessment of the risks and opportunities to those things which have social, environmental and economic value in the UK, from the current climate and future climate change, in order to help the UK and Devolved Governments identify priorities for action and implement necessary adaptation measures. The CCRA is required to identify, assess, and where possible estimate economic costs of the key climate change risks and opportunities at UK and national (England, Wales, Scotland, Northern Ireland) level. The outputs from the CCRA will also be of value to other public and private sector organisations that have a stake in the sectors covered by the assessment.

The CCRA will be accompanied (in 2012) with a study on the Economics of Climate Resilience7 (ECR) that will consider and, therefore, inform the UK and Devolved Governments about the costs and benefits of options for adaptation to climate change. This analysis will provide an overall indication of the scale of the challenge and potential benefits from acting; and, given the wide-ranging nature of possible interventions, will help to identify priority areas for action by Government on a consistent basis.

This will be followed by statutory adaptation programmes implemented by the UK Government and for Scotland, Wales and Northern Ireland through the DAs. These national adaptation programmes will set out:

objectives in relation to adaptation

5 http://www.defra.gov.uk/environment/climate/government/

6 http://www.legislation.gov.uk/ukpga/2008/27/contents

7 http://www.defra.gov.uk/environment/climate/government/

Page 31: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 3

proposals and policies for meeting those objectives

timescales

an explanation about how those proposals and policies contribute to sustainable development.

The CCRA analysis has been split into eleven sectors to mirror the general sectoral split of climate impacts research: agriculture; biodiversity and ecosystem services; business, industry and services; built environment; energy; floods and coastal erosion; forestry; health; marine and fisheries; transport and water.

1.2 Biodiversity and ecosystem services sector

For the purposes of the CCRA, this sector is defined as covering the biodiversity of terrestrial, freshwater, and coastal environments, with a summary of risks to the marine environment 8. In addition, the role of the natural environment in providing ecosystem services to sustain human social and economic welfare is incorporated. The sector report covers all of the UK but does not include Overseas Territories which are beyond the remit of the CCRA. Biodiversity refers to the variety of all living things: in the context of the sectoral analysis particular emphasis is directed at priority habitats and species as these identify key international responsibilities for the UK.

Following the Convention on Biological Diversity, the 2000 UN Conference on Biodiversity advocated the Ecosystem Approach as the key to holistic natural resource management (also referred to as Ecosystem-Based Management: EBM). Humans and cultural diversity are recognised as integral components of the ecosystem, with the key objective of EBM being to develop ‘integrated management of land, water and living resources that promotes conservation and sustainable use in an equitable way’.

Consistent with the Ecosystem Approach, ‘ecosystem services’ defines a concept that aims to identify the benefits that humans gain from the natural environment, beyond its intrinsic biodiversity value. It therefore provides a framework for considering the cross-sectoral links between this sector and others in the CCRA. Within this report, we have aimed to cover issues relating to supply of these services as provided by different ecosystems; by implication the demand side of ecosystem services is covered within the other sector reports. The CCRA Evidence Report integrates the evidence base around this broad framework.

Four broad categories of ecosystem services can be defined following the convention of the Millennium Ecosystem Assessment (MEA 2003) and as used in the UK National Ecosystem Assessment (UK NEA, 2011): Table 1.1. Provisioning services are the direct products of ecosystems and have been a major focus for human activity in the past. Regulating services act to mediate the interaction and transfer of biota and materials at multiple scales both within and between ecosystems and to humans. Cultural services are linked to the influence and contextual benefits of the environment for human welfare, therefore perceptions often differ among individuals and communities. Cultural issues are tightly bound to human values and behaviour, as well as to human institutions, and patterns of social, economic, and political organization. Underpinning each of these three main service groups are supporting services which are essentially the basic ecosystem functions that are responsible for life.

Biodiversity has a crucial role in maintaining a healthy functioning ecosystem and therefore is intrinsically associated with ecosystem services, although the mechanisms are complex and remain to be fully established. A diverse functioning ecosystem can 8 Marine biodiversity issues were primarily covered by the CCRA Marine Sector report

(Pinnegar et al., 2011).

Page 32: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

4 Biodiversity Sector Report

therefore cope with the loss of some species, unless they are particularly key components9, but the species itself will be unable to persist without the benefits such as nutrients, food, shelter and water provided by the ecosystem.

Table 1.1 Ecosystem Services

Supporting services

Necessary for the delivery of other ecosystem services

Soil formation, Nutrient cycling, Water cycling, Primary production

Provisioning

services

Crops, Livestock,

Game, Fisheries,

Water supply, Wild

species diversity

(genetic resources)

Regulating

services

Climate, Hazards,

Detoxification &

Purification,

Disease/pest

control, Pollination

Cultural

services

Aesthetic, Spiritual,

Inspirational,

Educational,

Recreation,

Tourism, Wild

species diversity

Ecosystem services

The benefits people get from ecosystems

Source: UK National Ecosystem Assessment (2011)

National Ecosystem Assessment (NEA) (UK NEA, 2011)

The NEA provided a comprehensive overview of ecosystem services in the UK, including their current status and trends, together with a future outlook. It explores the drivers of change impacting on ecosystems, and the services which flow from them to deliver a range of goods that we value individually and as a society. Broad habitats (aquatic and terrestrial) were used to provide a high-level framework through which ecosystems were characterised.

The NEA represents the first assessment of its type and acknowledges that there are currently many knowledge gaps and uncertainties. One of the more notable issues is that it was not possible to comprehensively quantify relationships between biodiversity and ecosystem services because of differences in knowledge across taxonomic groups relative to the functions and services they provide. Some services were also better characterised than others in the NEA, depending on the availability of suitable data. Cultural services were particularly challenging in this regard, and this is also attributable to the complex inter-relationships between biodiversity, culture and human well-being.

Nevertheless, the NEA developed an economic framework to provide indicative values for ecosystem services, and complemented this with non-monetary values for health and shared social benefits to represent their broader value to human well-being. One of the key findings was also that across the range of services provided by UK broad habitats, over 30% were assessed as declining, often as a consequence of long-term declines in habitat extent or condition.

9 These key species are sometimes known as ‘ecosystem engineers’ or ‘keystone species’

Page 33: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 5

1.3 Overview of the biodiversity and ecosystem services sector

UK biodiversity includes over 100,000 ‘known’ species and a diverse range of habitats. Much of the conservation interest is now included within a network of designated sites that include EU Natura 2000 sites (Special Areas of Conservation and Special Protected Areas), Ramsar wetland sites, Sites of Special Scientific Interest10, National Nature Reserves, and local designated sites. However, notable examples of biodiversity value exist outwith this network, including Local Nature Reserves. An amendment to the Habitats Directive in 2009 now requires the relevant authorities to report on habitat extent within the wider countryside in addition to Natura 2000 sites.

A particular attribute of this sector is the considerable amount of monitoring data available, providing a key source of information on recent change and variability. Notable examples include the statutory designated site reporting, Countryside Survey programme, Environmental Change Network (ECN), Phenology Network, National Plant Atlas, and citizen-science initiatives such as Nature’s Calendar, the Butterfly Monitoring Scheme, and the long-running bird monitoring initiatives involving the British Trust for Ornithology and other partners.

These observational data show a wide range of change in the natural environment. The largest threat to biodiversity is currently human-induced habitat degradation with the status of many of the priority habitats identified under the UK Biodiversity Action Plan currently described as ‘decreasing’ (Figure 1.1). The Countryside Survey programme has identified land use change and pollution as the major causes of this habitat degradation. As a consequence, climate change is typically acting as a further stress on ecosystems that are already under pressure from other drivers.

Ecosystems are inherently complex, and typically exhibit compound responses to change because of the many system inter-dependencies. This means that they often demonstrate complex (e.g. non-linear) rather than simple responses to change, making interpretation of change difficult. Nevertheless, statistical analysis of data from the ECN, and from recent studies such as BICCO-NET, has been able to separate a clear signal of climate change for some key species (e.g. Morecroft et al., 2009). One of the clearest demonstrations of change has been the shift in timing of key seasonal events such as budburst, flowering, egg-laying, and the arrival of migrant birds. The systematic study of these events (phenology) has collated a strong record of these changes across most parts of the UK.

10

Areas of Special Scientific Interest (ASSI) in Northern Ireland.

Page 34: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

6 Biodiversity Sector Report

Figure 1.1 Changes in the status (‘extent’ and ‘quality’) of UK BAP priority habitats: 1999 to 2008

Most climate change assessments to-date have focussed on the response of individual species or groups of species (particularly birds and butterflies). Rather less information has been derived on species interactions and habitat changes, and very little on ecosystem functions. To project large-scale species responses into the future, the most common approach has been the use of bioclimate envelope models. In addition, small-scale experimental studies have altered conditions (field or laboratory) to understand the effects of single or multiple climatic variables in controlled conditions. Modelling studies of population trends and species interactions have also been developed, despite considerable uncertainties. At the higher level, assessment of habitat vulnerability has tended to be based upon expert opinion, informed by key species’ responses (e.g. Hossell et al., 2000; Mitchell et al., 2007).

With regard to ecosystem services, the NEA used expert opinion to summarise the impact of a range of different drivers of change since 1940 (Figure 1.2). This highlights the increasing impact of climate change but that to-date it has been a lesser influence compared to other drivers, particularly habitat change and pollution. However, there are important geographic variations in this relative influence. Across different habitats the NEA suggested that the more pronounced effects of climate change had been on upland, coastal and marine habitats (Figure 1.3).

Page 35: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 7

Figure 1.2 Influence and trend of different drivers on ecosystem services Source: UK NEA, 2011

Page 36: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

8 Biodiversity Sector Report

Figure 1.3 Relative influences of drivers of change on UK broad habitats Source: UK NEA, 2011

1.4 Policy context

Biodiversity is a devolved matter in the UK, although matters that need a UK overview are coordinated by Defra at UK government level. The key organisations in regulation and delivery of statutory/scientific advice for implementation of biodiversity policy are Natural England, Countryside Council for Wales (CCW), Scottish Natural Heritage (SNH), Council for Nature Conservation and the Countryside (CNCC-Northern Ireland), the Forestry Commission and the Joint Nature Conservation Committee (JNCC). The Environment Agency in England and Wales, the Scottish Environmental Protection Agency in Scotland and the Northern Ireland Environment Agency in Northern Ireland also have regulatory and delivery functions relating to biodiversity.

JNCC is the statutory adviser to the Government on UK and international nature conservation. JNCC delivers the UK and international responsibilities of the four nature conservation agencies (i.e. Natural England, CCW, SNH, CNCC). There are also a significant number of NGOs which can be considered influential stakeholders in this sector including, for example, the Wildlife Trusts, RSPB, WWF, British Trust for Conservation Volunteers, National Trust, Game and Wildlife Conservation Trust, and the Woodland Trust.

The UK Government published “Biodiversity: The UK Action Plan” in 1994 in response to article 6A of the UN Convention of Biological Diversity. Subsequently, devolution has resulted in national biodiversity strategies for each of the four UK countries and, at the UK level, a strategic framework, which sets out a shared vision

Page 37: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 9

and approach to biodiversity conservation. Entitled, “Conserving Biodiversity – the UK Approach”, this framework, published in October 2007, also sets out the shared priorities for UK conservation and indicators to measure progress on biodiversity conservation. At the UK level, there are 1150 species and 65 habitats afforded conservation priority. These are captured in the separate national lists drawn up by the administrations in the four countries. The national biodiversity strategies in each of the four countries are shaped by international commitments. These include the historic agreement reached in 2010 by the parties to the Convention on Biological Diversity in Nagoya, Japan, to set a new global vision and direction for biodiversity policy; the obligations of other international conventions; and the decision of the European Union, also reached in 2010, to halt biodiversity loss and the degradation of ecosystem services. Some of the current conventions, legislation and frameworks that are most relevant to biodiversity in the UK are set out below. The extent to which they refer to climate change (implicit and explicit) varies. Conventions:

The UN Convention on Biological Diversity (Biodiversity Convention or CBD).

The Convention on the Conservation of European Wildlife and Natural Habitats (the Bern Convention).

The Convention on the Conservation of Migratory Species of Wild Animals (Bonn Convention or CMS).

The Convention on Trade in Endangered Species of Wild Flora and Fauna (CITES).

The Convention for the Protection of the Marine Environment of the North East Atlantic (The OSPAR Convention).

The Convention on Wetlands of International Importance especially as Waterfowl Habitat (Ramsar Convention or Wetlands Convention).

The Convention Concerning the Protection of the World Cultural and Natural Heritage (the UNESCO World Heritage Convention).

The European Landscape Convention (Council of Europe, 2000) as ratified by the UK government in 2006.

European Legislation and Policy Frameworks:

Directive 2009/147/EC (Birds Directive). This includes the designation of sites both on land and in coastal environments as Special Protection Areas (SPAs) for birds.

Council Directive 92/43/EEC on the Conservation of natural habitats and of wild fauna and flora (Habitats Directive), 1992. This includes the designation of Special Areas of Conservation (SACs) both on land and in offshore waters. Article 17 of the Habitats Directive also requires a regular cycle of reporting on habitats and species of international importance as defined in Annex 1 of the Directive.

Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 establishing a framework for Community action in the field of water policy' (Water Framework Directive or WFD). This requires member states to implement a series of measures designed to provide for ‘good

Page 38: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

10 Biodiversity Sector Report

ecological status’ in water bodies, and to carry out a programme of River Basin Management Planning.

The Common Agricultural Policy (CAP) provides a framework for support to the agricultural sector and rural communities. Under CAP Pillar I measures provide payments to farmers subject to the cross-compliance condition that land is maintained in ‘good agricultural and ecological condition’. CAP Pillar II measures provide support through agri-environment schemes and rural development programmes. The next round of the CAP (2014-2020) is currently being negotiated amongst member states.

Directive 2008/56/EC on establishing a framework for community action in the field of marine environmental policy - known as the Marine Strategy Framework Directive (MSFD). In particular this requires member states to implement measures to provide for ‘good environmental status’ in marine environments.

Directive 2004/35/EC on environmental liability with regard to the prevention and remedying of environmental damage (Environmental Liability Directive).

Regulation 812/2004 laying down measures concerning incidental catches of cetaceans in fisheries and amending Regulation (EC) No 88/98.

Strategic Environmental Assessment provides a regulatory framework for the statutory requirement to assess environmental impacts of large-scale plans, projects and strategies using a series of criteria to measure against a non-intervention baseline.

UK Legislation:

Wildlife and Countryside Act (1981 as amended). Equivalent for Northern Ireland is held within the Wildlife (Northern Ireland) Order 1985, Nature Conservation and Amenity Lands (Northern Ireland) Order 1985 and the Environment (Northern Ireland) Order 2002. Amendments of varying degrees have been made since 1981, for England and Wales, and for Scotland (Nature Conservation (Scotland) Act 2004), and for Northern Ireland (Wildlife and Natural Environment Act (NI) 2011).

The Offshore Marine Conservation (Natural Habitats, & c.) Regulations 2007.

UK Marine and Coastal Access Act 2009.

England and Wales

The Countryside and Rights of Way Act 2000 (CRoW Act 2000).

The Conservation of Habitats and Species Regulations 2010 (as amended 2011).

The Natural Environment and Rural Communities (NERC) Act (2006) established Natural England as the independent body responsible for conserving, enhancing, and managing the natural environment in England. It also provided the Welsh Government with new legislative powers and made amendments to the both the Wildlife and Countryside Act 1981 and the Countryside and Rights of Way (CROW) Act 2000, which further enhance provisions to biodiversity generally and SSSIs in particular. It also created a duty on all public bodies to have regard for biodiversity

Page 39: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 11

conservation when undertaking their functions; and for lists of species and habitats of principal importance for biodiversity conservation to be published for England and Wales.

Scotland

In Scotland the Habitats Directive is transposed through a combination of the Habitats Regulations 2010 (in relation to reserved matters) and the Conservation (Natural Habitats, &c.) Regulations 1994.

Nature Conservation (Scotland) Act 2004.

Marine (Scotland) Act 2010.

Northern Ireland

Nature Conservation and Amenity Lands (Northern Ireland) Order 1985 (as amended 1989).

Environment (Northern Ireland) Order 2002 (as amended by the Wildlife and Natural Environment Act (NI) 2011).

Wildlife (Northern Ireland) Order 1985, as amended by the Wildlife and Natural Environment Act (NI) 2011).

Conservation (Natural Habitats, etc.) Regulations (Northern Ireland) 1995 (as amended).

Under the Wildlife and Countryside Act 1981 (or Wildlife (Northern Ireland) Order 1985, Nature Conservation and Amenity Lands (Northern Ireland) Order 1985 and the Environment (NI) Order 2002), Sites of Special Scientific Interest (SSSIs) or, in Northern Ireland, Areas of Special Scientific Interest (ASSI), are sites that are identified for their flora, fauna, geological or physiographical features by the country conservation bodies, Natural England, the Countryside Council for Wales, Scottish Natural Heritage and the Council for Nature Conservation and the Countryside.

The Conservation of Habitats and Species Regulations 2010, in England and Wales; the Habitats Regulations 2010 (in relation to reserved matters) and the 1994 Regulations in Scotland and; the Conservation (Natural Habitats) Regulations (Northern Ireland) 1995 (as amended) in Northern Ireland transpose the EU Habitats Directive into National Law. Under these regulations, member states must designate Special Areas of Conservation (SAC) under the EU Habitats Directive and Special Protection Areas (SPA) under the Birds’ Directive. These designations form the larger, Natura 2000, EU conservation network.

Further statutory designations include:

Ramsar Sites are wetlands of international importance designated under the Ramsar Convention.

National Nature Reserves (NNR) - to protect wildlife and geology, and provide great opportunities for people to experience nature.

Local Nature Reserves (LNR) are places which have wildlife or geology of special local interest and are important to people.

Marine Protected Areas (MPA) are designated for the protection of marine biodiversity or natural and cultural resources.

Page 40: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

12 Biodiversity Sector Report

National Parks are designated to both conserve and enhance their natural beauty, wildlife and cultural heritage and to provide opportunities for the public to understand and enjoy these special qualities.

Areas of Outstanding Natural Beauty (AONB) are designated to conserve and enhance the natural beauty of their landscapes.

There are many other possible non-statutory designations including, biogenetic reserves, biosphere reserves, UNESCO World Heritage Sites etc among others. These designations constitute the array of protected sites for nature conservation in the UK.

The Carbon Plan for the UK defines the pathway to a low carbon economy which has important implications for land use change, and therefore indirectly for biodiversity and ecosystem services. To reach the goals in the UK Climate Change Act 2008, the Carbon Plan sets a greenhouse gas emission reduction target of over a third by 2020 and by 80% by 2050(compared to 1990 level emissions). Between 21% and 45% of heat supply to UK buildings will need to be supplied by low carbon sources by 2030.

1.4.1 Responses to climate change

As responsibility for biodiversity policy is devolved, there is a Biodiversity Group and Biodiversity Strategy for each country:

England Biodiversity Group - Biodiversity 2020: A strategy for England’s wildlife and ecosystem services11.

Scotland Biodiversity Committee - Scotland’s biodiversity – it’s in your hands12.

Wales Biodiversity Partnership - Environment Strategy for Wales13.

Northern Ireland Biodiversity Group - Northern Ireland Biodiversity strategy14.

The biodiversity strategies all take account of climate change as one of the most important factors affecting biodiversity and influencing policy development. As evidence to inform policy is limited for many species and habitats, a number of principle-based guidance documents have been written to help implement the biodiversity strategies, including:

Conserving Biodiversity in a Changing Climate: guidance on building capacity to adapt 15 published by Defra in 2007 on behalf of the UK Biodiversity Partnership. The guiding principles described in this document summarise current thinking on how to reduce the impacts of climate change on biodiversity and how to adapt existing plans and projects in the light of climate change.

11

http://www.defra.gov.uk/publications/2011/08/19/pb13583-biodiversity-strategy-2020/ 12

http://www.biodiversityscotland.gov.uk/pageType1.php?id=2&type=1&navID=27 13

http://wales.gov.uk/docs/desh/publications/060517environmentstrategyen.pdf 14

http://www.ni-environment.gov.uk/nibs2002.pdf 15

http://www.ukbap.org.uk/Library/BRIG/CBCCGuidance.pdf

Page 41: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 13

1.4.2 England

The Biodiversity 2020 Strategy for England sets out the ambition, priorities for action and the strategy for delivering them. It seeks to mainstream biodiversity into sectors that can have a significant impact on biodiversity, such as agriculture and land-use planning and development. It also contains a number of outcomes for 2020 which, amongst other things, seek to have 90% of all priority habitats in favourable or recovering condition, create 200,000 hectares of priority habitat, and at least 15% of very poor wildlife sites restored to help adapt for and mitigate against climate change.

The England Biodiversity Group recognise the risk of climate change impacts on biodiversity. In December 2008 they published the ‘England Biodiversity Strategy Climate Change Adaptation Principles: conserving biodiversity in a changing climate.16

England’s soil strategy published by Defra in 2009, Safeguarding our soils - a strategy for England, includes a section on building resilience of soils to the changing climate. Defra have also commissioned research on how soils are likely to be affected by climate change; Modelling the impact of climate change on soils using UK Climate Projections.

Defra commissioned the Lawton Review of England’s wildlife and ecological network to examine evidence on the extent to which designated sites represent a coherent and resilient ecological network capable of adapting to the challenge of climate change and other pressures. The resulting report ‘Making Space for Nature: A Review of England’s Wildlife and Ecological Network’ sets forth the argument that the UK needs a step-change in the approach to wildlife conservation; that long-term, large-scale habitat restoration, under-pinned by the re-establishment of ecological processes and ecosystem services is required; for the benefit of both people and wildlife (Lawton et al., 2010).

In addition, the Adapting to Climate Change Programme required each Government Department to produce Adaptation Plans whilst also acknowledging the importance of the natural environment for adaptation17,18. A key objective is to aim for conservation planning at a landscape scale to protect the best sites, ensure that they are connected to other similar sites and be present in a resilient patchwork of other habitats.

A further context is provided by the Government’s Structural Reform Plan for Defra19, which identified three key priorities, all of which are relevant to climate change and biodiversity, but one of which is particularly focussed on biodiversity – to “enhance and protect the natural environment, including biodiversity and the marine environment, by reducing pollution and preventing habitat loss and degradation”. Two of the underlying supporting actions identified under the heading of ‘helping communities and wildlife adapt to climate change’ are:

To assess the scope for actions to offset the impact of development on biodiversity.

Publication of the Natural Environment White Paper (NEWP), setting out measures to protect wildlife and promote green spaces and wildlife corridors. The NEWP seeks to build on the recently published National Ecosystem Assessment, and takes into account the UNEP report on ‘The

16

http://www.defra.gov.uk/environment/biodiversity/documents/ebs-ccap.pdf 17

http://www.defra.gov.uk/environment/climate/documents/taking-action.pdf 18

http://www.defra.gov.uk/environment/climate/documents/natural-environment-adaptation.pdf 19

http://ww2.defra.gov.uk/about/our-priorities/

Page 42: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

14 Biodiversity Sector Report

Economics of Ecosystems and Biodiversity study’ (TEEB) and the Nagoya agreement to coordinate biodiversity actions at international level. Some of the main actions are:

- Publication of the new Biodiversity Strategy for England (as mentioned above)

- Nature Improvement Areas. This takes forward the recommendation in the Lawton review for local partnerships to plan and deliver significant improvements for wildlife and people by sustainably using natural resources, restoring and creating wildlife habitats, connecting local sites and joining action across a large discrete area.

- Biodiversity Offsets. These are conservation activities designed to deliver biodiversity benefits in compensation for losses in a measurable way. Good developments incorporate biodiversity considerations in to their design but are still likely to result in some biodiversity loss. One way to compensate for this loss is by offsetting: the developer secures compensatory habitat expansion or restoration elsewhere.

- Natural Capital Committee. This will be established jointly with HM Treasury to advise Government on the state of English natural capital, and in particular when, where and how natural assets are being used unsustainably.

- Local Green Space Designation. A new local green space designation will allow people to protect the green areas that are important to them.

1.4.3 Wales

Maintaining Wales’ distinctive biodiversity is a key tenet of the Welsh Government’s Environment Strategy for Wales. The Strategy aims to provide a thriving, distinctive environment contributing to the economic and social wellbeing of Wales, and the health of its people by 2026. The strategy is split into themes:

Addressing climate change (addresses both mitigation and adaptation)

Sustainable resource use (includes soils and water)

Distinctive biodiversity, landscapes and seascapes

Our local environment (covers urban environment and public access)

Environmental hazards (principally pollution).

The Strategy is supported by action plans, which set out how the Strategy will be delivered. The latest, Environment Strategy Action Plan (2008-2011)20 sets out the Welsh Government‘s commitment to halting biodiversity loss and to seeing a definite recovery from the losses that have already occurred. The Strategy and its Action Plan are further supported by the Biodiversity Framework21, developed by the Wales Biodiversity Partnership. All of these strategic documents highlight the importance on adapting to and mitigating climate change impacts upon biodiversity and the wider environment.

20

http://wales.gov.uk/topics/environmentcountryside/epq/envstratforwales/actionplans/2ndactionplan 21

http://www.biodiversitywales.org.uk/legislation__guidance-20.aspx

Page 43: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 15

Following a public consultation in 2010, the Welsh Government produced a Natural Environment Framework – A Living Wales22, which adopts an integrated ecosystems approach to management of the natural environment. Climate change is identified as one of the four key drivers for developing the new Framework. The Framework aims to embed sustainable development as an objective in all policy and policy delivery, so that people in Wales can continue to benefit from ecosystem services in the face of climate change and other pressures such as those from human activities.

1.4.4 Scotland

Scotland’s biodiversity – it’s in your hands strategy document aims to conserve biodiversity for the health, enjoyment and wellbeing of the people of Scotland now and in the future. The key theme of the strategy is to reinforce the link between people and biodiversity. The strategy follows five ecosystem types: freshwater and wetland, lowland and farmland, marine and coastal, upland and woodlands. The strategy is supported by an indicators report to measure progress against the themes and action plans to implement the strategy. More recently, Scotland’s Climate Change Adaptation Framework identified further pressures on the ecosystem types from climate change. Areas for action to assist adaptation have already been identified in SNH’s Climate change and natural heritage - SNH's approach and action plan.

Scottish Natural Heritage has also produced “Applying an Ecosystem Approach in Scotland: A framework for Action”. A framework for delivering biodiversity actions at a larger scale, that also allows the needs of people to be incorporated as well. In addition, “Valuing our Environment: The Economic Impact of Scotland’s Natural Environment” has also been written, linking the importance of the natural environment to Scotland’s economy and highlighting the need to use it sustainably to secure Scotland’s future. The Soil Framework for Scotland (2001) and Scotland’s Soil Resource – Current State and Threats (2011) cover climate change impacts on soils and associated biodiversity. The National Planning Framework 2 supports the development of a National Ecological Network. The Land Use Strategy for Scotland ‘Getting the best from our Land’ provides a high-level framework for integrating these policy commitments within a series of common principles and proposals.

1.4.5 Northern Ireland

The Northern Ireland Biodiversity strategy document aims to create sustainable social progress, high and stable levels of economic growth and employment, and effective protection of the environment and prudent use of natural resources. The strategy set out how to deal with the 76 issues to biodiversity as set out by the Northern Ireland Biodiversity Group (NIBG) document Biodiversity in Northern Ireland: Recommendations to Government for a Biodiversity Strategy (2000)23 under the headings: agricultural systems and support; coastal and marine; management of freshwater use and management; construction and development; tourism and recreation; peatland management; and, introduced species and genetic material.

The second report of the NIBG, Delivery of the Northern Ireland Biodiversity Sector, reviews the progress made on the issues identified and responded to in the 2002 strategy. It also sets out the evidence for the impacts of climate change on Northern Ireland’s biodiversity.

22

http://wales.gov.uk/topics/environmentcountryside/consmanagement/nef 23

http://www.doeni.gov.uk/niea/biodiversityinnisection1.pdf

Page 44: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

16 Biodiversity Sector Report

The Wildlife and Natural Environment Act (NI) 2011 introduced a new statutory duty on public bodies to take action to further the conservation of biodiversity. This supplements the commitment to protecting and enhancing biodiversity contained in the Northern Ireland Sustainable Development Strategy and aims to raise the profile and visibility of biodiversity as a natural part of decision-making across the public sector in Northern Ireland.

NI and Ireland are co-operating on a project looking at invasive non-native species across the whole island building upon the NIEA Position Statement on Invasive Alien Species (2010)24.

1.4.6 Current initiatives that may influence future policy in the biodiversity sector

In the last decade there has been a shift towards a more integrated approach to thinking about biodiversity issues, whether it is through consideration of green infrastructure, landscape-scale approaches or ecosystem services. The evidence to support these approaches to policy has been limited to-date but is now being influenced by new developments including:

The UK National Ecosystem Assessment (see Section 1.2).

Natural England has developed the concept of Future Landscapes to help identify the likely impacts of climate change on landscape character. This is useful in helping to inform future policy and interventions within an integrated landscape approach, allowing ecosystem services provided by landscapes to be maintained. NGOs such as the RSPB (Futurescapes) and The Wildlife Trusts (Living Landscapes) promote similar initiatives.

Following the recommendations in the Review of Non-Native Species Policy (Defra, 2003), a GB Programme Board was established in March 2005 to deliver strategic consideration of the threat of invasive non-native species. It is supported by the independent Non-Native Species Secretariat. The Invasive Non-Native Species Framework Strategy for Great Britain (2008) provides the structure for key actions to address invasive non-native species problems, including some consideration of climate change25, although they suggest that in the longer term, further debate will be necessary from both a policy and a science perspective on the issue of colonisation by non-native species driven by climate change. The distinction between human-introduced “invasive” species and natural spread of species is important because the latter are intended to be incorporated within adaptation policy responses for ‘native’ biodiversity.

1.5 Structure of this report

This report describes the steps taken in the Biodiversity and Ecosystem Services Sector analysis. These steps include:

The generic CCRA methodology as adapted and applied for this sector.

Scoping and general characterisation of impacts (‘Tier 1’ assessment).

24

http://www.doeni.gov.uk/niea/niea_position_statement_on_invasive_alien_species.pdf 25

The Strategy follows the Convention on Biological Diversity by defining non-native invasive species as those that are introduced by human agency rather than natural movements.

Page 45: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 17

Identification of the most important impacts (the ‘Tier 2’ impacts).

For each Tier 2 impact, analysis of the current and future risk, including reference to ‘risk metrics’ that can allow further characterisation.

Evaluation of the impacts of climate change for selected climate change scenarios.

An assessment of the additional adaptation measures that may be required beyond those currently in process (the ‘adaptation deficit’).

Characterisation of the non-climate socio-economic components of change for the Tier 2 risks.

An assessment of organisational Adaptive Capacity in the context of its ability to reduce the severity of impacts. For this sector, this is also linked to principles to enhance natural adaptive capacity.

The report structure broadly follows the risk assessment steps as described in detail in the CCRA Method Report (Defra, 2010b).

Page 46: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

18 Biodiversity Sector Report

2 Methods

2.1 Introduction: CCRA framework

The overall aim of the CCRA is to inform UK adaptation policy, by assessing the main current and future risks (threats and opportunities) posed by the current climate and future climate change for the UK to the year 2100. The overall approach to the risk assessment and subsequent adaptation plan is based on the UK Climate Impacts Programme (UKCIP) Risk and Uncertainty Framework (UKCIP, 2003). The framework comprises eight stages as shown in Figure 2.1. The CCRA has undertaken the Stages 1, 2 and 3 as outlined below. Stages 4 and 5 will be addressed as part of a separate economic assessment, entitled the ‘Economics of Climate Resilience’, and the remaining stages will be implemented by the UK Government and Devolved Administrations. The framework presents a continual process that can adapt as new evidence and policy emerges; in the case of the CCRA the process will be revisited every five years.

Figure 2.1 Stages of the CCRA (yellow) and other actions for Government (grey)

Adapted from UKCIP (2003)

Stage 1 is defined by the aim of the CCRA project, to undertake an assessment of the main risks (including both threats and opportunities) posed by climate change that will have social, environmental and economic consequences for the UK.

Stage 2 established decision-making criteria for the study, which were used to inform the selection of impacts for analysis in Stage 3. These criteria are the social, environmental and economic magnitude of consequences and the urgency of taking adaptation action for UK society as a whole.

Stage 3 covers the risk assessment process. This involved a tiered assessment of risks with Tier 1 (broad level) identifying a broad range of potential impacts and Tier 2 (detailed level) providing a more detailed analysis including quantification and monetisation of some impacts. A list of climate change impacts was developed based on eleven sectors with further impacts added to cover cross-cutting issues and impacts which fell

Page 47: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 19

between sectors. This list of climate change impacts is referred to as the ‘Tier 1’ list of impacts. This list contained over 700 impacts – too many to analyse in detail as part of this first CCRA. A consolidated list of the highest priority climate change impacts for analysis was developed and referred to as the ‘Tier 2 list of impacts’. This report presents the risk assessment for Tier 2 impacts.

The background to the framework and the approach used for each of the first three stages is set out in more detail in the CCRA Method Report (Defra, 2010b). This chapter aims to summarise the CCRA method for the risk assessment stage (Stage 3 in the framework above) because this includes the specific steps for which results are presented in this report.

2.2 Outline of the method used to assess impacts, consequences and risks

The risk assessment presented in this report is the focus of Stage 3 in the CCRA Framework (see Figure 2.1). This was done through a series of steps as set out in Figure 2.2. These steps are explained in sections 2.3 - 2.7 below and are discussed in more detail in the CCRA Method report (Defra, 2010b).

The components of the assessment sought to:

Identify and characterise the impacts of climate change

This was achieved by developing the Tier 1 list of impacts, which included impacts across eleven sectors as well as impacts not covered by the sectors and arising from cross sector links (see Chapter 3 of this report).

Identify the main risks for closer analysis

This involved the selection of Tier 2 impacts for further analysis from the long list of impacts in Tier 1. Higher priority impacts were selected by stakeholder groups based on the social, environmental and economic magnitude of impacts and the urgency of taking action (see Chapter 3 of this report and Section 2.5 below).

Assess current and future risk, using climate projections and considering socio-economic factors

The risk assessment was done by developing response functions that provide a relationship between changes in climate with specific consequences based on analysis of historic data, the use of models or expert elicitation. In many cases in this sector, this was not possible, and a narrative approach was taken instead. Where response functions were possible, the UKCP09 climate projections and other climate models were then applied to assess future risks. The potential impact of changes in future society and the economy was also considered to understand the combined effects for future scenarios. (See Chapters 4 to 6 of this report and Section 2.6 below.)

Assess vulnerability of the UK as a whole

This involved:

i. a high level review of Government policy on climate change in the eleven sectors (see Chapter 1 of this report)

Page 48: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

20 Biodiversity Sector Report

ii. a high level assessment of the social vulnerability to the climate change impacts – this was not applied to Biodiversity and therefore is not included for this sector

iii. a high level assessment of the adaptive capacity of the sectors (see Chapter 7 of this report and Section 2.11 for an overview of the approach, below).

Report on risks to inform action

This report presents the results of the risk assessment for the Biodiversity and Ecosystem Services sector. The results for the other ten sectors are presented in similar reports and the CCRA Evidence Report (CCRA, 2012) draws together the main findings from the whole project, including consideration of cross-linkages, and outlines the risks to the UK as a whole.

Page 49: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 21

2.3 Identify and characterise the impacts

Step 1 – Literature review and Tier 1 analysis

This step scoped the potential impacts of climate change on the UK based on existing evidence and collating the findings from literature reviews, stakeholder participation through workshops, correspondence with wider stakeholders and soliciting expert opinion. This work developed the Tier 1 list of impacts (see Appendix 2). The Tier 1 impacts have not been analysed in detail; high level discussion of these impacts is provided in Chapter 3 of this report.

Figure 2.2 Steps of the CCRA method (that cover stage 3 of the CCRA framework: Assess risks)

Page 50: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

22 Biodiversity Sector Report

Step 2 – Cross sectoral and indirect impacts

The Tier 1 lists for the eleven sectors in CCRA were compared and developed further to elaborate cross-sectoral and indirect impacts. This was done by ‘Systematic Mapping’, which sets out a flow chart to link causes and effects in a logical process.

2.4 Assess vulnerability

Step 3 – Review of policy

Government policy on climate change develops and changes rapidly to keep pace with emerging science and understanding of how to respond through mitigation and adaptation. This report includes an overview of selected relevant policy in Chapter 1 as this provides important context for understanding how risks that are influenced by climate relate to existing policies. This information will be expanded in the Economics of Climate Resilience project and the National Adaptation Programme.

Step 4 – Social vulnerability

The concept of ‘ecosystem services’, which recognises goods and services from the natural environment that sustain human well-being, was incorporated into the approach for this sector to indicate the wider social and economic implications of change. Risks to biodiversity identified in Tier 2 are therefore also summarised with regard to the implications for ecosystem services, which also identifies key cross-sectoral links for further appraisal.

Some groups within society are more vulnerable to loss of ecosystem services than others. However, these relationships are typically multifaceted, particularly with regard to cultural services which depend on individuals rather than society as a whole. Therefore the evidence base is rather limited. Findings from the UK National Ecosystem Assessment are included where relevant.

Step 5 – Adaptive capacity

The adaptive capacity of a sector is the ability of the sector as a whole, including the organisations involved in working in the sector, to devise and implement effective adaptation strategies in response to information about potential future climate impacts.

For the Biodiversity and Ecosystem Services sector the concept was also taken to include ‘natural’ adaptive capacity, through which ecosystems have intrinsic properties to adjust to change and therefore have in-built resilience. This natural capacity and resilience has often been reduced due to factors such as land use change and pollution, which can increase detrimental impacts from climate change and have negative consequences for ecosystem services.

2.5 Identify the main risks

Step 6 – Selection of Tier 2 impacts

The Tier 1 list of impacts for each sector that resulted from Step 1 and Step 2 (see above) was consolidated to select the higher priority impacts for analysis in Tier 2. This prioritisation process was informed by a simple multi-criteria assessment based on the following criteria:

Page 51: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 23

the social, economic and environmental magnitude of impacts; for this sector, ‘Biodiversity’ consequences provided the environmental magnitude and ‘Ecosystem Services’ consequences provided the social and economic magnitude26.

likelihood of occurrence, including confidence in the evidence base.

the urgency with which adaptation decisions needs to be taken.

Each of these criteria were allocated a score of 1 (low), 2 (medium) or 3 (high) and the impacts with highest scores over all criteria were selected for Tier 2 analysis. The scoring for each sector was carried out based on expert judgement and feedback from expert consultation workshops (or telephone interviews). Checks were carried out to ensure that a consistent approach was taken across all the sectors. The results of the scoring process are provided in Appendix 2.

Step 7 – Identifying risk metrics

For each impact in the Tier 2 list, the generic CCRA method aims to identify one or more risk metrics. Risk metrics provide a measure of the consequences of climate change, related to specific climate variables or biophysical impacts. For this sector, although candidate risk metrics were identified, this approach was qualified by providing a broader characterisation of risk. Although considered a useful concept, risk metrics can potentially be misleading for the following reasons:

The response of one species or habitat is not necessarily indicative of others or of the whole ecosystem.

Most cause-effect relationships in the natural world involve many interdependencies, and are often non-linear through time, and highly variable in a geographic context.

Nevertheless, the risk metric approach can be useful by providing indicative information of potential future responses under a range of assumptions. Usually these assumptions are necessary because the complexity of the natural world means we have incomplete knowledge on the current inter-relationships within biodiversity and their links with ecosystem services. For these reasons, comparison between models and field evidence often shows significant differences because the models have made general assumptions that may not be appropriate for that field site.

2.6 Assess current and future risk

Step 8 – Response functions

This step aimed to establish how each risk metric varied with one or more climate variables using available data or previous modelling work. This was done by developing a response function, which is a relationship to show how the risk metric varies with change in climate variables. A more qualified approach was applied for this sector because of the problems in identifying risk metrics that could account for complex inter-relationships and the presence of a range of other drivers (e.g. land use change) that also account for the change in observational datasets.

26

Consequences were assessed across a range of Broad Habitats. For Biodiversity the score was based upon consequences for priority habitats and species, with the highest score taken to account for the concentration of biodiversity value in certain environments. For Ecosystem services the scoring was based upon the sensitivity of the wider ecosystem to change and an average value was used to allow for the more general flow of services to humans.

Page 52: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

24 Biodiversity Sector Report

Step 9 – Estimates of changes in selected climate change scenarios

This was generally based upon a broader review of literature rather than response functions because of the difficulties outlined above (Step 7 and Step 8).

Step 10 – Socio-economic change

It is recognised that many of the risk metrics in CCRA are influenced by a wide range of drivers, not just by climate change. The way in which the social and economic future of the UK develops will influence the risk metrics. Growth in population is one of the major drivers in influencing risk metrics and may result in much larger changes than if the present day population is assumed.

For all of the sectors, a broad consideration has been made of how different changes in our society and economy may influence future risks and opportunities. The dimensions of socio-economic change that were considered are:

Population needs/demands (high/low)

Global stability (high/low)

Distribution of wealth (even/uneven)

Consumer driven values and wealth (sustainable/unsustainable)

Level of Government decision making (local/national)

Land use change/management (high/low Government input).

The full details of these dimensions and the assessment of the influence they have on the Biodiversity and Ecosystem Services sector is provided in Chapter 5. Note that this step is different from Step 4, which considers how the risks may affect society; whereas this step considers how changes in society may affect the risks.

Step 11 – Economic impacts

Based on standard investment appraisal approaches (HM Treasury, 2003) and existing evidence, some of the risks were expressed as monetary values. This provides a broad estimate of the costs associated with the risks and is presented in Chapter 6 of this report. A more detailed analysis of the costs of climate change will be carried out in a study on the Economics of Climate Resilience27.

2.7 Report on risks

Step 12 – Report outputs

The main report outputs from the work carried out for the CCRA are:

The eleven sector reports (this is the sector report for the Biodiversity and Ecosystem Services sector), which present the overview of impacts developed from Tier 1 and the detailed risk analysis carried out in Tier 2.

The Evidence Report, which draws together the main findings from all the sectors into a smaller number of overarching themes.

Reports for the Devolved Administrations for Scotland, Wales and Northern Ireland to provide conclusions that are relevant to their country.

27

http://www.defra.gov.uk/environment/climate/government/

Page 53: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 25

2.8 Confidence levels

For the Tier 2 assessment the standard CCRA framework has been used to assess confidence across the suite of risks (Figure 2.3). This combines the quality of the evidence base for each risk with the level of consensus that the evidence provides.

Figure 2.3 Framework for the confidence levels

Page 54: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

26 Biodiversity Sector Report

3 Impacts and risk characterisation

In this chapter we provide a broad high-level characterisation of climate change impacts (Tier 1) in advance of a more detailed assessment of priority risks in Chapter 4. Evidence suggests that climate change is already having a significant impact on biodiversity as has been highlighted by a series of studies over recent years (e.g. Mitchell et al., 2007; Walmsley et al., 2007; Morecroft et al., 2009). For example, recent analysis in the BICCO-NET project has demonstrated how populations of some birds and aphid species have largely increased in response to higher temperatures, whilst beetles, butterflies and moths showed more complex responses, apparently benefiting from warmer summers but tending to decline after warmer winters. The more detailed data available for birds has indicated that population growth in species was promoted by increases in annual temperature, reductions in winter severity and improved conditions in the breeding season. However, the negative effects of summer warming suggest that ground feeding birds and long-distance migrants may be vulnerable to future change, possibly related to impacts on some prey populations.

Despite the considerable amount of observational data for biodiversity, there are often difficulties in establishing a clear ‘cause-effect’ relationship for climate drivers. This is a consequence of the complexity of ecosystems and also because of the concurrent pressures on biodiversity from other stressors such as land use intensification and pollution.

Land use intensification has occurred in both agriculture and forestry as a response to demands for improved yields and productivity. As emphasised by the recent Lawton report (Lawton et al., 2010), it has led to habitat loss and fragmentation with the remaining areas of semi-natural habitat often left as isolated portions within a wider landscape that is more hostile to biodiversity. These fragments are those usually covered by the ecological network of designated sites which afford a degree of protection against undesirable change. The negative impacts of pollution are probably now most starkly exemplified by the effects of eutrophication due to nitrogen enrichment from runoff of agricultural fertilisers and atmospheric deposition sources (originally from burning of fossil fuels and intensive livestock units). A series of studies and the Countryside Survey programme have demonstrated large-scale floristic change across the UK with a shift towards faster-growing plants at the expense of plants that prefer more nutrient-poor conditions. The widespread dispersal of atmospheric pollution means that impacts have occurred on designated sites in addition to the wider countryside.

3.1 High-level identification of climate impacts (Tier 1)

A wide range of high-level impacts were identified at a level above those that might be noted for specific habitats and species (Table 3.1). These include both the impacts of incremental changes in climate (including sea level rise) and also the more abrupt changes due to extreme events, such as drought, flood, or storm surge. This distinction is important for adaptation because extreme events in particular can combine circumstances to result in more abrupt change to ecosystems. This has implications not only for biodiversity but also the many benefits that humans derive from the natural environment.

Page 55: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 27

In Table 3.1 impacts have been grouped from species level (range shifts, phenology) up to species interactions and habitat change, and then to fundamental changes to ecosystem functioning. In general we know more about the responses of species (particular well-monitored species such as birds and butterflies) than for their interactions. Similarly we usually know proportionately less about change with regard to habitats and ecosystems. In the sections below (3.1.1 to 3.1.8), each of the impacts is introduced according to these groupings.

In addition to the direct effects of climate, a series of indirect impacts have also been distinguished in Table 3.1: these can occur through the positive or negative consequences of adaptive or non-adaptive responses in other sectors. This highlights some of the many interdependencies between this sector and others in the CCRA. These indirect effects are particularly apparent through the impacts of land use change or pollution.

This stage of the CCRA process does not necessarily distinguish between positive and negative aspects and hence the impact may often be characterised just in terms of ‘change’. This has the advantage that it does not assume a particular outcome which is often dependent on an array of other circumstances, and may therefore be highly uncertain. Where possible the outcome is inferred as ‘positive’ or ‘negative’ with respect to these inter-dependencies.

3.1.1 Range shifts

The distribution of species is associated with key bioclimatic factors that define its geographical range. With a changing climate, the range of these climatic factors will be altered which implies that a species may need to move to remain in the same bioclimatic zone unless it can adapt in situ. However, human modification of the landscape means that species may be unable to track their changing ‘climate space’. This phenomena may also be apparent on a smaller scale as species often occupy a particular niche favoured by certain microclimatic conditions (e.g. wetness) or that exclude competitors (e.g. cold or windy areas). Climate change therefore means that some species may be unable to find suitable microclimate. This can be particularly apparent in mountain areas where one natural response to a warming climate would be for species to move to higher altitudes to maintain the same bioclimatic regime. For those species that are already at the highest elevations (e.g. arctic-alpine species), this is not feasible, meaning that they would be forced to adapt in situ unless there is a managed intervention. However, a changing climate is not all bad news: it also means that there may be opportunities for new species in addition to being advantageous for some existing priority species, and this may enhance the biodiversity of the UK.

There is good observational evidence from some species that shows these impacts are already occurring.

3.1.2 Seasonal shifts and changes in phenology

In addition to the geographical changes, climate change brings temporal change, particularly through modification of the seasons. Species use environmental cues to synchronise their life and breeding cycles with favourable conditions for food and shelter. The timing of key events (e.g. budburst, flowering, egg laying, seed production, leaf changes) provides the basis for the study of phenology. A shift to earlier springs or later autumns implies a change in phenology, with changes in life cycles for many species, particularly insects because of their shorter cycles. In addition, as species have adapted to use different environmental cues, these may lead to asynchrony

Page 56: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

28 Biodiversity Sector Report

between a species breeding cycle and its food supply. A further consequence of the changing seasons is changes in species migration patterns, as migration has developed as an evolutionary adaptive response to a changing food supply through the seasons. With the availability of food now changing, then it can be expected that these adaptive responses will again be modified, particularly through species behaviour and natural selection. Migration patterns may therefore be modified in terms of their range (long-range or short-range migrants), favoured stop-over sites, and even with regard to whether a species actually migrates or not.

There is good observational evidence from some species that shows these impacts are already occurring, although interpretation is often difficult because of the complexity of species and environment interactions.

3.1.3 Changes in pests and diseases

Changes in the geographic range of species may also have consequences in terms of the greater prevalence of pests and diseases. These may have previously been constrained from persisting or spreading through the UK, despite their original introduction, due to the severity of the climate – particularly cold winters. Impacts may be particularly severe through an increased risk from invasive non-native species, as defined by those non-native (alien) species that pervasively enter and modify local ecosystems to the detriment of native biodiversity. In addition, a potential ‘wildcard’ impact is the increased risk from novel pathogens, which result from new recombinant variations of existing pathogens into a more virulent strain: a notable example is the recent H1N1 ‘bird flu’ virus. Invasive non-native species have the potential to seriously modify ecosystem functioning and hence could have implications for the services we obtain from ecosystems, such as clean water and pollination.

Observational evidence suggests that the record of entry of invasive non-native species is increasing, but this may be partly due to greater vigilance, in addition to an expansion in global trade and travel. Novel pathogens are effectively unpredictable due to the randomness of recombination between pathogens but the changing climate combined with other factors (e.g. increased globalisation) suggests that the potential for recombination is increasing which increases the background risk of a virulent strain developing.

3.1.4 Changes in species interactions and community structure

An inevitable consequence of differences in adaptive responses between species to range shifts and phenological change is that species interactions and communities of species will be modified. It seems likely that some species that have very specific niches and habitat preferences will be disadvantaged as environment conditions are modified, because they have a limited choice of alternative sites. This suggests that ‘generalist’ species may be favoured over ‘specialist’ species, as they are more adaptable to a wider range of conditions. In addition, changing species interactions due to changes in growth and survival rates may occur due to one species being better able to adapt to a changing food supply. In terms of food webs this would lead to changing interactions between trophic levels (i.e. across different steps of the food web) due to the alteration of predator-prey relationships. The structure of ecological communities may also be modified due to the impacts of a changing nutrient supply, as the changing climate interacts with changes in soil physical, chemical and biological processes to alter the availability of key nutrients required to regulate plant growth and soil microbes.

Page 57: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 29

It also seems likely that climate change will cause changes in genetic diversity of species, which will modify these interactions, as change inevitably modifies the factors that produce genetic combination and mutation. Another ‘wildcard’ impact would be the consequences of changing competition between plants that use C3 and C4 photosynthesis (C3 and C4 represent different structures of biomass carbon produced during photosynthesis). Most plants use a C3 structure but some have evolved to use a C4 structure (e.g. some members of the sedge family – Cyperaceae) which is more efficient in terms of water and nutrient use, but is also sensitive to the CO2 concentration of the atmosphere. As each of these factors change, then the interaction between the different plant groups would also change.

Observational evidence shows that some aspects of species interaction are changing, notably between generalists and specialists, and also through ‘phenological mismatch’ in which predator-prey interactions are modified by seasonal changes. This is often complicated by the influence of other drivers such as land-use change. However, for other impacts, particularly ‘wildcards’ there is much less observational evidence available and the main source of evidence is theory and models.

3.1.5 Geomorphological and hydroecological habitat change

The impacts of climate change are also manifest through land surface and hydrological processes that modify the abiotic and biotic components of ecosystems. Reductions in summer precipitation and increases in evaporation can be expected to lead to increased soil moisture deficits and drying with implications for the many semi-natural habitats in the UK that rely on reasonably high water tables all year round. Drier summers can also result in negative issues for aquatic habits due to low flow impacts and a reduction in dissolved oxygen supply for fish and other species due to higher temperatures. A related impact is that increased water temperature and stratification of water bodies may alter the supply of nutrients in addition to dissolved gases, with consequences for the distribution of many species, including the key role of plankton in the food chain. Conversely, an increase in winter precipitation would be likely to lead to increased waterlogging which may be a problem for habitats that require drier soils. An increase in quantity and intensity of rainfall may also be associated with increased soil erosion, particularly on areas of bare ground. The increase in sediment load and river-bed scouring can be responsible for the related impacts of high flow on spawning beds in aquatic habitats. In montane environments, a reduction in snow cover from milder winters would have negative implications for those species that use snow beds for food supply or camouflage.

Climate-related alteration of the flow and sediment regimes of rivers will act to modify to floodplain evolution with implications for associated aquatic and riparian habitats that occupy different components of the river and its floodplain. Similarly, changes in sea level and wave regime are associated with the changing evolution of coastal habitats, such as salt marshes and sand dunes, that each occupies a particular section of the littoral zone based upon sediment type, salinity and wave energy. This change may also be associated with saline intrusion of freshwater or brackish habitats. In both fluvial and coastal environments, change is also often heavily influenced by previous management interventions.

Floodplain, upland and coastal ecosystems provide an important regulating service by buffering against hazards such as flooding and erosion. Terrestrial systems also help maintain a reliable clean water supply. Change therefore has important implications for the sustainability of these services.

There is usually good observational evidence of these impacts, which become particularly apparent in anomalous years (e.g. wet or dry years), although knowledge of

Page 58: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

30 Biodiversity Sector Report

cause-effect relationships is usually complicated by other factors such as land use change and pollution. Knowledge of evolutionary changes in floodplain or coastal zones is often constrained to locations with good long-term datasets.

3.1.6 Habitat disturbance by extreme events

Ecosystems are particularly susceptible to the impacts of large-scale extreme events that may require long recovery times or even induce irreversible change. These extreme events are not always negative and can be beneficial in creating new habitat, particularly for pioneer species. However, for species that have very specific habitat requirements, this can potentially lead to a significant loss or gain of niche space for a variety of reasons. One example is wildfire risk which can result in major losses of priority habitats that have high biodiversity value. Another impact occurs from windthrow during storms which may have important consequences for priority woodland habitats. Extreme events can also aggregate a series of climate, geomorphological and hydroecological effects to result in major and abrupt change. Major drought events can therefore significantly exacerbate the soil moisture deficit and low flow impacts highlighted earlier (Section 3.1.5) and a major fluvial flood would exacerbate the high flow and waterlogging impacts (Section 3.1.5). Similarly, a major coastal flood or reconfiguration (due to managed intervention) would alter the large-scale distribution of freshwater, terrestrial, coastal margin and marine habitats in the affected area.

Extreme events occur sporadically, therefore our observational knowledge of them remains limited and restricted to specific locations. Much of the information about potential impacts is therefore derived from models developed to test sensitivity to different drivers.

3.1.7 Changes in ecosystem processes and functioning

The functioning of ecosystems is regulated through the combination of many different physical, chemical and biological processes. These processes are influenced by temperature and moisture availability; therefore they are also highly likely to be modified by climate change. Of fundamental significance here are changes in primary productivity, which is the production of biomass in the ecosystem by photosynthesis: primary productivity is likely to increase with higher temperatures and increased CO2 levels. This is related to changes in soil organic matter (carbon) which acts as a key ecosystem regulator, notably for the supply of nutrients. A warmer climate implies changes in soil microbial activity as the microbes are strongly influenced by temperature and moisture. This in turn is also associated with faster decomposition and changes in nutrient cycling as biomass will be broken down at a faster rate if sufficient moisture and air is available (i.e. soils are not too dry or waterlogged).

These changes will be critical for all components of biodiversity and, because they provide the supporting functions that underpin the flow of ecosystem services to humans, also have high relevance for human welfare.

Our knowledge of ecosystem functioning is very limited. We have good knowledge of some ecosystem processes, particularly from experimental sources, but our understanding of how processes interact remains limited as experiments are run in controlled conditions. Hence although there is often data available on key indicators of primary productivity or soil organic carbon, interpretation of current and future rates of change remains difficult. A key uncertainty is how the climate variables will interact with changing CO2 levels.

Page 59: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 31

3.1.8 Changes in indirect impacts

The direct impacts of climate variables on ecosystems can also be accompanied by more indirect effects that occur through adaptive or non-adaptive responses in other sectors, which may be either planned or unplanned (reactive). This is particularly exemplified by interactions between biodiversity and agriculture in the UK, because climate acts as a major driver for both sectors. In many locations, agricultural activities have (or have been) limited by climate restrictions such as a short growing season or soil wetness that will become a lesser constraint in a warming climate. This could lead to an increase in agricultural intensification, as defined by the amount of net primary productivity appropriated by humans and therefore removed from the ecosystem rather than recycled. This impact is likely to be most pronounced in present-day marginal agricultural areas. However, it is also possible that if these marginal areas experience increased precipitation at crucial times of the year then some areas could also be subject to agricultural land abandonment: this can have negative implications for those priority habitats and species that rely on a small amount of disturbance to maintain their niche. It is also possible that increases in precipitation rates and changes in land use result in an increase in water pollution risk and eutrophication. During eutrophication, high nutrient inputs, mainly from agricultural runoff, can cause algal blooms that deplete oxygen supply and are detrimental to aquatic habitats and species. Climate change will also interact with the impacts of atmospheric deposition in which elevated levels of pollutants such as nitrogen oxides, sulphur oxides and ozone already interfere with soil processes, species interactions and community composition, particularly due to acidification. Many semi-natural habitats are naturally nutrient poor and support slower-growing assemblages of plants that are adapted to these conditions. Therefore although sulphur oxides are declining, excess levels of nutrient nitrogen cause eutrophication with faster growing species outcompeting slower-growing species. These impacts could be further exacerbated in a warmer climate with higher CO2 levels.

Other indirect impacts include a potential increase in flood defence structures which can separate a river from its floodplain to the detriment of wetland habitats. A warming climate may also lead to increased societal water demand, notably for agricultural irrigation, which could lead to reduced low flows and lower water levels (depending on the role of environmental regulation). Finally, responses to climate mitigation measures could also have important implications for biodiversity, either positively or negatively. This issue is particularly associated with the recognition of species and habitats as dynamic features of the landscape. Mitigation measures and adaptation responses in other sectors will therefore have either positive or negative consequences depending on their appropriateness for a particular location and their role in allowing ecosystems to move and adapt to change.

Observational evidence for these indirect impacts is usually currently associated with the loss of biodiversity due to the primary effects of land use change and pollution. There is less information available to-date on how climate will interact with the other drivers and on the cross-sectoral impacts of climate change.

Page 60: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

32 Biodiversity Sector Report

Table 3.1 High-level impacts identified in Tier 1

Impact Type Specific Impacts Climate parameter

28

Consequences for Biodiversity

Evidence Base

RANGE SHIFTS species unable to track changing climate space

T,P,ET Loss of biodiversity and priority species

Medium

species unable to find suitable microclimate, including altitude

T,P, ET, wind

Loss of biodiversity and priority species

Medium

opportunities for new (priority) species

T,P,ET Increased biodiversity

Low

SEASONAL SHIFTS AND CHANGES IN PHENOLOGY

asynchrony between a species breeding cycle and its food supply

T Loss of priority species

Medium

change in life cycles (esp. insects)

T Loss/gain of priority species

Medium

changes in species migration patterns

T Loss/gain of priority species

Medium

CHANGES IN PESTS, DISEASES AND INVASIVE NON-NATIVE SPECIES

increased risks from pests

Tmin,P Loss of biodiversity and priority species

Medium

Increased risks from diseases

Tmin,P Loss of biodiversity and priority species

Medium

increased risk from novel pathogens

Tmin,P Loss of biodiversity and priority species

Low

CHANGES IN INTERACTIONS AND COMMUNITY STRUCTURE

generalists (eg:. ruderal spp.) favoured over specialists

CO2, T, P Loss of biodiversity and priority species

Medium

changing competition between C3 and C4 photosynthesis plants

CO2, T, P Habitat change; species loss/gain

Low

changing interactions due to differences in growth/survival rates

CO2, T, P Loss/gain in biodiversity

Low

changing interactions between trophic levels

CO2, T, P Loss/gain in biodiversity

Low

changes in genetic diversity

CO2, T, P Loss/gain in biodiversity

Low

impacts of changing nutrient supply

CO2, T, P Loss/gain in biodiversity

Low

28

T-Temperature, Tmin-Minimum Temperature, P-Precipitation, ET-Evapotranspiration, SLR-Sea Level Rise, CO2 – Carbon Dioxide Concentration, Rhu-Relative Humidity

Page 61: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 33

Impact Type Specific Impacts Climate parameter

28

Consequences for Biodiversity

Evidence Base

GEOMORPHOLOGICAL AND HYDRO- ECOLOGICAL HABITAT CHANGE

coastal evolution impacts on intertidal, grazing marsh etc.

SLR, wave Loss/gain in coastal habitats and species

High

floodplain evolution P Loss/gain in wetland habitats and species

High

increased water temperature and stratification of water bodies

T Loss of priority species

Medium

less snow cover T Loss of specialist montane species

High

high flow impacts on spawning beds

P Loss of priority aquatic species

High

low flow impacts via Biological Oxygen Demand (BOD)

T,P Loss of priority aquatic species

High

saline intrusion SLR Loss of habitat Medium

increased soil moisture deficits and drying

T,P,ET Habitat loss; priority species loss

High

increased soil erosion

T,P,ET Habitat loss High

increased waterlogging

P Habitat loss High

HABITAT DISTURBANCE BY EXTREME EVENTS

windthrow during storms

Wind Loss of woodland habitat

High

major coastal flood/reconfiguration

Wave, surge

Loss/gain of coastal habitats

Medium

major fluvial flood P Loss/gain of wetland habitats

High

major drought events

P, T Loss of priority habitats

High

Loss/gain of niche space

T,P, ET,Rhu Loss/gain of priority species

Medium

wildfire risk T,P,ET Loss of priority habitat and species

Medium

CHANGES TO ECOSYSTEM PROCESSES/ FUNCTIONING

changes in primary productivity

T,P, CO2 Implications for ecosystem integrity

Low/Medium

changes in soil organic carbon

T,P, CO2 (as above) Low

faster decomposition and nutrient cycling

T,P, CO2 (as above) Medium

changes in soil microbial activity

T,P, CO2 (as above) Low

Page 62: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

34 Biodiversity Sector Report

Impact Type Specific Impacts Climate parameter

28

Consequences for Biodiversity

Evidence Base

INDIRECT EFFECTS agricultural intensification (i.e. human use of NPP)

T,P,ET Loss of habitat and priority species

High

agricultural abandonment

T,P,ET Loss/gain of priority species

Medium

increased water pollution risk and eutrophication

P,T Loss of habitat quality and priority species

High

impacts of atmospheric deposition (e.g. N, SO2, O3)

T,P, CO2 Habitat loss and species change

Low

climate mitigation measures (positive/negative)

T (and others indirectly)

Loss/gain in habitat quality and priority species

Low

increase in flood defence structures

P Loss of wetland habitats

High

increased societal water demand

T,P, ET Loss of aquatic and wetland habitats

Medium

3.2 Identification of priority (Tier 2) risks

For the Biodiversity and Ecosystem Services sector, priority (Tier 2) risks were identified based on the multi-criteria method described in Section 2.5. The scoring of impacts is provided in Appendix 2. It was also informed by the principle of including, if possible, representative impacts from each of the eight impact groups discussed in Section 3.1. This would therefore facilitate a broad-based analysis of a range of diverse risks at different levels across the sector. Some impacts were ‘borderline’ and others either partly or primarily defined by other sectors but with an important biodiversity and ecosystem services component29. Therefore these are synthesised in Section 3.3 but are not subject to more detailed (Tier 2) analysis in Chapter 4. The methodology identified 16 priority risks that are analysed in more detail in Chapter 4 and Chapter 5. Table 3.2 presents the impacts and associated risk metrics that were selected for Tier 2 analysis.

The potential impacts of these risks vary in terms of their habitat specificity: Figure 3.1 provides a preliminary vulnerability assessment of BAP priority habitats to these particular risks. The scoring system used for this assessment for each habitat was based upon a combination of sensitivity and adaptive capacity (natural or managed) based on the following system:

V = Sensitivity * {(Natural Adaptive Capacity + Managed Adaptive Capacity)/2}

Sensitivity and the two different components of adaptive capacity were each scored from 0-3 (for a vulnerability assessment, high adaptive capacity is 0 or 1 and low

29

Biodiversity itself was not the primary consideration behind the analysis of the impacts identified in the other sectors; rather it was the consequence of the change in biodiversity leading to a change in the ecosystem services. For example, the ‘species shifting’ impact in the marine sector concentrated on the movement of commercially important species.

Page 63: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 35

adaptive capacity is 2 or 3). Hence the most vulnerable habitats that are highly sensitive to climate change and limited capacity to adapt are those that approach the maximum score of 9.

Table 3.2 Tier 2 primary risks

Identifier Risk Impact Group (‘Cluster’) Associated Sectors identified by CCRA

BD1 Increased soil moisture deficits and drying

Geomorphological / Hydroecological Change

Agriculture

BD2 Coastal evolution impacts on intertidal, grazing marsh etc.

Geomorphological / Hydroecological Change

Marine; Agriculture

BD3 Increased risks from pests Changes in pests, diseases and invasive non-native species

Agriculture, Forestry, Water, Health

BD4 Increased risks from diseases Changes in pests, diseases and invasive non-native species

Agriculture, Forestry, Water, Health

BD5 Species unable to track changing climate space

Range Shifts

BD6 Climate mitigation measures (positive/negative)

Indirect effects Energy

BD7 Major coastal flood/reconfiguration

Extreme Events Marine

BD8 Changes in soil organic carbon Ecosystem Processes / Functioning Forestry, Agriculture

BD9 Changes in species migration patterns

Seasonal Shifts / Phenological Change

BD10 Increased water temperature and stratification of water bodies

Geomorphological / Hydroecological Change

BD11 Generalists favoured over specialists (eg. ruderal spp.)

Changes in Interactions and Community Structure

BD12 Increased risk of wildifire Extreme Events Built Environment, Forestry, Agriculture

BD13 Increased water pollution risk and eutrophication

Indirect effects Water

BD14-16 Impacts of low flows Geomorphological/Hydrological change Water

BD15 Increased societal water demand

Indirect effects Water

BD16 Major drought events Extreme Events Water

Page 64: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

36

Figure 3.1 Preliminary vulnerability assessment of UK BAP habitats

Page 65: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 37

3.3 Synthesis of other key risks (terrestrial biodiversity)

In this section we synthesise further information on key risks that are relevant to this sector but are not analysed in detail in Chapter 4.

3.3.1 Land use change - impacts of agricultural intensification or land abandonment

This risk is an indirect consequence of autonomous (unplanned) or planned responses in the agriculture sector. Climate is an important influence on agriculture by providing the pattern of seasonal change for managing crops and livestock. The capability of the land to support different types of agricultural activity is highly variable across the UK and mapped by the Agricultural Land Classification (ALC) system (England, Wales, Northern Ireland) or the Land Capability for Agriculture (LCA) system (Scotland). These classification systems combine spatial data on climate, soils and topography to provide maps that show the suitability of the land for cropping, improved grassland and extensive agriculture. The best quality land tends to be used for arable and horticulture because it provides the highest returns, with intermediate land used for mixed farming and improved pasture, and the poorest quality land used for larger scale ‘rough grazing’. In this context, agricultural ‘improvement’ refers to the modification of the land cover and soils such that it can provide higher yields of food/energy crops or a higher stocking density of animals (e.g. by adding fertilisers, pesticides, lime or reseeding with faster-growing grasses).

An ultimate consequence of agricultural intensification is usually that a higher proportion of biomass produced by net primary productivity is removed from the ecosystem and that nutrients become depleted unless replaced by artificial fertilisers30. This can have negative implications for biodiversity, which together with the application of pesticides and herbicides, results in severe reductions in species abundance and diversity, unless remedial measures are applied. An example of intensification is the shift from spring-sown to autumn-sown crops that has occurred in many arable areas of the UK because of the potential for greater yields: the shorter fallow period in the field means a reduced food supply for wild species.

A changing climate means that the bioclimate factors associated with different agricultural activities will also shift. In Scotland, research has shown that the areas of prime agricultural land (as defined using the LCA system) may expand and that in marginal areas a greater proportion of land may be capable of agricultural ‘improvement’ where soils are suitable (Brown et al., 2008, 2011)31. This change in land use may be further encouraged due to concerns regarding food security. An increase in the area of land used for intensive agriculture could have further negative consequences for biodiversity, particularly in marginal areas that have high biodiversity value. The expansion of prime agricultural land may also be dependent on the

30

Intensive systems are not necessarily detrimental to biodiversity and of course provide a very significant ecosystem service through the supply of food and energy. In 2011, the UK Government Chief Scientist (John Beddington) framed the challenge as providing ‘sustainable intensification’ that can reconcile food and energy supply with the maintenance of wider ecosystem benefits in both the short and long term. 31

Defra project SP1104 (2010-2012) is currently carrying out a similar analysis of the impact of climate change on the ALC classes of England and Wales, but results for the future projections were not available during the CCRA analysis period.

Page 66: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

38 Biodiversity Sector Report

additional availability of irrigation water, because it tends to occupy the drier areas that could experience exacerbated moisture deficits during the future.

Marginal agricultural areas in some locations may also be at risk of land abandonment. This may result from high precipitation rates causing soils to remain wet or becoming wetter for key periods of the year in spring and autumn. When soils are at high water-holding levels (‘field capacity’) access to the land is very difficult and would be highly likely to cause damage to soil structure due to compaction. This can severely limit agricultural activities or make them highly dependent on favourable weather from year to year. Farming in marginal areas is therefore strongly associated with subsidies that provide a secure income through schemes of the EU Common Agricultural Policy (CAP). If subsidies decline or are ended as CAP is reformed, then a reduction in farming or even land abandonment becomes a possibility. A decline in headage payments for sheep in recent years has therefore been associated with a reduction in sheep numbers in the more marginal areas, which has been described as a ‘retreat from the hills’ (SAC, 2008). This has implications for biodiversity as some priority habitats and species are currently maintained by light livestock grazing. For example, this low-level disturbance allows grassland habitats to be maintained that would otherwise be likely to become scrub or woodland through time. Land abandonment would have mixed results for biodiversity: some species would gain and some would lose, but there could be significant consequences for some important and rare UK species that are present only in a few locations.

Biodiversity interactions with land use were also identified in other sectors, including Agriculture (Knox et al., 2012) and Forestry (Moffat et al., 2012), although specific risks were not analysed in detail.

3.3.2 Increase in fluvial flood defence structures

For a river in its natural state during high flow conditions a critical level is reached when some water leaves the normal watercourse and spreads out over a larger area which represents its floodplain. The variety of sediment left during floods and the ephemeral state of the floodplain produce a diverse range of habitats that support many important species. Man-made engineering structures are designed to protect people, properties, and land on the floodplain from the adverse effects of flood inundation by maintaining the river within a predictable course. However, these structures act to separate the river from its floodplain to the detriment of the associated riparian and wetland habitats.

An increased risk of flooding from climate change will require measures to reduce that risk to a level judged acceptable by society. If floodplains continue to be populated and cultivated, then it is very plausible that an adaptive response will be to increase the extent of flood defence structures. This would be particularly likely if society becomes more ‘risk averse’ in future and continues the current inclination to favour structural approaches to flood defence influenced by a reaction to provide visible responses to risk (Harries and Penning Rowsell, 2011). This acts against schemes that integrate natural flood management incorporating the river floodplain and wetland habitats to buffer against flood risk. If this happens the future of riparian habitats and the dynamic evolution of floodplain wetland habitats will be negatively impacted.

3.3.3 Changing species interactions due to differences in growth/survival rates

A warming climate and the seasonal shifts that are most apparent in earlier springs is associated with changes in the timing (phenology) of events such as flowering and egg-laying. In different species this will result in variations in their rate of growth and

Page 67: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 39

alter the competition between species. Ultimately this will result in changing survival rates between species and differences in their population size meaning that they may become more or less dominant against their competitors.

General findings that particular taxonomic groups seem to be benefiting or declining in association with a changing climate therefore also need to be contextualised against data for specific species in order to interpret the potential causes of that change. For example, as explained in detail in Section 4.2, analysis for upland birds suggests that golden plover (Pluvialis apricaria) have declined because of their preference for crane fly (Tipulidae) prey which have declined probably due to drier summers; however whimbrel (Numenius phaeopus), which has a more widely-variable invertebrate diet, has been less sensitive to these changes (Pearce-Higgins, 2010).

At present, although information on changes in growth and survival rates is widely available, particularly highlighting the benefits of milder winters for many species, the implications for species interactions remains rather uncertain.

3.3.4 Asynchrony between a species breeding cycle and its food supply

This issue is particularly illustrative of risk, uncertainty and complexity in the biodiversity sector. Climate change has already altered the phenology (seasonal timing of events) for many species by modifying the environmental cues that they use for migration, breeding and predation, ultimately influencing their demography and population dynamics. However, the influence of these changes through interactions between species across trophic levels (steps in the food web) may be much more profound.

Some evidence for this phenomenon (‘phenological mismatch’) has been reported across the UK (e.g. Thackeray et al., 2010) and Europe. A notable example is that of caterpillars hatching and then pupating too early compared with chick hatching of some insectivorous birds e.g. great tits (Parus major). This has led to less prey available for some woodland birds and therefore declines in their breeding success and abundance (Visser et al., 2006; Both et al., 2006). Similar issues have been reported for insectivorous birds related to the timing of breeding and the period of peak food availability (Visser, 2008). Recent evidence has also been presented regarding the interaction of the common cuckoo (Cuculus canorus) with its hosts based upon sixty years of data. This suggests that short-distance, but not long-distance, migratory hosts have advanced their arrival more than the cuckoo, with potential consequences for breeding of both cuckoo and hosts (Saino et al., 2010). The mismatch may explain the recent decline of cuckoo populations and observed local changes in parasitism rates of host species. A detailed analysis of phenological changes across trophic levels by Thackeray et al. (2010) suggests that terrestrial plants (primary producers) have on average advanced their phenology the most and that secondary consumers in particular have responded at rather slower rates. Again this highlights the potential for asynchrony, but the same analysis also showed a wide variation of responses across taxa related to factors such as local conditions and functional traits.

Disruption of these ecosystem relationships could cause major shifts in key functions that they maintain. However, finding evidence for or against disrupted relationships and their demographic effects is difficult because the necessary detailed observational data are rare or often provide only a partial picture. Moreover, we can often only speculate on how sensitive species will generally be to phenological mismatches when they do occur. It is quite possible that through behavioural change (the ‘plasticity’ of species) and natural selection that the phenology between species may through time become synchronous again. Also, we do not really know whether all levels in multi-trophic

Page 68: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

40 Biodiversity Sector Report

interactions across the food web will be affected at the same rate, and therefore whether synchronization can be maintained across the ecosystem under large-scale climate change. The rate of climate change is likely to be the key variable here, with the likelihood of asynchronous events increasing as the rate of change increases, and hence leading to an increased possibility of major ecosystem disruption. By implication, this could also have important consequences for many associated ecosystem services.

3.3.5 Impacts of changing nutrient supply and cycling

Nutrient cycling involves the movement and exchange of organic and inorganic matter between soil, water and atmosphere, regulated by the biological components of the ecosystem. It therefore acts as a fundamental ecosystem supporting service and has key interactions with other supporting services such as primary production, water cycling and soil formation.

Carbon- and nitrogen-based nutrients are particularly important for providing energy and building tissue, and therefore are basic requirements for plant growth and the decomposition of organic matter. The balance between carbon and nitrogen (C/N ratio) is a key measure of soil functioning. Soil carbon is further analysed as a Tier 2 risk (BD8) in Chapter 4. The slow conversion of atmospheric nitrogen into soil nitrates which can be accessed by plants means that many semi-natural habitats are nitrogen-limited. The variety of adaptive responses developed in these habitats has therefore been a major factor in developing their high biodiversity value.

The enrichment of semi-natural habitats by atmospheric nitrogen deposition from human sources and fertiliser application on managed land has significantly modified nutrient cycling in the last 50 years. Changes in vegetation growth and composition towards faster-growing species have modified the abundance and activity of soil organisms, with impacts on rates of nutrient cycling and primary production. The Countryside Survey (2010) reported an increase in the C/N ratio of topsoil across most UK terrestrial habitats, with the exception of arable areas. This indicates that nitrogen enrichment has increased plant production and the sequestration of carbon in plants and soil (Emmett et al. 2010).

There is a high degree of uncertainty about the interaction of climate change with nutrient cycling. Increasing temperatures will lead to increased decomposition rates but this is countered by elevated CO2 levels and atmospheric nitrogen deposition which will increase biomass production. The balance between increased productivity and increased decomposition will also be dependent upon precipitation patterns as the C/N ratio declines under drier conditions. Impacts will therefore also vary across terrestrial habitats depending on factors such as vegetation type, soil fertility, soil water conditions and soil type. In aquatic habitats, current levels of nutrient enrichment (eutrophication) from agricultural run-off will also be a factor. However, it seems highly likely that climate change in most situations will further accelerate the rate of nutrient cycling. The resultant change in nutrient availability will inevitably modify population dynamics and community structure, with implications not only for the biodiversity of priority habitats but also the many ecosystem services that nutrient cycling supports (UK NEA: Emmet et al., 2010).

It is quite possible that these changes will happen at a non-linear rate. Controlled experiments have shown that with rising CO2 levels, primary productivity of ecosystems tends to increase to a threshold level then begin to decrease. The reasons for this are not completely known but are associated with the influence of nutrient supply and other factors.

Page 69: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 41

3.4 Synthesis of other risks (marine biodiversity)

Marine biodiversity (outwith the coastal zone) was primarily included in the Marine and Fisheries CCRA sector report (Pinnegar et al., 2012) although biodiversity was not the main focus of that sector. Here we provide a synthesis of the main risks to marine biodiversity from that report. The UK Marine sector has been reviewed several times in recent years and these reviews provide much of the information used in this sector report of the CCRA. For example notable recent reports that have given an overview of maritime industries have included the Productive Seas Feeder Report, of Defra’s Charting Progress 2 assessment (Defra, 2010d) on the state of UK seas, the Scottish Government ‘Marine Atlas’ published in March 2011 (Scottish Government, 2011), Northern Ireland’s 2011 Devolved Administration Marine Sector Report ‘State of the Seas’ (Gibson, 2011) and the most recent OSPAR Quality Status Report published in September 2010 (OSPAR, 2010). The information provided below draws on information provided for Charting Progress 2 by each of the four UKMMAS ‘Evidence Groups’ (the Ocean Processes Evidence Group; Healthy-Biologically Diverse Seas Evidence Group; Productive Seas Evidence Group and Clean, Safe Seas Evidence Group).

3.4.1 Harmful algal blooms

Coastal and marine waters are impacted by nutrient enrichment (eutrophication) as a result of runoff of agricultural fertilisers and human or animal waste This encourages an increase in algal biomass rise that can become harmful when dominated by a single species or species group. The effects of such harmful blooms include overgrowth and shading of sea grasses, oxygen depletion of the water as a result of algal and bacterial respiration, suffocation of fish from stimulation of gill mucus production, direct toxic effects on fish and shellfish, and mechanical interference with filter feeding by fish and bivalve molluscs (Landsberg, 2002). The toxic affects can also extend to marine mammals and humans. In addition, many harmful algal bloom species are not efficiently grazed which results in a decreased transfer of nutrients to fish stocks, impacting on their population. The aggregation of impacts on fish and shellfish stocks, and potentially on bathing water quality and tourism, have important economic implications. A strong regional distribution is observed with the impacts being more regularly detected along the south and western coasts of Ireland and the coasts of Scotland.

Climate change may influence algal blooms through the following factors:

Temperature changes can alter ecological processes and species interactions.

Increases and decreases in sunlight due to changes in cloud cover and radiation which may affect toxicity and stratification.

Changes in precipitation may alter freshwater run-off and inputs of nutrients, sediment, and contaminants.

Alteration of wind and water circulation patterns influences geographic distributions of species and upwelling of nutrients.

Changes in stratification of the water column.

Increases in carbon dioxide availability affect phytoplankton productivity.

Ocean acidification influences species assemblages and processes or directly alters toxicity.

Page 70: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

42 Biodiversity Sector Report

3.4.2 Ocean acidification

The world’s oceans play an important role in reducing the contribution of CO2 to global warming by absorbing approximately half of the CO2 released by human activities since 1800 (Sabine et al., 2004). A consequence of this uptake is ‘ocean acidification’ due to the lowering of ocean pH as a result of the disassociation of CO2 in solution. Both modelling and observational studies suggest that the absorption of CO2 by the ocean has already decreased pH levels by 0.1 since 1750 (Orr et al., 2005), which is about 100 times the rate of change that has been detected in the geological record.

A decrease in pH may have negative impacts on a wide variety of species and ecosystems particularly calcifying organisms such as corals and shellfish that require a specific pH to produce calcium carbonate. Other studies have shown that functions other than calcification (such as metabolism, behaviour and immune responses) can also be affected by ocean acidification. This will lead to implications for marine food webs including higher species such as fish, and have associated socio-economic impacts (e.g. tourism reduction). However, the impacts of acidification on marine species and ecosystems are not fully understood. It may also have an impact upon the propagation of sound through the oceans and the rate of conversion between different nitrogen compounds, which could in turn impact upon phytoplankton growth and nutrient cycling.

Changes in pH can also affect the availability of trace metals, which may be necessary for plankton growth or in other cases (e.g. aluminium) may increase to toxic levels. Changes in plankton may lead to wider changes in ecosystem composition, structure and functioning with potential deleterious impacts on ecosystem goods and services.

As atmospheric concentrations of CO2 continue to rise, increased absorption by the oceans may accelerate oceanic acidification. However, projecting future changes in acidification is difficult due to the influence of many factors including rising temperatures and the changing rate of uptake of CO2 through biological processes.

3.4.3 Species range shifts

As noted earlier for terrestrial species, a warming climate induces shifts in the geographical range of species. This is more unconstrained for marine environments because of the lesser influence of man-made boundaries. The Marine Climate Change Impacts Programme (MCCIP) notes an abrupt ecosystem shift in the late 1990s which was most pronounced in parts of the NE Atlantic near the 9-10°C sea surface temperature isotherm. This isotherm represents a critical boundary between 'warm' and 'cold' water ecosystems and has moved northwards at an approximate rate of 22km/yr since the 1960s.

The best data on change in marine species comes from commercially important fish species where long data sets (90 years) have been used to highlight that some species at least appear to be shifting their distribution due to climate change, although pressures from overfishing are often dominant. Over the past 30 years, some fish distributions have moved northwards by 50- 400km (Perry et al., 2005), with coldwater species such as monkfish and snake blenny moving the furthest. Similarly, some species have moved into deeper waters at an average rate of about 3.5 metres per decade. Species such as salmon and eel which have life cycles in both fresh and marine waters have been shown to be particularly vulnerable to climate change (water temperature and river flow) with impacts on both the freshwater and marine phases. In addition, some species of toothed whales and dolphins are showing shifts in distribution, which may be linked to increasing sea temperatures.

Page 71: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 43

A northward shift in the distribution of many plankton species has also been recorded, typically by more than 10º latitude over the past 50 years. In the North Sea, the population of the previously dominant cold-water zooplankton species Calanus finmarchicus has declined in biomass by 70% since the 1960s. These significant changes in plankton abundance have had impacts through the food chain, contributing to the reduction in quality and abundance of species such as sandeels (Ammodytes marinus) which provide food for many seabirds and potentially some baleen whale species.

Future warming will inevitably further alter the range of phytoplankton and zooplankton, and because of their key role within marine ecosystems this will modify services such as oxygen production, carbon sequestration and nutrient cycling. By 2050, climate change modelling indicates that pelagic species (such as herring and anchovy) move northward by an average of 600km and demersal species (such as cod and haddock) by 220km. Other cold-water species such as some coral species and maerl are likely to be also affected (in combination with ocean acidification impacts) and their role as ‘ecosystem engineers’ may decline with associated impacts on the wider ecological community.

Warming will also impact on the pattern of marine currents, which redistribute warm and cold water, with consequences for the dispersal of fish eggs and larvae. It seems likely that winter and early spring spawners (such as cod and plaice) will experience poor larval survival, whereas warmer-water species (such as sprat) may benefit.

3.4.4 Invasive non-native species

A further implication of range shifts is that it increases the risk of biological invasions from non-native (alien) species. The introduction of non-native species to a marine ecosystem followed by their subsequent establishment may cause effects ranging from the almost undetectable to the complete domination and displacement of native communities. Recording suggests that an increasing number of non-native species are arriving in marine habitats, that some of these species are becoming more abundant, and that some are causing major ecological changes on both local and global scales (Ruiz et al., 1997; 2000). Of particular concern are notorious invasive non-native species such as the Northern Pacific sea star (Asterias amurensis), caulerpa seaweed (Caulerpa taxifola), and the American comb jelly (Mnemiopsis leidyi).

More than half the total number of introduced species in UK waters is attributed to shipping, whilst half of the non-native marine algae are attributed to deliberate introductions for mariculture. Invasive non-native marine plants have often spread quite rapidly, while invasive non-native invertebrates have tended to spread more slowly (Eno, 1997).

The distribution of non-native species in British waters indicates many more introduced species on the south and west coasts, particularly areas such as the Solent (probably due to the past and present volume of shipping) and the Essex coast (associated with oyster grounds). Climate change is likely to increase the incidence of invasive non-native species based upon their association with temperature; however the dominant influence is likely to be the rate of introduction which is also related to socioeconomic factors (e.g. shipping trade, mariculture).

3.4.5 Disease hosts and pathogens

In addition to invasive non-native species, range shifts associated with warming of the seas and oceans is likely to cause an increased risk from marine pathogens. Relatively

Page 72: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

44 Biodiversity Sector Report

small increases in temperatures may extend the distribution of zooplankton and therefore, increase the risk of infectious disease epidemics (Kuhn et al., 2005). In addition, physiological stresses induced by warming (and other factors such as acidification and pollution) compromise the resistance of host species and makes them more susceptible to an increased frequency of opportunistic diseases. In addition to the increased prevalence and virulence of diseases, they also provide a risk factor by their potential to induce new symptoms for host species. Ultimately, this can lead to modification of the ecosystem if the species at risk are key components of that ecological community.

Apparently, most new diseases are not caused by the spread of new micro-organisms but by known agents infecting new or previously unrecognised hosts (Harvell et al., 1999). For example phocine distemper virus (PDV) had previously been recognised in other species but has caused mass mortality in recent decades in harbour seals (Phoca vitulina) and grey seals (Halichoris gryphus). The disease was suggested to be transmitted to the new hosts through infected harp seals that had migrated toward Europe because of overfishing around Greenland. A further causal factor may be that some pollutants (e.g. organochlorides) impair the immune systems of host species.

Less is known about the impacts of pathogens on invertebrates, although increased incidence of disease in corals and oysters have been recorded worldwide. Higher marine temperatures have been correlated with the increased prevalence and transmission of disease in seafan corals in UK waters.

A warming climate is therefore liable to increase the risk of disease outbreaks by modifying the range of host species, by changing their susceptibility to disease, and by altering species interactions which may transmit disease to new hosts. Few epidemiological studies of marine organisms have been carried out, and these are often limited to taxa such as corals and oysters, therefore the wider consequences for biodiversity are likely to be very uncertain. The emergence of novel pathogens, as outlined for terrestrial environments, remains as a ‘wildcard’ and is associated with recombinant variations of existing pathogens that could lead to a more virulent strain.

3.4.6 Protected habitats and species

Impacts on coastal habitats and species have been highlighted as a major risk to biodiversity and are covered as a Tier 2 risk in Chapter 4 (Section 4.3). In addition, rocky shores are likely to experience a northward movement of benthic marine organisms, particularly if an increase in coastal defence structures facilitates greater connectivity between habitats. Changes in marine temperatures and seawater chemistry, and changes in ocean currents are likely to affect individual species associated with specific habitat types. In some instances the effects of such changes can be sufficient to significantly alter the community composition and structure of habitats. In the subtidal region, eelgrass beds, which are a UKBAP high priority habitat, may be affected by increased temperature and any potential increase in storminess and erosion.

Warming temperatures could result in increased stratification of surface waters. This would have a considerable impact on ocean productivity by reducing the upwelling of nutrients, with knock on effects for species at multiple levels of the food chain. The growth rate of corals and other reef organisms is affected by CO2 concentrations through ocean acidification, although much of the evidence for this comes from tropical habitats. Although surface waters would experience the greatest temperature rise in the near future, deep water coral reefs may also undergo significant changes.

Page 73: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 45

3.5 Cross-sectoral risks

Many of the risks examined for this sector are linked or interact with risks in other sectors (Figure 3.2). Ecosystems represent complex networks with multiple levels of organisation and interdependencies between biotic and abiotic components. The CCRA Systematic Mapping exercise therefore inevitably involved considerable simplification as most cause-effect relationships are often contingent on many other factors. For this reason, these relationships are often geographically-specific or dependent on antecedent conditions with lag times between stimulus and response.32 More abrupt change can happen during extreme events when a key threshold is crossed and a step change in ecosystem processes can then occur.

With these caveats in mind, some of the main cross-sectoral links relevant to this sector can be highlighted. Some of the key risks in this sector, such as species being unable to track range shifts (BD5), soil moisture deficits (BD1), invasive non-native species (BD3/4), coastal evolution (BD2/7) and water-related issues (BD10/11/14/15/16), are seen to be have particularly strong links with other sectors.

3.5.1 Land use

Firstly, there is an important interaction between AGRICULTURE and BIODIVERSITY as intensive agricultural production systems can lead to biodiversity loss unless appropriate management measures are implemented. Hence a change in agricultural practices (including their intensity and timing) in response to a changing climate could potentially reduce food and shelter for farmland species leading to a reduction in biodiversity. In marginal upland areas, the presence of low-intensity agriculture is beneficial to many species and if a changing climate (e.g. increased seasonal wetness) reduces agricultural capability leading to land abandonment, then this could be detrimental for biodiversity. The exact consequences for species and habitats in both examples would be dependent on local circumstances and any changes would involve ‘winners’ and ‘losers’ but the implications for UK biodiversity obligations could be very significant. Similar issues regarding production activities and biodiversity relate to FORESTRY and BIODIVERSITY. In addition, climate-related improvements in the quality of marginal land together with the development of 2nd and 3rd generation bioenergy crops encouraged by schemes in the ENERGY sector could have implications for existing biodiversity in these areas. The EU Common Agriculture Policy and policy targets for renewable energy both act as major drivers for land use, and the need to reconcile production systems with environmental quality.

3.5.2 Water

Particularly for aquatic and wetland habitats, there are important interactions between WATER resources and BIODIVERSITY with regard to both the quantity and quality of water available. A trend to warmer drier summers implies lower streamflows and water levels, a reduction in dissolved oxygen, and consequent impacts on ecosystems. This could be further exacerbated due to increased demands for water by people and industry, leading to greater abstraction of water from rivers and lakes. Furthermore, extreme drought events can result in a combination of factors whereby the supply-demand balance becomes highly unbalanced and competition for water is exacerbated: these future events could lead to irreversible change for some ecosystems that are

32

Ecosystems are therefore characteristic of complex adaptive systems including features such as non-linear behaviour, hysterisis, multiple thresholds, and dynamic equilibria.

Page 74: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

46 Biodiversity Sector Report

already stressed. In addition to the direct effects of climate change, impacts on water quality and quantity will also be strongly dependent on changes in land use (AGRICULTURE; FORESTRY). Any changes in the release of major pollutants could combine with lower flows in summer and potentially higher flows in winter to modify the distribution and extent of pollution to water bodies.

3.5.3 Flood risk and flood/coastal defence

Important links also exist between FLOODS AND COASTAL EROSION and BIODIVERSITY related to the presence of fixed defence structures that protect people, properties and land from flooding and erosion. Maintaining these structures in situ for coastal locations during rising sea levels means that intertidal habitats on the seaward side of the defences are vulnerable to increased erosion and inundation: this issue of ‘coastal squeeze’ is further explored within Tier 2 risks BD2 and BD7. Conversely, if the defence structures are removed or realigned inland, some non-marine habitats and species that have low saline tolerance could be adversely affected. In fluvial environments the presence of fixed defences often leads to the separation of the river from its floodplain. Changes in flow regime of rivers due to wetter winters or more intense rainfall imply an increased flood risk potential (possibly further modified by land use changes in AGRICULTURE and FORESTRY) that could lead to an increase in fixed defences to the detriment of floodplain wetland habitats and species. Decisions with regard to the status of defences in coastal or fluvial situations have important cross-sectoral implications for flood/erosion risk hence intersect with the BUILT ENVIRONMENT, HEALTH, BUSINESS, ENERGY, TRANSPORT and AGRICULTURE sectors.

3.5.4 Built environment and infrastructure

Within the BUILT ENVIRONMENT, the presence of ‘greenspace’ or ‘green infrastructure’ can have important benefits for BIODIVERSITY by maintaining and enhancing habitat variety for a range of species. With temperature rises likely to be exacerbated by ‘urban heat island’ effects, greenspace can provide a valuable regulating role by moderating temperature change, particularly during extreme events. The presence of vegetation and functioning soils can also slow drainage and runoff rates, reducing FLOOD risk from changing precipitation patterns. Finally, although interactions between a changing climate and atmospheric pollutants are uncertain, greenspace can also moderate adverse effects by scavenging and storing air pollutants. These multiple benefits suggest that green infrastructure can have a key role for maintaining HEALTH and WELLBEING as the climate changes in urban areas.

3.5.5 Adaptation and mitigation

A further issue that highlights cross-sectoral and indirect impacts is related to climate mitigation schemes (highlighted by Tier 2 risk BD6). These schemes can potentially have positive or negative implications for BIODIVERSITY depending on location and therefore need to take full account of both the present and future status of the natural environment and conservation objectives to avoid unintended indirect consequences during future decades. In addition to the potential for large-scale renewable ENERGY developments related to wind, tidal and hydro power, the development of 2nd or 3rd generation bioenergy schemes could be substantially expanded through interaction with the AGRICULTURE and FORESTRY sectors. Bioenergy may be also favoured by increased temperatures and radiation levels improving yield potential. However, large-

Page 75: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 47

scale development of energy crops could have consequences for WATER resources because of their increased water demand compared to other land uses33, with implications for BIODIVERSITY as explained in 3.5.2 in addition to impacts due to habitat modification. With good planning, significant potential exists to integrate mitigation schemes within green infrastructure (see Section 3.5.4 above) to provide multiple benefits from ecosystem services (e.g. flood risk alleviation; improvement of air quality) in addition to the development of habitat networks (e.g. for woodland).

3.5.6 Wildfire risk

Impacts from wildfires were also highlighted by other sectors including AGRICULTURE, FORESTRY and the BUILT ENVIRONMENT, with the likelihood of large-scale events extending across sectors. This Tier 2 risk is further evaluated in Section 4.12.

3.5.7 Invasive non-natives, pests and diseases

Increased risk from pests and diseases were also highlighted by the AGRICULTURE, FORESTRY and MARINE AND FISHERIES sectors, identifying the need for a cross-sectoral approach. Impacts on biodiversity are further evaluated in Section 4.4. In the Forestry sector report, loss of production due to pests and diseases was assessed as a Tier 2 risk (metric FO1).

3.6 Ecosystem services

The concept of Ecosystem Services provides a framework in which to further consider these cross-sectoral interactions, by recognising the many direct and indirect benefits that people obtain from the natural environment (Table 3.3). The provision of food and energy from biomass and the supply of clean water are amongst the obvious benefits. However, ecosystems provide a wide range of services including those associated with the regulation of hazards (notably the role of wetlands in reducing flooding and erosion), the regulation of pests and diseases, pollination, and the maintenance of soil, water and air quality. Cultural services are often less tangible but relate to issues associated with aesthetics, inspiration, education and heritage.

The impacts of climate change on biodiversity and ecosystem processes therefore have fundamental relevance for human society because of the services they provide to sustain human wellbeing. Using this framework, Table 3.3 identifies key services that are provided by this sector to other CCRA sectors. Whether these other sectors have recognised an explicit or implicit ‘demand’ for this service depends on progress in establishing a reciprocal process to maintain the flow of this service from the natural environment. For example, the water industry has recognised the benefits of restoring natural ecosystems such as wetlands as a means of securing an improved supply of clean water.

33

This is a consequence of the higher evapotranspiration rates and fast growth rate of these crops

Page 76: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

48 Biodiversity Sector Report

Table 3.3 Key ecosystem services provided to other sectors

Ecosystem Service Sector

Water purification and regulation of water quality

WATER; HEALTH (and others indirectly)

Regulation of water flow and water quantity

FLOODS; WATER; ENERGY (hydro); AGRICULTURE

Regulation of erosion/sedimentation FLOODS; TRANSPORT

Slope stabilisation TRANSPORT

Regulation of air quality (e.g. scavenging of pollutants)

BUILT ENVIRONMENT

Phytoremediation of contaminated land BUILT ENVIRONMENT

Pollination AGRICULTURE; FORESTRY

Pest and Disease control AGRICULTURE; FORESTRY; HEALTH

Nitrogen Fixation Regulation AGRICULTURE

Climate Regulation via Carbon Storage ALL

Climate regulation via micro-climate BUILT ENVIRONMENT; AGRICULTURE

Crop and Livestock production AGRICULTURE; ENERGY

Timber and fibre production FORESTRY; ENERGY

Aquaculture MARINE; AGRICULTURE (freshwater)

Protection from dust and noise BUILT ENVIRONMENT

Culture (e.g. education, spiritual benefits, inspiration, identity)

BUILT ENVIRONMENT (cultural heritage); HEALTH; BUSINESS

Landscape and amenity benefits, (e.g. tourism, recreation, greenspace)

BUILT ENVIRONMENT (cultural heritage); HEALTH; BUSINESS

NB. Many additional indirect and secondary benefits also accrue from ecosystem services, such

as via AGRICULTURE/FORESTRY, WATER or TRANSPORT, to BUSINESS or to HEALTH.

Page 77: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

49

Figure 3.2 Cross-sectoral linkages with other sectors in the CCRA

Page 78: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

50 Biodiversity Sector Report

4 Sector risk analysis (Tier 2)

4.1 Introduction

In this section, analysis and further information is provided for each of the Tier 2 risks identified in Chapter 3. Most impacts defined for this sector cannot at present be subjected to a robust national-scale quantitative analysis of risk using the climatic projections defined under UKCP09. Partly this is because of uncertainties in the climate response and partly because the probabilistic data in UKCP09 was not designed with the spatial coherence between variables necessary to model geographic changes in impacts on regional or national scales34. Furthermore, risks to biodiversity are often further contingent on other influences, both current and historic, including biophysical factors (e.g. soils) and human factors (e.g. land use systems).

A broader discussion of risks is therefore provided to contextualise the analysis including case studies that can provide a varied or more detailed regional or country-based perspective. A detailed risk assessment of all habitats and species is clearly beyond the scope of the CCRA, therefore reference is also made to case studies that highlight specific habitat vulnerabilities (see Figure 3.1). The main objective of the analysis is to characterise any change in risk from present to future, either qualitatively or quantitatively. This is used to infer any ‘adaptation deficit’ that would imply the need for further action beyond current adaptation strategies. Social and economic drivers of change are introduced in Chapter 5 in order to distinguish the relative importance of climate and non-climate influences for each type of risk.

As identified in Chapter 3, some risks are cross-sectoral and have also been identified in other CCRA sectors. However, analyses in other sector reports have not specifically addressed risks to biodiversity or the potential changes to ecosystem services.

For each metric a scorecard is given at the start of each section to indicate the confidence in the estimates given and the level of risk or opportunity. Confidence is assessed as high (H), medium (M) or low (L). Risks and opportunities are scored either high (3) medium (2) or low (1) (shown to the right). These are given for the lower (l), central (c) and upper (u) estimates for the 2020s, 2050s and 2080s. Further information is provided in Appendix 5. Where estimates are uncertain, or no data is available, this is stated in the scorecard.

M Confidence assessment from high (H) to low (L)

3 High opportunity (positive)

2 Medium opportunity (positive)

1 Low opportunity (positive)

1 Low risk (negative)

2 Medium risk (negative)

3 High risk (negative)

34

The non-probabilistic UKCP09 Spatially Coherent Projections only become available during the later stages of the CCRA analysis. For this reason, previous spatial analysis using the UKCIP02 scenarios has also been included in the assessment to indicate potential geographic variations in risk factors.

Page 79: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 51

4.2 Increased soil moisture deficits and drying (BD1)

Summary

Analysis has highlighted the potential for significant contraction of important habitats and species in both upland and lowland locations due to increased soil moisture deficits. Case studies are provided for beech woodlands, blanket bog and upland birds.

Although there is a high level of agreement that moisture deficits are projected to increase and have significant impacts on many existing habitats, the evidence base remains at medium level. This limited evidence means we have MEDIUM confidence in the assessment of this risk. In addition to bioclimate envelope models, a wider range of predictive tools are required for both species and habitats (and at multiple scales), including improved validation of these tools against observations of change in different habitats.

l c u l c u l c u

BD1Increased soil moisture deficits

and dryingM 1 1 1 2 2 2 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.2.1 Introduction

Water availability is a key influence on the distribution of species and habitats, and hence ecosystem functioning. Many species have physiological adaptations (e.g. via plant roots or leaves) that have developed to take advantage of particular environmental conditions, and therefore a change in those conditions will require further adaptation that will favour some species and habitats over others.

Most habitat groups were identified as sensitive to changes in soil moisture in the Tier 1 assessment. For the purposes of the analysis, two UK BAP Priority Habitats have been selected for further analysis based upon available information and the climate adaptation issues they identify: these are Lowland beech woodland and Blanket bog (Figure 4.1). These habitats also provide differing examples of the risks to lowland and upland habitats respectively and are discussed in the following sections. The risks to species are highlighted through upland birds.

Page 80: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

52 Biodiversity Sector Report

Figure 4.1 Distribution of presence of lowland beech/yew woodland and blanket bog per 10km grid-square. For reference annual maximum soil moisture

deficits are shown (1971-2000 mean) Source data: JNCC; UK Met Office

POTENTIAL SOIL MOISTURE DEFICIT (PSMD) PSMD provides a measure of the balance between precipitation and potential evapotranspiration (ET) and hence of wetness (lower values) and dryness (higher values). As a measure of ‘potential’, it assumes unlimited soil water. The pattern of actual ET and soil moisture will be much more complex due to the interactions of different soil and vegetation types. However, PSMD is commonly used in bioclimatic studies as it can provide a reliable large-scale reference surface. The reference surface is by convention usually based on a land cover of grass. Maximum PSMD values give an estimated magnitude for the largest deficit during the year and also provide a reasonable indication of the period of deficit as in general deficits are longer where they are largest.

4.2.2 Current risk status

Lowland beech woodland

The European Beech (Fagus sylvatica L.) is usually the dominant species within two priority habitats identified by Annex 1 of the EU Habitats Directive:

Page 81: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 53

Asperulo-Fagetum beech forests which occur on neutral or calcareous soils (notable examples include the beech hangers of the Chilterns, Cotswolds and E. Hampshire).

Atlantic acidophilous beech forests which occur on acid soils (notable examples exist in Epping Forest and the New Forest).

These beech woodland habitats are typically associated with thin soils having a relatively low available water capacity (AWC) that increases their vulnerability to drought. Soils with a higher AWC can accommodate larger potential soil moisture deficits and provide more available water to plant roots before reaching a critical threshold for water stress. However, for habitats developed on chalk, beech trees may be able to access additional moisture stores by capillary pressure through the porous bedrock.

The sensitivity of five tree species to increased soil moisture deficit has been considered in the FORESTRY CCRA sector report (Moffat et al., 2012) under metrics FO2 (Loss of productivity due to drought) and FO4 (Change in tree species suitability). The focus of FO2 is on productivity and timber yield rather than biodiversity conservation, but it does highlight defoliation issues linked to soil moisture deficits for two native species: beech and oak; metric FO4 is considered further in Section 4.2.3. In addition, beech was chosen by Defra as one of the UK Climate Change Indicators (indicator 26)35 using Forestry Commission data on beech foliage cover between 1987 and 2001 that was correlated with mean July rainfall from the previous year. This metric can provide a simple quantitative indicator of the climate sensitivity of tree health (Figure 4.2) although it does not relate directly to tree mortality.

In their study of the response of a particular stand of beech in Wales to the drought of 1975/76, Peterken and Mountford (1996) found that drought was implicated in the deaths of 12 out of a stand of 92 beech, with 6 trees dying within a couple of years and the remainder in a subsequent state of decline leading to death up to 14 years later. Mortality is associated with splits in the bark and increased susceptibility to pathogens such as Biscogniauxia nummularia and Armillaria (Green and Ray, 2009).

More recent data have been presented by Stribley (2005) and reviewed by Broadmeadow et al. (2009) who note the association between synchronous heavy seed production over large areas (‘masting’) and climate stress in beech; this is linked to a reduction in resources for leaf growth and lower foliage cover. Similarities are also noted with observations of European decline in beech growth after the 2003 drought (Lorenz et al., 2008).

35

http://www.ecn.ac.uk/iccuk/indicators/26.htm

Page 82: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

54 Biodiversity Sector Report

R2 = 0.4493

0

10

20

30

40

50

60

70

0 20 40 60 80 100 120 140 160

Average rainfall in England and Wales the previous July (mm)

% b

ee

ch

tre

es

wit

h >

25

% l

es

s f

oil

ag

e

tha

n n

orm

al

Figure 4.2 Correlation between beech foliage cover and previous July monthly rainfall

Blanket bog

Blanket bog occurs in the cooler wetter areas of Britain, where peat has accumulated not only in wet basins but also draped extensively over the surrounding undulating topography to form a complex variety of mire habitats. The UK contains about 15% of global blanket bogs and therefore has a special responsibility for protection of this priority habitat type. Large areas of blanket bog have been designated as protected areas because of their importance for priority species, notably breeding birds and invertebrates. All peatland habitats, including blanket bog, are primarily dependent on the depth of water table, with peat accumulation taking place only when the water table is approximately level with the surface over the long-term (Parish et al., 2008).

Sphagnum bog moss species are key species acting as ecosystem engineers for water retention and ultimately active peat formation (paludification). Sphagnum species are non-vascular, accessing water either directly by rainfall or via capillary rise from the water table. When the water table falls in dry conditions, capillary water is restricted and Sphagnum growth can be severely limited.

The full extent of blanket bog habitat in the UK is uncertain as until recently multiple definitions existed and detailed mapping was mainly carried out only for designated sites. A review for the IUCN UK Peatland Programme has suggested that only 18% of blanket mire is currently in a natural or near-natural condition, and of the remainder,16% is eroded, 16% is afforested and 40% is modified (Littlewoood et al., 2010). The same review has also noted that although species in peatland areas show mixed trends, the majority of those designated as UK BAP priority species have declining populations.

Climate change has been suggested as a trigger for peat erosion, notably through increased drought frequency, but changing land use practices and the impacts of atmospheric deposition have also been implicated (e.g. Billet et al., 2010; Clark et al., 2010a). Nevertheless, because of the key role of local hydrological patterns in vegetation development, blanket bog is highly likely to be sensitive to long-term changes in the water table. In addition to its importance for biodiversity, blanket bog has particular significance for climate change mitigation because of its role as a carbon sink (Worrall et al., 2009). UK peatlands (including lowland bog and fen) are estimated to contain 5200Mt of carbon, of which the majority, 4500Mt is in Scotland (Dawson and

Page 83: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 55

Smith, 2007). In pristine condition, active peat bogs can accumulate up to 0.7 tonnes of carbon per hectare per annum (Holden et al., 2007).

Upland birds

Many species of subarctic breeding birds are at the southern limit of their range in the uplands of the UK where they occupy blanket bog, grassland and heathland habitats and feed on invertebrates. Crane flies (Tipulidae), in particular, provide the most important prey group for most of these birds, with large populations favoured by wet damp soils and pools. These invertebrates are sensitive to changing weather conditions, and a shift to warmer drier summers can have negative impacts on their population size. Analysis by Pearce-Higgins et al., (2009) has shown a link between the breeding success of golden plover (Pluvialis apricaria) and the abundance of crane fly prey for their offspring. A trend to increased soil moisture deficits in summer during recent decades has had an impact on prey abundance and therefore on the population of the golden plover. This association seems to dominate over changes in the phenology (seasonal timing) of golden plover breeding and cranefly emergence in the spring. Significant statistical relationships (negative correlations) with climate data were only achieved by relating cranefly populations with August temperatures of the previous year and golden plover populations with August temperatures from two years previously (Figure 4.3).

Figure 4.3 Correlation between August temperature of 2 years earlier (mean daily maximum temperature) and golden plover population (log-ratio of change

between consecutive years). Negative changes in population are associated with higher temperatures

Source: Pearce- Higgins et al. (2009)

Similar associations have been noted by Pearce-Higgins (2010) across a range of upland birds and the relationship was used to define a climate sensitivity index for 17 species: the most vulnerable species were identified to be dunlin (Calidris alpina) and snow bunting (Plectrophenax nivalis). Less sensitive species included whimbrel (Numenius phaeopus) which has a more widely-variable invertebrate diet, and could be defined as more ‘generalist’ in its preferences (see BD11, generalists favoured over specialists, Section 4.11). This analysis also suggests that for these birds winter mortality is becoming less of a dominant factor and that summer conditions are

Page 84: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

56 Biodiversity Sector Report

becoming a stronger influence on population dynamics (although warmer winters will also act to modify breeding distributions, see Austin and Rehfisch, 2005).

Soil moisture deficits have also been linked to changes in lowland invertebrates and bird species. An increased prevalence of drought causes worms to burrow deeper in the soil, making them less available as prey, and this has been linked to the decline of the song thrush (Turdus philomelos) (Peach et al., 2004).

4.2.3 Potential impact of climate change

Lowland beech woodland

Future projections of climate suitability for beech suggest that the greatest changes may be seen in the south and east of England (Broadmeadow et al., 2005) where most of the beech woodland priority habitats are located (Figure 4.4). This analysis was based upon the Ecological Site Classification (ESC) model of Forest Research using a combination of climatic and soil variables. These findings are consistent with results from the MONARCH project that also implied beech could lose climate space in the south and east of England (Harrison et al., 2001). The CCRA Forestry sector report (Moffat et al., 2012, drawing on Reed et al., 2009) includes an analysis of changes in future productivity for beech, amongst other species, which indicate for South East England a decline in yield class (relative to the maximum achievable for the UK) from 0.67 in the baseline to 0.64 for 2050s Low Emissions, 0.46 for 2050s High Emissions and 0.15 for 2080s High Emissions (all references to the UKIP02 scenarios). These are translated into reductions in the suitable productive area for beech in this region from a baseline of 64ha to 61ha (2050s Low), 44ha (2050s High) and 14ha (2080s High) respectively.

Although these model projections suggest a decline in beech, differences in soils and topography also imply considerable variation in response with beech being unlikely to be lost completely in these locations (Mitchell et al., 2007). The MONARCH2 project could not find conclusive evidence of a major climate change influence on woodland communities of the E. Hampshire beech hangers (Berry et al., 2005). As beech trees are widespread across drier parts of continental Europe, a key issue may be the genetic variability of the species. Although the genetic pattern is rather uncertain, it is more likely to be higher and provide more resilience to change in larger communities that have less habitat fragmentation. However, this may be further modified as the beech tree is often planted in the UK which could cause variations from the expected pattern.

Page 85: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 57

Figure 4.4 Indicative maps of suitability for beech under UKCIP02 low and high emissions scenarios for 2020s, 2050s and 2080s

Source: Broadmeadow et al. (2005) [NB Maps do not include the additional effects of increased CO2, pests and diseases, or extreme events]

Blanket bog

Depending on the methodology and data used, climate envelope models can produce varying future projections of species or habitat change. A combined analysis using nine different models to define the bioclimate envelope for blanket bog has shown that the climate space associated with peatlands is projected to shrink (Clark et al., 2010b; Figure 4.5). All except one of the models showed a reduction, with some indicating it would be larger than others. This reduction in the climate envelope implies that more than 50% of the peatland area is projected to become vulnerable to change by the 2050s, with western Scotland as the least vulnerable area. This does not necessarily imply that the peat will be lost in other areas, as this will be dependent on a range of local factors, but drier vegetation types are projected to become more common and active peat formation reduced. Tree invasion of bogs as a consequence of summer drought could also locally lead to increased water loss through transpiration and higher heat absorption enhancing the drying effect on the bog surface (Mitchell et al., 2007).

Bioclimate models provide general tools that indicate the potential for changing climate space. Actual patterns of change are usually also heavily influenced by other factors such as species interactions, thereby these projected changes require further validation.

A shift to a drier regime for blanket bog habitats would have important implications for biodiversity through species loss but also lead to the underlying peat soils becoming more unstable. Evidence derived from those sites that lost vegetation cover in the past

Page 86: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

58 Biodiversity Sector Report

(e.g. from pollution or grazing pressure) suggests that the underlying peat is then at a much higher risk of erosion and loss of carbon (Orr et al., 2008; Lilly et al., 2009).

Figure 4.5 Future projections of blanket peat distribution using a range of bioclimate envelope models and the UKCIP02 low and high emissions scenario

Source: Clark et al. (2010b) [Published by permission of Inter Research]

Upland birds

Based upon the analysis of golden plover and cranefly populations outlined in Section 4.2.2, Pearce-Higgins et al. (2009) developed a deterministic model which can be extrapolated to provide projections of future sensitivity of climate change (Table 4.1). This shows that a continued increase in summer temperatures implies a considerable risk of extinction for the golden plover at this site within the next 100 years, particularly when the temperature rise exceeds 4ºC. Other bird species are also considered to be at risk due to the decline in invertebrates that prefer cool, damp conditions (Pearce-Higgins, 2010). The less sensitive species tend to have a wider variety of diet including prey species that are less sensitive to drier conditions.

Future projections should be interpreted with caution. It is possible that a change in prey availability might induce behavioural change in some species: this adaptive response would reduce their vulnerability. In addition, it is possible that changes in life cycle phenology, such that the emergence of insect prey is asynchronous with the breeding cycle of predators such as upland waders (either later or earlier), may become increasingly important in the future, hence potentially increasing vulnerability. These issues are further discussed in Section 4.8 (changes in species migration patterns) and Section 4.11 (generalists favoured over specialists).

Page 87: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 59

Table 4.1 Estimated golden plover population declines at Snake Summit and probabilities of extinction for a range of temperature scenarios by 2100

Source: Pearce-Higgins et al. (2009)

4.2.4 Implications for other habitats/species

Risks will be apparent for many other habitats beyond the two highlighted here. The higher transpiration demands of trees means that other woodland habitat types are likely to be vulnerable to drier conditions, both directly and indirectly through water resource pressures (see Chapter 5). Calcareous grasslands, which have high species diversity, are suggested to show changes in community structure as the changing conditions favour xerophytic (arid) species (Duckworth et al., 2000). In the coastal zone, dune slack communities are also strongly dependent on the local water table and its hydrological regime: recent modelling work in NW England has projected large falls in the water table (>1m by 2100 using the UKCIP02 Medium-high Emissions scenario) which would make many of these communities unviable (Clarke and Sanitwong Na Ayutthaya, 2010). With regard to species, the impacts of reduced soil moisture availability during current drought periods have been shown to affect a range of insects and plants, particularly those adapted to cold-moist conditions (Morecroft et al., 2002).

Changes in soil moisture regimes have important links to other risks for biodiversity. Increased deficits are associated with an increase in frequency of wildfires (BD12, Section 4.12) and the climate stress imposed on some priority and dominant species can lead to increased vulnerability to pests and diseases (BD3 and BD4). Seasonal shifts may also modify the rates of soil mineralization and nutrient cycling, although the interaction with temperature and CO2 means the outcome is very difficult to predict.

4.2.5 Implications for ecosystem services

Projected impacts also have implications for ecosystem services. For example, peat bogs and other wetland habitats are particularly important in providing regulating services, including for water flow (notably reduced flood peaks), water quality and carbon storage. In recent decades, increased concentrations of dissolved organic carbon (DOC) in water have been attributed to peat erosion (see BD8, Section 4.7), requiring expensive treatment to reduce discolouration and potential risks to human health. Peatlands and habitats such as the beech hangers of Southern England are also recognized for their cultural landscape qualities as well as their biodiversity value. The UK NEA estimated the amenity value of all wetland broad habitat types (including coastal) at ca. £1.3 billion p.a.

Page 88: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

60 Biodiversity Sector Report

Soil moisture is a key influence on supporting ecosystem services such as primary production and nutrient cycling. Therefore a shift towards drier soils has implications for the overall functioning of many ecosystems. Existing knowledge of these interactions is limited but suggests that there are key thresholds beyond which a change in soil moisture leads to major step change in ecosystem processes.

4.2.6 Implications for adaptation

Both of the highlighted habitats also raise important issues for adaptation. The climate space for beech woodlands is inferred to move north and this would also allow it to potentially colonise or be planted upon soils of higher AWC that are less drought-prone. There are certainly opportunities for development of beech habitat outside of its current distribution (d’Erlanger, 2001; Wilson, 2006), although establishment of a functional ecosystem would require significantly more effort. However, in areas further north in the UK, beech is often characterised as ‘non-native’ and management plans can target its active removal (Wesche, 2003). Potential adaptation measures within its existing range include establishing new native woodland on soils with higher AWC and the introduction of more southerly provenances that may be better adapted to summer water deficits.

Sympathetic management of blanket bog can increase resilience to climate change, such as the blocking of drains to raise the water table and the exclusion of detrimental land management practices such as excessive burning and overgrazing (LIFE Peatlands Project, 2005; Bain et al., 2011). In some locations, peatland restoration could re-establish vegetated surfaces with diverse ecological communities. However, the topographic variability of the habitat means that this is likely to be highly site-specific rather than a universal solution to enhance biodiversity, carbon storage and water quality.

4.3 Coastal evolution impacts on intertidal, grazing marsh etc (BD2) and Major coastal flood / reconfiguration (BD7)

Summary

Coastal evolution represents long-term incremental and episodic change, whilst a major coastal flood would occur as an abrupt extreme event or possibly due to a large-scale reconfiguration of the coast.

The level of analysis has been restricted by availability of UK-level data, with regard to both the geographic extent of the assessment and BAP habitats included. Projected loss due to coastal erosion is quite significant for some habitats, notably saline lagoons (maximum loss of 20% habitat), whilst erosion losses for others are much more modest.

Defra project CR0422 suggests that in England an average of around 4-6% of selected terrestrial and freshwater habitats in the coastal floodplain could be lost due to flood inundation (High Emissions scenario), although a significant portion of this represents habitat that is already at risk under current climate conditions. This represents quite a risk to the BAP habitat in the coastal floodplain. It would certainly not be straightforward to attempt to recreate a similar amount of habitat elsewhere that has similar characteristics but which is not exposed to the same risk. It is likely that there will also

Page 89: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 61

be commensurate gains in other more saline tolerant habitats if inundation does occur.

The analysis suggests that we have MEDIUM confidence in assessing the risk of incremental change (BD2: medium consensus with a medium evidence base) and MEDIUM confidence for extreme events or large-scale change (BD7: medium consensus with a medium evidence base). Considerable information is available on the current risk in some coastal areas, particularly East and Southern England that are experiencing the accumulated effect of multiple pressures for change. However there have been rather few systematic assessments of future risk to coastal biodiversity in the UK. This represents a significant knowledge gap.

Similar findings were highlighted during reviews for the Marine Climate Change Partnership (combined findings of Mieszokowska, 2010; Rees et al., 2010).

4.3.1 Introduction

l c u l c u l c u

BD2Coastal evolution impacts on

intertidal, grazing marsh etc.M 1 2 2 2 2 2 2 3 3

BD7Major coastal flood /

reconfigurationM 1 2 3 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

Coastal habitats have high biodiversity value, supporting a high number of species relative to their extent: for example, sand dune, coastal shingle and maritime cliffs in England support 148 UK BAP species within an estimated total extent of 31,200ha (Webb et al., 2010). Ecological characteristics of the coastal zone are strongly inter-related to its geomorphological profile through the interaction of biota, sediments and landforms. Some coastal systems are very resilient and their structural dynamics allows them to adjust to prevailing conditions. Most beaches fall into this category being drawn down in storms and recovering during more benign conditions. Other systems are more sensitive and once disrupted may not recover or move to a different geomorphological state (Carter and Woodroffe, 1994). The onset of overwashing to transform a freshwater lagoon to a saline lagoon is a good example of this. The complex dynamics of the UK coast depend on the interaction of sea levels (and any isostatic adjustment through land movements), sediment supply and climate forcing (winds, waves, tides and river discharges; rainfall and temperature).

The two impacts identified for more detailed examination (BD2 and BD7) both relate to impacts on coastal habitats but distinguish between how these impacts come about. Long-term coastal evolution (BD2) is the combined influence of a range of events including gradual change and storms36, which collectively influence the transition from marine to terrestrial habitats. Extreme storm events also have the potential to cause extensive saline inundation with implications for predominantly terrestrial habitats that are immediately behind the natural or defended shoreline37: this impact is examined as BD7.

36

Both gradual change and storms may result in both erosion and inundation from flooding. 37

High probability, high frequency events also have the potential to influence coastal habitats.

Page 90: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

62 Biodiversity Sector Report

(i) Evolutionary impacts on coastal margin habitats (BD2)

The UK coastline comprises areas that are predominantly stable (e.g. those comprising hard rock formations) and those that are either eroding or accreting (e.g. soft rock coast, shingle, beaches etc.). These processes often occur in close proximity to one another. The result of erosion can be the loss of both marine-influenced habitats (e.g. mudflats, saltmarsh) and terrestrial habitats inland (e.g. coastal grazing marshes, reedbeds, fens etc). Conversely, accretion of sediment can result in the creation of suitable areas for new habitats such as saltmarsh or coastal vegetated shingle.

Coastal defences by their very nature disrupt the natural progression from marine to terrestrial habitats. For better or worse, this has created numerous anomalies around the coast where the marine suite of habitats are limited by the defences and the freshwater habitats are extended. The latter make a significant contribution to the structure and function of the overall contemporary habitat mosaic but would not be present on a natural coastline, being an artefact of the defences. Under changing conditions and in particular sea level rise, this leads to the additional problem that the defences constrain the shoreline geomorphology and associated habitats from responding in the most appropriate way.

In a natural environment, sea level rise typically causes coastal landforms to migrate landwards; a process often referred to as ‘rollover’. Where this is constrained by coastal defence structures a phenomenon known as ‘coastal squeeze’ may occur leading to the loss of habitat, such as saltmarsh, on the seaward side of the defence structure.

Sea level rise under climate change is anticipated to lead to an increase in the rate of erosion of some areas and may also lead to an increase in the rate of accretion in others. This would speed up the long-term reconfiguration of some coastlines in the UK that is already occurring in many areas. Hence, habitats will be both lost and created, including marine and terrestrial components. Habitat creation can also take place through anthropogenic activities, such as managed realignment, a process generally carried out through the deliberate breach of existing flood defences.

(ii) Inundation impacts due to coastal flooding (BD7)

The UK coastline also comprises numerous areas that are at risk of coastal flooding under extreme events. Coastal flooding can occur as a result of direct tidal inundation resulting from breaches of the natural coastline. It can also occur due to breaches of coastal defences as these have been designed to cope with events up to a particular magnitude (e.g. 1:200yr). Should an event occur that is bigger than this (e.g. a significant storm surge), overtopping or a breach may occur, potentially leading to flooding of the areas behind the defence. Coastal flooding can also be exacerbated by fluvial flooding where high tides coincide with fluvial flood events.

Although a serious breach of the natural coastline or of flood defences can result in a major coastal reconfiguration, such events are rare. The risk due to this type of change has not been evaluated but has occurred during a number of major storms over the last century, notably during a major storm surge in 1953. It remains as an ongoing risk that is likely to occur more frequently as sea levels rise (Nicholls et al., 2007); therefore we evaluate the impact on habitats behind the shoreline due to changes in the frequency of inundation.

The majority of affected habitats would be terrestrial or freshwater habitats that have variable but generally limited ability to tolerate saline inundation. It is likely that expansion of one habitat may be at the expense of another, particularly where intervention is made through managed coastal realignment. The key measure of risk is the impact of losses, or shifts in habitat types, on the ecological function of the coastal ecosystem as a whole.

Page 91: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 63

4.3.2 Current risk status

(i) Evolutionary impacts on coastal margin habitats (BD2)

The current risk is examined in terms of rates of erosion and accretion. A European-scale study of coastal erosion (Eurosion, 2004) found that over 17% of the coastline of the UK coast was experiencing erosion (Table 4.2). These figures indicate that England has the greatest proportion of coastline (30% of the total length) subject to erosion in the UK and that it is the most extensively defended (46% is protected by artificial structures and beaches), with significant erosion also experienced by Wales, Scotland and Northern Ireland.

The Foresight coastal flooding project found that 28% of the combined English and Welsh coast was experiencing erosion rates greater than 10 cm/year (Evans et al., 2004). Some areas experience significantly greater erosion rates: for example, soft cliff areas of Suffolk typically erode at 1-4 m/year and up to 7m/year at locations such as Dunwich and Covehithe (Doody et al., 2006). At Covehithe, Benacre and the Easton Valley in Suffolk, priority habitats of European significance (saline lagoons, freshwater reedbeds) are under serious threat from these rates of erosion and similar situations occur at other locations around the UK. Risk assessment for the National Trust ‘Shifting Shores’ study (Table 4.3) found that 169 of their sites were at risk from erosion (over 608km in length) with 126 sites currently at risk from tidal flooding (4040ha in total).

Coastal erosion data were also analysed in the Futurecoast project (Burgess et al., 2002). Rates of change of high water and low water were mapped around the coast of England and Wales to investigate the changing profile of the shoreface. Analysis found that the majority of the overall coastline had steepened (61%), whilst 33% had flattened and 6% had experienced no changes. This widespread steepening has occurred relatively evenly around the west (64%), south (66%) and east coasts (58%) (Taylor et al., 2004).

In North West Scotland, coastal habitat systems characterized as ‘machair’ have been identified as particularly vulnerable (Angus et al., 2010). Machair is low grassy land on calcareous, often shell-rich coastal sand forming a mosaic of rich and diverse floral habitat that supports breeding and overwintering birds of international importance. The land also has an important socio-economic value because it provides relatively fertile soils in an otherwise inhospitable location. Machair is found only in North West Scotland (ca. 70% of the global habitat) and Western Ireland meaning the UK has important international responsibilities to protect and maintain it. Much of the machair is not only low-lying, but in parts of South Uist appears to occupy an altitude below High Water Mark (Angus, 2009). The main machair areas are separated from the foreshore by systems of coastal dune ridges that provide protection from the sea, but in places the dunes have been removed by erosion. Machair systems are therefore vulnerable to coastal erosion and flooding, particularly during extreme events: most recently, a severe storm in January 2005 led to substantial marine inundation.

Page 92: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

64 Biodiversity Sector Report

Table 4.2 Coastal erosion and coastal defences in the UK [NB. Figures in the last 2 columns show that some coasts with artificial defences are eroding]

Coast Length (km) Length of Coast Eroding (km)

Length of Coast with Defences and Artificial

Beaches (km)

Country

England 4,273 1,275 (30%) 1,947 (46%)

Wales 1,498 346 (23%) 415 (28%)

Scotland 11,154 1,298 (12%) 733 (7%)

Northern Ireland 456 89 (20%) 90 (20%)

UK Total 17,381 3,008 (17%) 3,185 (18%)

England by Region

North East England 297 80 (27%) 111 (37%)

North West England 659 122 (19%) 329 (50%)

Yorkshire and the Humber 361 203 (56%) 156 (43%)

East Midlands 234 21 (9%) 234 (100%)

East of England 555 168 (30%) 382 (69%)

South East England 788 244 (31%) 429 (54%)

South West England 1,379 437 (32%) 306 (22%)

Source: Living with coastal erosion in Europe: Sediment and Space for Sustainability, PART II – Maps and statistics (2004), p. 21. Available from http://www.eurosion.org/index.html

Table 4.3 Current coastal erosion and flood risk for National Trust sites in England and Wales

Region/country Kilometres of Trust coast affected by

erosion

Hectares of Trust land at risk of

flooding

South West 279 852

South East 44 467

East of England 45 1837

London - 1

North West 9 70

Yorkshire 12 1

North East 52 26

Wales 167 786

Total 608 4040

Source: National Trust (2005)

(ii) Inundation impacts due to coastal flooding (BD7)

Inundation of terrestrial/freshwater habitats with brackish/ saline waters has the potential to cause changes in species composition, impacts on growth rates, and changes in food webs and ecosystem functioning (e.g. loss of invertebrates, amphibians and fish that are important for bird populations). The magnitude of the effect on habitats depends on a range of parameters including the differences in salinity experienced, frequency/duration of flooding, and season. For example, impacts

Page 93: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 65

may be moderated by freshwater flushing following heavy rainfall. In each case, species and communities will react differently depending on their salinity tolerance.

The joint Defra/Environment Agency NEOCOMER38 project (Defra, 2006a) calculated potential losses of habitats from coastal flooding in Natura 2000/SSSI/Ramsar sites to be over 32,000ha. This comprised 2,400ha of inland waterbodies and lagoons, 14000ha of wet grassland, 700ha of drier grassland and 14,000ha of bogs, marshes and swamps. Although not quantified by NEOCOMER, climate change is likely to exacerbate this vulnerability with sites flooded sooner or more regularly than predicted from current standards of protection at the sites. Furthermore, this study did not include an assessment of the sensitivity of habitats to frequency/ duration of inundation, but assumed that all inundation would result in loss.

Defra project CR042239 developed matrices describing the sensitivity of priority BAP

habitats to saline inundation based on scientific data, empirical observations and expert opinion (Defra, 2011a; example matrix: Figure 4.6). Together with outputs from the National Flood Risk Assessment (NaFRA) model this was used to estimate the extent of habitat in England at risk of irreversible damage and loss from current and future climate change (via sea level rise) in the 1:1000yr tidal and tidal/fluvial floodplains. The methodology incorporates the resilience of some habitats to inundation (from frequent and extreme events) or the risk of irreversible damage and loss of habitat type. Estimating replacement habitats is not a simple task, although this is one key element in understanding the overall level of risk, or opportunity.

Project CR0422 mapped selected habitats in the 1:1000 year coastal floodplain for England including grazing marsh, saline lagoons, eutrophic standing waters (lakes and ponds), lowland fens, reedbeds, deciduous woodland, lowland raised bog, and purple moor grass and rush pastures40. Over half of the resource of coastal and floodplain grazing marsh, reedbed and saline lagoons, and over 30% of lowland raised bog is located within the coastal floodplain. An average of about 4,600 ha (4.2% of the total41) of the total selected BAP habitat are at risk of loss under climatic conditions prevailing currently (2010). This represents 81% of the total selected BAP habitat at risk under the most extreme climate change scenario that was assessed (which was 2080s, medium emissions, with degraded coastal defence condition). Of the areas at risk, 86% is coastal floodplain grazing marsh. Future risk is considered in Section 4.3.3ii. Similar data is not currently available for other habitats, although local-level data is available from some locations42.

Similar country-level studies are not currently available for Wales, Scotland and Northern Ireland meaning that the level of risk or opportunity remains uncertain. However an assessment for the UK BAP has estimated saltmarsh losses at 100ha/yr and mudflat losses at 100-150ha/yr. There are important regional differences in risk across the UK with examples in North West England and the Dyfi estuary (Wales) of recent coastal accretion rather than retreat (Mieszkowska, 2010).

38

Neocomer: National Evaluation of the Costs of Meeting Coastal Environmental Requirements 39

Developing Tools to Evaluate the Consequences for Biodiversity of Options for Coastal Zone Adaptation to Climate Change

39 Defra project CR0422 ran concurrently with the CCRA and its

findings have been partially incorporated. 40

National-scale habitat datasets were found to have major errors regarding distribution and extent of priority BAP habitats hence the use of those habitats with highest confidence in any one location. 41

A total of 109,020 ha of habitat is mapped as present in the coastal floodplain 42

For example, saltmarsh loss in Essex has been estimated at 70ha/yr (English Nature, 2004).

Page 94: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

66 Biodiversity Sector Report

Value Assigned Risk

1.0 Absolutely certain the habitat will be lost

0.9

0.8 Very likely that the habitat will be lost

0.7

0.6

0.5 Equally likely that the habitat will be lost or will recover

0.4

0.3

0.2 Very unlikely that the habitat will be lost

0.1

0.0 Absolutely certain the habitat will recover

Figure Note: An assessment was made by the project team and experts at a workshop of the likelihood of habitat loss for each combination of flood frequency and duration in the matrix, for each BAP habitat. For Coastal and Floodplain Grazing Marsh two matrices were derived, one for the grazing marsh ditches and one for the grassland as it was expected that these two would be affected differently. Using these matrices enabled the sensitivity of each of the habitats to saline inundation to be incorporated into the modelling analysis.

Figure 4.6 Likelihood matrix for grazing marsh grassland Source: Defra (2011a)

Characterisation of the current risk is also presently being reappraised due to new data (tide gauges, satellite monitoring and intertidal sedimentation rates) that suggest the rate of relative sea level rise may have accelerated in recent decades (Rennie and Hansom, 2010; Teasdale et al., 2011). This is particularly important for Scotland: the prevailing view had been that this was an area of lesser risk with the exception of peripheral areas (e.g. the Northern Isles and Western Isles which cover the majority of machair habitats), due to rising land levels from glacio-isostatic ’rebound’ exceeding rates of sea level rise. However, the new data cover a relatively short time period and further work is necessary to ascertain whether the increased rate of sea level rise is a long-term trend or a presently-unexplained component of natural variability.

Page 95: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 67

4.3.3 Potential impact of climate change

(i) Evolutionary impacts on coastal margin habitats (BD2)

Lee (2001) used a simple model to predict areas of habitat change over the next 50 years within three different environmental designations (SAC, SPA and Ramsar) in England and Wales. The predicted changes were based upon a review of available Shoreline Management Plans (SMPs) with regional workshops, and assumed that current and projected plans would be implemented over the long-term. The analysis adopted rates of sea level rise based on Defra guidance and broadly linked to the UKCIP02 projections. Habitat gains occurred from managed realignment and accretion, whilst losses occurred due to coastal squeeze, managed realignment and erosion. This showed that there would be a net loss of coastal dry land, wetland and open water habitat of approximately 4,000 hectares from protected sites in England and Wales over the next 50 years. He also noted that there could be a net gain of intertidal habitats (saltmarsh and mudflat/sandflat) of some 2,220 ha, although this is based on the assumption that much of the gains come from managed realignment programmes to offset the losses due to coastal squeeze and erosion of the unprotected coast.

The habitat datasets from Defra project CR0422 were used in the CCRA to model the projected losses from coastal erosion using UKCP09 data for three epochs (2020s, 2050s and 2080s) excluding additional coastal flooding. Coastal erosion rates were derived from the Futurecoast project (Defra, 2002) and modified based upon four different rates of sea level rise. Results are tabulated in Appendix 3, with the following conclusions:

Regional variability:

- The North West England and North East England are projected to experience considerably fewer losses overall than South West England and East England. By the 2080s coastal grazing marsh is projected to lose a between 16-58ha in East England, 25-38ha in South West England and 3-8ha in South East England43.

Habitat variability:

- Under the total projected habitat losses, deciduous woodland is projected to lose the greatest area of habitat: between 98-130 ha for South West England and 82-140ha for South East England by the 2080s. Coastal floodplain and grazing marsh and saline lagoons lose between 16-58ha and 6-51ha respectively in East England (the greatest loss for a region) by the 2080s. Fen, purple moor grass and rush pasture and reedbed habitat all lose <20ha per region by the 2080s.

- The greatest proportion of habitat lost is projected for saline lagoon habitat; between 2% and nearly 20% of the resource is lost in East England by the 2080s. Between 7-10% loss of reedbeds is modelled for North East England. By comparison, coastal grazing marsh, and fen, purple moor grass and rush pasture, and deciduous woodland all lose less than 1% of the total area of habitat resource, even under the most extreme projection.

Comparison between flooding risk and erosion risk for selected BAP habitats:

43

Those habitats that experience smaller losses are also those habitats that have less total cover in England as well.

Page 96: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

68 Biodiversity Sector Report

- The loss of selected coastal BAP habitats due to increased flooding as a result of sea level rise is estimated to be an order of magnitude greater than the losses from coastal erosion alone. Many of the BAP habitats investigated will not only be impacted by coastal erosion, but by coastal flooding (see BD7 below).

This assessment only used a selection of priority BAP habitats and does not take account of marine-related habitats (e.g. salt marsh, coastal vegetated shingle) that may also be lost due to erosion. In addition, other habitats not included are dry lowland habitats, such as coastal sand dunes, eutrophic standing waters (ponds and lakes) and lowland heathland. The total amount of BAP habitat loss due to erosion is therefore very likely to be higher.

Quantification of habitat losses for Wales, Scotland or Northern Ireland is not available at present. However, in Wales a previous assessment by Pye and Saye (2005) based upon a median range sea level rise projection (0.41m by 2100) suggested that several important sites could experience net loss of dune habitat, notably Morfa Dyffryn, Newborough Warren, Whiteford Burrows and Kenfig. Perhaps equally importantly, the same study highlighted that if sediment supply rates are maintained at current high rates then a net gain may be experienced at some sites, including Laugharne-Pendine, Morfa Harlech and Ynyslas. In Scotland, dunes systems fringing the outer Firths were highlighted by Pethick (1999) as candidate sites that may experience future remobilisation due to increased rates of sea level rise.

(ii) Inundation impacts due to coastal flooding (BD7).

Defra project CR0422, incorporated risk analysis of irreversible loss to selected habitats within the 1:1000 year coastal floodplain of England due to coastal flooding, based upon different climate change projections and defences being either maintained or allowed to degrade. The area at risk of loss increases from approximately 4,600ha by 13.5%44 to 23% to a maximum average loss of approximately 5,600 ha (ranging from 5,160-5,630 ha) or 5.3% of the total area (medium emissions with defences degraded).45

Projected losses due to climate change are therefore likely to be considered significant in the context of both maintaining the national resource of certain habitats and also maintaining the extent of the coastal variations of these habitats. 40% of the Biodiversity Action Plan (BAP) habitats in the coastal floodplain occur within Sites of Special Scientific Interest (SSSI) and 33% of this is within SSSI that also have one or more of the following designations: Special Area of Conservation (SAC), Special Protection Area (SPA) and Ramsar site. Whilst there is some uncertainty over the quality of the habitats that may be lost where they occur outside designated sites, it is expected that in the cases of designated sites, good quality habitat may be lost.

It is clear that there is wide variability in projections of habitat at risk from coastal flooding and climate change in England (and Wales in the case of Lee, 2001). Different approaches have been used and different assumptions and timescales applied. All studies suggest that areas of existing habitat will be lost as a result of coastal inundation exacerbated by climate-induced sea level rise. This may be offset by the development of alternative habitats, in many cases establishing a more natural shoreline transition. Estimates of such habitat ‘gains’ have not been made.

44

2100 medium emissions with defences maintained 45

Note that this is the average loss that could be experienced. This does not translate to the loss that will be experienced every year but rather the average loss that may be experienced at any point in time given the climate change scenario. Actual losses over a year or a decade may be higher or lower.

Page 97: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 69

Analysis of potential habitat losses or gains resulting from coastal inundation has not yet been undertaken for Scotland or Northern Ireland, but risk areas have been identified46, particularly zones which are currently experiencing sediment starvation and therefore prone to accelerated erosion and inundation (e.g. the inner Firths in Scotland).

Estuaries feature a complex interaction of fluvial and coastal influences; changes in this interaction will modify the patterns of erosion and accretion that will affect the current distribution of intertidal habitat. Changes in estuary shape may also influence tidal dynamics: for example, it has been tentatively suggested that erosion and widening of estuary mouths could lead to a shift from fine-grained to sandy sediments in some inner estuaries. If this does occur it is very likely to have a strong influence on the presence of invertebrates and hence on bird populations (Austin and Rehfisch, 2003).

An increase in erosion and saline intrusion would be likely to lead to displacement of the vulnerable areas of machair habitat in North West Scotland highlighted in Section 4.3.2, leading to replacement by saltmarsh, intertidal sand flats or saline lagoons. The combination of biophysical and human factors that produced machair systems over previous millennia means that they have evolved in dynamic situations but there is usually limited scope for inland migration of the system due to the rising topography. Protection in situ therefore usually provides the only option.

4.3.4 Implications for ecosystem services

In addition to their high biodiversity value, coastal habitats and landforms provide a range of ecosystem services, most notably their role in regulating flood and erosion hazards through attenuation of water levels (e.g. Moller et al., 2001). The UK NEA has estimated that saltmarsh could bring a capital cost saving of £2.17 billion on sea defence in the UK. It also calculated the value of coastal wetlands at £1.5 billion annually in terms of the role they play in buffering the effects of storms and in controlling flooding. Active saltmarsh can also provide a valuable contribution to carbon storage and sequestration: for example, Shepherd et al. (2007) have reported carbon sequestration rates of 0.44-1.7t/ha/yr from the Blackwater estuary. Coastal ecosystems also provide important landscape, amenity and recreational benefits that contribute to social and economic well-being. The UK NEA has estimated the total value of ecosystem services from UK coastal habitats to be £48 billion. Hence, the wider societal value of maintaining coastal margin habitats and of creating new coastal wetland through managed realignment schemes can potentially provide cost-effective alternatives to structural approaches to coastal defence, depending on the location.

4.3.5 Implications for adaptation

There remain some significant gaps and uncertainties from studies completed to date. However, all studies projected a significant loss of habitat within the coastal zone. The magnitude of loss will be related to the rate of future change and the associated unplanned and planned adaptation responses, particularly the location of the coastal defence line and any planned realignment.

Analysis suggests that the projected losses vary between habitat type and across geographical regions. Some geographic locations and habitats are particularly important because of the number of BAP priority species that they support. If the results of the analysis are considered solely on the basis of BAP habitat extent within the intertidal area and coastal floodplain then the outcomes are significant by

46

The SNH Shorelook study has broadly identified risk areas In Scotland.

Page 98: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

70 Biodiversity Sector Report

themselves. Hundreds of hectares of coastal BAP habitat are projected to be lost in the medium to long term as a result of erosion, and the losses will be compounded by the effects of coastal flooding. However, a further important consideration is that some locations, particularly eastern coastal estuaries of the UK, are vitally important for biodiversity. The analysis suggests that these areas may be most affected by coastal change. As discussed in BD9 (Section 4.8), eastern coastal locations often have a richer habitat with evidence to suggest that birds prefer to overwinter there during milder conditions.

4.3.6 Further analysis required

Important gaps remain in our knowledge of coastal change. In some cases, basic information on the extent and distribution of habitats could be improved to provide a more consistent UK inventory and baseline. Large-scale modelling of erosion and flood risk needs to be systematically developed at national-scale, with necessary simplifying assumptions tested at local level to better understand sensitivity of habitat change under different projections of future climate change. Further work should also include the specific sensitivities of estuaries in combining fluvial and coastal change, and acting as sites of particularly high biodiversity value. In addition, more emphasis is required on the process of habitat creation and the succession process between habitats under different assumptions of change.

It is also important to note that the risk assessment has not been able to fully evaluate risk BD7 which is characterised by extreme events that could overcome the artificial or natural defences of the coast, causing inundation and large-scale movement of coastal sediments. A full risk assessment would require detailed hydrodynamic, geomorphological and hydroecological modelling of vulnerable locations. Some of this modelling is now being developed to better understand integrated impacts of flood risk on human populations and settlements, therefore it would represent a logical extension to incorporate the dynamic response of natural ecosystems.

4.4 Increased risks from pests, diseases and invasive non-native species (BD3 and BD4)

Summary

There is LOW confidence in the assessment of this risk, due to the high variability between individual species and diseases, and the low evidence base for impacts on biodiversity. However, there is a general high consensus that this risk could significantly increase due to climate change because of the climate sensitivity of many species. In particular, a shift to warmer winters implies a greater survival rate for invasive non-native species and pathogens that have previously been constrained by low temperatures. Much of the risk evaluation to-date has been based upon impacts for human health or productivity in forestry and agriculture, and there is therefore a strong requirement to extend this to a broader evaluation of risks to biodiversity and wider ecosystem services.

Case studies are provided for Parrot’s-feather, Phytophthora ramorum and Chytridiomycosis, although in each case the future change in risk remains uncertain.

Page 99: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 71

l c u l c u l c u

BD3

and

BD4

Increased risks from pests and

diseasesL 1 2 2 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

NB. For legislative and other purposes, reference is made to ‘invasive alien species’ (IAS) or invasive non-native species.

47. Information on Invasive Non-Native Species is maintained by the

GB Non-native Species Secretariat at Fera. A joint initiative between Northern Ireland and the Republic of Ireland has recently established the Invasive Species Ireland website and database.

4.4.1 Introduction

Impacts on biodiversity from invasive non-native species, pests and diseases scored highly in the Tier 1 assessment because of the potential consequences, despite the uncertainties. Here we have grouped them together because some of the background information and consequences are similar. In this context, pests represent either native or non-native (alien) organisms that cause nuisance value through damage to native species or ecosystems. Diseases are micro-organisms (pathogens such as bacteria, fungi or viruses) that cause harm when transmitted to a particular host, such as Dutch elm disease in trees and distemper virus in marine mammals. Invasive species are those that substantially modify an ecosystem and that, by displacing key species and modifying functions, can cause substantial damage to the environment, economy or social systems. The legislative framework particularly highlights the risks brought about by human introduction (deliberate or accidental) through its definition of invasive non-native species (INNS) because preventing initial introduction is the key step from a risk management perspective: after introduction, the risk can become much harder to contain. Risks have the potential to be inter-linked: pests can act as hosts for the transmission of disease, such as ticks spreading Lyme Disease; or grey squirrels (Sciurus carolinensis) transmitting squirrel pox disease to the detriment of native red squirrels (Sciurus vulgaris).

The additional risks from climate change are likely to develop through complex interactions and pathways with human introductions either in the UK or elsewhere. Some commentators (e.g. Hulme et al., 2008) have suggested that the basic division employed by the CBD for INNS based on human introduction may become too simplistic in this context, as, rather than being dichotomous, there is a continuum in the degree of human intention attributable to most pathways. For example, human introduction to one region may be followed by ‘natural’ climate-mediated spread into neighbouring regions. Ultimately, however, this issue emphasises the importance of international collaboration and vigilance to avoid the initial introduction. In addition, it is possible that climate change may mean some apparently benign non-native species that are currently present in the UK (i.e. previous introductions) are enabled to become more ‘invasive’ by further favouring them against native species.

47

Under the definitions of the Convention on Biological Diversity (CBD: as adopted by the Invasive Non-Native Species Framework Strategy for GB), invasive non-native species are identified as those introduced by human agency. Article 8h of the CBD recommends ‘‘each Contracting Party shall, as far as possible and appropriate, prevent the introduction of, control or eradicate those alien species which threaten ecosystems, habitats or species’’. Most European states have a further commitment ‘‘to strictly control the introduction of non-indigenous species (Bern Convention on the Conservation of European Wildlife and Natural Habitats) and both the ‘‘Habitats’’ and ‘‘Birds’’ Directives of the European Union also contain provisions to ensure alien introductions do not prejudice the local flora and fauna.

Page 100: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

72 Biodiversity Sector Report

The severity of the risk at different levels (species, habitat, ecosystem function or ecosystem service) depends upon many interacting variables which cautions against over-generalisation. Examples of such variables include:

Land-use and land management

Host/vector interactions

Population dynamics (birth/death rates, immigration/emigration) of the pest or disease as well as the host

Dispersal ability (by the vector either vertebrate or invertebrate, wind or water borne)

Predator/prey dynamics

Interspecific competition and interaction with similar species

Life cycle complexities (seasonal development and requirements of different life stages)

Abiotic factors (including climate)

Many pests are generalists in their habitat preference which may also give them a competitive advantage (see BD11, Section 4.11).

Invasive non-native species have been highlighted as a particular issue of concern due to recent increases (Figure 4.7). In England, an audit found 2721 non-native species living in the wild (English Nature, 2005) but most of them have not had noticeable negative impacts. A small minority have caused perceptible harm, including Grey Squirrel (Sciurus carolinensis), ‘Sudden Oak Death’48 (Phytophthora ramorum), Japanese Knotweed (Fallopia japonica) and Signal Crayfish (Pacifastacus leniusculus). In the absence of their native ecological communities, the non-native species may also be without native predators, parasites, or competitors.

In the context of climate change, non-native species may become more suited to the changed climate than native species by removing temperature-constraints or competitive ability. For example, Pacific oyster (Crassostrea gigas) was allowed to be farmed in UK estuaries as waters were believed to be too cold for it to reproduce but this species has now spread to adjacent intertidal habitat areas: the eastern Channel coast has been worst affected but the risk is also now apparent in the western Channel and in the vicinity of oyster farms in Scotland (Maggs et al., 2010).

Similar issues apply to native pest/disease species. In some cases these may become more competitive in existing or new environments due to climate change. Other species may be affected through a variety of causes, including competition for resources, predation and grazing, parasitism or pathogenesis, impacts on habitat structure or impact on ecosystem function (alteration of water table, soil properties, food production) (Manchester and Bullock, 2000).

Hybridization may occur between non-native species and related native species (resulting in changes to genetic constitution and in phenotype). This hybridization may modify the ability of native species to adapt to their local environment. A shift in climate can lead to changes in geographical distribution, increased overwintering, changes in population growth rates or generation number, extension of the development season and changes in the synchrony of the pest with other species (Porter et al., 1991) and affect the characteristics described above. As a result of this complexity it is rarely a single variable that will cause an invasive non-native species or pest to increase in destructive ability.

48

‘Sudden Oak Death’ is nomenclature adopted from North America but may not necessarily be descriptive of impacts in the UK. As noted in Section 4.4.2, other trees appear more susceptible.

Page 101: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 73

With regard to the risk from disease, analysis is usually based upon a standard framework as highlighted in the Foresight study ‘Infectious Diseases: Preparing for the Future’ (Wright and Woodhall, 2006), this framework distinguishes disease sources, pathways, drivers, and outcomes. Climate change, one of the key drivers identified, has the potential to impact on disease sources and pathways directly; for example, by changing the range of vector species or increasing disturbance events that facilitate disease. Habitats and communities of hosts may also be modified with changes in climate, thus influencing the diseases that can occur. Arthropod vectors are particularly likely to be impacted by climate, seeing a change in latitudinal or altitudinal range. Ultimately, change may bring unfamiliar hosts, diseases and vectors together, thus bypassing the natural and managed defence systems that have evolved over time (Bayliss, 2006). Most disease outbreaks occur when a disease moves to a new host. However, the potential emergence of novel pathogens through recombination (as exemplified by the avian influenza virus H5N1) was also identified in the Tier 1 CCRA assessment: it is not possible to characterise this risk type in any further detail because of its intrinsic uncertainty.

Figure 4.7 Change in widely-established invasive non-native species for Great Britain: 1960-2008

Source: Defra (2010c)

4.4.2 Current risk status

As noted earlier, generalisations regarding climate sensitivity of biodiversity to this risk are difficult due to the great range of species responses and the many interactions across the ecosystem. Current risk-based assessments for a series of non-native species that could have major ecological consequences are provided in Table 4.4 based upon a standard risk-scoring system developed by Fera (including uncertainty).

Page 102: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

74 Biodiversity Sector Report

Table 4.4 Species currently identified as ‘High Risk’ as defined by the standard risk methodology of the GB non-native species secretariat

Species Current

Risk Status

Entry Establishment Spread Impact Overall

Uncertainty

Water Fern (Azolla filliculoides)

Medium

Sika Deer (Cervus Nippon)

Low

Australian Swamp Stonecrop (Crassula helmsii

Low

Carpet Sea Squirt (Didemnum vexillum)

Low

Zebra Mussel (Dreissena polymorpha)

Low

Chinese Mitten Crab (Eriocheir sinensis)

Low

Japanese Knotweed (Fallopia japonica)

Low

Giant Knotweed (Fallopia sachalinensis)

Medium

Floating Pennywort (Hydrocotyle ranunculoides)

Low

Curly Waterweed (Lagarosiphon major)

Low

Water Primrose (Ludwigia spp.)

Low

Parrot’s-Feather (Myriophyllim aquaticum)

Low

Signal Crayfish (Pacifastacus leniusculus)

Low

Red Swamp Crayfish (Procambarus clarkii)

Low

Topmouth Gudgeon (Pseudorasbora parva); TMG

Low

American Bullfrog (Rana catesbeiana)

Low

Ris

k

High

Med

Low

Low Med High

Uncertainty

Colour coding of each attribute reflects a combination of scores for intrinsic risk and for uncertainty following the schema to the left.

Overall uncertainty for each species is given in the final, right-hand, column.

Full details of the risk assessment for each species (and other of lower risk) available at www.nonnativespecies.org

Page 103: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 75

Climate-related case study examples are provided for an invasive non-native species, Parrot’s-feather, and two diseases, Phytophthora ramorum, and Chytridiomycosis.

Case Study I: Parrot’s-feather (Myriophyllum aquaticum)

Parrot’s-feather is an invasive non-native aquatic plant that originates from South America and is normally deliberately introduced by human agency through horticulture. It may then be discarded accidentally into natural habitats or also possibly transferred by large wildfowl. Dense infestations can develop that exclude native species or cause flooding in slow flowing channels. It has no known natural enemies in the UK and appears to experience very little direct competition from other species. Habitats at most risk are natural ponds and slow-flowing canals or rivers. Although the increased aquatic cover may benefit some species, severe changes to physical and chemical characteristics of water bodies, and the shading out of algae that serve as the basis of the aquatic food chain, can cause serious ecosystem disruption. Observations have shown that in eutrophic (nutrient-enriched) coastal or brackish water conditions it can displace native species. In Guernsey a consequent reduction in native biodiversity has been recorded (David, 2003 in Varnham, 2006). It currently threatens the UK BAP sedge species Cyperus fuscus, which only occurs at six UK sites.

The current distribution of this species is mainly in southern England, possibly reflecting climatic suitability, or availability of propagules, or proximity of suitable native habitats adjacent to more densely populated areas (largest sources). Water chemistry and nutrient conditions do not appear to be important control factors indicating the potential for it to become much more widespread.

In its current southern UK locations, Parrot’s-feather survives most winters but evidence from continental Europe suggests low temperatures and continued exposure to frost/ice in harsh winters are key limiting factors. Inability to store phosphate in rhizomes during winter may also limit its distribution in colder areas with lower nutrient supply, but overwintering in eutrophic ponds is possible as continued phosphorous supply can compensate in the following spring (Sytsma and Anderson, 1993). Significant costs are associated with control of this species, either by mechanical control, manual control or application of herbicides.

Case Study II: Phytophthora ramorum

Phytophthora is a genus of exotic fungus-like pathogens that causes damage to trees, shrubs and other plants. P. ramorum has caused widespread death of trees in the USA, where it is commonly known as Sudden Oak Death. In Europe, including the UK, P. ramorum has been found mainly on container-grown Rhododendron, Viburnum and

CURRENT RISK STATUS is defined by a conditional probability approach, hence high risk status requires likelihood of entry to be ‘medium’ or ‘high’ for other 3 factors to take effect [in sequence: establishment, spread, impact]

ENTRY – Likelihood of introduction

ESTABLISHMENT – Likelihood of suitable conditions that favour persistence

SPREAD – Rate of dispersal and lack of containment

IMPACT – Severity of consequences: ecological, economic, social.

Page 104: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

76 Biodiversity Sector Report

Camellia plants in nurseries49. Many habitats may be suitable for its survival, including woodland and managed gardens, parks and public green spaces (Sansford et al., 2009). The first recording in the UK was made in 2002 and until 2009 fewer than 100 trees had been found with the infection. Usually these trees are found close to infected plants, particularly Rhododendron ponticum. When P. ramorum infects rhododendron it produces an especially large amount of spores (the part of the life cycle that allows the organism to spread and infect other individuals). Most oaks that are infected do not support much spore production because the leaves are rarely infected (bark infections do not produce spores), and it is believed that native oaks are resistant to the disease with very few infections. Therefore, the presence of leaf susceptible hosts contributes highly to the spread and survival of the pathogen (Defra, 2004). In 2009, the first recorded instance, globally, of the infection on commercially-grown larch tree leaves was recorded. Larch produces five times the spores produced on rhododendron, resulting in considerable concern about further increased spread of this pathogen.

For biodiversity, there is serious concern regarding the spread and impact of P. ramorum, and similar species, on bilberry (Vaccinium myrtillus: blaeberry in Scotland and winberry in Wales) and other heathland plants (Sansford et al., 2009)50. Heathland habitats are important for biodiversity and there is currently ongoing work by Fera to understand control options beyond the current strategy of clear felling of woodland (Defra, 2009). The risk level from this disease causing pathogen is currently considered to be high and the possible damage could be extensive if control is not possible. Research in the USA suggests that P. ramorum can be spread by wind, rainsplash and in water courses or by vectors such as dogs, humans and possibly birds. The Forestry Commission, with Forest Research, Defra and Fera, are monitoring spread and damage.51

Studies into the sensitivity of P. ramorum to climate variables, specifically temperature, humidity and water potential, have identified optimal levels for growth at different stages of the organism’s life cycle (Defra, 2004, 2007a; Tooley et al., 2009). Work has also been done identifying impacts of extremes in temperature and moisture (Defra, 2004 and 2007a) and its recovery ability (Tooley et al., 2008). Collectively, these studies indicate that P. ramorum optimally survived at above 93% humidity or between 20 – 30°C. The temperature range that it is able to survive is wide, only temperatures below -25°C and above 40°C killed the organism after short periods of time and although lack of moisture has a greater impact, sporulation still occurred at 63% humidity. These findings suggest that the UK has a very suitable climate.

The potential distribution of P. ramorum has been modelled for Europe using different climate-based risk-mapping approaches (Figure 4.8).

49

Datasheet for Phytophthora ramorum. Fera. Available from: http://www.fera.defra.gov.uk/plants/plantHealth/pestsDiseases/documents/pram.pdf 50

See also http://www.forestry.gov.uk/website/forestry.nsf/byunique/infd-5ubesn 51

See http://www.forestry.gov.uk/website/forestry.nsf/byunique/infd-5ubesn .51

A map of the current outbreak in larch, showing the concentration of reported/suspected infections in SW England and Wales (but also other sites in western UK) is available at: http://www.forestry.gov.uk/website/forestry.nsf/byunique/infd-86ajqa

Page 105: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 77

CLIMEX Match Index

– a method used to match the climate variables from one area (two locations in Oregon and California, USA, where large-scale outbreaks have occurred) to European climate variables (in this case, specifically the UK), to identify areas where P. ramorum might have the potential to flourish.

Risk Ranking method

–uses four weighted environmental conditions (precipitation, mean maximum temperature, relative humidity and mean minimum temperature) to model where the risk of the organism being able to survive is highest. The original work by Meentemeyer et al. (2004) also included a host-species index, but this level of data is not available for the UK.

CLIMEX Ecoclimatic Index

– combines a “Growth Index” (measure of species’ response to temperature and moisture and potential for growth during the favourable season) with four stress indices (hot, cold, wet and dry that describe the probability of the population surviving through unfavourable seasons) to give a measure of suitability of the location for the target species.

Figure 4.8 Risk of P.ramorum derived from three different models Source: Sansford et al. (2009), p. 121.

This modelling suggests that the higher risk areas are in the moister west of the UK, especially South West England and Wales. However, there are often large regional differences between the models, none of which consider the availability of suitable habitat (or host species). The CLIMEX map (right) includes optimal growth climate variables and stress indicators that mean western Scotland is no longer highlighted as a high risk zone, probably due to the colder temperatures. The three models are best considered together, to highlight that the whole of the UK is potentially suitable for P. ramorum, but that the current risk is highest in the moister west and milder south of the UK. This is consistent with the incidence of disease outbreaks to-date. Further work also needs to be extended to include the availability of host species.

Fera also suggests that many other factors act to confound prediction. The complexity and sensitivity of disease-host interaction, differential sporulation between species, transmission, population dynamics, and effectiveness of control measures, all interact to define the extent and impact of P. ramorum.

Case study III: Chytridiomycosis

This is a potentially fatal frog disease now found in the UK, although its full extent is uncertain and the subject of extensive survey work. The causal fungus (Batrachochytrium dendrobatidis) was almost certainly introduced by human agency. It has been identified as a major cause of amphibian extinctions (e.g. golden toad) and catastrophic population declines in many species across North, Central and South America. It seems to be temperature limited (Garner et al., 2005) and climate change

Page 106: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

78 Biodiversity Sector Report

has been recognised as a causal factor enhancing the ability of the fungus to spread and/or induce disease (animals can carry the fungus without showing disease symptoms).

4.4.3 Potential impact of climate change

Case study I: Parrot’s-feather

No future projections are currently available for the potential distribution of Parrot’s-feather. However, one of the indirect effects of increased water temperatures (see BD10, Section 4.9) is likely to be the increased potential for overwinter survival of this invasive non-native species. As low temperatures seem to be a key control (this could be an explanation for its current distribution only in the south of the UK), then it is reasonable to infer that there is a significant risk of the species spreading north and invading more aquatic habitats. Changes in water flow regimes will also be a key influence, as the plant is at its most aggressive in still water. By covering large areas of aquatic habitat, Parrot’s-feather can effectively smother a water body and by reducing light, oxygen and nutrients available to other species then its likely spread has important implications for priority species and habitats.

Case study II: Phytophthora Ramorum

There is little uncertainty that suitable host species and climate are widespread across Europe (Sansford et al., 2009) and encompass the whole of the UK (Defra 2004, 2007a; Sansford et al., 2009). Various stages of the P. ramorum life cycle have also been shown to be correlated with temperature, moisture and extreme and sustained heat; cold and drought can severely damage or kill the organism (Defra, 2004, 2007a; Tooley et al., 2009). However, confounding factors as mentioned previously mean that experts are not certain about the future risk in relation to climate change. Considerable research is currently being conducted on the epidemiology and control of this pathogen. The EU has carried out a standard Pest Risk Analysis (Sansford et al., 2009), to contribute toward EU-wide emergency phytosanitary measures. It is possible that a shift to drier summers with lower humidity levels may limit the spread of this disease but this could also be countered by increased prevalence of milder wetter winters.

Case study III: Chytridiomycosis

Analysis by Bosch et al. (2006) using a 28-year meteorological time-series from a temperate alpine area of Spain has shown a significant association between epidemic years and specific climatic variables. Climate change at that location is moderating the naturally severe cool conditions, with shorter milder winters producing elevated temperatures and humidity values that are believed to favour the fungus. Inter-annual variations in these parameters associated with the North Atlantic Oscillation correlate significantly with disease epidemics, such that the presence of occasional severe winters reduces the prevalence of the disease. The climate ‘envelope’ for this disease is inferred to be expanding across Europe to include much of the UK and an increased frequency of milder wetter winters is therefore associated with an increased risk of outbreaks.

Page 107: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 79

4.4.4 Implications for ecosystem services

In terms of ecosystem services, the greatest impacts from invasive non-native species, pests and disease are likely to be where provisioning ecosystem services, such as food or fibre provision, are affected. As a result much of the information on the level of risk is focussed on pests and pathogens that affect agriculture, forestry and human health, such as Phytophthora ramorum, Lyme disease (Lyme borreliosis) or the veroa mite that impacts on bee populations and pollination services. This does not discount that regulating services (e.g. flow regulation) or cultural services (e.g. amenity value) may also be affected when a large-scale outbreak or infestation occurs. Much less information is available for the wider impacts on biodiversity and ecosystem function.

However, there is growing awareness of the role of biodiversity in providing an ecosystem service by mediating the spread of pests and diseases, particularly vector-borne diseases. A recent review by Keesing et al. (2010) found that reduced biodiversity was associated with increased disease transmission of 12 diseases (including West Nile fever and Lyme disease) across various ecosystems. The reasons for this are not clear but are apparently associated with the ‘dilution effect’ associated with the availability of suitable hosts as vectors for disease (Johnson and Thieltges, 2010). More species-rich communities are likely to support a higher proportion of unsuitable (low competency) hosts for the vector; the vectors encounter more of the low competency hosts reducing the infection of disease from host to new vector and thus fewer possible vectors become carriers of the disease. Keesing et al. (2010) also suggest that maybe species with low rates of reproduction, or that have invested heavily in disease immunity, are lost first when ecosystems are degraded by human actions; conversely species with high reproductive rates or have invested less in immunity, persist to provide an increased number of disease hosts. The same authors also note that there is also the possibility that areas of high biodiversity also could provide a higher probability of pathogen emergence, but studies have provided contradictory results on this issue.

4.4.5 Implications for adaptation

Invasive non-native species, pests and pathogens may be directly affected by climate, their lifecycles may be regulated by temperature or moisture or they might have threshold values under / over which they are not able to survive. Diseases can have complex development cycles, interacting with various hosts and vectors and so the characteristics of one disease may not necessarily be relevant for another. This makes analysing the overall risk in the natural environment very challenging indeed. However, examples such as Parrot’s-feather and Phytophthora demonstrated high natural adaptability with no known enemies. The jump in Phytophthora to commercial forestry and heathland species and the vast increase in sporulation ability on Japanese Larch highlight the potential for risk multipliers. Human actions can exacerbate the prevalence of invasive non-natives, pests and diseases by transfer of infected individuals or lack of control of suitable hosts. The key requirement for adaptation is therefore an increased awareness of the threat in order to reduce the risk of introduction and transmission, combined with high vigilance and surveillance. Modelling studies need to be further developed to reduce uncertainties regarding the main risk areas. Natural England has recently piloted a rapid screening technique for invasive non-native plants (Commissioned Report NECR053: Thomas, 2010).

The Invasive Non-Natives Species Framework Strategy for Great Britain recognises that the most cost-effective and least environmentally damaging approach to solving the problems caused by invasive non-native species is through prevention of introduction into the wild, rapid response and early intervention.

Page 108: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

80 Biodiversity Sector Report

As explained in Section 5.4, non-climate factors such as technological innovation, transport and trade of natural resources, and international regulation, will also have a fundamental role in determining which of many other potential risk species is elevated to critical level (Tait et al., 2006). This risk therefore clearly requires a cross-sectoral approach. Important interactions are recognised with other CCRA sectors, notably FORESTRY and AGRICULTURE, but also WATER and FLOODS for aquatic species, and HEALTH due to changes in human exposure. Analysis of risk metrics in the Forestry CCRA sector report (Moffat et al., 2012) were developed for red band needle blight and green spruce aphid with regard mainly to impacts on productivity.

4.5 Species unable to track changing climate space (BD5)

Summary

There is a high level of consensus and a good evidence base that range shifts are occurring and will most likely be further manifest in the future, and that the current landscape will constrain many species from tracking these range shifts. The recent Lawton review has highlighted the magnitude of these impacts, with considerable observational evidence of the consequences of habitat loss and fragmentation in hindering natural adaptive responses. Additional microclimatic effects also affect this response, particularly due to the influence of altitude in mountain areas. Confidence regarding this risk is HIGH although considerable further work is required to develop an improved knowledge base on the mechanisms by which different species change their range. This will allow the additional influence of species interactions to be included within risk assessments and a more complete assessment for different ecosystems.

l c u l c u l c u

BD5Species unable to track

changing 'climate space'H 1 2 3 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.5.1 Introduction

The geographical range of a species is associated with climate parameters that broadly define a ‘bioclimate envelope’ or ‘climate space’. This concept can be used, with caution, to infer the potential large-scale impacts of climate change. A warming climate generally implies a northward shift in species distribution so that they can maintain their climate space. In addition, at local level some species are adapted to particular microclimate conditions (e.g. humidity, shade, or wind) therefore smaller-scale changes can also be important. For example, in mountainous areas, an altitudinal gradient can be important in influencing local ecological variations [see Box on ‘Topography and Altitudinal Gradients’].

However, the survival of any species also relies on biotic factors such as:

Available and suitable habitat

Community and trophic (predator-prey) interactions with other species

Page 109: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 81

Reproduction and life cycle

Ability to disperse or transit to new locations

Genetic adaptability and the ability of a species to modify its key traits (‘phenotype plasticity’) in response to changes in the environment.

For these reasons, bioclimate envelopes can only provide an indication of the vulnerability of species to climate change. Species most likely to be affected by a change in climate space are those where the climate effect is compounded by the biotic factors, notably those that have limited dispersal ability, limited genetic adaptability, and occupy rather specialised ecological niches (see BD11, Section 4.11). However, the main factor defining this risk is the barriers to movement and dispersal imposed by human modification of the landscape through habitat fragmentation.

4.5.2 Current risk status

Observed changes in species’ ranges have already been recorded from the UK, and in some cases linked to a changing climate. As a broad generalisation across a range of taxonomic groups, warmth-loving southern species are expanding their ranges northwards and cold-adapted northern species are retreating at their southern limits (Warren et al., 2001; Hickling et al., 2006; Franco 2006; Morecroft et al., 2009). However, variations in species response within taxonomic groups can exceed those between groups, indicating the key role of ecological traits associated with habitat requirements, dispersal ability and life cycle (Hickling et al., 2006). In particular, species that lack the ability to disperse through landscapes with discontinuous habitat fragments have been unable to shift their range. In some cases (e.g. amphibians), species ranges have therefore contracted. At smaller scales, species movements to find suitable microclimate have also been observed: for example, studies of the silver spotted skipper butterfly (Hesperia comma) suggest a shift in habitat preference towards cooler, taller grasslands (Davies et al., 2006).

A recent review of the network of ecological sites in England by Lawton et al. (2010) evaluated its suitability to cope with climate change. It highlighted the current fragmentation of the network which has produced sites that have rather limited connectivity and are too small to accommodate change. Outside of the site network, land use has often been specialised to meet the requirements of agriculture or forestry, with reduced capacity to support biodiversity, especially in lowland areas (Hannah et al., 2007; Foresight Land Use Study, 2010). There are often physical barriers to movement of species due to built infrastructure. As a consequence, although species’ range shifts have been recorded, there are many counter examples that highlight that other species have not been able to adapt to change, whether due to climate, land use or other drivers.

4.5.3 Potential impact of climate change

Detailed analysis of changes in climate space for a range of species has been carried out by several studies using bioclimate models. MONARCH3 produced simulated climate space maps for current and future species distributions using the UKCIP02 scenarios (Berry et al., 2007; Walmsley et al., 2007). From this analysis, 32 UK BAP priority species which have provided robust validation against observed European distributions are highlighted in Table 4.5 using four categories; Gain, Loss, No Change and Shift. The results need to be interpreted with caution and are indicative of the changing climate space rather than actual distributions, but they provide a summary of expected general trends.

Page 110: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

82 Biodiversity Sector Report

The BRANCH project successfully modelled and validated 386 species, primarily vascular plants. Table 4.6 presents an assessment of the likelihood, based on climate space alone, of species expanding in range within the UK, or moving into the UK from continental Europe (Berry, 2007 unpublished). A third of the species in Table 4.6 are not native which could have important implications for the management of the ecological network in the UK. A further assessment of likelihood of extinction is provided in Table 4.7 based upon those species that are projected to lose 90% or more of their UK climate space and for which the future bioclimate envelope is not projected to overlap with the current envelope.

A series of other studies have also used bioclimate envelope approaches to model changes in climate space and all have reported simulated range shifts northwards (e.g. Araújo et al., 2006; Huntley et al., 2008). This approach has been used to provide indicative maps and risk profiles in Europe for both birds and butterflies (Huntley et al., 2007; Heikkenen et al., 2010). Although the simulations have only covered a proportion of the 1,150 BAP species within the UK, the general finding is that most species would be required to change their range due to climate change.. Range shifts may also occur for non-native species highlighting the risk from invasive non-native species reaching the UK (see BD3/4 – Section 4.4).

Hill et al. (2002) analysed distribution records for 52 British butterfly species and found that species with a northern and/or upland distribution disappeared from low elevation sites and colonized sites at higher elevations during the 20th century. Modelling of future distribution of 35 species for the period 2070-2099 using contemporary climate change models (i.e. circa 2002) suggested that most northerly distributed species would have little opportunity to expand northwards and would disappear from areas in the south, resulting in reduced range sizes. Southerly distributed species would have the potential to shift northwards, resulting in similar or increased range sizes. However, 30 of the 35 species modelled are thought to have failed to track recent climate changes due to a lack of suitable habitat, and when this is taken into account, the model projected 65% and 24% declines in range sizes for northern and southern species respectively.

Although validation work has shown a reasonable correspondence with current observations of change for some species (Araújo et al., 2005; Green et al., 2008), bioclimate envelope modelling is not an equally-valid approach for all species. It is less appropriate for those that are rare or endemic with limited current distributions. Current distributions of some species may not be primarily climate-related but the result of centuries of land use change. In addition, as the range of species change, different effects tend to dominate at opposing margins, with the cold (or uphill) margin influenced by tolerance to a stressful climate and the warm (or downhill) margin dominated by biotic interactions (Brooker, 2006). Furthermore, it has also been suggested that for mountain areas the envelope models could significantly underestimate the potential change due to the aggregation of climate data at coarse-scale resolutions and potential disregard of the intraspecific adaptations that montane species have developed to their local environment (Trivedi et al., 2008). Ecological complexity and the variety of species adaptive responses means that general projected responses are often contradicted by counter examples (Willis and Bhagwat, 2009).

Page 111: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 83

Table 4.5 Overview of general results from MONARCH3 based upon projected changes in UKCIP02

Category GAIN LOSS NO CHANGE SHIFT

Definition of category

Species that gain substantial potentially suitable climate space and show no significant loss.

Species that show significant loss of potentially suitable climate space and no significant gains.

Species that show no significant gain or loss of climate space.

Species that show a shift in potentially suitable climate space.

No. of species 15 8 3 6

Names of species birds: stone-curlew, corn bunting, turtle dove; butterflies: pearl-bordered fritillary, marsh fritillary, silver-spotted skipper, heath fritillary, Adonis blue; mammals: greater horseshoe bat, lesser horseshoe bat; plants: stinking hawk’s-beard, red-tipped cudweed, broad-leaved cudweed, red hempnettle, small-flowered catchfly.

Birds - skylark, common scoter, black grouse, capercaillie, song thrush; Plants - Norwegian mugwort, twinflower, oblong woodsia.

Birds Tree sparrow, linnet Plants shepherd’s needle

Invertebrate - stag beetle; Mammal - Barbastelle bat; Plants - tower mustard, cornflower, cut-grass, floating water

plantain.

Description All 15 species have a southern distribution in Britain and Ireland at present and are at the northern edge of their ranges.

Six have predominantly northern distributions within Britain and Ireland and, within the UK, their strongholds are Scotland or upland habitats. Five of these species are at risk of losing all climate space and becoming extinct within Britain and Ireland.

These species show no significant change because their present climate space covers most of Britain and Ireland; they are also widespread within Europe.

They all show a northward shift in potentially suitable climate space within Britain and Ireland.

Source: Walmsley et al. (2007)

Page 112: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

84 Biodiversity Sector Report

Table 4.6 The likelihood of European species gaining over 50% new potential suitable climate space from 2020s to the 2080s UKCIP02 HIGH scenarios

Likelihood of expansion Common name

Exceptionally unlikely Wild service tree, Turtle dove

Very unlikely Silver-studded blue, Dormouse, Whitebeam, Chalk milkwort, Wood spurge

Unlikely Yellow-bellied toad, Grey-headed woodpecker, Mastic tree, Valonia oak, Narrow-mouthed whorl snail, Marsh gentian, Lady's mantle

As likely as not Southern damselfly, Shrubby seablite, Red-tipped cudweed, Narrow-leafed ash, Oleander, Aleppo pine, Stone curlew, Flowering ash, Olive, Wheatear, Portuguese oak

Likely Hop hornbeam, Spiny broom

Virtually certain Bristle bent, Midwife toad, Purple emperor, Creeping marshwort, Box, Nightjar, Dwarf sedge, Chequered skipper, Cetti's warbler, European fan palm, Steppe grasshopper, Zitting cisticola, Western whip snake, Lily of the valley, Stinking hawks beard, Middle spotted woodpecker, Black woodpecker, Little egret, Reed bunting, Herb Robert, Silver-spotted skipper, Icterine warbler, Wryneck, Sand lizard, Wood white, Wood lark, Large copper butterfly, Adonis blue, Chalkhill blue, Marbled white, Granville fritillary, Heath fritillary, European bee-eater, Common vole, Swallowtail, Herb Paris, Wall lizard, Oxlip, Downy oak, Agile frog, Pool Frog, Greater horseshoe bat, Lesser horseshoe bat, Shore dock, Fire salamander

The following likelihood terminology has been used in these tables. These are based on the IPCC likelihood terminology (IPCC, 2007). Likelihood terminology

Virtually certain Very likely Likely As likely as not Unlikely Very unlikely Exceptionally unlikely

Likelihood of the occurrence of outcome

>99% probability 90-99% probability 66-90 %probability 33-66% probability 10-33% probability 1-10% probability <1%

Source: Berry (2007)

Page 113: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 85

Table 4.7 The likelihood of extinction of species based upon loss of climate space from 2020s to the 2080s UKCIP02 HIGH scenarios

Likelihood of extinction

Common name

Virtually certain Hawkweeds, Shetland pondweed, Woolly willow, Whorl snail

Likely Bittern, Red-tipped cudweed, Slender naiad

As likely as not Meadow pipit, Twite, Pied flycatcher, Narrow-headed ant, Dune gentian, Wryneck, Red-backed shrike, Twinflower, Fen orchid, Small cow-wheat, Wheatear, Grey partridge, Narrow-mouthed whorl snail, Oblong woodsia,

Unlikely Yellow-necked mouse, Stiff sedge, Prickly Sedge, Scottish scurvy-grass, Scottish wood ant, Black-backed meadow ant, Wood crane's-bill, Sea pea, Lax-flowered sea-lavender, Red-necked phalarope, Roseate tern, Capercaillie, Round-mouthed whorl snail

Very unlikely Water vole, Reed bunting, Southern wood ant, Common scoter, Bird cherry, Yellow marsh saxifrage

Exceptionally unlikely Bullfinch

The following likelihood terminology has been used in these tables. These are based on the IPCC likelihood terminology (IPCC, 2007). Likelihood terminology

Virtually certain Very likely Likely As likely as not Unlikely Very unlikely Exceptionally unlikely

Likelihood of the occurrence of outcome

>99% probability 90-99% probability 66-90 %probability 33-66% probability 10-33% probability 1-10% probability <1%

Source: Berry (2007)

4.5.4 Implications for ecosystem services

Differential shifts in species’ ranges will result in changes to species interactions and community composition. This dynamism is an essential component of ecosystem resilience allowing adaptation to change. However, if a species that is integral to ecosystem integrity and functioning is lost then the system becomes vulnerable to major disruption or loss. This has implications for ecosystem services, including any provisioning or culture value that an individual species may have supplied in the affected area.

4.5.5 Implications for adaptation

The inability of species to disperse and track their changing climate space will be highly likely to lead to the loss of biodiversity, particularly where species that have specialist habitat requirements are outcompeted by species that are more ‘generalist’ (see BD11, Section 4.11). The necessary planned adaptation response, as highlighted by Lawton et al. (2010) is to develop a more coherent and resilient ecological network that features sites that are larger, better managed and more connected to facilitate species movement. However, measures to improve the general ‘permeability’ of the landscape beyond the ecological site network will probably be equally important in an uncertain future. Upland areas that have a wide variety of local climatic gradients will also be particularly important in maintaining core zones of high habitat and species diversity. For some montane species that already occupy mountain-top locations, a loss of their current niches may imply that translocation is the only remaining option. Measures to enhance landscape connectivity may have the secondary consequence of improving permeability for invasive non-native species, pests and diseases (Section 4.4), emphasising the need for greater vigilance and risk management of these species.

Page 114: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

86 Biodiversity Sector Report

Topography and altitudinal gradients

Topography is a key influence on local climate patterns. Hence in mountainous areas (notably Scotland, Wales, northern England), an altitudinal gradient can be important in determining the local distribution of a species because average temperatures reduce with altitude (Berry et al., 2005, Mitchell et al., 2007, Hopkins et al., 2007). With regard to climate change, this can produce local-scale variations to the more general latitudinal (i.e. south-north) shifts in climate space. Species will generally respond to a warming climate by moving up an altitudinal gradient, either retreating from stressful warm conditions (including increased competition) or expanding into areas that were once too cold for their survival. Research by Franco (2006) indicates that local climate warming has been of comparable importance to habitat loss in driving local extinctions of northern species of butterflies in Britain over the past few decades. Future modelling in the MONARCH project indicated that for seven out of 12 of the species investigated, their climate space will move with altitude and their abundance would shift up-slope, especially under the higher emissions scenarios.

However, climatic variables other than temperature may be more important for some species: for example, Crimmins et al. (2011) have highlighted observations in North America of species moving downhill in response to climate change, due to the changing water balance being the dominant influence (see Section 4.2). Aspect can modify temperature gradients as south-facing slopes intercept more solar radiation than north-facing slopes. Wind exposure can also provide a major restriction on the upward expansion of some species. Climate may also be a lesser influence: Britton et al. (2009) found that lichen species richness declined in the alpine zone of Scotland, but that this response was probably due to the effects of nitrogen pollution. At the downhill edge of a species range, biotic interactions may become the driving force, rather than the physiological stress imposed by climate at its upper limit (Brooker, 2006).

In areas of high topographic diversity, where high mountains are dissected by deep valleys, a wide variation of microclimatic conditions exists (especially at locations that are also adjacent to the coast). This provides a wide variety of ecological niches and these areas will therefore continue to be areas of high biodiversity value. However, it is highly likely that some BAP species and habitats in upland areas will be highly vulnerable to change because they are unable to track their climate space uphill due to unsuitable soils or unfavourable exposure. For montane species that currently inhabit the higher areas of the mountains and specialist environments such as snow-bed communities, they simply have ‘nowhere to go’. In these cases, it appears that both natural and managed adaptation options are severely limited, and translocation to montane areas further north (in the UK or elsewhere) may be the option of last resort.

Page 115: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 87

4.6 Climate mitigation measures (positive/negative) (BD6)

Summary

This impact is not directly caused by a direct change in climatic conditions but indirectly through responses to reduce the human greenhouse gas (GHG) emissions that cause anthropogenic climate change. There is a rather low evidence base for this risk (mainly because of its new status) and hence a limited consensus on the potential consequences, especially as these are highly dependent on location and the current land uses that the schemes are replacing. It is therefore assigned LOW confidence in the assessment. There is a need to ensure that more information is collated on impacts at the landscape scale in addition to site level, and that impacts are assessed for both present and future climate change. These emerging findings should then be incorporated within procedures for Environmental Impact Assessment.

l c u l c u l c u

BD6Climate mitigation measures

(positive/negative)L Too uncertain

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.6.1 Introduction

Climate mitigation measures are, of course, an essential component of climate change policy and will in the long term help reduce the impacts on global biodiversity. Climate mitigation policy is already acting as an important driver for land use change in the UK and this change will almost certainly become more apparent in the landscape through plans to meet ambitious GHG emissions reductions targets. While this is a very different risk it is one highlighted by stakeholders as an issue which should be considered within the CCRA, although it is also important to recognize that with good planning there could also be benefits for biodiversity.

Mitigation measures presently identified as being of most concern to the biodiversity sector are renewable energy projects (although other schemes are also likely to have an impact in the future). Such projects include wind turbines, tidal schemes, hydroelectricity and bioenergy sources. These are promoted by government policy and economic incentives, such as Renewable Obligation Certificates and Feed-In Tariffs to contribute to GHG emissions reduction targets.

Like any development on sites with biodiversity value, renewable energy projects can have a negative effect on priority species and habitats, depending on the technology and the location. Wind turbines, for example, are known to cause deaths among some species of birds and bats, notably due to the direct consequences of collision with the turbine blades. Hydroelectricity schemes primarily affect riparian and fluvial habitats by altering flow regimes. Studies of biomass crops have shown both positive and negative impacts on biodiversity depending upon both crop and its management. The direct impact of wave and tidal generators will be on marine or estuarine habitats. The challenges of finding balance in the trade-off between mitigating future climate impacts and localised impacts on biodiversity was particularly apparent in response to the

Page 116: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

88 Biodiversity Sector Report

proposed Severn Barrage tidal energy scheme52 and its potential effect on the biodiversity of the Severn Estuary, which is a designated SSSI, SAC, SPA and Ramsar site (Black and Veatch, 2007). A Defra project is currently underway to review the evidence and provide an assessment and suggestions to help formulate integrated policies for low carbon energy and biodiversity under different scenarios for delivery of greenhouse gas emissions cuts.

4.6.2 Current risk status

The development approval process for proposed energy generation schemes includes an assessment of the potential impacts (including cumulative effects) on biodiversity. However, the identification of this as a risk within the CCRA reflects concern from stakeholders that, despite the Environmental Impact Assessment (EIA) process, alternative energy schemes are being approved, or may be approved in the future, that could have a detrimental impact on species, habitats and ecosystems. In response to EU Directive 2009/28/EC, the UK has committed to sourcing 15% of all of its energy from renewable sources by 2020, which means that at least 35% of electricity will need to be generated by renewable sources. Hence, it is likely that there will be an increasing number of sites proposed for development.

Evaluating the risk posed to biodiversity from alternative energy projects at a national scale is difficult because of a scarcity of suitable examples and their recent appearance in the landscape. The only alternative energy source to be developed at something approaching national scale is wind energy and there has been little research on the effects of wind turbines on biodiversity at anything but the local scale or a single breeding cycle (Pearce-Higgins et al., 2009). A significant problem faced by such research is that it is often difficult to ascertain whether any changes observed are due to the impact of the alternative energy scheme itself, or other factors, such as changes in climate, or natural fluctuations in population size.

The majority of onshore large wind farms are currently located in windier upland landscapes with semi-natural habitats. Concern has been expressed about the potential impacts of wind farms on a range of species, notably birds and bats, in a number of ways, such as collision, displacement due to disturbance, acting as a barrier to movement, and habitat loss. The potential impacts on birds, particularly from blade-strike, have been recognised for many years, although awareness of danger to bats (notably due to lung damage from the drop in air pressure near turbines) has been relatively recent and previously underestimated (SNH 2000; Drewitt and Langston 2006; Natural England 2009). In addition, developments on organic soils have implications for carbon storage and climate change mitigation (Holden et al., 2006; Nayak et al., 2008).

The Royal Society for the Protection of Birds published a ‘bird sensitivity map’ for Scotland in 2006 to assist developers and local planning authorities select wind farm sites that were less likely to have an impact on bird populations. Areas of relative sensitivity were identified mainly by buffering known breeding distributions of priority bird species incorporating foraging ranges, collision risk and sensitivity to disturbance (Bright et al., 2006). The map, shown in Figure 4.9, indicates that wind turbines, if constructed, are more likely to impact upon bird populations in North West Scotland, with the Highlands, Western Isles and Northern Isles being particularly sensitive.

52

A feasibility study, published in 2010 concluded that there was not a strategic case for public investment in this scheme at the time (DECC, 2010)

Page 117: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 89

Figure 4.9 Bird sensitivity map for Scotland at 1km scale Note: Sensitivity is based upon the distribution of 16 species of bird of conservation importance

and the distribution of Special Protection Areas (SPAs) Source: Bright et al.(2006)

The other renewable energy source that has expanded significantly recently is bioenergy (although as yet with a small land footprint at UK level), particularly through the development of energy field crops, short rotation coppice (SRC) and forestry by-products. The impacts of bioenergy crops have been reviewed by Booth et al. (2010) who highlighted inter alia the following issues:

Impacts on biodiversity depend on the use of native or non-native species, and the provision of additional habitat diversity.

Removal of brash and coarse wood material from existing forestry removes habitat, ground cover and food resources for small invertebrates, mammals (e.g. bats) and fungi leading to changes in community structure and loss of species which utilised this material. The UK Forestry Standard and its Forests and Biodiversity Guidelines provide recommendations for good forest management practice to retain deadwood. It is likely that adherence to these recommendations would be a sustainability requirement for biomass.

Microbial biomass and most soil fauna groups increase in abundance and diversity after afforestation, particularly decomposers, compared to arable land.

Forestry or SRC expansion is likely to be at the detriment of rare birds adapted to open habitats. However SRC in farmland can provide increased

Page 118: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

90 Biodiversity Sector Report

cover and food for rare species, notably through increased invertebrate diversity.

Hence, the consequences for biodiversity strongly depend on the land use that it replaces and the implications for changes in habitats and species diversity (Haughton et al., 2009). Hence, beneficial effects may accrue from the development of schemes in the lowlands that enhance farmland diversity, whereas there is a risk of increased ecological disruption in the uplands where semi-natural habitats are most common.

4.6.3 Potential impact of climate change mitigation

The planning system53 does take into account the current distribution of priority species (CLG, 2007), although no explicit consideration is given to potential future distribution of species following climate change. National planning policy includes a requirement for local planning authorities to take into consideration “the effect of development on biodiversity and its capacity to adapt to likely changes in the climate”. This requirement is however confounded with respect to future distribution of species by the difficulty of predicting what the effects of climate change on a particular species population might be (as discussed in BD5, Section 4.5) over the lifetime of a scheme (wind farms, for example, are typically given planning approval for a duration of 25 years).

Although it is expected that a large proportion of additional wind power will come from new offshore wind farms, this still implies there will be a need to construct more onshore wind farms. As shown in Table 4.8 and Figure 4.10, the total capacity of wind energy projects that have been consented but not yet operational exceeds the capacity of projects that are already operational. However, the planned capacity increase does not equate to a doubling of the number of wind turbines as new turbines tend to be much larger in capacity (and size) than older turbines. It should also be noted that all of the consented projects have undergone EIA (or environmental appraisal for smaller schemes).

Table 4.8 Onshore wind energy generation in the UK (MW) in 2010

Stage England Scotland Wales Northern Ireland

Total

Operational 795 2,364 380 310 3,848 Under Construction 157 882 35 30 1,104

Consented 1,172 1,796 234 288 3,490 In Planning 1,088 4,125 1,323 794 7,331

Source : British Wind Energy Association, http://www.bwea.com/

The UK Biomass Strategy54 identifies the potential

to increase (by 1 million tonne dry wood) recovery from currently unmanaged and managed woodlands and other sources of wood waste products.

to increase the area dedicated to energy crops to 350,000 hectares by 2020 bringing the total land availability for biofuels and energy crops to 1 million hectares (17% of UK arable land).

As many of the current biomass schemes have been developed only on a small scale, then it is not certain how these issues will scale up to landscape level. A potential issue related to expansion is that, due to the increased transpiration demands of fast-growing 53

Note that the planning system is in a period of significant change. CLG published the draft National Planning Policy Framework for consultation in August 2011, which sets out principles that local councils and communities must follow to ensure that local decision making is consistent with nationally important issues, including climate change. 54

A new Bioenergy Strategy is currently being developed and is expected in late 2011.

Page 119: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 91

trees, soil moisture deficits may be further increased (see BD1: Section 4.2) in combination with the direct effects of climate change, to the detriment of local habitats and species.

Figure 4.10 Current and planned wind farms in the UK Source British Wind Energy Association http://www.bwea.com/ukwed/google.asp

4.6.4 Implications for ecosystem services

Bioenergy sources such as the growth of biomass crops are an important ecosystem service, whereas other sources of renewable energy such as wind, tidal and hydro all have important indirect interactions with the natural environment. These interactions clearly indicate the need to adopt a systems-approach in considering the sustainable supply and demand for renewable energy sources, relative to the distinctive features and assets of the local environment.

Page 120: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

92 Biodiversity Sector Report

4.6.5 Implications for adaptation

Although robust systems are currently in place to assess the potential ecological impacts of proposed renewable energy schemes, the scale of future expansion suggests that the cumulative effects of many schemes could become a problem and therefore these systems may need a strategic review. In particular, as risk BD5 (Section 4.5) has identified, many species will require additional space beyond the existing fragmented ecological network to track their changing climate space. There is a possibility that new energy schemes will compromise this requirement further and introduce additional competition for land (e.g. with agriculture) that hinders the development of robust habitat networks.

The risk here is therefore currently characterised by a combination of ‘unknowns’: lack of a spatial strategy for future renewable energy expansion, and uncertainty over the cumulative effects of many schemes and distribution of future biodiversity. Evidence from current schemes is still being appraised over timescales suitable to derive robust conclusions, with the impacts of wind farms on bats and birds being major sources of concern to-date. With good strategic planning and management, and development of appropriate schemes in the right locations, then this aspect of change could become an opportunity for adaptation. This would also include the additional benefits for climate mitigation through maintenance and additional sequestration of soil organic carbon (BD8).

4.7 Changes in soil organic carbon (BD8)

Summary

A qualitative assessment is made for this risk which is a key measure of the content of organic matter in soils. This organic content is a key regulator of ecosystem dynamics, including the supply of nutrients and water, and the release of the greenhouse gas CO2.

There is a medium evidence base for this risk but this evidence is often contradictory, partly as a result of different analytical methods. Hence the level of consensus on current and future change is low. Assessment of this risk therefore has LOW confidence because of fundamental uncertainties regarding the key processes. More systematic research is required to monitor and model these processes across a range of land use systems.

l c u l c u l c u

BD8 Changes in soil organic carbon L 1 2 2 1 2 3 1 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.7.1 Introduction

There is limited evidence for the responses of soil systems to climate change although it is considered a major threat for multiple soil processes and properties (Towers et al., 2006; Schils et al., 2008). Any threat to soils should be a major source of concern since soils maintain effective functioning in all terrestrial ecosystems and support the delivery of many vital ecosystem services (e.g. crop production, water regulation, carbon

Page 121: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 93

sequestration, water purification, etc.). Soils are also an important global reserve for biodiversity, including many BAP species, such as fungi, bees and ants.

Soil organic carbon is a surrogate measure for soil organic matter which is a fundamental ecosystem property produced and maintained by the dynamic interactions of soil biodiversity with vegetation, climate and other drivers. In turn soil organic matter is the primary energy source for soil organisms and therefore changes in the composition or amount of soil organic matter can in turn alter the composition and activity of soil organisms. All components of soil biodiversity are considered to be at risk from climate change. Although there is little field-based evidence, it has been suggested that the most vulnerable groups are soil bioturbators, species-poor macro-faunal shredders of organic matter, specialized bacteria such as nitrifiers and nitrogen fixers, and plant-symbiotic mycorrhizas (Brussaard et al., 1997). Impacts on these components may also act to constrain the maintenance and development of above-ground habitats and biodiversity, and to reduce the ability of soils to support associated ecosystem functions, in particular the role of soils in mitigating climate change.

As highlighted in BD1 (Section 4.2) and the recent report of the IUCN Commission (Bain et al., 2011), UK peatlands are particularly at risk from climate change. Although a reduction in climate suitability may prevent new peat formation, it is not clear how it will affect the survival of existing peat. Evidence from lowland peat suggests that drying under warmer temperatures will lead to losses of soil organic matter either through erosion or decomposition. The areas most vulnerable to climate change are those where the peat has been degraded through loss of vegetation and structure. Field research in Europe indicates that the restoration of degraded peatlands may take decades to re-establish carbon sequestration and that climate change may extend this delay significantly (Samaritani et al., 2010).

The paleoenvironmental record clearly shows that soil organic matter is sensitive to changes in climate. However relationship between soil carbon and climate are very complex (Figure 4.11) and typically confounded by other factors (Billett et al., 2010). This complexity is well-illustrated by the continuing discussion on recent changes in soil organic carbon, as described below.

Figure 4.11 Conceptual model of changes in soil organic carbon linked to climate

Source: CLIMSOIL project (Schils et al., 2008)

Page 122: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

94 Biodiversity Sector Report

4.7.2 Current risk

Soil organic carbon (SOC) for Great Britain was estimated by Dawson and Smith (2007) to be 9800±2400Mt (6900Mt in Scotland and 2800Mt in England and Wales), although recent research suggest this may be an overestimate. Hence, recent reappraisal of SOC in peatlands (blanket bog, lowland raised bogs and fen) has estimated that they contain 2300Mt of carbon, of which the majority, 1620Mt is in Scotland (Bradley et al., 2005; Chapman et al., 2009; Billet et al., 2010). SOC in Northern Ireland was estimated at 400Mt, with 42% of this found in peat soils (Cruickshank et al., 1998). By comparison, it has been estimated that the vegetation carbon stock of Great Britain is significantly smaller, approximately 113.8±25.6Mt (Milne and Brown, 1997), with vegetation in Northern Ireland containing an additional 3.8–4.4Mt (Cruickshank et al., 1998, 2000). This huge carbon store in peat puts into context the importance of risk to blanket bog peatland habitat as discussed in BD1 (Section 4.2).

A study by Bellamy et al. (2005) suggested that there was a decline in soil organic matter in all soils of England and Wales, particularly in peat soils from the 1970s to 1990s at an average rate of 0.6%/yr. This study therefore predicted an estimated loss of UK soil carbon in the region of 13Mt per annum. However, no more than 10% of losses described by Bellamy et al. (2005) have been attributed to climate change (Kirk et al., 2010; Smith et al., 2007a). In reviewing current evidence and model projections, Smith et al. (2007a) suggest that land use change and management have been the more significant historical drivers of change in soil organic matter while climate change will likely become more significant over time. Several subsequent large-scale surveys have not found equivalent consistent losses across UK or Europe from the 1970s to the present day (e.g. Carey et al., 2008; Kirby et al., 2005; Hiederer and Durrant, 2010). A lower projected loss of ~0.1%/yr due to direct climate change would be consistent with work reported elsewhere (e.g. Ise et al., 2008). Other work even suggests UK and European soils are a net carbon sink (Janssens et al., 2003). Important differences may exist between land use types: evidence from UK field surveys have not corroborated the losses recorded by Bellamy et al. (2005) for grasslands, semi-natural and native habitats but do corroborate losses from agricultural soils (Emmett et al., 2010). This highlights an urgent need to resolve the compatibility of large-scale surveys, and associated historical data, which have used different sampling and analytical strategies.

Soil erosion is also an important issue (both water and wind erosion), especially during extreme events (Lilly et al., 2009), and may be linked to greater rainfall intensity in recent decades (Jenkins et al., 2009). However, increased erosion rates can also be indicative of changing land use and management practises (e.g. more autumn-sown crops). In the long term, erosion in excess of soil formation rates is likely to reduce soil functionality whilst the sediment in runoff can damage aquatic habitats.

4.7.3 Potential impact of climate change

Impacts will be determined by multiple climate and non-climate factors interacting at different spatial and temporal scales. Although increased atmospheric CO2 concentrations could potentially increase plant productivity and hence biomass and SOC, a general rise in temperature is likely to increase biomass decomposition rates and conversion of SOC to CO2 emissions and dissolved organic carbon (DOC) in water. In wetland areas, the drier summers projected by UKCP09 imply decreased CH4 emissions due to reduced anaerobic conditions. Increased dryness would also suggest a greater susceptibility to loss of soil organic carbon through fire (Ciais et al., 2005) – see BD12, Section 4.12. Conversely, the trend to wetter winters, particularly following drier summers would increase the potential for soil erosion, especially in peatland

Page 123: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 95

areas due to instability and even peat slides. More intense and/or sustained rainfall events in combination with other factors such as land cover change (e.g. more autumn-sown crops; conversion of grassland to arable) is likely to lead to changing patterns of soil formation and erosion, but interpretation of this is confounded by inherent climate variability and long-term lag effects from past land-use change and atmospheric pollution. A key parameter is changes in soil moisture levels, particularly through feedbacks with vegetation and CO2, but this parameter is currently inadequately represented in climate models and hence in future projections (e.g. not included in UKCP09).

The current state of knowledge regarding SOC response to climate change is summarised in Table 4.9. Uncertainties over the present-day interactions between climate change, atmospheric pollution and soil functioning (including soil C dynamics) limit the capacity to predict how DOC fluxes, GHG emissions or SOC stocks will be affected by future climate change. There is some evidence that elevated CO2 levels have increased soil organic matter in forest soils (Lichter et al., 2008). However enhanced soil nutrient levels and water availability could negate this or reduce soil organic matter (Heath et al., 2005; Garten et al., 2009). A key uncertainty in resolving these interactions is the feedbacks between soils and vegetation (Dieleman et al., 2010).

Table 4.9 Current anticipated responses of soil carbon to direct and indirect effects of climate change

Note that ‘uncertainty’ refers to the direction of soil carbon response: uncertainties are high throughout.

Source: CLIMSOIL project (Schils et al., 2008)

4.7.4 Implications for ecosystem services

SOC is a key measure of soil quality. Changes to the formation and turnover of soil organic matter has fundamental implications for supporting ecosystem services such as nutrient cycling and primary production, and therefore to many final ecosystem services on which humans depend. For example, changes in regulation of the release of CO2 and other greenhouse gases from soils are important factors influencing the

Page 124: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

96 Biodiversity Sector Report

rate of climate change. As well as binding and buffering release of nutrients and chemicals, organic matter regulates water retention and infiltration and therefore has a role in mediating against flood risk and supplying clean water. Nutrients and water are key requirements for plant growth; therefore organic matter also supports provisioning services such as crop and timber production. In semi-natural and native ecosystems, soil organic matter quality, quantity and dynamics are major determinants of ecosystem functioning and therefore changes to soil organic matter have profound implications for habitat maintenance and restoration.

Evidence from the increase in DOC in water suggests that many organic soils in the uplands are experiencing changes in their composition. Much of the UK’s drinking water comes from catchments dominated by organic soils. This has placed additional pressure, and costs, on DOC removal at water treatment plants over the last 25 years. Current evidence suggests that climate change is probably only playing a small role in these increases at present (Monteith et al., 2007). A number of causal factors have been suggested, most notably a reduction in Sulphur dioxide (SO2) deposition along with continued nitrogen enrichment (Monteith et al., 2007; Evans et al., 2006). There is no indication as yet that increases DOC fluxes have resulted in losses of soil organic matter stocks.

4.7.5 Implications for adaptation

The likely interaction between the effects of climate change, atmospheric pollution and the land-use pressures highlighted above indicate that measures to reduce existing pressures on soil stocks are likely to increase their overall resilience. This will act to reduce losses of SOC. For example, in peatlands, reduced disturbance and the blocking of man-made drainage ditches could restore water table levels and reduce dissolved organic carbon losses although this may be partially offset by increased CH4 emissions. Measures to reduce the risk of soil erosion include maintaining an active vegetation cover by avoiding overgrazing, and minimising ploughing on steep slopes. The environmental limitations of primary productivity could potentially be overcome by improving nutrient availability and water balance to enhance the carbon sequestration capacity of soils under increased atmospheric CO2; however, the scale and uncertainties of such intervention suggests it is unlikely to be a viable policy option. The primary focus for adaptation actions should be on protecting existing carbon stocks, particularly active peatlands that already sequester large amounts of carbon.

The risk of changes to SOC is linked to other impacts from climate change, particularly changes in soil moisture (BD1, Section 4.2), although climate change mitigation measures, such as the construction of wind farms (BD6, Section 4.6), and an increase in fire risk (BD12, Section 4.12) could also have important adverse effects.

Page 125: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 97

4.8 Changes in species migration patterns (BD9)

Summary

A wealth of observational evidence clearly indicates that migration patterns in birds are changing and therefore very likely to continue to change in response to a changing climate. This will occur in combination with other factors, such as habitat change, changing species interactions, and possibly behavioural change, therefore future predictions are difficult.

The evidence base for changes in migration patterns is high and the level of consensus is generally high, therefore there is HIGH confidence that change will occur. Action is required to ensure the UK ecological network and the wider landscape can facilitate this natural adaptation to change.

l c u l c u l c u

BD9Changes in species migration

patternsH 1 2 3 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.8.1 Introduction

Many species have evolved adaptive strategies based upon migration from one location to another at certain times of the year or during a particular period of their life cycle. This risk evaluates the consequences of changes in these seasonal migratory patterns. In reality, the changing length of days (photoperiod) acts as the proximate environmental cue that stimulates hormonal and behavioural changes that result in migration. This cue is related to the ultimate factors behind the migration, which is the decline in food supply, but has developed (probably through natural selection) to initiate migration before the food supply has declined to a critical level and the animal has insufficient energy to migrate successfully.

Although a variety of taxonomic groups have developed migration patterns, the majority of research and observational data relates to birds, although migratory behaviour also occurs in bats, fish and insects. For birds, there are different types and ranges of migration. Some species of bird from less temperate regions overwinter in the UK due to the mildness of the winter climate, particularly in coastal areas where food is available in intertidal and coastal margin habitats, and depart in the spring back to their breeding grounds. In addition, some upland breeding birds of the UK migrate to coastal areas in the winter to avoid the more severe upland winters. Other species of birds that are native to the UK avoid the British winter entirely and migrate south to return in the spring to breed. Migratory species that return to the UK in the spring have long been celebrated in culture and folklore as a herald of the new season.

In general, observations suggest that migratory bird species are shifting their over-wintering grounds to higher latitudes and arriving earlier in the spring in response to climate warming (Lehikoinen et al., 2004). This is of particular concern because such changes may not be synchronous with the ecosystems and communities upon which the birds interdepend at the migration sites. This in turn has wider ramifications for

Page 126: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

98 Biodiversity Sector Report

community structure and the functioning of the ecosystem (Wormworth and Mallon, 2006).

4.8.2 Current risk status

Development of a Wintering Waterbird Indicator based upon an aggregate of 46 species has allowed annual reporting on changes to Britain’s overwintering bird population: the most recent data from Eaton et al. (2010) are shown in Figure 4.12. The indicator shows a recent gradual decline in numbers after the steady rise that occurred from 1975 to the late 1990s, although still remaining above the levels of the 1975 baseline. Both the wildfowl and wader indicators are decreasing, with the wader index reaching its lowest level since the 1992/93 winter.

The reason for this decrease is not fully understood, and it is uncertain whether it is a result of climate-mediated range shifts (see BD5, Section 4.5) or a general decline in population levels. Results from other parts of Europe have suggested that it may be partly attributable to ‘short stopping’, whereby an increased proportion of a waterbird population is able to winter closer to its breeding area (usually further east or north) due to milder winters (Austin and Rehfisch, 2005; Maclean et al., 2008; Visser et al., 2009).

Figure 4.12 UK wintering waterbird indicator Source: Eaton et al., 2010

For other migratory species, populations of some medium-distance migrants (e.g. chiffchaffs, blackcaps) are increasing, but population declines in many long-distance migrants have been recorded (e.g. cuckoo, nightingale). These trends may relate to factors (e.g. food supply) acting on the breeding grounds, on the wintering grounds or during the migratory journey itself (Hewson and Noble, 2009). For this reason a simple cause-and-effect relationship is difficult to determine; although climate change is implicated, the exact mechanisms often remain elusive. A trend to warmer winters has led to some species (e.g. chiffchaff) preferring to over-winter in the UK rather than migrate. For other species, the changes could be attributed to habitat changes or

Page 127: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 99

species interactions beyond the UK. An unusual example of this is recent observations in the Arctic of polar bears that are now predating the nests of Barnacle geese (which overwinter in the UK).

There is an extensive literature on the changes in arrival and departure dates of breeding migratory birds (e.g. Crick, 2004; Robinson et al., 2009). Previous work for the Defra UK Climate Change Indicators has highlighted the annual arrival date of the swallow (Hirundo rustica) as a useful indicator of climate change. Long-term Bird Observatory data is available that record its arrival from over-wintering habitat in southern-Africa. Comparison between arrival date of swallows and mean February-April temperatures using the Central England Temperature series is shown in Figure 4.13 (Huin and Sparks, 1998). An alternative more recent record of first swallow arrival dates is available from the UK Phenology Network Nature’s Calendar website, which is based upon public observations. Arrival dates in this dataset are on average of 18 days later than the Bird Observatory data probably because of the extra number of sites, especially of sites inland. The same trend is apparent in both datasets but with a higher correlation in the Bird Observatory data which again is probably due to the lesser number of observation sites. Recognition of a clear trend is important given the range of confounding factors: migration patterns are influenced by climatic conditions at the location of departure and en-route, as well as the destination (e.g. Sparks and Tryjanowski, 2007).

80

85

90

95

100

105

110

115

120

125

130

2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0

mean temperatures in Feb-April in Central England

da

te s

wa

llo

w w

as

fir

st

se

en

in

Bri

tain

from defra indicator data

from natures calender

Figure 4.13 Correlation of swallow arrival time (from 1959 – 2002) with mean temperature for February to April

Data: Central England Temperature index

Correlation between temperature and arrival time of the swallow has also been examined in Scotland using data from the Scottish Bird Report for the years 1968-2000 (Sparks et al., 2006). This study found that the highest correlations existed when considering swallow arrival date and the mean temperature of the month of arrival, showing trends which are statistically significant.

There is also evidence of changes in favoured sites for migrant birds. Austin and Rehfisch (2005) studied the changes in non-breeding distribution in nine wader species and found that in seven of them a lower proportion of the population overwinter in the south west of Britain during warmer winters (Figure 4.14). Estuaries on the east and south coasts of Britain have muddier sediments than those on the west coast and thus

Page 128: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

100 Biodiversity Sector Report

support a higher biomass of the invertebrate prey of waders. The suggested explanation is that milder conditions in the east have reduced the threat of cold-weather-induced mortality, and when combined with the better foraging opportunities in the east, it has not been necessary in recent years to fly west to overwinter in less favourable conditions. Observations also suggest large scale declines of migratory passerines that appears linked to climate change, either due to impacts on wintering grounds, on migration, or on their breeding grounds in the UK.

Source: Austin and Rehfisch (2005)

Figure 4.14 Average minimum temperatures in eastern England and proportions of UK populations of sanderling and ringed plover overwintering in

south west of Britain from 1974-1998 Source: Austin and Rehfisch (2005)

4.8.3 Potential impact of climate change

Indicator data for the swallow highlight that there is a relationship between arrival data and temperature, but that this relationship is complex and also dependent on other influences on the migratory route. However it seems reasonable to infer that a continued temperature rise will lead to further advances in arrival dates for migratory species with differential responses amongst species as highlighted by the discussion on range shifts in BD5 (Section 4.5).

Underlying these changes are adaptive shifts in species behaviour as their environment is modified. In some cases, species will modify their migration patterns by travelling shorter distances to overwinter as recent observations seem to suggest (Visser et al., 2009). For example, Scandinavian blackcaps are now observed to be overwintering in the UK rather than further south. This will shift the foci of migration routes and change inter-species population dynamics at both the old and new locations. The final complicating factor is habitat change, particularly the possible loss of coastal habitats (as highlighted by BD2/BD7, Section 4.3) which are vital for many overwintering bird species. It may be that due to either the climate warming or the loss of existing habitats on the migration route that this would actually mean species have to travel further in future to migrate, and the end result could be that migratory behaviour

Page 129: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 101

may cease completely in many species (Guttal and Couzin, 2010). This will modify interactions with existing resident species and in turn could result in changes to the wider ecosystem.

A key topic of concern could be the ‘phenological mismatches’ (see Section 3.3.4) that develop as the behaviour and timing of species changes at different rates, both between migrants and sedentary species, and between long-distance and short-distance migrants (Saino et al., 2010). Environmental cues related to climate and food supply have evolved through varying mechanisms for different species, and these will be further transformed through behavioural cues within the social interactions of different species. It has been hypothesized that certain key individuals within the population may recognise the proximate environmental signal for migration and that others follow through social interaction (Guttal and Couzin, 2010). Whatever the mechanism, these traits can have an impact on breeding success and through natural selection they will influence the future viability of the species.

4.8.4 Implications for ecosystem services

The collation of public data through citizen science initiatives such as Nature’s Calendar clearly shows that that many people recognise this as an important and relevant issue. This perceived importance can be attributed not only to the direct implications for biodiversity but because the arrival and departure of particular species represent important cultural symbols and signals of the changing seasons. Any notable changes in these seasonal events can therefore have implications for the relationship between cultural and natural heritage at particular locations in the UK which has often acted to define key features of the local landscape.

Changes in the distribution of some key species could also have implications for local provisioning services. This is particular exemplified by any changes in the migratory behaviour of fish such as eel and salmon (see Section 4.9) which both provide food stocks but also contribute to local culture. For example, the start of the salmon ‘season’ is an important cultural event on several rivers in the UK.

Ultimately changes in distribution and abundance of migratory species that have a key role in regulating other species in the ecosystem (either as prey or predators) will also have wider implications for ecosystem structure and potentially wider functions and services.

4.8.5 Implications for adaptation

Changes in migration patterns will inevitably lead to both ‘winners’ and ‘losers’ and with changing regional dimensions. From a policy perspective, the shift in distribution of migratory species within the UK could be highly significant as the qualifying features for which the sites were designated move to new locations. Maclean et al. (2008) have suggested that for waders, some species will continue to use their existing overwinter sites whilst others will move to new sites, indicating that if this important group is to continue to receive protection, the site network will need to further expand. The potential shift in distribution between west and east coast sites also implies the need for a more flexible network, possibly with buffer zones around existing sites to allow for both variations in interannual and longer-term change. In addition, shifts in the migratory behaviour of species will be better accommodated in the landscape through reductions in other pressures; for example, permitted hunting seasons for wildfowl (under EU law) may need to be altered to take into account changes in the migratory annual cycle.

Page 130: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

102 Biodiversity Sector Report

4.9 Increased water temperature and stratification of water bodies (BD10)

Summary

The analysis highlights the risk from changes to the thermal regime of water bodies, including streams, rivers and lakes. Species with low tolerance limits will be most affected and as aquatic communities tend to contain species with varying thermal tolerances, there is an increased likelihood of changes to community structure and ecosystem function.

This risk is assessed with MEDIUM confidence based upon a medium level of consensus and a medium evidence base. Further work is required to better understand the range of interactions between thermal regime and aquatic ecosystems, and to link these with both interannual variability patterns and long-term trends indicated by climate projections.

l c u l c u l c u

BD10Increased water temperature

and stratification of water

bodies

M 1 2 3 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.9.1 Introduction

Changes in water temperatures follow the same pattern as atmospheric temperatures across the seasons, although extremes are buffered within water. Climate change will therefore have a strong influence on water temperatures. In addition to warming of the water surface, sustained warming increases temperatures throughout the water column, although this is an outcome highly dependent upon many factors including inflow and outflow rates, evaporation, and depth.

Any change in temperature of a water body can be important because of the impact it can have on freshwater ecosystems. For example, potential impacts on the aquatic environment include (but are not limited to) changes in lake thermal stratification, change in water quality and changes to aquatic habitats. Further, water temperature influences aquatic organisms in a variety of ways, including growth rate, metabolism, reproduction, distribution, behaviour, and tolerance to parasites/diseases and pollution (Webb 1996, Caissie 2006, Whitehead et al., 2009). Species that are currently close to thermal or other environmental thresholds are likely to be particularly at risk. Depending upon the magnitude and rate of the change, and the sensitivity of aquatic habitats and species, this could impact on key ecosystem functions. A change in thermal regime can also have implications for water quality (see Section 4.10), particularly the prevalence of algal blooms.

Stratification of a water body may be diurnal, seasonal or permanent depending on the location of the lake as well as the water depth within it. Stratification occurs when the warmer surface layer (epilimnion) of a lake floats on a colder, denser layer beneath (hypolimnion). Studies have shown that changes in stratification can result in changes to the distribution of dissolved oxygen, nutrients and any toxins within the water column (Nickus et al., 2009) which could then have impacts on water quality and the ecological

Page 131: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 103

status of the water body. Oxygen levels within lakes are also strongly influenced by the water temperature and thermal structure (Adrian et al., 2009). Generally, the respiration of bacteria within the hypolimnion of a lake continues after stratification occurs and therefore oxygen stores within the bottom layer are decreased while carbon dioxide levels are increased (Moss 1998) which has implications for species at these depths.

Increases in temperature and thermal stratification are also an issue for marine environments. Linkages to species movements, invasive non-native species and harmful algal blooms have been identified and were discussed in Section 3.4 and in more detail in the MARINE AND FISHERIES CCRA sector report (Pinnegar et al., 2012).

4.9.2 Current risk

Surface water temperatures can be correlated to regional air temperatures and exhibit a clear upward trend attributable to climate change (Hammond and Pryce, 2007) 55. It has been suggested that a warmer surface layer (epilimnion) may also increase stratification within UK water bodies, although many other factors have an influence. Changes in air temperature, solar radiation, wind, humidity, and precipitation will not only impact the temperature profile of a lake, but also influence the thermal stability of the water and associated mixing pattern (Nickus, 2007, Nõges et al., 2009).

It has also been shown that water temperature can have a direct impact on fish growth. Elliott and Hurley (1997) developed a model for Atlantic salmon (Salmo salar), assessing the impacts of different water temperatures on growth rates. Under controlled conditions, they found that fish in waters regulated at around 16°C had the highest growth rate. Temperature limits for growth were measured to be between 6°C and 22.5°C. Elliott and Elliott (2010) have also reviewed the critical temperatures for survival of Atlantic salmon, brown trout (Salmo trutta) and Arctic charr (Salvelinus alpinus) concluding that salmon is the most able to tolerate higher temperatures with an upper limit of 30-33°C. There is also evidence that warmer temperatures are influencing salmon growth through taking fewer years to reach smolt stage in northern Scotland (Gurney et al., 2008).

A climate change vulnerability index for lakes in Scotland (based upon latitude, altitude and mean depth) has been developed by Winfield et al. (2010), indicating that southern lakes at low altitude and with shallow waters have a high vulnerability and are more likely to get warmer than deep, mountainous lakes in the north. This assessment corresponds closely with sites that show the largest decline in Arctic charr populations since 1990, providing preliminary data that rising water temperatures might be an important influence on species decline.

With regard to invertebrate species, analysis by Durance and Ormerod (2007) of 25 years (1981–2005) of data from headwaters at Llyn Brianne (Wales) found a significant warming trend but that significant responses among macro-invertebrates were confined to pH-neutral streams. This suggests that acidification, in impacted streams, overrides climatic effects on macro-invertebrates by simplifying assemblages and reducing richness. Climatic processes might, nevertheless, exacerbate acidification or offset biological recovery in some headwater streams.

The climate sensitivity of the physical, chemical and biological properties of water bodies has been demonstrated through their association with interannual variation of the North Atlantic Oscillation (NAO), which influences the dominant UK weather

55

Hammond and Pryce (2007) identified “an overall lack of good quality, long-term water temperature monitoring data with which to investigate” observed increases in the temperature of many rivers in England and Wales.

Page 132: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

104 Biodiversity Sector Report

patterns from year to year (particularly winter). For example, George et al. (2004) have shown clear links between the NAO and the concentrations of nitrates (negative correlation) and phosphates (positive correlation) in Cumbria. Similarly, the timing of ice cover and melting on lakes has a clear inter-annual variation and has been suggested as a good indicator of regional climate change for freshwater systems because of the long-term records available56.

4.9.3 Potential impact of climate change

Future temperature increases (along with other climate change effects such as increased frequency of summer drought and winter flood) are likely to have significant effects on the growth rate of freshwater fish, such as trout, salmon and grayling in UK rivers (Webb and Walsh, 2004; Davidson and Hazlewood, 2005). It has also been suggested that along with impacts to in-stream habitats, warmer waters may influence some physiological characteristics of native fish. Although most British fish currently live well within their thermal limits, the native brown trout and the greyling (Thymallus thymallus) are both known to be limited by high water temperatures. Two cold water lake fish species, the whitefish (Coregonus lavaretus) and Arctic charr, are likely to be particularly vulnerable (Arnell, 1998).

The risk to salmon is identified as greatest in southern England with those rivers inferred to experience the highest temperatures. However, salmon is a migratory species and, as identified by BD9 (Section 4.8), the pattern of migration for many species is changing in response to seasonal shifts and habitat modification. This suggests that salmon will face additional risks (Graham and Harrod, 2009), especially due to thermal and salinity changes in the oceans. This issue also highlights the varying tolerance of fish species to change at different stages of their lifecycle.

Water temperature changes also have implications for lake phytoplankton due to likely increase in primary productivity. Modelling work by Elliott et al. (2006) suggested a potential for cyanobacteria to dominate the phytoplankton community and that this dominance was at its greatest when high water temperatures were combined with high nutrient loads. Similarly, changes in stratification and mixing may have impacts on plankton abundance and composition, with the occurrence of massive blooms of buoyant cyanobacteria, lower oxygen concentrations and lower light levels considered more probable in a warmer future climate (Mitchell et al., 2007). Evidence from the warm summer of 2003 suggests that mesotrophic to weakly eutrophic lakes are likely to be most strongly impacted by intensified stratification from warmer temperatures (Nickus, 2007). This may counteract management and restoration efforts undertaken in the past to mitigate anthropogenic nutrient enrichment (eutrophication) of these habitats.

Temperature changes may also modify the populations and growth rates of aquatic plants and invertebrates; for example, seasonal advances in the life cycles of dragonflies (Odonata) have been described by Hassel et al. (2007). In lowland Scottish lakes, increased water temperatures have shown a significant correlation with increased spring densities of aquatic invertebrate grazers (e.g. water flea species: Daphnia.) (Carvalho and Kirika, 2003). Analysis of macro-invertebrate abundance in spring suggests a potential decline by 21% for every 1°C rise (Durance and Ormerod, 2007). Although many core species could persist if temperature gains reached 3°C, the rarer taxonomic groups (5–12% of the species pool) would risk local extinction.

56

One of the original Defra Climate Indicators was the number of ice days observed on a bay of Lake Windermere.

Page 133: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 105

Elevated water temperatures also increase the risk from invasive non-native species such as Parrots’-feather or zander (Stizostedion lucroperca) which are currently limited by low temperatures (see Section 4.4).

4.9.4 Implications for ecosystem services

The analysis highlights that species with low tolerance limits will be most affected. Aquatic communities contain species with varying thermal tolerances, therefore the differential impacts between species implies an enhanced prospect of changes to community structure and ecosystem function. Aquatic ecosystems are important for the range of ecosystem services that they provide including natural filtering, water purification, and the regulation of oxygen supply and other nutrients.

In addition a range of provisioning and cultural services are associated with these systems, particularly associated with fish, which provide important social and economic benefits for some areas [see Box on Salmonid fisheries].

Salmonid fisheries

Salmon is the focus for economically-important sport and commercial fisheries.

Factors known to affect freshwater growth of Atlantic salmon (Salmo salar) include temperature, food availability, and density (Elliott and Hurley 1997; Grant et al., 1998). A relationship between mean age at smolt stage and the combination of temperature and day length has been demonstrated across the range of Atlantic salmon by Metcalfe and Thorpe (1990). Climate change is therefore associated with a shorter time for salmon to reach smolt stage. Experimental evidence has led to qualitative suggestions that waters with current average temperatures below 16-18°C will show increased populations as a result of increasing temperatures (e.g., Langan et al., 2001), whereas waters with higher current temperatures will show population decreases (e.g. Langan et al., 2001; Swansburg et al., 2002). However, this will also be dependent on other factors, such as the continued availability of food.

In the UK and elsewhere, the number of fish returning annually to spawn (comprising the great majority of the female breeding stock) has shown a dramatic downward trend over the past 40 years which is associated with increased marine mortality. The life cycle of the salmon, which involves both freshwater and marine components, is therefore vulnerable to climate change at various stages.

In addition to their economic value, salmon have an important cultural role on many UK rivers. This includes the association between their migratory behaviour and the pattern of seasons (sometimes celebrated in a yearly ceremony), and the tradition or folklore that has developed based upon sport fishing.

4.9.5 Implications for adaptation

Adaptation options to increase the resilience of aquatic ecosystems to climate change have been reviewed by Wilby et al. (2010) and include:

increases in the shading of vulnerable reaches through tree planting, particularly in headwaters

in-stream habitat modification to create thermal refugia such as pools

the release of cooler hypolimnetic water to compensate flows

Page 134: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

106 Biodiversity Sector Report

selective re-introduction or translocation of genetic material from areas with higher temperatures

removal of physical barriers to migration and dispersal.

However, there is a lack of evidence on how to implement these options in practice. An additional adaptation strategy involving translocation is currently being trialled for the cold-water species vendace (Coregonus albula) by moving populations to more resilient lake locations (Winfield et al., 2008); however the success of such schemes will also be dependent on appropriate habitat and species interactions.

Adaptation strategies related to thermal regime will need to be integrated with other measures to respond to geomorphological and hydro-ecological changes in water bodies, including the impacts of water quality (BD13, Section 4.10) and water quantity (BD14/15/16, Section 4.13).

4.10 Water quality and pollution risk and eutrophication (BD13)

Summary

Impacts on water quality will occur as a result of the interaction of warmer temperatures and pollution, particularly from land use. These impacts may be exacerbated by lower summer flows and runoff due to increased rainfall intensity. Changes in water quality could further modify species and community composition in aquatic habitats, and hence ecosystem functioning.

This risk is assessed with MEDIUM confidence based upon a medium level of consensus and a medium evidence base. Further work is required to better understand the diverse ecological impacts of spatial interactions between diffuse pollution sources and climate variables throughout catchments and across the variety of UK river basins.

l c u l c u l c u

BD13Water quality and pollution risk

and eutrophicationM 1 1 2 1 2 3 1 2 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.10.1 Introduction

Negative impacts on water quality are associated with industrial, domestic and land use sources. In recent decades, impacts from industry have declined whilst diffuse pollution in rural areas, mainly from agriculture has increased. Diffuse pollution is difficult to regulate because of the multiplicity of potential sources. By contrast, wastewater treatment plants, which can act as point sources of pollution, are regulated through monitoring of their discharges. In headwater streams within semi-natural habitats, nutrient enrichment and acidification from atmospheric deposition is a continuing issue because of the normal low-nutrient status of these habitats.

Nitrogen and phosphorous loads from fertiliser runoff have reached critical levels for many water bodies, often associated with a shift towards agricultural intensification, and with arable land often occupying large proportions of a catchment (although other

Page 135: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 107

factors can also contribute). Nitrate Vulnerable Zones have been declared in many regions to protect quality of drinking water and the EU Water Framework Directive requires that the UK implements measures to ensure water bodies meet ‘good ecological status’. The consequence of elevated levels of nitrogen and phosphorous nutrients in water bodies is eutrophication and the risk of harmful algal blooms that deoxygenate the water and have other detrimental effects for aquatic ecosystems.

Climate change could impact on water quality, both directly and indirectly. An expansion in intensive agriculture in some catchments due to the beneficial effects of a warming climate (see Section 3.3.1) could increase diffuse pollution loads. Higher temperatures (Section 4.9) together with higher nutrient loads could further increase the risk of algal blooms. Heavier rainfall events may increase the rate of runoff of pollutants, whilst lower flows in summer (Section 4.13) may result in less dilution and greater concentrations of pollutants.

A change in nutrient status, dissolved oxygen concentration (DOC) or toxins in water bodies can all have detrimental impacts. For example, growth measurements of fish can be related to environmental conditions indicating that poor water quality, environmental stress and limited food sources will result in low growth (Elliott and Hurley, 1997). Reduced dilution effects and increased organic pollutant concentrations during low flows can combine with the higher biological oxygen demand (BOD) during such episodes to exacerbate stresses on aquatic ecosystems due to lower DOC levels.

4.10.2 Current risk

Under the Water Framework Directive, the Environment Agency (England and Wales), Scottish Environmental Protection Agency, and Northern Ireland Environment Agency are assessing the quality of hydrological units within each river basin district of the UK. Overall status is assessed as either: high, good, moderate, poor, bad. The summary maps illustrate current pressures on water resources, of which the dominant factor is land use rather than climate change. Those units classed as ‘bad’ have ecosystems that are severely stressed and at risk of irreversible damage, particularly during extreme events.

In many upland areas the impacts of atmospheric deposition through acidification and nutrient enrichment are a problem. Therefore research often shows that although a warming trend is present in the water, it is often not the dominant influence. For example, the 25 year (1981–2005) analysis of invertebrates at Llyn Brianne (Wales) by Durance and Ormerod (2007) found that temperature-based responses were confined to pH-neutral streams. This analysis suggested that acidification, in impacted streams, overrides climatic effects on macro-invertebrates by simplifying assemblages and reducing richness.

4.10.3 Potential impacts of climate change

Qualitative assessment in the WATER CCRA sector report (Rance et al., 2012) analysed the percentage of rivers with a net decline in Ecological Status linked to changes in river flow [metric WA9]. The assessment, based on expert elicitation, indicates that there is projected to be ‘very low’ to ‘low’ consequences for the 2020s with a decline in status of up to 10% of rivers. In the long term (2080s) with larger reductions in summer flows more than half of rivers in England and Wales may be affected by a decline in status. However, this analysis did not consider changes in water quality outcomes due to the indirect effects of climate change on land use change, or the impacts of population growth, both of which are likely to be significant in the medium to long-term for some districts. If these districts also contain areas of high

Page 136: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

108 Biodiversity Sector Report

biodiversity value (for example, water bodies designated under Natura 2000) then even with a small overall reduction in water quality, the ecological consequences could be rather more severe.

Impacts could also become more pronounced during extreme events. Analysis by Cox and Whitehead (2009) for the River Thames using the UKCIP02 scenarios shows that by the 2080s increased BOD and reduced DOC will occur but that this would not be an issue for ecosystems under normal circumstances. However, during algal bloom events, large diurnal variations in DOC occur and this could exacerbate ecological stress, particularly if these episodes increase in frequency and duration. Problems will therefore be most apparent in poorer quality rivers during summer low-flow conditions.

With the potential for increased storm events from climate change, especially in summer, the frequency with which combined sewer overflows (CSO) discharge highly polluted waters into receiving water bodies could increase. Analysis in the WATER CCRA sector report (Rance et al., 2012) has developed a metric to indicate a change in the spill frequency from CSO outfalls.

4.10.4 Implications for ecosystem services

Poor water quality is detrimental to the effective functioning of aquatic ecosystems including supporting services such as nutrient cycling and oxygenation. A surplus of nutrients during eutrophication mean that these nutrients are not cycled through the system and the dynamic balance of the ecosystem is destabilised, particularly when toxic algal blooms occur. This has implications for the purification and supply of clean water. Impacts are also apparent for fish stocks which provide an important economic, social and cultural service in key locations, and for other important species that provide ecosystem benefits. Poor water quality can also cause issues with human recreational and amenity use of water bodies, including water sports and tourism, to the detriment of local communities.

As discussed in Chapter 6, the benefits of inland wetlands for water quality have been estimated by the UK NEA to be as high as £1,500 million/year. Furthermore, the same assessment indicates that planned river quality improvements may generate additional values up to £1,100 million/year.

4.10.5 Implications for adaptation

The EU Water Framework Directive (WFD) provides the structure for policy initiatives and regulation, requiring that Good Ecological Status is defined by “...allowing only a slight departure from the biological community which would be expected in conditions of minimal anthropogenic impact.” Member states are required to provide status reports on a six-year cycle and to implement River Basin Management Plans. Climate change is implicitly included within this framework by acting with other stressors such as land use change and pollution to potentially alter ecological status. However, reference to a ‘natural’ baseline’ in the WFD provides some difficulties as many water bodies in the UK are no longer in their natural state and climate change involves the influence of large-scale atmospheric influences on water levels and river flows, which may require a reappraisal of the viable baseline conditions. However, the WFD provides a high-level framework in which strategies to reduce other stressors such as land-use change and pollution can be developed in order to allow water bodies to successfully adapt to changing climatic conditions (Wilby et al., 2006). The key challenge for water quality is to reduce the risk from diffuse pollution through measures either at their source (e.g. timing of fertiliser application; precision farming) or before they enter water bodies (e.g. buffer strips). The effectiveness of these alternative

Page 137: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 109

adaptation measures in varying environmental and socio-economic contexts requires further assessment including, for example, in different land use systems and catchment contexts.

4.11 Generalists favoured over specialists (BD11)

Summary

Generalist species have a greater ability to naturally adapt to a changing environment as they have less specific habitat requirements than more specialist species. If generalist species expand at the expanse of specialist species, this has implications for biodiversity. Examples are provided for birds and butterflies, together with the wider implications for ecosystem services through potential impacts on pollinators.

There is a MEDIUM level of consensus regarding the extent of this risk but the evidence base is LOW, being limited to a few taxonomic groups and dominated mainly by the detrimental impacts of land use change rather than climate change. Confidence is therefore assessed as LOW.

l c u l c u l c u

BD11Generalists favoured over

specialistsL 1 2 3 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.11.1 Introduction

Species that are more ‘generalist’ in their behaviour (i.e. those which can exploit a wide range of environmental conditions, habitats, or food types) are generally considered to be more adaptable than species that have more ‘specialist’ requirements. The specificity shown by specialist species can make them more vulnerable to environmental changes, such as habitat loss or changes in climate, which impact on their current ecological niches; in addition they may have limited ability to exploit new niches.57 Many specialist species are associated with a natural or semi-natural land cover that retains a wide diversity of habitats, whereas generalist species can often roam over the wider countryside including both semi-natural habitats and landscapes that have been modified by agriculture or forestry. Therefore, in theory at least, as climate change becomes more pronounced, then generalist species may be expected to expand at the expense of more specialist species.

4.11.2 Current risk status

Butterflies are often used as environmental indicators because they are representative of many other insects, occur in a variety of different habitats and respond rapidly to changes in conditions. The UK Butterfly Monitoring Scheme (UKBMS) has produced a multi-species index of butterfly abundance that includes comparison of generalist and

57

It should be noted that although the terms specialist and generalist suggest a dichotomy, in reality there is a continuum from highly specialised species which require very specific environmental conditions, to highly generalist species which are able to survive in a wide variety of habitats, with the majority of species lying somewhere between the two.

Page 138: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

110 Biodiversity Sector Report

specialist species (Freeman, 2009). This index provides the basis for one of the UK Biodiversity Indicators which is updated annually (Defra, 2010c). Analysis of the index (Figure 4.15) shows that there are often large year to year fluctuations in populations, particularly of generalist species, which can usually be attributed to the weather of that year, indicating general climate sensitivity. A long-term trend is more difficult to find, as after a significant decline following 1976 (an anomalously warm year), the index shows no significant changes58.

Another important indicator of change is the population of farmland birds. Data for England (based upon 19 species) are maintained by Defra as one of a series linked to the Biodiversity Strategy, and provide a similar distinction between specialist and generalist species (Figure 4.16). The data show a 60% decline in specialist farmland bird species between 1970 and 2007. By comparison, generalist species have increased by 8% over the same period. The declines in specialist species are believed to be predominantly associated with changes in agricultural practice and land management techniques, but may be considered to be indicative of the lower adaptability of specialist species to respond to environmental changes compared with generalists.

Similar conclusions have been derived from analysis of species change in woodland habitats. For example, Kirby et al. (2005) have provided evidence that specialist woodland flora declined between 1971 and 2001 at the expense of more generalist species.

Figure 4.15 Populations of butterflies in the UK: 1976-2007 Source: Biodiversity Indicators in your pocket 2010, Defra (2010c).

Data: CEH, Butterfly Conservation.

58

UK data may obscure inter-country differences. Defra indicators for both woodland and farmland butterflies suggest significant declines in both specialists and generalists. See: http://www.defra.gov.uk/environment/biodiversity/documents/indicator/200904f1b.pdf http://www.defra.gov.uk/environment/biodiversity/documents/indicator/200904a1b.pdf

Page 139: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 111

Figure 4.16 Farmland bird populations in England This indicator shows the changes in farmland bird populations in England from 1966 to 2009

expressed as an index with 1966 = 100. Source: Defra (2011c)

The limited ability of specialist species to adapt to changes in the environment has led to many declining in population or distribution to such an extent that they have been included on the UK Biodiversity Action Plan list of priority species. For example, the bittern Botaurus stellaris, which requires large reedbeds in which to breed, or the Capercallie, Tetrao urogallus, which has its habitat in Caledonian pine forests.

The success of generalist species in adapting to environmental changes can lead to further pressure upon specialist species through competition or predation, leading to further population declines. For example, the use of willow tit (Poecile montanus) holes by both great tits (Parus major) and blue tits (Parus caeruleus) has been implicated in the long term decline (c. 80% in the last 40 years) of the former species (Gregory et al., 2002). Both of the latter species are able to utilise a variety of man-made habitats, including farmland, parks and gardens, and populations of both species increased between 1967 and 2002, by 32% in the case of blue tits, and 71% for great tits (Baillie et al., 2005).

4.11.3 Potential impact of climate change

Relatively few studies have examined differences in the effects of climate change on generalist species compared with specialist species. For many UK species there is limited data on historic population size and distribution which could be used to develop models, with the notable exceptions of birds and butterfly species. As discussed above, data for these two groups are apparently dominated by other influences, notably land-use change and habitat fragmentation. These factors have tended to over-ride climatic trends although inter-annual variation in weather patterns appears to have an influence on the butterfly data.

Analysis of specialist versus generalist species is intrinsically linked to the relationship between habitat loss and climate change. Consequently, the effects of climate change on specialist species are compounded by restrictions on habitat availability due to both past and future habitat loss. Specialist species are typically found in semi-natural habitats, for example mature woodland, heathland or wetland, which are often

Page 140: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

112 Biodiversity Sector Report

restricted in size and distribution and can be highly fragmented due to historic habitat loss. With changes in climate, as existing habitat becomes unsuitable, specialist species may be unable to move to remaining areas of optimal habitat due to a lack of intervening sub-optimal habitat (see BD5, Section 4.5). In some cases, climate change may result in a complete loss of suitable habitat; for example, ecological niches within woodland, heath and bog habitats due to increased soil moisture deficits (BD1, Section 4.2) or wildfire (BD12, Section 4.12). If the species is unable to adapt to these changes in habitat, this may result in its extinction.

In contrast, more generalist species are less restricted by habitat loss and are likely to be able to move through areas of sub-optimal habitat to reach remaining patches of optimal habitat. Highly generalist species will not be restricted by habitat type and will be able to take advantage of the new niches provided by the changing environmental conditions.

The magnitude of the effects of climate change on a species will also depend on the rate of change. Rapid changes in environmental conditions and habitat will limit the ability of species to adapt to the new conditions. Thus specialist species are likely to be disproportionately affected by more rapid changes in climate that induce large range shifts (BD5, Section 4.5).

There have been a number of studies on British butterflies which have looked at the combined effects of habitat loss and climate change. Menéndez et al. (2006) found that although theory predicted that recent climate warming should have driven increases in species richness in Britain, species richness had only increased by a third of that predicted. This change was primarily due to northwards expansion of southerly generalist species so local butterfly assemblages have become increasingly dominated by species that occupy widespread habitats and were able to respond quickly. The response of southerly specialist species was constrained by population losses due to habitat destruction, the patchy distribution of potential habitats, and limited dispersal capacity. The authors suggest that it may be decades or even centuries before the species richness and composition of biological communities adjusts to the current climate, and that some specialist species may not be able to respond at all, unless habitat restoration takes place. Warren et al., (2001) also found that range expansions by butterflies in Britain since 1982 have mainly been confined to generalist mobile species.

An important caveat to this general interpretation is provided by the phenomenon of ‘species plasticity’, through which adaptive variations develop in response to environmental signals. We do not know how important this will be, but it is noteworthy that some specialist species are becoming more generalist in their response (Thomas et al., 2001; Braschler and Hill 2007). For example, the Silver-spotted Skipper Butterfly (Hesperia comma) has begun to breed in a wider range of grassland types in England, mirroring its behaviour further south in Europe, due to increased temperatures (Davies et al., 2006). In terms of the conservation prospects of these species, this can be regarded as a beneficial side-effect of climate change.

4.11.4 Implications for ecosystem services

A decline in specialist species reduces the overall species richness of ecosystems and is believed to reduce overall ecosystem resilience to change (although evidence for this is currently limited). The most visible manifestation of this is provided by impacts on pollinators [see Box on Pollinators and Pollination services] where climate change appears to be interacting with a range of other stresses to cause Colony Collapse Disorder and other negative consequences, with knock-on effects for pollination services. A reduction in overall system resilience can also reduce the resistance to

Page 141: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 113

invasive non-native species and diseases, which has implications for both biodiversity and human welfare.

Pollinators and pollination services

During the last 20 years, habitat losses and intensification of agriculture are believed to be responsible for a 54% decline in honey bee colony numbers in England meaning that more than 50% of our landscapes now have fewer species of bees and hoverflies than in 1980. Intensification appears to change the community composition of bees within agricultural ecosystems by negatively impacting on the least resilient species and reducing overall diversity. When combined with the effects of climate change and pathogens, this has major implications for the stability of pollination services to both biodiversity and agriculture.

Some 84% of European crops and 80% of wildflowers rely on insect pollination with the value of pollination to UK agriculture estimated at £440 million per year (13% of the total value of agriculture). Over the last 20 years, the area of crops dependent on insect pollination has increased by 32% in England. In the agriculture sector, the cost of replacing bee pollination with hand pollination is greater than the total market value of the crops, at over £1.5 billion per year. The value of pollinators and pollination services to wild flowers and for recreational/ cultural services is unknown but expected to be non-trivial.

Source: UK NEA (2011) and references therein

4.11.5 Implications for adaptation

The limited ability of specialist species to adapt to changing environmental conditions is compounded by the restricted area and fragmented nature of the semi-natural habitats they typically depend upon. Furthermore, even those studies that have reported on generalist butterflies suggest they may take decades to adjust to new climate conditions, depending on the magnitude and rate of change. As identified by risk BD5 (Section 4.5), adaptation will be facilitated by an ecological network that provides larger, more connected sites with better quality habitat. Potential negative impacts on many species, particularly pollinators, provide a strong economic rationale to provide a diversity of flower-rich meadows, field margins, road verges, gardens and parks, in order to better accommodate change.

4.12 Increased risk of wildfires (BD12)

Summary

Climatic conditions that would promote an increase in wildfire risk such as higher temperatures, lower summer rainfall and drier soils are projected to increase. Increased frequency of wildfires could result in large changes in habitat extent and species populations in vulnerable locations. These impacts have occurred in extreme dry conditions during recent years.

There is a high consensus that that wildfire risk is likely to increase. However, the evidence base is currently rather limited (medium level) and therefore the confidence level for this risk is overall MEDIUM. Until the introduction in 2009 of a new database of fire events (the Incident Recording System), data was not systematically collected for the magnitude, extent, and other characteristics of wildfires. A full understanding of the potential frequency and scale of events at national scale, including affected habitats, therefore remains a knowledge gap.

Page 142: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

114 Biodiversity Sector Report

l c u l c u l c u

BD12 Increased risk of wildfires M 1 1 2 1 2 3 2 2 3

Metric

codeMetric name

Co

nfi

den

ce Summary Class

2020s 2050s 2080s

4.12.1 Introduction

Increased risk of wildfire, including both small-scale and large-scale fires, may have a profound effect on habitats and species in fire-prone areas. This risk has the potential to result in the local extinction of species if they are unable to find new habitat. Wildfires can also increase rates of erosion and have a detrimental effect on water quality, and hence on aquatic ecosystems. In addition to the potential consequences for biodiversity, this topic has been identified as a priority risk for FORESTRY, AGRICULTURE, and BUILT ENVIRONMENT. In addition the implications for human HEALTH, the loss of landscape amenity and disruption to BUSINESS activities and TRANSPORT mean this is a genuine cross-sectoral risk (McMorrow et al., 2005).

4.12.2 Current risk

Statistics on outdoor fires in England and Wales were selected by Defra as one of the original UK Climate Change Indicators (indicator 16).59 Original statistical analysis used a comparison between fire data for 1984-2001 and mean summer rainfall, June to August. The risk of outdoor fires was shown to be significantly higher in drier years (correlation co-efficient with precipitation of -0.74).

Further analysis has used more recent data held by the Department for Communities and Local Government (CLG). This shows a strong association between the frequency of fire and the ‘heatwave’ years of 1995, 2003 and 2006 (shown as the red ellipses on Figure 4.17).

Since 2009, fire reporting in the UK has used the Incident Reporting System (IRS), which will improve the level of detail and standardisation of data collection (Gazzard 2009). Significant additional detail collected for each fire under the IRS includes the location (grid reference), the extent, and whether the fire was in a national park (CLG 2009). The introduction of a record of extent will greatly improve understanding of the current level of risk from wildfire as before the IRS there was no means of easily determining the spatial scale of this risk (e.g. frequency of the larger-scale events).

The seasonal timing of wildfire also has important implications for the resulting consequences. Increased fuel loading as a result of dry vegetation, either in spring before growth, or in late summer during a drought, can provide an additional risk factor. With regard to forest fires, in spring these are generally surface fires, whereas in summer they can extend below the surface. When organic soils, particularly peat, are affected by fire, the damage can become extensive in depth and extent, because of the large fuel supply and difficulties of suppression.

59

http://www.ecn.ac.uk/iccuk/indicators/16.htm

Page 143: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 115

Figure 4.17 Outdoor fires in the UK (1995 – 2008) and heatwave years Source: Gazzard R, 2010, ‘Development of wildfire statistics and risk impacts in the United Kingdom’, Fire and Rescue Statistics User Group (FRSUG). http://www.frsug.org/reports/

4.12.3 Potential impact of climate change

A range of established fire danger indices exist, including the McArthur Forest Fire Danger Index (FFDI), the Canadian Forest Fire Weather Index (FWI) and the US National Fire Danger Ratings System (NFDRS) (Dowdy et al., 2009). The Met Office has recently begun modelling of climate change effects on fire danger using the FFDI, as previous experience indicated that it correlated well with actual fire events in the UK (Met Office, 2005). The Met Office has also developed an operational Fire Severity Index (FSI) for the UK, with background information provided by Natural England and FSI forecasts provided by the Met Office.60

The FFDI combines the probability of a fire starting, rate of spread, intensity, and difficulty of suppression into a single index derived from climate variables and soil moisture content. Within the FFDI scale, a value of 1 means fire will not burn while 5 to 12 is a ‘moderate risk. Climate change analysis is based upon work originally carried out in the Amazon basin (Golding and Betts, 2008) that has been updated for the UK through use of the 11-member Regional Climate Model (HadRM3) ensemble associated with UKCP09.

The results of this analysis (Figure 4.18) need to be interpreted with caution as they provide only change in annual mean values; it would be assumed that the likely values would be higher in the summer months. In addition, only one climate model has been used, although all the simulations show an increase in the FFDI for the 2080s, especially in the south of the UK. The coarse resolution of the soils and land cover data also probably mean that the fuel (biomass) component is poorly represented. To provide an indication of how these national-scale results may affect land management at the local scale, model results (% difference) have been extracted for national parks across England, Scotland and Wales (Table 4.10), highlighting the higher risk rating for all areas in the future with the largest changes in the south.

60

See http://www.naturalengland.org.uk/ourwork/enjoying/places/openaccess/fireseverity.aspx; http://www.metoffice.gov.uk/weather/uk/firerisk/

Page 144: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

116 Biodiversity Sector Report

Figure 4.18 FFDI Index for 1980s and 2080s using HadRM3 climate model (ensemble mean; Medium emissions).

Source: Richard Betts, Met Office

Table 4.10 McArthur Forest Fire Danger Index applied to UK national parks

Country National Park Mean % Difference

2080s – 1980s

England Lake District 30% North Yorkshire Moors 30% Northumberland 30% The Broads 30% Yorkshire Dales 30% Exmoor 40% Peak District 40% Dartmoor 40 – 50% New Forest 50% South Downs 50% Scotland Loch Lomond and The Trossachs 30% Cairngorms 30 – 40% Wales Pembrokeshire Coast 30 – 40% Brecon Beacons 30 – 40% Snowdonia 40 – 50%

Source: Richard Betts, Met Office

These figures relate only to average change, meaning that they give no indication of the possible impact of changing variability and extreme events such as heat waves. There is a clear association between outdoor fires and warm dry conditions, although it is not clear how much of this is related to the background fire risk or the increased opportunity for ignition (for example through increased visitor numbers). More detailed regional analysis of the wildfire risk in the Peak District by Albertson et al. (2010) suggested little change in the short term but potentially major changes in fire regime in the medium to long term. Also highlighted in the same regional study was an important interaction with the changing seasonal pattern of vegetation growth/decay which might lead to significant changes in the length and intensity of the ‘at risk’ season.

The effects of fire on priority species will vary depending on such factors as fire tolerance (flora), capacity to escape or find shelter from fire, and ability to find sustenance in a burnt landscape (fauna). Some habitats will be more tolerant to fire

Page 145: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 117

events than others, with some grassland and heathland habitats possibly benefiting from well-planned controlled burning to prevent colonisation by woody species. A particularly vulnerable habitat is Lowland Heath which is of particular significance for rare British reptile species such as the sand lizard (Lacerta agilis) and the smooth snake (Coronella austriaca). In addition, peatland areas have high vulnerability when dry: as discussed under BD1 (Section 4.2) and BD8 (Section 4.7), such habitats are a significant carbon sink. Once ignited, peat habitats can burn and ‘smoulder’ for an extended period and are difficult to extinguish. Where there is damage as a result of fire, carbon stored within the peat can be released back to the atmosphere and the ability of the habitat to act as an effective carbon sink is severely constrained. It is also possible that the original vegetation communities, such as blanket bog or wet heath habitats, cannot be re-established.

4.12.4 Implications for ecosystem services

In addition to the risks to biodiversity and human health, wildfire can be detrimental to ecosystem services such as carbon storage, soil quality, protection from soil erosion, water flow regulation and water purification. These impacts can be particularly severe in peatland areas and release large amounts of carbon into watercourses and into the atmosphere (see BD8, Section 4.7). Some semi-natural habitats, particularly wetland habitats can provide natural resistance to the propagation of fire, whereas large commercial woodland plantations with limited species diversity are specifically at risk of large-scale fire.

4.12.5 Implications for adaptation

In addition to the background risk factors from warm dry conditions and availability of fuel supply, a wildfire event also requires a trigger. Although there are natural causes of wildfire in the UK (e.g. lightning), these are extremely rare, with the vast majority of events being due to the actions of people (either accidentally or deliberately). This raises the prospect of increased use of closure orders to restrict access to high risk areas (e.g. heath and moor) that are important for biodiversity and ecosystem services (e.g. water resources, carbon storage). This will have implications for access and recreation. Other management measures may include controlled burning, grazing or mowing to remove fuel. Schemes that maintain high water tables and reduce soil moisture deficits (e.g. blocking of drainage ditches) may also assist in improving resilience to fire. Systematic provision of fire ponds may help in constraining the risk when wildfire outbreaks do occur. More systematic collection and analysis of data from wildfires throughout the UK is required to improve modelling and risk assessment tools.

Page 146: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

118 Biodiversity Sector Report

4.13 Impact on water quantity: low flows (BD14); increased societal water demand (BD15); major drought (BD16)

Summary

A change in water supply due to climate change is likely to be accompanied by increased demands for water from other sectors, particularly agriculture. This combined impact will act to modify flow regimes and water levels, which is likely to reduce the capacity of aquatic ecosystems to adapt to change, particularly when aggregated with changes in water quality and thermal regime. The most pronounced risk occurs during extreme drought events when there could be major biodiversity loss and some ecosystems may experience irreversible change in the absence of a more precautionary approach to water abstraction.

This overall risk is assessed with MEDIUM confidence based upon a good evidence base and a medium level of consensus. The confidence rating takes into account the methods, data and consensus issues related to the magnitude and direction of change. Since there is not complete consensus on the hydrological modelling of low flows, the confidence falls into the medium category, rather than the high category despite the evidence available for the negative consequences for ecology of past events. However, further work should be carried out to provide a more detailed assessment of changes in future variability of supply and demand, to evaluate the implications of both incremental change and extreme events for biodiversity and ecosystem services.

l c u l c u l c u

BD14/

15/16Impacts on water quantity M 1 2 2 2 2 3 2 3 3

Metric

codeMetric name

Co

nfi

de

nc

e Summary Class

2020s 2050s 2080s

4.13.1 Introduction

Three impacts identified during the Tier 1 assessment are here combined with regard to their impact on low flows and water levels. Firstly, the direct impact of a changing climate on water supply and flow regimes through modified rates of precipitation and evaporation. Secondly, the indirect consequences that may occur from an increased human demand for water which may cause additional stress on the natural environment. Thirdly, the compound effects that occur from both direct and indirect effects when they are exacerbated from extreme events during major drought episodes. These impacts have also been evaluated using risk metrics in the CCRA WATER sector report (Rance et al., 2012) but this analysis was not conducted from a biodiversity perspective.

The frequency, magnitude and duration of low flows are known to have impacts for aquatic communities. Low flows reduce effluent dilution capacity and can lead to large increases in nutrient and organic pollutant concentrations (see BD13: section 4.10). Under the EU Water Framework Directive, the concept of ‘environmental flow’ was established as a key indicator measure for water bodies to meet the overall characterisation of ‘good ecological status’. Many UK rivers have regulated rather than

Page 147: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 119

natural flows and a key challenge presented by climate change is to modify regulatory regimes to maintain healthy ecosystems.

4.13.2 Current risks

The high variability in flow regime from year to year means that long-term trends in flow are hard to detect (Wilby, 2006). Analysis of near-natural catchments by Hannaford and Marsh (2006) found no significant long-term decline in summer flows in the UK, but this excluded those catchments that have been modified or subject to major abstractions. Datasets of water demand collated by water companies, show per capita consumption is sensitive to mean monthly temperature and that it rises during heatwave periods. Trends in water demand are upwards, particularly from agriculture due to requirements for increased irrigation to improve the quality of produce.

The impact of low flows on biotic communities has been studied extensively in particular catchments (e.g. chalk streams), with some species such as algae benefiting at the expense of others, including some invertebrate communities. If the period of low flow is infrequent, systems can recover but a sustained drought can mean recovery times are much longer. The UK has experienced several extended periods of widespread drought in recent years, including 1990, 1995, 2003 and 2006 (Marsh et al., 2007). During summer droughts, 25-33% of flow in some rivers in Central and Southeast England can be comprised of sewage effluent. Drought conditions increase nitrogen mineralisation and sulphur oxidation, leading to the release of nitrate and sulphate into surface waters which can further impact on water quality (Whitehead et al., 2009). Storms that terminate droughts act to flush nutrients or generate acid pulses in acidified upland catchments. Upland systems are particularly vulnerable to these events because of their normal low-nutrient status.

4.13.3 Potential impacts of climate change

Analysis in the WATER CCRA sector report (Rance et al., 2012) used an aridity index [metric WA1] to measure future change against a baseline reference ‘dry’ year: 2003. For the UKCP09 2050s medium emissions scenario central estimate (p50% level), average aridity is projected to be higher than that of the 2003 aridity score. For all three emissions scenarios in the 2080s, the p50% aridity score is projected to be higher than the aridity score for 2003. In addition to these average changes, these projections indicate an increased frequency of drought conditions.

The same sector report (Rance et al., 2012) also assessed changes in flow levels. Results vary across river basin regions by between -1 to -70 % (depending on location and UKCP09 projection) for the 2080s compared to the baseline. For all regions under the central estimate (Medium emissions p50) flows are projected to decline from -12 to -35%. These results are consistent with Environment Agency project SC070079 which used the UKCIP02 scenarios: based upon the med-high emissions scenario, this project found that the total annual river flow in England and Wales could drop by as much as 10–15% by 2050 (Environment Agency, 2008). In terms of mean monthly flows, a decrease was projected during the summer and autumn months of around 50%, with a fall of up to 80% in some areas. The magnitude of change varies slightly with other emissions scenarios, but the pattern and timing of changes in river flow were found to follow a similar pattern.

With regard to water demand, the CC:DEW project (Downing et al., 2003) estimated that the impact of average climate change by the 2050s using the UKCIP02 scenarios was to increase domestic demand by 2-4%, industrial and commercial demand by 4 -6% and agricultural demand by 26%. Similar results were found using UKCP09 in the

Page 148: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

120 Biodiversity Sector Report

WATER CCRA sector report using a statistical relationship between demand and mean monthly temperatures [metric WA4].

Also in the WATER CCRA sector report (Rance et al., 2012), river basin regions were analysed in terms of sites meeting WFD Environmental Flow (EF) indicators [metric WA7]. In the near term (2020s) the assessment indicates that a large proportion of rivers in England and Wales, 37% (within a range of -8% to 69%) for the Medium Emissions scenario, will fail to meet environment flow targets and consequently a large proportion of abstractions licences may be regarded as unsustainable. In the longer term (2080s) this analysis showed that more than two thirds of rivers would fail to meet current targets. Changes in flow regime would combine with temperature rise and changes in dissolved oxygen to have serious impacts on the life cycles and interactions of many aquatic organisms, including invertebrates, amphibians, fish and birds.

The PRINCE project (Conlan et al., 2007) investigated the impact of changing flow regime on aquatic habitats and indicated that many streams could also have a more variable hydrological regime with lower low flows and higher high flows, although the dominant trend is a decline in flow. Specific impacts depend upon the species under consideration. For salmon, 2050s simulations suggest a reduction in suitability for all three life stages, with a dramatic reduction for nursery habitat and smaller reductions for spawning and rearing habitats. For brown trout, no major changes were found which most likely reflects the greater tolerance of brown trout to reduced flows, in terms of the interaction between the flow and the bed topography. This emphasises the important contextual influence of particular local river geometries and individual organism habitat preferences in mediating the impact of climate change. These results also suggest that general regulatory rules based upon flow alone are likely to be highly misleading.

4.13.4 Implications for ecosystem services

Changes in flow and water levels will interact with impacts on water quality (BD13) and thermal regime (BD10) to modify the functioning of aquatic ecosystems. Resulting changes to supporting services such as nutrient cycling and oxygenation have implications for a range of regulating, provisioning and cultural services that these systems provide. This includes water purification and the regular supply of clean water together with impacts on fish stocks and other aquatic species. Some water bodies such as chalk streams in southern England are celebrated for their cultural heritage in addition to the natural heritage value. Therefore the potential drying-up of watercourses could have significant implications for wider landscape amenity and cultural value, including recreational and navigational access. The UK NEA (2011) estimated climate change-induced losses of water availability at £350 million to £490 million/yr.

4.13.5 Implications for adaptation

The high likelihood of decreasing summer flows indicate increased competition for water resources in future, particularly in areas where water demands are already high and increasing. Policy tools are available to plan for and regulate abstractions through water resource management plans and catchment abstraction management plans. The Water Framework Directive provides the high-level structure for river basin management, and although climate change is not explicitly included, it has an overall objective to provide for ‘good ecological status’. A key challenge will be to develop and clarify key WFD indicators, such as environmental flow, to ensure that they provide an acceptable level of protection to aquatic ecosystems, including a necessary buffer to protect from extreme climatic events (notably droughts).

Page 149: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 121

5 Socio-economic change

5.1 Introduction

Each of the Tier 2 risks assessed in Chapter 4 will also be influenced by social and economic changes. In this chapter we assess the relative influence of climate and non-climate factors for each risk, both with regard to biodiversity and the wider implications for delivery of ecosystem services.

As context, Figure 5.1 is a summary of the current pressures on UKBAP priority species and habitats compiled from the 2005 reporting round when lead partners identified the top five threats in plans over the next five years (Defra 2006b). This identified 15 threats in addition to climate change (here described as ‘global warming’). Habitat loss, infrastructure development, changes in management practices, climate change and invasive non-native species were identified as the top five threats to priority habitats. Habitat loss and degradation (particularly from agriculture and other land management) continues to be a significant threat for a high proportion of species. Woodland management and habitat succession changes are also of concern for some species. Climate change was identified as an emerging threat for a high proportion (47%) of habitats, with infrastructure development recognised as a key emerging threat for both species and habitats (particularly on the coast).

Figure 5.1 Current and emerging threats faced by priority habitats and species as identified by UK BAP plans

Source: Defra (2006b)

To evaluate the relative future influence of socio-economic factors (extending from now until 2080s), six key drivers were distinguished at a generic level within the CCRA. These dimensions of change are used to qualitatively asses the role of non-climate factors in modifying the risk for those impacts identified in Tier 2 (Chapter 4). The six drivers are:

i) Population needs/demands (high/low)

This is based upon changes in population size and distribution (geographically and demographically) and the pressure the population forces onto the country in terms of housing, education etc. One extreme is that there is a high degree of demand on

Page 150: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

122 Biodiversity Sector Report

natural, economic and social resources (demand exceeds supply); the other is that demand is very low (supply exceeds demand).

ii) Global stability (high/low)

This driver of change is based upon world events that could increase or decrease global stability (e.g. war, natural disasters, economic instability). The extremes are higher global stability (with little pressure on Governments and people) compared to today, and lower global stability (with a high degree of pressure on Governments, or people that outweigh other priorities) compared to today.

iii) Distribution of wealth (even/uneven)

This dimension considers the distribution of wealth amongst the British population; the extremes being whether it is more even compared to today or more uneven (with a strong gradient between the rich and poor) compared to today.

iv) Consumer driven values and wealth (sustainable/unsustainable)

Globalisation and consumerism are the primary drivers here, specifically movement towards or away from consumerism values. The extremes are that consumers prioritise their time for working and the generation of wealth, with a focus on the consumption of material market goods and services compared to today; or consumers reduce the importance of work and wealth generation in favour of leisure and less materialism, with a focus on the consumption of non-market goods and services such as conservation and recreational activities in green spaces.

v) Level of Government decision making (local/national)

This driver relates to how centralised policy making is on adaptation; the extremes being whether there is a completely centralised policy compared to today; or whether there is a very small central Government input and high degree of localism in decision making compared to today.

vi) Land use change/management (high/low Government input)

These dimensions relate to aspects of urbanisation versus rural development. The extremes are that looser planning restrictions might increase development in rural areas (development on the green belt, renewable energy in the uplands etc.) compared to today, versus tighter planning which might increase urban development (more brown field sites) compared to today.

Commentary is provided as to the relevance of each socio-economic dimension for each Tier 2 risk (see summary: Table 5.1), together with a brief discussion of potential interactions with climate change impacts and adaptation. Scenarios were also developed during the UK NEA using different but similar drivers and analysed in terms of the consequences for ecosystem services (see Table 5.1).

Page 151: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 123

Table 5.1 Influence of socio-economic drivers on biodiversity risks

BD Metric Population needs/ demands

Global stability

Distribution of wealth

Consumer driven values and wealth

Gov. decision making

Land use change/ management

1 Increased soil moisture deficits and drying

2/7 Coastal evolution impacts and major coastal reconfiguration

3/4 Increased risk from pests, diseases and invasive non-native species

5 Species unable to track changing climate space

or

6 Climate mitigation measures (positive/negative)

8 Changes in soil organic carbon

or

9 Changes in species migration patterns

10 Increased water temperature and stratification of water bodies

11 Generalists favoured over specialists

or

12 Increased risk of wildfire

or

13 Water quality and pollution risk and eutrophication

14,15, 16

Impacts on water quantity

or

Relevant Relevant and a stronger driver of change than climate [NB Relative importance compared to climate change may be dependent on rate and magnitude of change]

5.2 Increased soil moisture deficits and drying (BD1)

This risk occurs as a direct impact of climate on the annual soil water balance. Future climate projections imply that this balance will change with a greater prevalence of warmer drier summers that will be accompanied by reduced precipitation and increased evapotranspiration rates over much of the UK.

However, soil moisture levels can also be strongly influenced by land use change and land management practices. This is particularly apparent through drainage systems: large areas of the UK have agricultural field drains (lowlands) or ditch gullies (uplands) that were installed to remove surplus water and to reduce soil moisture levels. These drainage systems were intended to improve access to the land and general productivity, and in the past were usually grant-aided.

With a changing climate, the combination of climate and land use is likely to exacerbate soil moisture deficits in many areas during the summer. In winter, future projections of increased precipitation may mean that field drains in agricultural areas may continue to be important. However in summer, increased deficits could be detrimental to BIODIVERSITY, AGRICULTURE and FORESTRY by reducing water availability. Good

Page 152: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

124 Biodiversity Sector Report

quality soils with significant amounts of organic material (see BD8, Section 4.7) are more resilient and maintain water availability for longer periods in dry conditions. Land management practices that can have an impact on soil moisture levels include irrigation, crop rotation patterns, grazing (e.g. stock levels) and moorland burning, which can adversely modify vegetation and the organic content of the soil that provides water retention. Many semi-natural habitats are already under stress from atmospheric deposition of pollutants that have altered chemical properties and mineralization rates in soils (see BD13, Section 4.10).

Additional important socio-economic drivers are population demands, which is closely linked with land demand and the pressure for land use change and intensification; and government decision making, which could act to regulate demand and protect vulnerable areas.

5.3 Coastal evolution impacts and major coastal reconfiguration (BD2 and BD7)

The primary sensitivity here is to coastal change occurring as a result of sea level rise, but during extreme events high water levels can be accentuated by storm surge and large waves increasing the risk of large-scale coastal flooding.

As a considerable proportion of the UK coastline has artificial defences (see Section 4.3), loss or gain of habitats in these areas will be strongly influenced by the rate of sea-level rise in conjunction with decisions to ‘hold the line’. This policy decision may be influenced by several socio-economic drivers. In particular, it will be strongly influenced by population demands and land use change, as artificial defences are created to protect settlements, infrastructure and high-quality farmland. These drivers in turn are mediated by consumer values and the distribution of wealth, related to perceptions of what we expect from our coastal zone. A trend towards an increased profile for ‘food security’ in policy agendas and for the population density in some coastal areas of the UK (e.g. South East England) to increase is highly likely to lead to demands for greater protection for coastal settlements, infrastructure and farmland.

These issues represent a key interaction between BIODIVERSITY and other sectors, including FLOODS AND COASTAL EROSION, BUILT ENVIRONMENT, TRANSPORT, ENERGY, and AGRICULTURE. However, the coastal zone also provides a wide range of ecosystem services, including benefits accruing from its high landscape quality, wildlife-related tourism, recreation, seafood, and natural hazard regulation, and a shift towards a greater societal recognition of these benefits could potentially be reconciled with a long-term transition towards a more natural coastline.

A further key influence is the scale of government decision making. The legacy of previous decisions on the coast has shown that local-scale reactive intervention can have highly detrimental consequences on adjacent sections of coast. Coastal evolution is determined not only by water levels but also by the availability of sediment supply. Conversely, rigid national-level strategies could not account for important regional differences in coastal change related to the rate of sea-level rise (and factors such as changes in the wave regime) and past interventions.

Page 153: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 125

5.4 Increased risk from pests, diseases and invasive non-native species (BD3 and BD4)

As the spread of these organisms is predominantly mediated by human agency, this risk is dominated by socio-economic factors. Climate change acts as a background factor that is likely to further increase the current risk, especially due to an increased prevalence of milder winters and enhanced survival rates of invasive non-native species in the UK.

Socio-economic drivers related to globalization and consumer-driven values/wealth are the key factors that have lead to the increased risk from invasive non-native species, with greater import of goods increasing the probability of ‘stowaway’ organisms reaching the UK. Globalization has also been highlighted as likely to lead to an increase in emergent infectious diseases due to increased rates of pathogen recombination (Anderson et al., 2004). Analysis by Hulme (2007) has shown a significant positive correlation between the number of non-native species in European countries and national GDP across a range of taxonomic groups, indicating the link between increased consumer wealth and increased trade in exotic goods that can harbour non-native species. This is exemplified by the expanding range of plants used in gardens and horticulture.

The likelihood of non-native species becoming invasive to native ecosystems is related not only to climate but also other factors such as land use. Population demands and associated changes in land use that lead to simplified, less diverse ecosystems (as for example through agricultural intensification for food provision) increase the propensity for a non-native species to become invasive by reducing the number of natural competitors or predators. In this context there are important cross-sectoral links with AGRICULTURE and FORESTRY. The natural service that ecosystems provide by regulating pests and diseases, as opposed to the use of pesticides, is being increasingly recognised (e.g. UK NEA – Smith et al., 2011).

The scale of government decision making is also an important factor. Effective biosecurity measures require coordination and regulatory controls at both national and international level. Currently agencies at international, EU, UK and devolved administration level provide information and risk assessments on invasive non-native species, pests and diseases. The CCRA identifies an increasing need for climate change to be factored into these risk assessments, and for more work to understand key controls on their spread.

5.5 Species unable to track changing climate space (BD5)

This risk is primarily a combination of climate change and land use change drivers. There are some examples of species not being able to track their changing climate space for natural reasons, notably those in montane areas of the UK that cannot move further north or uphill, or those with very limited dispersal ability. However, in most cases the restriction on species’ movement presently occurs through habitat loss and fragmentation. In general, this has created a landscape where natural species dispersal is often restricted (Lawton et al., 2010).

Land use change is driven by factors such as population demands and consumer values that can result in greater demands from the land, a conversion of semi-natural habitat to ‘improved’ agricultural land or forestry plantations, and general intensification of land use. Intensification usually results in a considerably simplified ecosystem that

Page 154: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

126 Biodiversity Sector Report

has little food or shelter for non-productive species, except possibly in field margins. The reduction in natural soil nutrients is compensated for by the application of artificial fertilizers but the end product is generally a reduction in soil quality. All of these factors mean that intensive production-based landscapes provide very poor habitat. Areas of undegraded habitat (such as protected areas) can often become ‘islands’ within a wider landscape that is more hostile to many species.

In addition to the reduction in size of habitat patches, the loss of connectivity between habitats can provide a severe barrier to species’ movement and dispersal. Depending on the rate of climate change, resultant decline in populations and potential extinctions are likely if further land use intensification occurs, and a particularly notable risk would be the loss of pollinators due to the benefits that they provide for crop production.

The level of government decision making is also clearly a key factor by either encouraging a coherent and robust ecological network or not. A purely local agenda may create landscape and habitat diversity which would be good for the viability of some species, but risks not being adequately ‘joined up’ to allow dispersal through the wider landscape.

5.6 Climate mitigation measures (positive/negative) (BD6)

Ultimately this risk is related to public perception and the role of regulation and markets in delivering sustainable climate mitigation schemes. Climate may have an influence in that a higher rate of temperature rise and increased frequency of extreme events might trigger stronger action. Key socio-economic factors are population demands and consumer values as increased demand for energy would equate with the need for an increased provision of supply (see ENERGY CCRA sector report – McColl and Angelini, 2012). As many renewable schemes have a pronounced land/water ‘footprint’, a trend towards a greater number and scale of schemes has the risk that it impacts negatively on important areas for biodiversity, present and future.

Government decision making also has an important role by regulating supply and demand, and developing an energy strategy through large-scale co-ordination rather than a multitude of ad hoc schemes. The viability of and planning approval for alternative energy projects depends to an extent upon government subsidies and financial incentives plus public/political support for such projects. Attitudes to energy-saving technologies and reduced dependence on fossil fuels will also be important. At a global level, a change in stability that impacted on global protocols and targets could act to undermine or reinforce EU or national targets.

Climate mitigation schemes could also have important interactions with ecosystem services, depending on the scheme and location. Hydro-power can have impacts on aquatic biodiversity and associated activities such as fishing that are important to local economies. Bioenergy schemes could be either developed through plantations of fast-growing non-native species that potentially impact on services such as water quality and soil quality, or they could be based upon native species which could benefit biodiversity, soils and water, and provide additional benefits through cultural services, including recreation.

The level of investment in renewable energy projects is also determined by a range of other factors, including energy prices, provision of public subsidy, likelihood of gaining development consent (e.g. regulatory requirements) and technology. All these are likely to change as increased emphasis is placed upon meeting emissions reduction targets.

Page 155: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 127

5.7 Changes in soil organic carbon (SOC) (BD8)

There has been considerable debate regarding recent changes in SOC with the consensus view that land use change has been the dominant factor rather than climate change (Section 4.7). This relative attribution may change in the future, depending on the rate of climate change and the interaction of climate variables, soil moisture (see BD1, Section 4.2) and CO2 levels, but the complexity of these interactions means that there is considerable uncertainty in the future outcome. A key confounding factor for interpretation of change is also the amount of atmospheric pollution and its interaction with soil carbon, including lagged effects over several decades. Levels of sulphur (S) deposition are declining and projected to continue to decline in the future but nitrogen (N) deposition continues to be a problem and likely to continue to be so in the future. The lagged effects are important because declining levels of S have been linked with increased levels of dissolved organic carbon DOC to watercourses (Monteith et al., 2007). Important cross-sectoral links exist with WATER, AGRICULTURE and FORESTRY.

The same drivers influencing trends in land use and land management described for BD1 (Section 4.2) will be important, notably population demand and consumer attitudes. The scale of government decision making and the role of regulation and markets will also be significant drivers. If a significantly higher price for carbon emissions becomes established, then peatland or wetland restoration and other C sequestration schemes are likely to receive more investment by land managers. This would have co-benefits for biodiversity by retaining water in organic matter and buffering future increases in soil moisture deficits (BD1, Section 4.2). The ecosystem service benefits of C storage and sequestration in peatlands and wetlands for climate regulation is more complex: benefits in terms of C emissions would be offset by an increase in CH4 emissions through anaerobic decomposition until these ecosystems reached an equilibrium state.

5.8 Changes in species migration patterns (BD9)

Climate has a clear direct influence on species phenology and hence on seasonal patterns of migration, but this occurs in combination with socio-economic factors. Land use is again a key influence as migratory species, especially birds, require suitable habitat at their migration sites. Many of these locations, particularly on the coast, are currently designated as Natura2000 sites and receive statutory protection (although vulnerable to the issues highlighted in BD2 and BD7, Section 4.3) but locations outside the site network are more vulnerable to land-use pressures that are linked with population demands and consumer values. The location of sites may become an increasingly important issue if migratory routes alter with climate change, including both new sites and the presence of different species at existing sites, requiring an adjustment of the designating criteria for that site.

Particularly for long-distance migrants, global stability is an important factor as conditions at either the overwintering site or breeding site, or on the migration route, may be substantially altered by changes in stability. This would be an especially prominent issue if international agreements no longer continued to be upheld. Decision making at all levels of government is required in order to ensure coordination of sites within an international network. At local level, the cultural value of migrating species as symbols of the changing seasons may result in a greater recognition of the links between natural and cultural landscapes, for example through local nature reserves.

Page 156: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

128 Biodiversity Sector Report

5.9 Increased water temperature and stratification of water bodies (BD10)

This risk is also primarily a combination of climate change with land use pressures, as climate-driven pressures on thermal regime will interact with water quality (Section 5.10) and water quantity (Section 5.13). The direct impact of climate change will be related to the change in life cycles and availability of dissolved oxygen for respiration. The socio-economic effects will be experienced through land use pressures related to population demands and societal attitudes towards sustainable land use.

5.10 Water quality and pollution risk and eutrophication (BD13)

Studies of the effect of changing water thermal regimes on aquatic ecosystems have found that impacts are often highly dependent on concentrations of key pollutants such as nitrogen and phosphorous (Section 4.10). These pollutants can be traced to either atmospheric deposition from anthropogenic sources or runoff from AGRICULTURE, or less commonly from FORESTRY. Hence land use is the dominant influence, possibly compounded in some areas by population growth and increased discharges of effluent.

The scale of government decision making is important with regard to site protection. Not all of the areas of aquatic ecosystems with priority habitats are included in the Natura2000 network. However, at the river basin scale, an important interface exists with the Water Framework Directive (WFD) which has identified key criteria for ‘good ecological status’ and is a major policy driver in the WATER sector. Synergies between the WFD and Habitats Directive, especially with regard to climate change, remain to be fully established.

With regard to the wider ecosystem service benefits, the role of freshwater fish and angling can be important contributors to the local economy and have a broader cultural value (e.g. during the fishing ‘season’). Water bodies are also widely acknowledged for the landscape amenity value and an increase in toxic blooms from a changing thermal regime could impact on that amenity. Both of these factors could shape societal attitudes towards sustainable land and water use in the future.

5.11 Generalists favoured over specialists (BD11)

This risk has a very similar profile to BD5 (Section 4.5) with regard to climate and non-climate risks. Natural adaptive responses of species are currently being restricted through habitat loss and fragmentation from land use pressures, and this is dominating over climate responses, except for some species that are more generalist in terms of their habitat preferences. These land use pressures can be related to increasing population demands. Loss of habitat is primarily an issue for the AGRICULTURE sector but for some areas (e.g. SW Scotland) the FORESTRY sector has also reduced landscape and habitat diversity. Measures to restore this diversity, particularly at the local level, provide the key adaptation response but this is also associated with government decision making regarding any future changes to the ecological site network.

A reduction in overall biodiversity, resulting from generalist species benefiting at the expense of specialist species, would have important implications for the cultural benefits obtained from wildlife. However, beyond this, as biodiversity underpins most or all ecosystem services to a greater or lesser extent, the potential for disruption to the

Page 157: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 129

ecosystem and the loss of key regulatory and supporting services, such as pollination and nutrient cycling, could have more long-term consequences.

5.12 Increased risk of wildfire (BD12)

Outdoor fires in the UK are almost always started, either accidentally or deliberately, by human agency. Climate is a background risk factor which may increase in the future due to increased soil moisture deficits (BD1, Section 4.2) and even potentially due to the potential increase in lightning strikes tentatively inferred by UKCP09. The dominant factors are however likely to remain socio-economic, notably through an increasing population with possibly greater leisure time, and a greater number of visits to sensitive areas. This may also be associated with a change in societal values or wealth. Land use practices can also have an influence, with burning of moorland in particular still employed as a traditional management practice.

Fire risk is a cross-sectoral issue. CLG (2008) identified numerous other links between socio-economic factors and wildfire, for example across major infrastructure (various types of transportation, wayleaves and buildings), environmental (conservation and heritage designations), social (recreation and access, landscape value, health and wellbeing) and economic (food, fibre and biofuel, field sports and tourism) assets. Depending upon the scale of damage, species and habitats affected, fires have the potential to impact upon many ecosystem services. Of particular importance are the potential impacts upon carbon storage (which may be particularly severe if large-scale fire occurs on peatlands), water purification, and cultural services associated with landscape amenity, recreation and ‘sense of place’.

5.13 Impact on water quality: low flows, increased societal water demand, major drought (BD14/15/16)

For these risks, climate change will interact strongly with socio-economic change. Climate directly influences water supply and indirectly influences water demand which increases in warmer, drier periods. In addition, socioeconomic influences will alter demand-side pressures. Land use change towards increased intensification to meet requirements for ‘food security’ or ‘energy security’ may increase demand for water, particularly if a trend towards more consumerist lifestyles leads to greater demand for high-value crops that increase demand for irrigation water.

The role of government decision-making will also be crucial. The Water Framework Directive identifies a series of indicators for river basins to meet ‘good ecological status’. These include the concept of ‘environmental flow’ which represents minimum flows necessary to maintain viable ecosystems in a catchment. If these regulatory instruments do not continue to be implemented at catchment level, then long-term ecological degradation could occur, with implications for water purification and other ecosystem services.

5.14 Social vulnerability Issues

The welfare of some people and groups within society has vital links to the natural environment. These links may be vulnerable to present and future change, although the connections are often complex and dynamic.

Page 158: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

130 Biodiversity Sector Report

These relationships include increased exposure of vulnerable groups to natural hazards (e.g. flooding, erosion) when the natural buffer provided by ecosystems is lost (e.g. loss of wetlands or woodland). They also include consequences from changes to the supply of clean water, which can be particularly important for those that rely on private water supplies. In addition, people and communities that particularly depend for their livelihood on primary industries (e.g. crops, livestock, timber, genetic or ornamental products) can be severely affected by impacts that directly or indirectly affect those ecosystem services. These impacts can include pests and diseases, loss of water supply or soil quality, changes in pollination, and the natural control of fire. Finally, cultural benefits from the natural environment (e.g. utilisation of greenspace, local identity and ‘sense of place’) can be eroded through direct or indirect change: these are frequently intangible benefits that are of key significance to certain groups who may be particularly adversely affected by their loss.

These issues are currently a source of ongoing research following publication of the UK NEA (2011) but at present firm evidence on current and future vulnerability of individuals and groups is limited.

Scenarios in the UK National Ecosystem Assessment

The UK NEA developed six scenarios, each with two different climate scenarios (high and low). Based upon the scenario storylines, ‘expert’ judgement was used to project how land would be used throughout the UK and consequent changes in ecosystem services. Five ecosystem services were evaluated: i) agricultural food production; ii) net change in greenhouse gases from land use; iii) open-access recreation; iv) urban greenspace amenity; and (v) ‘biodiversity’ (assessed using birds as indicator species). Results for the scenarios are summarised in Figure 5.2 and as follows:

1. Go with the Flow: This is a scenario projection based upon present-day trends, targets and ideals. Overall agricultural incomes and recreational values rise, but an increase in greenhouse gas emissions damages climate change regulation and urban greenspace amenity declines.

2. Green and Pleasant Land: This is a preservationist view of the future with a focus on maintaining national landscape quality. A decline in agricultural intensity leads to reduced incomes from farming but gains in reducing greenhouse gas emissions and improved urban greenspace and recreational values of land.

3. Local Stewardship: This scenario features an emphasis on local-level decision-making and identity. Agricultural incomes, recreation and urban greenspace amenity all improve but the level of greenhouse gas emissions increases slightly.

4. National Security: Under this scenario, world conditions force an emphasis on national-level energy resources. Agricultural incomes increase and a growth in renewable energy means that greenhouse gas emissions fall. However, the prioritisation on production means urban greenspace values decline and social values are negative.

5. Nature@Work: This view of the future is based upon the prioritisation of multifunctional landscapes. Farmland and farm incomes decline, but greenhouse gases are significantly reduced. Recreational and greenspace values are also substantially improved.

6. World Markets: High economic growth with limited barriers to trade dominates in this scenario. Agricultural output is significantly improved but as a consequence greenhouse gas emissions rise. Large losses in urban greenspace value occur as the production value of land is seen as most important.

The scenario analysis in the NEA showed the difference that non-market values from

Page 159: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 131

ecosystem services can have. If only market values are taken into account then storylines that emphasised national self-sufficiency or economic growth resulted in the largest economic gains in the short to medium-term due in particular to increased food and energy production. Conversely, if all monetised values (including non-market values) are taken into account then the scenarios that emphasised environmental awareness and ecological sustainability resulted in the largest economic gains to society, particularly in the long term (Table 5.2). The assessments also revealed significant spatial differences across the UK for each ecosystem service analysed, meaning there would be important regional implications of these changes.

Figure 5.2 Projected trends in ecosystem services for the six scenarios explored by the UK NEA

Source: UK National Ecosystem Assessment (2011)

Table 5.2 Summary and ranking of six future (2050) scenarios for GB based upon four ecosystem services (‘biodiversity’ values were not monetised)

Source: UK National Ecosystem Assessment (2011).

Page 160: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

132 Biodiversity Sector Report

6 Monetisation

6.1 Summary

Climate change adaptation decisions that are designed to reduce climate change risks inevitably involve making trade-offs concerning the use of scarce economic resources. To the extent that economic efficiency is an important criterion in informing such decision-making, it is useful to express climate change risks in monetary terms, so that they can be:

Assessed and compared directly (using £ as a common metric) and

Compared against the costs of reducing such risks by adaptation.

For the CCRA, a monetisation exercise has been undertaken to allow an initial comparison of the relative importance of different risks within and between sectors. Since money is a metric with which people are familiar, it may also serve as an effective way of communicating the possible extent of climate change risks in the UK and help raise awareness.

A tentative summary of the evidence for economic impacts from the risks covered in this report is provided in Table 6.1, mainly based upon a qualitative assessment of the evidence at UK level and extrapolation from limited quantitative case studies carried out to-date. The UK NEA has demonstrated that even baseline knowledge of the relationship between ecosystem services and biodiversity is often rather uncertain, highlighting the need for caution in interpreting this information. Assessing the risks of future change is therefore very challenging, especially as climate will also combine with socio-economic change to produce further interactions that may often be linked to key thresholds and to produce non-linear responses.

Table 6.1 Summary of economic impacts based upon informed judgement of existing evidence base (climate change only – central estimate [50% value];

socio-economic factors following current trajectories; current levels of adaptation)

Risk 2020s 2050s 2080s Estimation method

Confidence ranking

Notes

BD1 Increased soil moisture deficits and drying

Low to Medium

Medium Medium to High

Impacts on ecosystem services

L Extrapolation of results from 2 BAP habitats

BD2 & BD7 Coastal evolution impacts and major coastal reconfiguration

Medium Medium to High

Medium to High

Impacts on ecosystem services

M Extrapolation from case studies and baseline valuation in UK NEA

BD3 & BD4 Increased risk from pests, diseases and invasive non-native species

Low to Medium

Low to Medium

Low to High

Impacts on ecosystem services

L Highly uncertain with potential for ‘surprises’

BD5, BD9 & BD11 Species movement and

Medium Medium to High

High Impacts on ecosystem services

L Assumes current levels of habitat

Page 161: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 133

Risk 2020s 2050s 2080s Estimation method

Confidence ranking

Notes

diversity

fragmentation

BD6 Climate mitigation measures (positive/negative)

Low Low To Medium

Low to High

Impacts on ecosystem services

L Assumes current scale of development with limited adaptation

BD8 Changes in soil organic carbon (SOC)

Low to Medium

Low to High

Low to High

Impacts on ecosystem services

L Likely to be non-linear with threshold effects

BD12 Increased risk of wildfire

Low Low to Medium

Medium Impacts on ecosystem services

L Extrapolation from case studies

BD10/13/14/15/16 Water-related risks

Medium Medium to High

Medium to High

Impacts on ecosystem services

L Extrapolation from case studies and baseline valuation in UK NEA

Marine Biodiversity -Algal blooms -Acidification -Invasives -Species movement - Diseases

Low Low Medium Low Low to Medium

Medium Medium Medium Low to Medium Low to Medium

Medium Medium High Medium Low to High

Impacts on ecosystem services

L See Pinnegar et al. (2012) for further details

Key: Low £1-9million/yr; Medium £10-99million/yr; High £100-999million/yr

Monetisation Uncertainty Ranking:

Ranking Description Colour code

High Indicates significant confidence in the data, models and assumptions used in monetisation and their applicability to the current assessment.

Medium Implies that there are some limitations regarding consistency and completeness of the data, models and assumptions used in monetisation.

Low Indicates that the knowledge base used for monetisation is extremely limited.

6.2 Introduction

UK Government appraisal practice is increasingly moving towards an approach in which impacts on biodiversity are described in monetary terms rather than just through physical impacts to be considered alongside monetised impacts. Although environmental economics has had a long-established role within decision making, the adoption of an ecosystem approach within the last few years, including the characterisation and valuation of ecosystems services has encouraged this shift in

Page 162: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

134 Biodiversity Sector Report

practice. The concepts have also been further developed through the UK National Ecosystem Assessment (2011; Box 6.1) and internationally by a large initiative on The Economics of Ecosystems and Biodiversity (TEEB, 2010). Nevertheless, there is often a paucity of data at present for many ecosystem services, with the subject being the topic of considerable new research61, necessitating a cautious approach. During the next cycle of the CCRA it is very likely that improved methodologies and data will become available that link economic valuation with an improved understanding of ecosystem functions and services. The fundamental challenge is to ensure that policy appraisals fully capture the benefits from ecosystem services as well as the costs of the impacts of policies on those services (Defra 2007b; Fish et al., 2011).

Many people recognise that biodiversity has intrinsic value (i.e. a non-humanistic right to exist) rather than just a utilitarian value to humans but this judgement is typically personal and highly subjective. Objective valuation methods therefore aim to derive extrinsic (instrumental) values for the benefits from biodiversity and ecosystems by identifying their human welfare benefits (Defra, 2007b). The rationale for this is that by including these values within national economic frameworks and decision making processes, it can be used to justify the financial resources required to sustain, restore or enhance ecosystem services and therefore make substantially better decisions regarding land, marine and natural resource use in order to benefit both present and future generations.

Ecosystem services are directly paid for if the benefit is a tangible material product, such as food or timber, which are traded as ‘goods’ on markets. However, most benefits from ecosystem services are an improvement in the condition or location of things of value to human wellbeing (public good benefits), rather than directly traded products with an explicit market value. Failure to take account of public good aspects of biodiversity and other externalities (i.e. those beyond the direct market value) may lead to biodiversity being undervalued in a market-orientated world. The consequence of this is that appropriate action is not taken to conserve biodiversity for the benefit of society as a whole.

Box 6.1 Headline findings from the UK NEA

The amenity value of all wetland types, including coastal, is around £1.3 billion/yr.

Biodiversity pollination services are estimated at £430 million/yr.

The value of UK fish landings is about £600 million/yr, while that of aquaculture (fish and shellfish farming) is around £350 million/yr.

The total value of net carbon sequestered currently by UK woodlands is estimated at £680 million/yr.

Loss of stored carbon through emissions from degraded peatlands is currently estimated to cost -£130 million/yr.

The average value of flood risk reduction from wetlands (inland and coastal) is estimated at £1279 per hectare per year, although in some locations (particularly coastal sites) this will be considerably higher.

Modelling suggests that a medium-sized (5km2) woodland site can generate recreation values of between £1,000-65,000/yr depending purely upon its location relative to population centres.

61

Major current developments in valuation include the Natural Capital Initiative and the Valuing Nature Network.

Page 163: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 135

Willingness to pay (WTP) estimates of the non-use (existence) value of terrestrial biodiversity range from £540-million to £1,262 million per annum and for marine biodiversity, estimates of around £1,700 million/yr have been reported.

Legacy values for biodiversity are estimated at around £90 million/yr.

A wide range of health and well-being benefits associated with greenspace have been reported over and above those induced by increased exercise. These include psychological, emotional and mental health benefits, reduced stress and increased quality of life.

6.2.1 Total economic value and associated valuation techniques

The concept of Total Economic Value (TEV) can be employed to capture both the ‘use’ and ‘non-use’ values of ecosystem services (Defra, 2007b; TEEB, 2010; Figure 6.1). The use value components include:

Direct use values, where individuals make actual use (consumptive or non-consumptive use) of particular species and habitats, such as in agriculture and forestry. Tourism can also benefit from non-consumptive use values associated with charismatic species: for example, Scottish Natural Heritage has estimated the value of sea eagles to the local economy of the island of Mull at ca. £5million/yr.

Indirect use values, which are unrelated to current use but are nonetheless linked to the location including regulating services such as carbon sequestration by soil and vegetation, or regulation of water and soil quality.

Option values, where individuals are willing to pay for potential benefits in the future. These values might be related to future visits for recreational purposes. Alternatively, they can be associated with ‘bioprospecting’ in which there is an anticipation of new pharmaceutical applications from particular species62. Such values can also be considered analogous with ‘insurance’ values that aim to prevent any changes in the ecosystem that could bring irreversible negative consequences for human wellbeing. Even if an ecosystem or its components are currently believed to generate no immediate value now, its option value may still be significant63.

62

A notable example of bioprospecting is the potentially significant treatments for Alzheimer’s Disease being derived through galantamine extracted from daffodils (Narcissus pseudonarcissus) and snowdrops (Galanthus nivalis). With the growing dementia population and naturally variability in galantamine levels between species, conservation of Narcissus species can be demonstrably linked with very important social and economic benefits (JNCC, 2011). 63

Currently, environmental economists interested in valuing resilience of ecosystems regard it not as a property but as natural capital (stock) yielding a ‘natural insurance’ service (flow) which can be interpreted as a benefit amenable for inclusion in cost benefit analysis.

Page 164: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

136 Biodiversity Sector Report

Total Economic Value

Use values Non-use values

Actual valueOption

value

Direct

use

Indirect

use

ConsumptiveNon

Consumptive

Philanthropic

value

Altruism to

biodiversity

Bequest

value

Altruist

value

Existence

value

Crops,

livestock,

fisheries,

wild foods,

agriculture

Recreation,

spiritual/cultural

well-being,

research,

education

Pest control,

pollination,

water regulation

and purification,

soil fertility

Future use of

known and

unknown

benefits

Satisfaction of

knowing that

future generations

will have access to

nature’s benefits

Satisfaction of

knowing that

other people

have access to

nature’s benefits

Satisfaction of

knowing that a

species or

ecosystem

exists

Market prices

Cost methods

Production

functions

Market analysis

Cost methods

Hedonic pricing

Contingent

valuation

Replacement costs

Mitigation costs

Avoided costs

Contingent valuation

Contingent election

Figure 6.1 Components of total economic value and associated techniques used to elicit values for biodiversity and ecosystems

Adapted from TEEB (2010)

Non-use values accrue through individuals who derive a benefit from knowing that some valued component of biodiversity and ecosystems will continue to exist at some location including:

existence values, which reflect the fact that individuals may value the 'existence' of key components of biodiversity regardless of any aspirations for current or future use;

altruistic values, which might arise because individuals wish others in the current generation to also have these benefits (i.e. to provide intragenerational equity);

bequest values, where individuals aim to ensure that future generations will also be able to enjoy the benefits and opportunities (i.e. to provide intergenerational equity).

Valuation techniques aim to establish either an individual’s willingness to pay (WTP) for an ecosystem service (or to avoid its degradation) or willingness to accept (WTA) compensation for the degradation or non-improvement of an ecosystem service. Depending on the service, and its association with use and non-use values, different methods can be applied to provide the valuation. These include the following techniques (with the exception of ‘market prices’ these are all examples of non-market valuation):

Market prices estimate the value of traded ecosystem goods and services (e.g. timber, crops, fish) that are traded in formal markets. They can also potentially be applied to some regulating services (e.g. to value the reductions in damage to property caused by flooding).

Page 165: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 137

Cost methods are based on damages due to the loss of an ecosystem service, or expenditure to prevent that damage, or the cost of replacing the ecosystem service altogether. For example avoided costs can be used to assess the value of some regulating services, including water treatment (due to water purification services) or avoided expenditures on flood defences (due to water regulation services).

Revealed preference methods, such as travel or access costs, which aim to identify a WTP to obtain a ‘service’ (e.g. for recreation), or alternatively those that are expressed through property or land values (hedonic pricing).

Stated preference methods are typically obtained during surveys or choice experiments to determine people’s willingness to pay for ecosystem services in hypothetical markets.

Deliberative and participatory valuation methods, including group-based deliberative monetary techniques, citizen’s juries, and deliberative multi-criteria assessment. These are an innovation beyond conventional economic approaches to include the shared and social benefits that society obtains from the natural environment (see Section 6.1.2 below).

Benefits transfer, whereby values from one location are applied to another location for which primary valuation data are not yet available. This approach can be applied using data based upon any of the other methods. However, it is often limited by the site-specific context of many of ecosystem services and by the availability of suitable data.

Non-use values are a critical component of valuation for biodiversity and ecosystems, but provide difficulties for conventional economic approaches because they do not provide outputs measurable in terms of productivity or of observable behaviour. In particular, the use of survey-based stated preference methods (such as Contingent Valuation and Choice Modelling Experiments), to elicit ‘existence values’ has been contested. The validity and reliability of these estimates, derived in hypothetical ‘markets’, may contradict with the likelihood and feasibility of projected outcomes from a natural science perspective (Sagoff, 2011). Nevertheless, studies have suggested that these existence values may account for a substantial proportion of the overall cultural services people receive from protected areas (CJC Consulting, 2004; Eftec, 2006).

6.2.2 Limitations of valuation

Different valuation techniques have their strengths and weaknesses related to issues of spatial and temporal scale, uncertainty, non‐linearity and thresholds. Knowledge about

the multiple benefits of ecosystem services often requires integrating quantitative and qualitative evidence derived under different philosophical, theoretical and methodological assumptions. In general, information is often more robust for provisioning services because as tradable commodities they typically have direct ‘use’ values (e.g. crops or water). However, information is usually less robust for most regulating services and generally poor for cultural services. Applied valuation techniques appropriate for high experience goods and services may not necessarily be valid for low experience ones for which stable preferences have yet to be formed (Bateman et al., 2011a). Human cognitive limitations can therefore undermine monetary valuation (particularly non-use estimation through stated preference techniques in hypothetical ‘markets’) because individual survey respondents can have difficulties in framing the issue against appropriate reference points in the ‘real’ world (Turner et al., 2003).

Page 166: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

138 Biodiversity Sector Report

It is important to recognise that monetisation can only provide a partial characterisation of the benefits from biodiversity and ecosystems. Most practitioners agree it should be used to complement a broader range of factors used in decision making rather than be the sole focus. The complexity of inter-relationships and inter-dependencies within ecosystems mean that monetisation and even quantification can only cover a fraction of the benefits to humans (Figure 6.2). At a fundamental level, ecosystems act as the life-supporting systems for the planet, therefore their inherent value is effectively infinite. The UK NEA and other studies aiming to apply economic valuation techniques have circumvented this issue by considering marginal changes in ecosystem goods and services, as defined by incremental changes in unit value of stock (e.g. per hectare of habitat). However, the application of marginal valuation has its limits: these include the implications of ecological threshold effects in complex systems that mean decisions become irreversible or reversible only at prohibitive costs. Such ‘tipping points’ can mean that small changes can lead to proportionately much larger impacts, resulting in non-marginal behaviour, and even system collapse (Cornell, 2011). The marginal unit value only therefore applies within certain ‘safe’ limits and may not necessarily produce robust values for policy appraisals in situations of high uncertainty and/or where non-use values have a wide diversity of interpretations (Figure 6.3).

A further limitation of conventional economic tools, such as cost-benefit analysis (CBA), is that they are based upon individualistic concepts of human welfare. This can act to exclude the broader benefits to human well-being and quality of life from the natural environment through collective meaning and significance. Social, health‐related, and

cultural interaction with ecosystems is related to factors such as social norms, altruism, ethics, aesthetics, and existence value (Fish et al., 2011). CBA and other techniques can therefore tend to overlook issues of distribution, equity, fairness and social justice (Sagoff, 2011). The UK NEA has recognised the multiple dimensions of ‘value’ by emphasising that the full range of ecosystem services to human well-being is provided by economic benefits, health benefits, and shared/social benefits (e.g. spiritual or aesthetic appreciation).

For ecosystems, total system value is always greater than total economic value (Turner et al., 2003) therefore analysis of individual components as ‘services’ or ‘goods’ can be contrary to measures designed to enhance system integrity and resilience. These latter measures aim to assess and conserve the holistic structure, processes and functions of ‘healthy’ evolving ecosystems, whilst recognising that our knowledge of these and of key thresholds for change often remains limited. Unfortunately, most direct evidence at present is usually negative in that it is related to foregone benefits when a service has been lost or degraded.

Page 167: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 139

Figure 6.2 Schematic illustration of the role of monetary valuation within the wider context of quantitative and qualitative knowledge

Source: TEEB (2008)

Figure 6.3 The limits to monetary valuation for complex systems and where multiple values prevail

Source: TEEB - Brondizio et al. (2011)

6.2.3 Valuation of UK biodiversity: costs and benefits

In lieu of more detailed valuation, the cost of managing biodiversity is sometimes taken as an indication of the value that society holds for biodiversity. This is based on the assumption that targets and legal mechanisms introduced to support biodiversity are supported by the public through the political process. Total projected annual costs for biodiversity management in the UK, based on achieving the targets in the BAP (2010-2015) and for managing protected areas, has been estimated at just over £1 billion a year, consisting of £837 million for the BAP (Christie et al., 2011), £217million for protected areas and £63 million for the marine environment (GHK, 2011a). Current

Page 168: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

140 Biodiversity Sector Report

expenditure on the UK BAP is £564million per annum, meaning there is currently a shortfall in resources required to deliver BAP targets. These cost estimates partly overlap with analysis conducted for the Lawton Review (Lawton et al., 2010) which estimated that the total annual costs of establishing a coherent and resilient ecological network in England are likely to lie in the range of £600 million to £1.1 billion per annum. By implication, extending this estimate to cover Scotland, Wales and N. Ireland would increase costs to approximately £1.5 billion to £2 billion per annum.

These high-level exercises have also used valuation techniques based upon ‘choice experiments’ to identify public preferences for well-defined ecosystem services64 when varying outcomes are offered at different ‘prices’; this was complemented by surveys of expert opinion to identify added value of these services to habitat management. The resulting values provide an indication of the benefits that people obtain from biodiversity which can then be assessed against costs. These benefits were established (based on current levels of provision) at £1.4billion from the BAP (Christie et al., 2011) and £960million for the SSSI network (GHK, 2011a). In both cases, benefits considerably exceed levels of expenditure (£564million for the UK BAP and £111 million for SSSIs in England and Wales) and suggest that benefits from biodiversity are currently undervalued in national accounting systems. The analysis of the SSSI network also suggested a WTP estimate of a further £769 million to secure additional benefits from all SSSIs being in favourable condition. For reference, currently-protected sites have 3.3 times the level of biodiversity expected from sites outside the network and 1.8 times more storage of carbon (Eigenbrod et al., 2009). Further high-level information is provided by estimates of the wildlife, landscape and carbon benefits of the Higher Level Stewardship Scheme (using a different methodological approach) which have been estimated at £590million/yr (Fera and Newcastle University 2010).

Analysis for the England Biodiversity Strategy suggests that the benefits of moving from current levels of biodiversity to those identified within the 2020 mission statement for England (see Chapter 1, including targets to increase the total area of priority habitats by 200,000ha) could be worth in the range of £294 million/yr through the wider availability of ecosystem services, against additional costs of £100 million/yr. Similarly, in Scotland, a study of the costs and benefits of Natura 2000 sites estimated that the amount the public was willing to pay to protect these sites was £210 million per year; benefits were estimated to outweigh costs by a ratio of 7:1 (Jacobs, 2004). Of these WTP values in Scotland, less than 1% was accounted for by the value of using the sites and 99% was related to non-use values. GHK (2011a) also noted that a large proportion of the WTP values for SSSIs arose from a general public concern to protect the site rather than to visit the site.

Valuation of the benefits from SSSIs (GHK, 2011a) has also indicated that the broad habitats estimated to deliver the highest annual value of ecosystem services (under current funding levels) were heathland (£320 million), bogs (£195 million), fen, marsh and swamp (£101 million) and intertidal mudflats and saltmarsh (£76 million). When the value of services provided per hectare of habitat was analysed, the highest values (assuming the same level of funding) were provided by sand dunes and shingle (£1377/ha), heathland (£1141/ha), intertidal habitats (£1107/ha), bogs (£1035/ha), and broadleafed woodlands (£1002/ha), and the lowest by inland rock (£200/ha) and coniferous woodland (£237/ha). Valuation of the UK BAP (based upon current spend) estimated that the highest value BAP habitats at UK level were blanket bog (£607m for 2,209,000ha), upland heath (£145m for 981,500ha), native woodland (£258m for 1,059,200ha), and improved grassland (£172m for 5,206,000ha) (Christie et al., 2011); however this does not include the high values that some limited-extent habitats provide in a small area (e.g. coastal habitats; lowland heath; limestone pavement).

64

These ecosystem services included the direct ‘provision’ of both familiar and unfamiliar species, as well as wider benefits of access to nature, and water and climate regulation.

Page 169: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 141

Restoration or re-creation costs can provide indicative estimates of the implications of climate-changed induced changes (Table 6.2), although in most cases restoration or re-creation would take many years to establish a functional ecosystem. These costs can be considered as either an indication of societal willingness-to-pay to maintain or expand habitat coverage and their ecosystem services; or as necessary requirements to avoid infraction proceedings associated with the obligations from international agreements (e.g. EU Habitats Directive). There may also be important issues to consider regarding replication of the original range of services and benefits provided by the habitat which are not necessarily covered in these indicative figures. For instance, the loss of wetland habitats could mean loss of services related to flood defence and water quality within a river basin; costs would therefore be required to also include an alternative method of reducing flood risk and improving water quality at that location (which might be expensive engineering solutions).

Table 6.2 Indicative habitat restoration and creation costs when established through management agreements

Restoration unit costs (£/ha) Creation unit costs (£/ha)

Upland habitats 2151 7382

Lowland heathland 8530 11791

Lowland grassland 10168 11293

Woodland 7776 7436

Wetlands 9436 11072

Coastal 4509 48758

Source: GHK (2011b) NB These annual costs could be reduced by ca. 50% through land purchase schemes although

this would entail a large initial payment to acquire the land.

In addition to providing a baseline valuation of ecosystem services, the UK NEA conducted scenario analyses, combining land use change and climate change to provide initial estimates for future projections in five key ecosystem services (Bateman et al., 2011b). An exploratory assessment for biodiversity was developed through indicator bird species using a diversity index and woodland bird communities, although species were not necessarily those of conservation interest, and valuation was not provided for this index. The other services assessed and valued were agricultural food production, greenhouse gas emissions, recreation and urban greenspace amenity. Reference is made to the valuation of changes in greenhouse gas emissions specifically for soil organic carbon in Section 6.2.6; the other services are related to the biodiversity sector more indirectly and are not considered further.

6.3 Specific risks

6.3.1 Increased soil moisture deficits and drying (BD1)

A shift to drier summers would be very likely to have negative consequences for some priority habitats and species, particularly for wetland areas. This risk was assessed in Section 4.2 through case study examples for blanket bog, beech woodland, and upland bird species (golden plover etc.). The interaction of climate and non-climate factors mean it is not possible to define an unequivocal relationship between climate and habitat or species loss, but modelling studies have identified indicative future changes based upon key assumptions. These can be used to further explore the implications for loss of ecosystem services.

Page 170: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

142 Biodiversity Sector Report

With regard to blanket bog, valuation for the UK BAP (Christie et al., 2011) estimated that this habitat provides a ratio of benefits to maintenance costs of the order of 16:1 when wider ecosystem services are considered. In addition to their high biodiversity value, peatlands provide key regulatory services (e.g. carbon storage, regulation of water quality/quantity) and have significant cultural and amenity value. The NEA estimated that emissions from peatlands currently cost £130million/year. Bioclimate modelling of blanket peat by Clark et al., (2010b) has suggested that ~50% of the current area is vulnerable to change. This reduction in extent could have major implications for ecosystem services, particularly for carbon storage.

Unfortunately, process models do not currently give a robust indication of the rate of change of carbon flux with climate change in the UK. Warmer drier conditions will increase decomposition rates and emissions of carbon dioxide but drier peatlands may emit less methane, so the net greenhouse gas flux is hard to predict. In eroding peat catchments, annual carbon losses of up to 100 tonnes per km2 have been reported. Assessment of climate-influenced land use change on carbon stocks in the UK NEA has suggested potential losses of approximately 37% (113 tC/ha) by 2060 (Bateman et al., 2011b). A study by Eftec for Natural England (2009) found that the value of reduced carbon losses from the restoration of an SSSI blanket bog at Bleaklow in the Peak District was estimated at £0.4 million over 50 years. As a test of sensitivity, using the DECC 2010 central price for carbon of £52 per tonne, a loss of 10% of the 2300MtC stored in UK peatlands by 2080 would imply a net cost of £170million/year. Further work is required to establish if this scale of loss is credible. In addition to the loss of stored carbon, active peat bogs in pristine condition can sequestrate carbon at up to 0.7 tC/ha/yr (Holden et al., 2007), therefore the potential loss of these benefits needs to be also included within further analysis of vulnerable areas.

Additional costs from degradation of blanket bog may accrue due to impacts on water quality, unless remedial measures can reduce the impact. Avoided cost calculations can be made of the benefits of reducing water colouration problems by blocking drains to reduce peat decomposition rates. These are only currently available on a catchment basis, with indicative costs (e.g. from the United Utilities / RSPB Sustainable Catchment Management programme) indicating that peat restoration could avoid water treatment costs of £1-2m/year. Clearly, if peat bogs lose the ability to provide this service then treatment costs will be unavoidable. Increased water colouration has been identified as an issue for water companies in at least 65% of catchments. Further costs may be associated with loss of amenity value, but these are again difficult to quantify, not least because it will depend on the replacement land cover. In aesthetic terms, a transition to wet heathland may not be viewed as negatively by visitors as a shift towards an eroded, unvegetated and degraded peat landscape.

With regard to loss of beech woodland, the main services affected would be loss of carbon storage/sequestration and of cultural services related to recreation and amenity value. Bioclimate modelling by Broadmeadow et al., (2005) has suggested that by the 2050s beech woodland in southern England could be particularly vulnerable to habitat loss. A 10% loss of this habitat (the UK BAP estimates the extent of lowland beech and yew woodland at ca.30,000ha) assuming an average carbon density of 78.2t/ha (Milne and Brown, 1997) would represent a loss in value of ~£12million based upon the DECC 2010 central carbon price (£52 per tonne). Creation of new habitat elsewhere in less drought-prone areas and on less drought-prone soils would be a feasible adaptation strategy. Replacement costs for establishment of beech woodland are not available, but a standard value for creation of new woodland has been established at £8000/ha (GHK, 2011b).

Monetisation of losses to particular species are very difficult to quantify in the absence of specific data associated with ‘use’ values (e.g. tourism revenue from birdwatching). Decline of distinctive upland wetland species, in some cases to the level of possible

Page 171: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 143

extinction by 2100 (e.g. golden plover: Pearce-Higgins et al., 2009), will clearly have a detrimental impact on the cultural and amenity value of these areas that may be reflected in visitor numbers. In addition, these species have a significant non-use (existence) value to many people.

6.3.2 Coastal evolution impacts and major coastal reconfiguration (BD2/BD7)

Coastal habitats provide a broad range of benefits in addition to their intrinsic biodiversity value, including protection against flooding and erosion, regulation of water quality, high diversity of wild species, and amenity value. The average value that each of these ecosystem services provides has been estimated by the UK NEA (Table 6.3) together with the marginal values for benefits gained or foregone through habitat gain or loss. These values will vary significantly depending on the location of the site, particularly the relative geographic position with respect to people and settlements that would obtain that benefit (Bateman et al., 2011b). In general, measures to maintain habitat at existing locations (where this is possible) tend to provide a more cost-effective response than seeking to develop replacement habitat elsewhere. This is particularly demonstrated by saltmarsh and mudflat habitats: recent re-evaluation of schemes by the Environment Agency indicates that replacement costs for these habitats are ca. £50,000 per ha, which is significantly higher than for other habitats (Table 6.2). However, at some locations the maintenance of habitat in situ may be considered unviable and the only alternative is to seek compensatory habitat elsewhere under the terms of the EU Habitats Directive.

Based upon the analysis undertaken by the UK NEA, the estimated gross value of coastal wetlands for buffering the effects of storms and in controlling flooding has been assigned a notional value of £1.5 billion annually (Bateman et al., 2011a). However, to-date there has been no national-level assessment of the relative benefits of natural against man-made defences. Instead, evidence is provided by case studies of local schemes that have estimated the avoided costs from man-made defences when natural schemes are present, although these appraisals have tended not to quantify the full range of ecosystem benefits. For example, it has been estimated that an 80m width zone of inter-tidal habitat fronting sea walls can save £2600-4600 per metre in sea defence costs (King and Lester, 1995). At some locations, investing in habitat restoration can therefore provide a cost-effective alternative to the maintenance or upgrade of hard engineered defences as a response to sea-level rise.

Table 6.3 Ecosystem service valuations for coastal wetlands

Average additional value of service where

present(£/ha/yr)

Marginal value of service per new unit area

(£/ha/yr)

Flood control and storm buffering

3730 2498

Wild species diversity 2786 1866

Water quality 2676 1793

Water supply 16 12

Amenity and aesthetics 2080 1394

Source: Bateman et al. (2011a) Default value where no services are present is £1856ha/yr.

Coastal ecosystems can also store and sequester carbon, particularly in saltmarsh habitats through the interaction of vegetation with fine sediment particles. It has been estimated that this could provide an additional value of £60–600/ha/yr (Beaumont et al.,

Page 172: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

144 Biodiversity Sector Report

2010). Furthermore, this sequestration role also acts to store other pollutants that would otherwise reduce water quality, particularly in estuarine environments. The value of coastal habitats in providing food is another service where data is currently lacking: often these ecosystem goods have a high-value or premium price based upon quality and diversity.

Specific local data is available through economic appraisal of managed realignment schemes. These appraisals have shown that such schemes can be cost-effective, particularly when based upon a long-term analysis period, but again that spatial location is critically important in determining the ecosystem services generated, the human beneficiaries, and the relative benefits against costs (e.g. Humber estuary and Blackwater estuary - Turner et al. 2007; Luisetti et al. 2011). It has been also suggested that larger schemes may provide less benefits in marginal economic terms (Bateman et al., 2011a); however, this assertion needs to be reconciled with minimum interventions required from a biophysical perspective to meet the primary goals of flood risk reduction and creation of habitat for priority species.

6.3.3 Increased risk from pests, diseases and invasive non-native species (BD3/4)

Initial estimates of the total costs of invasive non-native species have indicated that the annual cost associated with yield loss and control of plant pathogens is £400 million in Great Britain (CABI, 2010). The same study has estimated a total cost for all invasive non-native species of £1.7 billion/year; Japanese knotweed is identified as one of the most problematic species with an annual cost of £179million. Early proactive action may be particularly cost-effective for this risk: for example, it is estimated that proactive measures to eradicate water primrose from waterways and aquatic habitats would cost £73,000 but this could be considerably less than the estimated £242 million that it would cost if the plant was to become widely established as it has in countries like France and Belgium.

For some invasive non-native species, control costs are primarily to protect native biodiversity rather than agriculture, forestry or aquaculture. These include the signal crayfish, where much of the control work is to protect native crayfish, and for which total costs are estimated at £1.5million/yr. Similarly, control measures for the carpet sea squirt (Didemnum vexillum) are primarily for conservation reasons, with individual case study estimates of control measures indicating a cost of £100,000 per location. The CABI (2010) report estimated the total annual direct costs of invasive non-native species to biodiversity to be at least £38million (£20.6miilion for conservation, £17.4million for research).

Although it seems likely that the costs associated with protecting priority habitats and species from this risk will increase due to climate change, due to the interaction of climate and non-climate factors it is not possible to quantify this additional risk. It will vary according to the risk factors identified in Section 4.4 (introduction, establishment, spread, and impact) and the virulence of specific invasive non-natives. With regard to the benefits provided by biodiversity in regulating the spread of invasive non-natives, pests and diseases, the evidence base for this is expanding but currently rather limited (e.g. Bianchi et al. 2006), and valuation data are not available to indicate how it may reduce damage or remediation costs.

Page 173: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 145

6.3.4 Risks to species movement and diversity (BD5 /BD9 / BD11)

Risks associated with changes of biodiversity due to species being unable to disperse through the landscape (BD5), shifts in migratory patterns (BD9), and the loss of specialised niches (BD11) are here considered together. These risks are primarily related to the need for habitat availability, diversity and connectivity. As explained earlier, valuation data for individual species is usually very limited unless obtained from case studies for particular locations and for charismatic or harvestable species (e.g. fish, wildfowl).

The Lawton Review estimated that the total annual costs of establishing a coherent and climate-resilient ecological network in England are likely to lie in the range of £600 million to £1.1 billion/year. This would act to moderate the more severe impacts of climate change on biodiversity, although some individual species loss is probably inevitable.

These estimated costs need to be balanced against the additional benefits that accrue from the ecological network (both through ‘use’ and ‘non-use’ benefits). These include the role of priority species and habitats in ensuring a healthy functioning ecosystem and through the provision of ecosystem services to humans. Both the recent reviews of the SSSI network (GHK, 2011a) and the BAP (Christie et al., 2011) have demonstrated that monetary benefits are considerably higher than costs in providing a landscape that sustains and enhances the human welfare benefits from the natural environment (Section 6.1) A further factor to be considered is the possible infraction costs if species or habitat loss was considered to be in breach of obligations under EU directives for designated sites: for example, the role of the SPA network for protecting migratory birds is particularly important in an international context.

With regard to measures to maintain species and habitat diversity, insect pollinators provide a key exemplar because of the wider ecosystem service that they provide. The value of ecosystem pollination services have been estimated by the UK NEA at £430 million/yr and it has also been inferred that about 20% of the UK cropped area comprises pollinator-dependent crops. A high proportion of wild flowering plants depend on insect pollination for reproduction therefore it also provides a supporting service that helps maintain the integrity of ecosystem functioning. In addition to the costs accruing from loss of species that are currently harvested, a reduction in species and genetic diversity also impacts on the ‘option value’ provided by any future known or unknown benefits from particular species (e.g. ‘bioprospecting’ in the pharmaceutical industry).

6.3.5 Climate mitigation measures (positive/negative) (BD6)

This risk is defined by mitigation schemes that do not take appropriate account of a changing climate; it is assumed that present-day impacts on biodiversity are adequately accounted for through Environmental Impact Assessment. Costs and benefits associated with the integration of adaptation measures into current and proposed schemes for climate mitigation will vary dependent on context and spatial location. As such schemes are at an early stage then data is currently very limited. However, it seems highly likely that incorporation of adaptation into these schemes from their inception will be more cost-effective than applying remedial measures at a later stage when the climate has changed. The UK Government is currently introducing sustainability standards for biomass used for heat and electricity to accompany those that apply for transport biofuels.

Page 174: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

146 Biodiversity Sector Report

In England, the Natural Environment White Paper emphasises that, in moving towards the low carbon economy, the Government has a key role to ‘work with others to establish a research programme to fill evidence gaps about impacts on the natural environment of the level of infrastructure needed to meet 2050 objectives, in particular with respect to the cumulative and indirect effects’. Such information throughout the UK will enable a more strategic approach to assessing the co-benefits from GHG emissions reductions together with energy security, affordability and protection of the natural environment, rather than considering mitigation and adaptation requirements separately. In this context, a notable example is the co-benefits from maintaining levels of soil organic carbon (see Section 6.2.6) or of enhancing the stock of native woodlands.

6.3.6 Changes in soil organic carbon (SOC) (BD8)

There has been limited research to-date that has attempted to value the many ecosystem services provided by soil (Harris et al., 2006). Organic content is a fundamental property of soils that is associated with biodiversity, nutrient cycling and water cycling, together with final ecosystem services such as soil quality, crop production, water regulation/purification and climate regulation (via carbon storage). Much of the valuation work to-date has focussed on the carbon storage value, linked to its importance for climate change mitigation and including the much smaller proportion of carbon stored in vegetation. For example, based upon analysis of the costs and benefits from the SSSI network, GHK (2011a) inferred from choice experiments a WTP estimate for climate regulation of £89/household/yr.

Estimates have been made of modifications to soil carbon content due to different land use changes (Harris et al., 2006): for example, if drainage of previously undrained upland peat takes place, the loss of stored carbon is valued at £2,619 per ha (£105/ha/yr)65. As highlighted in Section 6.2.1, the impact of climate change on the loss of carbon storage or on emission/sequestration rates is currently subject to high scientific uncertainty. In addition to its direct effects (via temperature and precipitation), climate change is also likely to be a factor in influencing land use change in conjunction with socio-economic drivers, particularly in marginal upland areas. Further losses are also possible due to changes in soil erosion rates.

In England, research for the Natural Environment White Paper has estimated that soil degradation as a result of erosion by wind and water, and the loss of soil organic matter and compaction costs the economy at least £150-250 million per year. Potential additional losses due to climate change (direct and indirect) would exacerbate this and could lead to a general decline in soil quality that affects other ecosystem services such as crop and timber production. The UK NEA assessed future changes in soil organic carbon (as a component of greenhouse gas emissions) using a range of six socioeconomic scenarios via land use change combined with climate change. Results varied between an upper limit of 7.5 MtCO2e/yr (‘World Markets’ scenario) and a lower limit of -3 MtCO2e/yr (‘Nature@Work’ and ‘Green and Pleasant Land’ scenarios) to 2060 in the lower climate change projection (UKCP09 low emissions); positive figures indicate net emissions and negative figures indicate net sequestration (Bateman et al., 2011b). Using the DECC central value for the price of carbon, this produces a range in valuations from a cost of £390 million/yr to a net benefit of £156 million/yr, this variation indicating the key role of land management rather than uncertainty over the influence of climate change.

65

For this assessment the ‘social costs of carbon’ at £70 per tonne was used with a discount rate of 3.5% per annum over 60 years

Page 175: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 147

6.3.7 Increased risk of wildfire (BD12)

Again limited data are available for this risk, and these are based upon case examples rather than a systematic national assessment. There are two distinct size-duration categories of fire; small ‘4-hour’ fires of typically less than 15 ha, and large ’3-day’ fires of over 350 ha. From a preliminary study of fourteen fires, the former category has been estimated to cost Fire and Rescue Services around £15,000, and the latter £210,000. Such figures exclude helicopter costs, damage to property, livelihoods and other ecosystem services, therefore they represent a significant underestimate. Suppression costs for the April 2003 fire on Bleaklow in the Peak District (which covered 7.4km2) were estimated at approximately £500,000. The Moors for the Future partnership have so far restored 4.3 km2 of the affected site at a total cost of approximately £1,235,000, leaving another 3 km2 of the area to be restored; cost per hectare has been estimated at £2,900. In addition to the re-establishment of habitat, land cover restoration after wildfire will provide ecosystem services such as carbon sequestration (although this effectively replaces the lost carbon), improvement of water quality to its previous state, and restoration of amenity value, including for recreation and hunting.

As highlighted in Section 4.12, habitats at highest risk from wildfire include lowland heath and upland heath. Heath vegetation itself is estimated to take 20 years to re-establish its full maturity, and in the meantime is likely to be invaded by bracken and grasses which can lead to the complete elimination of heath without further managed intervention. Restoration or recreation costs for lowland heath range from £8530-11791/ha (Table 6.2) giving an indication of the cost implications that large-scale damaging fire events incur.

6.3.8 Water-related risks (thermal regime, quantity, quality) (BD10, BD13-16)

These risks are inter-related with regard to their impacts on biodiversity and ecosystem services because changes in thermal regime and water flow can influence water quality. In addition, water quality can be impacted by changes in land use and pollution levels, which can also be affected indirectly by climate change. Detrimental impacts on aquatic habitats can have consequences for the ecosystem services they provide which include: the provision, purification, and regulation of water supplies; fisheries; amenity and recreational value. With regard to water quality, considerable research is currently in progress to establish costs and benefits associated with the provision of ‘good ecological status’ (GES) in water bodies as defined by the requirements of the EU Water Framework Directive. However, as GES is defined according to a reference level associated with the pristine condition of the water body, the impact of climate change is not yet explicitly factored into these assessments. Modification of river flows, water levels and thermal regime by a changing climate will shift the baseline reference condition of the water body regardless of human-influenced inputs to water quality, but these impacts remain uncertain and valuation data for climate change is not currently available.

Direct benefits of ecosystem services for water quality can be estimated by avoided treatment costs: for example, benefits with regard to avoided discolouration costs have been reported as £5million over 10 years from one catchment (Bateman et al., 2011a). Similarly, a contingent valuation study found that average WTP per household, in order to avoid one day of discolouration problems, was £5.40 (Bateman and Georgiou, 2010). However, these direct benefits are probably significantly exceeded by the non-market benefits.

Page 176: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

148 Biodiversity Sector Report

The UK NEA reviewed the available evidence and identified the water quality benefits of inland wetlands to be currently as high as £1.5 billion/yr (Bateman et al., 2011a). Furthermore, it has been estimated that the non-market benefits obtained by improving all water bodies to GES in England and Wales could provide additional benefits of 1.14 billion/yr (Bateman et al., 2011a). The average value of inland wetland ecosystems for providing improved water quality has been estimated at £436/ha/yr (Table 6.4), although this will vary depending on location. With regard to river systems, benefits obtained from improvements will depend on the difference relative to current quality, with average benefits varying from £15.6/km (improved from low to medium quality), £18.6/km (from medium to high quality) and £34.2/km (from low to high quality) (Bateman et al., 2011a). The UK NEA also concluded that the costs associated with changing agricultural land use in order to reduce nutrient loadings into rivers are substantially smaller than the benefits which consequent reductions in diffuse water pollution would bring66. Evidence to support this general finding is provided by detailed cost benefit analyses in specific catchments: for example, work in the Humber river basin (Fezzi et al., 2010; Bateman et al., 2011a) showed that, after controlling for other factors, significantly more visits are made to rivers with higher water quality, and that in the Aire catchment the benefits of improving water quality to GES were of the order of £12.5 million/yr compared with land use change costs of just over £5.5 million/yr.

The general public appear to consider water quality a priority issue, and if necessary express a willingness to pay for measures to sustain or enhance water quality. A recent report for both England and Wales inferred a WTP value for improvements to establish High Quality Status (equivalent with GES) by 2015 of £44.5/household per year and £167.9/household per year (Nera and Accent, 2007). These findings are broadly similar to those derived by GHK (2011a) in their analysis of the benefits of SSSIs which estimated a WTP for the water regulation benefits provided by ecosystems as £66/household/yr.

Table 6.4 Ecosystem service valuations for inland wetlands

Average additional value of service where present(£/ha/yr)*

Marginal value of service per new unit area

(£/ha/yr)

Flood control and storm buffering

608 407

Wild species diversity 454 304

Water quality 436 292

Water supply 2 1

Amenity and aesthetics 339 227

Source: Bateman et al. (2011a) Default value where no services are present is £303/ha/yr

Similar issues apply to assessments of water quantity and the requirements to maintain minimum ‘environmental flows’ under the WFD. Costs of regulating abstraction (e.g. for agriculture and industry) need to be balanced against the market and non-market services that functioning ecosystems provide. The UK NEA estimated that climate change-induced losses of water availability on ecosystem services could be in the range of £350-490 million/yr (Bateman et al., 2011a). The impacts of low flows occur not only through increased biological oxygen demand for fish and other important species, but also in terms of reduced dilution of pollution in rivers. WRC (1999) identified that recreational values provided by angling per person per fishing trip are often substantial, varying by the type of fish expected to be caught, and that the WTP

66

It did however also note that these costs would be concentrated on rural communities, while benefits are

distributed across all of society, both rural and urban.

Page 177: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 149

differs according to the water quality, which itself impacts on likely fish catch. Average WTP for improvements in water quality from ‘Poor’ to ‘Good’ were estimated to be as much as £10/person/trip.

At present, prices charged for abstraction (varying from £0.003 to £0.06/m3 for abstracted raw water, through to £1.50/m3 for metered, treated, potable water piped to households) reflect the cost of managing the licensing system rather than the true value of water which can encourages inefficient use. The marginal value for treated water ranges from £0.50/m3 to £1.20/m3, whilst the marginal value for irrigation water ranges between £0.23/m3 to over £1.5/m3 for irrigated potato and salad crops in eastern England (e.g. Moran and Dann 2008; Knox et al., 1999; Morris et al., 2004).

The thermal regime of the water body (including any stratification impacts) will also interact with changes in quality and quantity to modify ecosystem services. Measures to reduce this impact include extension of riparian woodlands to provide shading and to moderate temperature extremes. Although valuation data for this is not available and restoration/replacement costs can be high (see Table 6.2 for indicative costs), riparian woodland can also provide a range of co-benefits that extend to carbon sequestration, amenity value, and flood protection.

Wetlands also provide additional benefits through their role in moderating high flows in rivers. The average additional value of flood risk reduction from wetlands is estimated at £608 per hectare per year (Table 6.4), although in some locations with vulnerable populations this value will be considerably higher.

6.3.9 Marine biodiversity

A baseline economic assessment of ecosystem services from UK marine biodiversity has been provided by Beaumont et al. (2006, 2008; Table 6.5), although it was only possible to provide explicit monetisation for 8 of the 13 services. This previous work gives an indication of the large range of direct and indirect benefits obtained from the marine environment.

Page 178: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

150

Table 6.5 Ecosystem services provided by marine biodiversity

Source: Beaumont et al., 2006

Page 179: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 151

Harmful algal blooms

The costs of algal blooms have been valued in a number of studies using either damages to fisheries or contingent valuation and choice experiment studies (see Pinnegar et al. 2012 for a more detailed discussion). Based on data up to 1998, the ECOHARM (2001) study estimated economic impacts from algal blooms in the EU not just on commercial production (e.g. aquaculture) but also by using contingent valuation to assess wider impacts, including for recreational users and tourism through negative consequences for the environment. Values of between €215 and €1,524 million were elicited for the whole of Europe to provide a baseline estimate without the additional impact of climate change. The UK is likely to represent a significant proportion of this total, based on population and available coastline. In the absence of more detailed data on the impacts of climate change, Pinnegar et al. (2012) inferred potential economic costs within a Low magnitude band (i.e. £1 to £9 million annually ) but with low confidence due to the high uncertainty in changes to the baseline.

Ocean Acidification

The impacts of changes in ocean chemistry remain highly uncertain due to the complexity of ecosystem processes and this has implications for any monetary assessment. The implications for calcification processes in marine species have been predominantly considered through those with direct commercial value, notably shellfish and crustaceans, rather than for wider biodiversity and ecosystem services. Four of the ten most valuable marine species in the UK are calcifying shellfish or crustaceans, with Nephrops (scampi) accounting for a quarter of the total value of fish landed in the UK in 2007. In addition, aquaculture of shellfish (mainly oysters and mussels) is annually worth £20 million in England and Wales (2006), £5 million in Scotland (2007) and more than £3.5 million in Northern Ireland.

However, to-date there is limited data and interpretation of potential consequences such as reduced growth, impaired reproductive output or increased mortality, even in commercial species which would be reflected in market prices due to reduced fisheries yields and productivity. A preliminary assessment of annual costs for the 2020s, 2050s and 2080s for ocean acidification impacts on commercial species has been developed by Cefas (Pinnegar et al., 2012; Table 6.6). The wider costs for marine biodiversity remain rather speculative at present but are likely to be at least an order of magnitude above these.

Table 6.6 Mid-point estimates of the direct costs of ocean acidification on commercial species (£million per annum) (Pinnegar et al., 2012)

Period Shellfish Mollusc Aquaculture Total

2020 4.4 1.2 0.4 6.0

2050 23.5 6.2 2.0 31.7

2080 50.6 13.3 4.4 68.3

Invasive Non-native species

Climate change is likely to increase the prevalence of marine invasive non-native species although the overall risk will also be contingent on socioeconomic factors. Pinnegar et al. (2012) evaluate the risk from 9 non-native species but noted that economic data are rather limited, and this is particularly apparent for biodiversity and ecosystem services. As an exemplar, eradication of the Carpet Sea Squirt from those UK marinas where it is present was estimated at £2.4 million, and if this species were to continue to spread over the whole of the UK, then the overall cost of eradication could rise to £72million (CABI 2010). Based upon this limited data, Pinnegar et al.,

Page 180: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

152 Biodiversity Sector Report

(2012) provide an indicative economic cost of Medium magnitude (i.e. £10 to £99 million/year).

Changes in species distributions and range

Analysis by Pinnegar et al. (2012) suggests that by 2100 climate change may mean that seabird species such as the great skua and arctic skua may no longer breed in the UK. Furthermore, the range of black guillemot, common gull and Arctic tern may all shrink significantly to the extent that breeding colonies would only persist in Shetland and the most northerly tips of mainland Scotland; currently, they also exist further south in the UK.

Potential shifts in the distribution of major seabird colonies could have important economic consequences in terms of ecotourism which provides large visitor numbers at favoured sites. A summary report by the RSPB (2010) gives an indication of typical local incomes directly attributable to seabirds from four reserve locations for 2009: Bempton Cliffs, over £750,000; South Stack, £223,000; Mull of Galloway, £115,000; Rathlin Island Reserve, over £115,000. Beyond these locations there are many other sites, maintained by the RSPB and other organisations in the UK, that have high visitor numbers because of the quality of experience provided by marine biodiversity. Changes in species distributions, due either to ecosystem disruption, or range shifts or modifications to migration patterns, are therefore tentatively suggested to have Low to Medium economic consequences.

Diseases and pathogens

Quantitative data is not currently available for this risk from a biodiversity perspective. Pinnegar et al. (2012) assessed the risk of marine diseases to human health, particularly through the consequences of marine Vibrio outbreaks, which although currently having a low probability, when they do occur they can have serious consequences. They provide a qualitative assessment of economic costs that increases from Low in the 2020s (i.e. between £1 – 9 million/year) rising to a potential Medium risk in the 2050s and beyond (i.e. between £10 and 99 million/year). In the absence of further information, consequences for marine biodiversity are assessed within this same monetary range.

Page 181: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 153

7 Adaptive capacity

7.1 Overview

Adaptive capacity considers the ability of a system to design or implement effective adaptation strategies to adjust to information about potential climate change, to moderate potential damages, to take advantage of opportunities, or to cope with the consequences (Ballard, 2009, after IPCC, 2007). This can be considered as having two components; the inherent biological and ecological adaptive capacity of ecosystems and the socio-economic factors determining the ability to implement planned adaptation measures (Lindner et al, 2010). Considering adaptive capacity is essential for adaptation planning and the CCRA project has included work in this area that will contribute to the ongoing Economics of Climate Resilience study and the National Adaptation Programme. The CCRA work on adaptive capacity focuses on structural and organisational adaptive capacity and this chapter provides an overview of the assessment approach. The subsequent sections of this chapter provide an overview of the findings from other work on adaptive capacity in the biodiversity sector that has been carried out.

The climate change risks for any sector can only be fully understood by taking into account that sector’s level of adaptive capacity. Climate change risks can be reduced or worsened depending on how well we recognise and prepare for them. The consequences of climate change are not limited to its direct impacts. Social and physical infrastructure, the backdrop against which climate change occurs, must also be considered. If such infrastructure is maladapted, the economic, social or environmental cost of climate impacts may be much greater; other consequences could also be considerably more detrimental than they otherwise might have been. Avoiding maladaptation is one outcome of high adaptive capacity; high adaptive capacity lowers the negative consequence of climate impacts. Conversely, low adaptive capacity increases the negative consequences.

7.2 Assessing structural and organisational adaptive capacity

The methods used for assessing structural and organisational adaptive capacity in the CCRA are based on the application and further development of the PACT framework67. The work included a preliminary literature- and expert interview-based assessment of all eleven sectors in the CCRA. This was followed by more detailed analysis for the following sectors:

Business, Industry and Services (focusing on the finance sector)

Transport (focusing on road and rail)

Built Environment (focusing on house building)

Health

Biodiversity and Ecosystem Services

Water

67

PACT was developed in the UK as one of the outcomes of the ESPACE Project (European Spatial Planning: Adapting to Climate Events) http://www.pact.co/homehttp://www.pact.co/home.

Page 182: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

154 Biodiversity Sector Report

Structural adaptive capacity

The extent to which a system is free of structural barriers to change that makes it hard to devise and implement effective adaptation strategies to prepare for future impacts. This covers issues such as:

Decision timescales: This considers the lifetimes of decisions, from their conception to the point when their effects are no longer felt. The longer this period is, the greater the uncertainty as to the effects of climate change impacts. Cost-effective adaptation becomes harder. Potential climate impacts also become more extreme over longer timescales. This means that a greater scale of adaptation may need to be considered, and that the barriers to adaptation resulting from 'lock-in' to maladapted processes become more pronounced (Stafford-Smith et al., 2011). Adaptive capacity is therefore lower, and maladaptation more likely, when long-lasting decisions are taken.

Activity levels: This considers what opportunities are there for adaptation, and on what scale. The frequency with which assets are replaced or created determines how many opportunities there will be to take action which increases adaptive capacity.68 In addition, when a lot of asset replacement and/or new investment is expected, there will be more chances to learn from experience, which increases adaptive capacity.

Maladaptation: This evaluates the effect of decisions already made on adaptive capacity. Long-term previous decisions which have reduced adaptive capacity are often difficult or expensive to reverse. Such decisions were made either before climate change was recognised as an issue, or more recently as a result of poor organisational capacity. Such maladaptation makes implementing effective strategies much harder.

Sector (or industry) complexity: This refers to the level of interaction between stakeholders within an industry, or with outside industries and groups, that is required to facilitate effective decision-making. Complexity is higher (and adaptive capacity lower) when many stakeholders are involved in decision-making and when their agendas (e.g. their financial interests) differ substantially.

Organisational adaptive capacity

Organisational adaptive capacity is the extent to which human capacity has developed to enable organisations to devise and implement effective adaptation strategies. The framework used to assess this recognises different levels of adaptation, from entry level (‘Engaging’) to advanced levels (‘Pioneering’ and ‘Leading’), all of which may be needed for effective adaptation. Effective adaptation requires decision-making that takes account of an uncertain future and avoids locking-out future options that might be more cost-effective if climate impacts become more severe, or arrive more rapidly, than expected. The PACT framework used to assess this recognises different levels of adaptation. This framework is arranged in a hierarchy of ‘Response Levels’ (‘RLs’), as set out below, of increasing capacity69. These levels do not supersede one another; instead, each one builds on the experiences and practices built up in the previous response level. Organisations may need to be active on all levels for an effective

68

This differs from ‘Decision timescales’ because investment in a sector is not continuous but varies over time, with periods of high investment being followed by periods of little or no investment. 69

The PACT framework contains six response levels: those cited are the most relevant to the adaptation field.

Page 183: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 155

adaptation programme. An RL4 organisation focused on breakthrough projects still needs to be stakeholder-responsive, for example.

RL1: Core Business Focused: At this level, organisations see no benefit from adapting; if change is required of them, it should both be very straightforward to implement and also incentivised, e.g. through ‘carrots’ and ‘sticks’.

RL2: Stakeholder Responsive: At early stages of adaptation, organisations lack basic skills, information, processes and also skilled people; they need very clear advice and information plus regulations that are straightforward enough to help them get started.

RL3: Efficient Management: As organisations begin to professionalise adaptation, they become more self-directing, able to handle short term impacts up to 10 years (Stafford-Smith et al., 2011). They need professional networks, best practice guidelines, management standards, etc.

RL4: Breakthrough projects: When impacts beyond 10 years need to be considered, organisations may need to consider more radical adaptation options. As well as high quality support from scientists, they may need support with the costs of innovation.

RL5: Strategic Resilience: Adapting a whole region or industry for long-term climate impacts of 30 years or more requires lead organisations to develop very advanced capacity that is able to co-ordinate and support action by a wide range of actors over programmes that are likely to last for many years.

7.3 Natural adaptive capacity

Within the ongoing research on adaptive capacity for the CCRA, the work has focused on the organisational and structural capabilities of human systems. However, for the Biodiversity and Ecosystem Services sector, this is an incomplete perspective. The term ‘adaptive capacity’ is also used to describe the capacity of an ecosystem to adapt to changes in its external environment. Such capacity is related to factors such as diversity within species, biodiversity across species, ecological niches and habitat availability, and ultimately to the structure and functioning of an ecosystem. With regard to the influence of a changing climate at species level, vulnerability or resilience of a species is strongly related to key traits, particular its degree of specialized requirements, genetic variability, reproduction rate, dispersal ability, and physiological or behavioural adaptability (Table 7.1).

Adaptation to change is a fundamental feature of species, characterised by their specific traits and genetic properties, and their integration with the wider ecological community. This can allow ecosystems to cope with, modify and buffer changing environmental conditions whilst still providing the same range of ecosystem functions. This inherent capacity to adapt to change also needs to be recognised as a key component of resilience. In many cases, natural adaptive capacity acts to buffer the worst excesses of change, whilst its limits often define the limits for human adaptive capacity unless we contemplate establishing artificial life-support systems.

Page 184: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

156 Biodiversity Sector Report

Table 7.1 Physiological and life history traits that may make a species more or less vulnerable or resilient to climate-related disturbances

Adapted from Steffen et al. (2009)

7.4 Interactions between natural and organisational adaptive capacity

It should therefore follow that maximum adaptive capacity will be realised when organisational adaptive capacity acts to enhance the natural adaptive capacity already present within an ecosystem. At the very least, adaptation should seek not to further erode natural adaptive capacity by requiring that it occurs within small isolated blocks detached from the remainder of the landscape. As identified by the Lawton report in England (Lawton et al., 2010), the current ecological network is not considered to be adequate to provide robust resilience against a changing climate. Therefore, the key requirement is to ‘make space for nature’ in order to facilitate the adaptation process, whether in the coastal zone, agricultural lowlands, or uplands. An increase in the number, size and connectivity of protected sites within the ecological network would be a key asset in building natural adaptive capacity to buffer future change.

A further interaction is identified through the concept of ‘adaptive management’. Ecosystems are typically complex, multi-scale, dynamic phenomena. Our knowledge of their functions and services is incomplete and guided by data from the past which is only likely to provide a partial guide to their future responses. Hence, the future has an element of inherent uncertainty and is likely to involve unexpected surprises. Through adaptive management and a more systematic observation and collation of change on the ground it should be possible to better contextualise change and learn from it. This requires a better use of organisational resources to share this knowledge and use it to maximise the potential of natural adaptive capacity, and to integrate its benefits into other sectors. It also requires a more flexible approach that is regularly reviewed, rather than fixed conservation objectives that may prove unviable in the long term.

Species at least risk Species at most risk

Physiological tolerance to a broad range

of factors such as temperatures, drought

and flooding

High degree of phenotypic plasticity

(ability to change observable

characteristic or trait)

High degree of genetic variability

Short generation time (i.e. life cycle) and

short time to sexual maturity

High fecundity (reproductive ability)

‘Generalist’ requirements for food, nesting

sites, etc

Good dispersal capability

Broad geographic range

Narrow range of physiological tolerance to

factors such as temperature, drought and

flooding

Low genetic variability

Long generation times and long time to

sexual maturity

Specialised requirements for other

species (e.g. for a disperser, prey species

or pollinator) or for a particular habitat that

may itself be restricted (e.g. a particular

soil type)

Poor dispersers

Narrow geographic ranges

Page 185: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 157

7.5 Enhancing adaptive capacity

The England Biodiversity Strategy identified a series of high-level principles to facilitate the adaptation process (Figure 7.1). These emphasise the need to strategically plan and implement cross-sectoral actions to enhance ecological resilience and accommodate change, based upon best available knowledge.

The EMBEDS project has considered criteria to measure progress for these principles, together with an understanding of barriers to implementation and guidance on overcoming these obstacles (Berry et al., 2011). The Biodiversity sector generally has a very high awareness of climate change impacts, combined with a strong perception that these risks may be severe and escalating.

However, a series of barriers acts against implementation of the principles. These include current organisational structures across government and the many agencies that have interests in or requirements to deliver objectives that are wholly or partially related to biodiversity. This structural issue is often particularly apparent from national to regional to local scale, and can hinder opportunities to share information. Related to this is limits on human resources to scope and implement adaptation actions, together with an often more pressing requirement to deal with short-term requirements. If commitments have already been made, this can mean a reluctance to revisit them, or a reluctance to change habitual procedures, particularly where they have been established through cross-agency or cross-sectoral agreement. In other cases, revisiting adaptation requirements could mean potential conflicts with other pressing objectives (e.g. meeting climate mitigation targets).

Figure 7.1 Key principles for biodiversity to adapt to climate change Source: Smithers et al. (2008)

Page 186: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

158 Biodiversity Sector Report

EMBEDS also identified that the legislative contexts, within which all biodiversity partners must function, suggest that the adaptation principle ‘accommodate change’ may be especially challenging to implement in practice. This legislative legacy may be responsible for difficulties in coordinating action at the landscape scale, which is particularly necessary to ensure the best opportunities for priority species and habitats. A related issue which also tends to encourage a ‘wait and see’ approach is uncertainties on the timing and severity of impacts of climate change and hence differences in perceived priorities. There is often reluctance therefore to change established practices, many of which are based upon the traditional conservation objectives of restoration of species and habitats from the historic past at a particular location.

Reliance solely on past conservation objectives is particularly problematic in terms of the degree of future change expected (climate and non-climate), which will not allow a return to the past, and also the long time frames required in developing adaptation measures. Some habitats, such as certain types of woodland or blanket bog, require a long period of time to become established; therefore action is required now to prepare the ground for the future. In addition, habitat creation is not an exact science. As managed coastal realignment schemes have demonstrated, the distribution and extent of new habitat, such as saltmarsh or mudflat, is dependent on many factors (known and unknown). Therefore time needs to be built into these schemes in order that they can adjust and adapt to changing circumstances.

In summary, the actions required to enhance adaptive capacity include mechanisms to enable better integration and information sharing across sectors and from national to local level. This includes schemes to disseminate best practice, examples of successful adaptation, and exemplars of the benefits of collective action. Legislative barriers need to be challenged to ensure that adaptation is kept as a ‘live’ issue and that coordinated guidance can be implemented at different levels of governance. The need for coordination and integration to enhance adaptive capacity is also manifest in terms of strategies to establish climate-resilience at local level and a resilient site network at national level. A series of projects have recently investigated the ecological resilience of the site network (see Section 8.3) but further work to assess this in conjunction with social and organisational indicators of adaptive capacity would allow a more robust evaluation of biodiversity vulnerability in the UK.

Page 187: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 159

8 Discussion: implications for adaptation

In this chapter we discuss the range of impacts identified in the context of both climate change and socio-economic influences, and how the risks may be moderated by adaptive capacity. Key measures of adaptive capacity will be provided by the ability to implement the principles discussed in Chapter 7 (see Figure 7.1), namely to enhance ecological resilience, accommodate change, and develop cross-sectoral approaches, to enable action now based upon best-available knowledge. Risks are summarised according to specific categories related to the interface of climate change adaptation with other policy frameworks. This is consistent with the findings of Chapter 5 that governance and decision making has a key role in influencing successful outcomes for the Biodiversity and Ecosystem Services sector by moderating demands for natural resources and encouraging sustainable land use that can accommodate both climate and non-climate changes.

8.1 Risks

8.1.1 Soils (including risks BD1 and BD8)

The maintenance of soil quality, including processes of nutrient cycling and soil formation, is a fundamental ecosystem service that links biological activity with physical and chemical processes. Changes in soil moisture (BD1, Section 4.2) or soil organic carbon (BD8, Section 4.7) could significantly modify these interactions. Most soils usually naturally have a high level of resilience but quality has been in decline, mainly due to land use change. This undermines its ability to accommodate change. Our knowledge of key processes is also rather limited although the importance of key constituents such as organic matter is well-known.

Soils policy is cross-sectoral and defined by strategic frameworks at UK and DA level, which clearly identify the threats from climate change. Adaptation actions have also been identified and trialled at pilot sites, but have yet to be implemented on a large scale because they require cooperation with land use sectors (AGRICULTURE, FORESTRY). They include schemes to help maintain moisture levels in drier summers, such as temporary blocking of field drains in agricultural lowlands or permanent blocking of drainage gullies in the uplands. Complex interactions exist with carbon storage, CO2 emissions and other greenhouse gas emissions which have very important implications for climate mitigation measures. The reduction of dissolved organic carbon and discolouration of water supplies is also an important issue for the WATER sector. Although a full economic assessment is currently unavailable, the prospect of multiple cross-sectoral benefits through improvements in soil quality implies that schemes will be cost-effective in many situations. Future reform of the CAP may be an important socioeconomic and policy driver for land use change.

8.1.2 Coastal zone (including risks BD2 and BD7)

Coastal habitats host a rich and diverse flora and fauna, in addition to their importance for overwintering migrant birds. The coastal zone is naturally dynamic and will adjust its profile in response to sea-level rise and wave energy. In many locations, however, the

Page 188: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

160 Biodiversity Sector Report

presence of fixed coastal defences severely restricts this natural adaptive capacity and has left a landscape that requires managed interventions. In some cases, terrestrial habitats are protected by defences whilst on the seaward side habitats are threatened by ‘coastal squeeze’. Defences have also reduced the supply of sediment which is increasing erosion rates at vulnerable locations.

Our knowledge of coastal processes is improving and the development of Integrated Coastal Zone Management, Shoreline Management Plans (SMPs) and Coastal Habitat Management Plans recognises the need for cross-sectoral approaches over long time scales. Development of river basin management plans under the requirements of the WFD could also be instrumental in developing resilience of coastal freshwater sites. In England, the Floods and Water Management Act 2010 provides for the use of natural processes to manage flood risk and to undertake environmental enhancements to realise the benefits from natural processes. The key issue for adaptation is often the degree of ‘lock in’ to the current approach (as agreed across different sectors) which usually favours a ‘hold the line’ policy rather than managed realignment which would restore a more natural coastal profile.

The OMREG (Online Managed Realignment Guide70) database currently identifies 51 UK projects, carried out between 1991 and 2009, that have created new habitat or compensatory habitat to offset losses elsewhere (Figure 8.1). These include exemplar schemes at Wallasea Island, Essex; Snape, Suffolk; Frampton Marshes, Lincolnshire; and Nigg Bay, Scotland (see Townend et al., 2010 for further background). Non-governmental organisations, such as the Wildlife Trusts and the RSPB, are also working with the statutory agencies to create habitat that will potentially compensate for any losses at the coast. The net effect of these schemes is to reduce the risk of coastal inundation in some areas, where managed realignment has taken place, and/or to have a ‘habitat bank’ in readiness to compensate for future losses.

Figure 8.1 Cumulative habitat area created in the UK for all completed managed realignment schemes (orange bars) and net habitat area created (blue

bars – accounting for habitat compensation schemes) Source: OMREG http://www.abpmer.net/omreg/

70

Managed by the consultancy ABP Marine Environmental Research

Page 189: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 161

Despite these schemes, even current levels of change imply considerable loss of coastal habitats, regardless of the additional impacts from accelerated rates of sea level rise. This has major implications for the UK’s obligations to meet the requirements of the Habitats Directive and to support internationally-important populations of over-wintering birds. There are also major socioeconomic implications: the UK NEA estimated that saltmarsh habitats provide £2.17billion in benefits from reducing flood risk and that coastal wetlands provide £1.5billion in benefits by buffering against storms.

A natural adaptation response of coastal habitats is the possibility for one habitat type (e.g. saltmarsh) to succeed another (e.g. grazing marsh), either as a result of coastal inundation events and/or coastal reconfiguration. However, many protected areas have been designated with static site boundaries and some conservation targets, such as those linked with the BAP, can act against such changes. As identified by the Lawton report (Lawton et al., 2010 p.79), “climate change will test the current management regimes of many of the sites within the [ecological] network, and may require new approaches to setting conservation objectives and new habitat management techniques and standards”. To deliver successful long-term outcomes, habitat replacement schemes will therefore need to be matched to the right locations taking appropriate account of the underpinning soil, geology, hydrology and topography. In dynamic environments, such as the coastal zone or floodplains, this may require additional flexibility to maximise the structure and function of the suite of habitats most suited to a given geomorphological environment.

8.1.3 Pests, diseases and invasive non-native species, (including BD3 and BD4)

These risks are currently dominated by socio-economic factors that encourage their introduction, but climate change is modifying the background factors that favour their persistence and spread. Resistance amongst native communities is often hindered due to other stressors they are experiencing (e.g. eutrophication). The risk is particularly pronounced in wetlands and aquatic ecosystems. Existing knowledge suggests a more biodiverse landscape increases the resilience to invasive non-native species (i.e. as an ecosystem service) but more work is required on this issue.

Knowledge of generic risks is generally good and risk assessment has become a standard procedure, however the risk from each species requires a specific case assessment which may be time-consuming until sufficient knowledge is collated. Effective biosecurity measures require coordination and regulatory controls at both national and international level. Currently agencies at international, EU, UK and devolved administration level provide information and risk assessments on invasive non-native species and diseases. The CCRA identifies an increasing need for climate change to be factored into these risk assessments, and for more work to understand key controls on their spread. The cross-sectoral benefits of action to control these risks (including for AGRICULTURE, FORESTRY, WATER and HEALTH) indicate that they are also supported by economic and social criteria.

8.1.4 Species movements, migration patterns, and interactions (including BD5, BD9, BD11)

Many species naturally adapt to climate change by moving to reach more suitable ‘climate space’. These natural responses are severely restricted by habitat loss and fragmentation which makes dispersal through the landscape much more difficult, and that suitable climate space may not necessarily be accompanied by suitable habitat. These restrictions tend to favour species that are more generalist in their habitat

Page 190: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

162 Biodiversity Sector Report

preferences. In addition, species under stress with declining populations have difficulties in dispersing. Similarly, evidence suggests migratory patterns are changing with favoured sites shifting so that some species move to new sites or their abundance changes at a specific site; in some cases (e.g. chiffchaff), the species has ceased its migratory behaviour due to warmer winters.

The protected site network provides the capacity for species to adapt but as identified by the Lawton Review (Lawton et al., 2010) this ecological network is generally not large enough or connected to accommodate change. The 2011 England Biodiversity Strategy has identified a target of 200,000ha of new priority habitat but in some locations the quality of habitat also requires to be improved (e.g. through Nature Improvement Areas). Adaptation measures are also required that promote permeability to species movement in the wider landscape. In this context, there are clearly very important interactions with AGRICULTURE and FORESTRY in order to develop integrated strategies that can provide habitat diversity, together with strategic schemes such as buffer zones, corridors or stepping stones. At some sites, qualifying features for designation may require more regular review in order to accommodate change, particularly for migratory species. In principle, the EU Habitats Directive should be flexible enough to handle this change but further obstacles may exist within the planning system and through requirements for cross-sectoral co-operation. For all these reasons, adaptation action to-date has been rather limited.

Loss of ecological functioning in the landscape due to species loss can have implications for ecosystem services such as water purification (8.1.5), flood alleviation, soil quality (8.1.1) and air quality. Current knowledge suggests that a further consequence of habitat loss and fragmentation is negative impacts on pollination, which is now further compounded by climate change. Further evidence is required but there are important economic implications for AGRICULTURE if this ecosystem service is further disrupted. As with soil quality (Section 8.1.1), CAP reform is a key policy driver that could facilitate a more proactive approach to cross-sectoral adaptation linking landuse with biodiversity and wider ecosystem benefits.

8.1.5 Water (including BD10, BD13, BD14-16)

Climate-related impacts on water temperature and quantity will interact with existing pressures, particularly the impact of pollution on water quality and over-abstraction. The combination of these pressures at present suggests that there is very limited capacity to accommodate future change at vulnerable locations. Aquatic ecosystems provide important ecosystem services such as water purification and fisheries that could also be at risk.

Positive action is now being focussed on these risks at catchment level through the implementation of the Water Framework Directive, although climate change adaptation remains to be fully included within its objectives. Knowledge of the interaction of climate change with water quality remains limited and this hinders targeted adaptation measures. Cross-sectoral measures are being implemented at a large scale with the WATER sector but have barriers to implementation at the farm scale within AGRICULTURE. The interface between the WFD and Habitats Directive, especially with regard to climate change, remains to be fully established. In England, the 2011 Biodiversity Strategy has developed a vision for wetlands which can provide a specific focus for habitat creation targets and nature improvement areas in terms of both biodiversity and wider water resource benefits.

Page 191: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 163

8.1.6 Wildfire (BD12)

Climate change implies an increased risk from wildfire although at present human agency remains the main risk factor. Some vulnerable species and habitats, notably on lowland heath, have limited capacity to tolerate an increased frequency of wildfire. Knowledge of the current level of risk is improving due to new reporting systems. Control of the fire risk may require a greater role for intervention during extreme conditions, for example through public access exclusion orders, closer regulation of burning on moorland and agricultural land, or education and awareness schemes. The ecological, social and economic implications of habitat loss due to fire in the UK remain to be fully established.

8.1.7 Renewable energy and climate mitigation schemes (including BD6)

Large-scale expansion in renewable energy schemes is already a reality and expected to increase in the future due to climate mitigation targets and ‘energy security’ concerns. In the absence of strategic planning these schemes could further hinder the resilience of the natural environment and its ability to accommodate change. Conversely, they could enhance this adaptive capacity with good planning that matches schemes well with their local context. Environmental Impact Assessment is currently a key component of the planning process but this does not include future changes in the medium to long term, sometimes due to uncertainty. Procedures therefore need to be further refined to accommodate future change using risk-based measures, including the viability of the scheme within a future climate. Implementation of the sustainability criteria defined by the EU Renewable Energy Directive may also be expected to enable a longer-term perspective. For bioenergy, a cross-Whitehall study is expected in 2012 that identifies the scale and sources of biofuel and biomass planned for the UK, including a sustainability assessment.

8.1.8 Marine biodiversity

A range of risks to marine biodiversity have been summarised in Chapter 3 and described in more detail in the MARINE CCRA sector report (Pinnegar et al., 2012). Resilience within the marine environment has often been reduced due to pressures such as over-fishing and pollution, therefore reducing these stresses will allow improved ability to accommodate change. The complexity and scale of marine processes mean that our knowledge is limited. A cross-sectoral approach is promoted and encouraged by the EU Marine Framework Directive which should facilitate a more systematic assessment of risks, but the need for a coordinated international approach makes strategic adaptation planning difficult. Therefore adaptation actions have remained rather limited to-date.

8.1.9 Cross-sectoral risks

All of the risks show important interactions between climate and socio-economic drivers, with land use change as the strongest interaction as evidenced by current patterns of habitat fragmentation and species/habitat loss. Other crucial drivers include population demands, societal values, technological development, and the level of government decision-making. A healthy natural environment requires that demands on natural resources are sustainable and that we recognise its multiple benefits within coordinated cross-scale decision making (local to national and international). All risks

Page 192: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

164 Biodiversity Sector Report

also show key interactions to other sectors, either directly, or through the maintenance of ecosystem services that support human well-being, notably those that support food and energy production, regulation of environmental hazards such as flooding, and cultural identity. There are also potentially highly-significant feedbacks to the climate system through changes in carbon stocks in soils and vegetation, and hence from GHG emissions or sequestration. This array of inter-relationships emphasizes the importance of a cross-sectoral approach to biodiversity and ecosystem services. The role of CAP reform and its implementation through targeted schemes is likely to be a significant policy driver for delivering integrated land use management for multiple benefits.

8.2 Costs and benefits

Valuation techniques can provide costs and benefits for both the use and non-use (existence) values of biodiversity and for a range of ecosystem services. However, these techniques are sometimes contested and the subject of considerable debate. For example, techniques often exclude shared-value cultural benefits (e.g. spiritual benefits) which mean that they provide only a minimum assessment of true value. Furthermore, some commentators have suggested that stated preference methods, even those derived from group-based elicitation, downplay an understanding of the role of biodiversity in ecosystem functioning. Valuation of the climate change costs and benefits with regard to biodiversity and ecosystem services is therefore complex and at an early stage of development: current values should therefore be seen as indicative and certainly not definitive. Nevertheless, the UK NEA provided willingness to pay (WTP) estimates of the non-use value of terrestrial biodiversity ranging from £540-£1,262 million per annum and for marine biodiversity, estimates of around £1,700 million per annum.

The UK NEA also highlighted the following benefits from the natural environment:

The total value of net carbon sequestered currently by woodlands is estimated at £680 million/year.

Pollination services for crop production are estimated at £430million/ year.

Water quality benefits of inland wetlands may be as high as £1,500 million/year with planned river quality improvements generating values up to £1,100 million/year. This may be negated by climate change-induced losses of water availability that are valued at £350 million to £490 million/year.

Costs associated with changing agricultural land use to reduce pollution to water courses are substantially smaller than the benefits which consequent reductions in diffuse water pollution would bring.

The amenity value of all wetland types (including coastal) is ca. £1.3 billion /year.

Economic valuation in the NEA also suggests that a modestly-sized nature and recreation site can generate values of between £1,000 and £65,000 p.a., depending purely upon location. Similarly, the international TEEB study (2010) showed that protected natural areas can deliver economic returns that are 100 times greater than the cost of their protection and maintenance.

It is highly likely that there will be significant further developments in valuation of ecosystem services in the next round of the CCRA. Policy commitments (e.g. the England Biodiversity Strategy; Environment Strategy for Wales) identify a need to

Page 193: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 165

better reflect the full benefits from the natural environment within decision-making processes, but this will be reliant on the application of improved techniques to elicit individual and collective ‘values’ for non-market public goods.

8.3 Risk in wider context

The CCRA analysis should be seen in the context of a series of other initiatives and recent studies that collectively define the evidence base for adaptation.

A review of the scientific evidence for the impacts of climate change on biodiversity was provided by Mitchell et al., (2007) for the England Biodiversity Strategy. The review summarised and qualitatively ranked the impacts based upon expert opinion, including both direct and indirect impacts. An assessment of the quality of the evidence was also provided. Although only available for England, this report has provided a significant source of evidence for the CCRA sector report. No equivalent systematic assessment of the impacts at UK level has been developed since a review by Hossell et al. (2000).

With regard to both monitoring of change and analysis of current trends, the Countryside Survey programme and the Environmental Change Network (ECN) are important initiatives. The Countryside Survey periodically analyses vegetation changes at UK level: the most recent survey (2007) interpreted most change to be the result of land use change and pollution (particularly eutrophication from atmospheric deposition in upland areas), rather than climate change, although there are significant interactions. The ECN provides a network of sites with detailed long-term monitoring and has detected trends in flora, fauna, water quality, soil quality etc; however, it is often difficult to attribute these trends because climate change has occurred concurrent with other drivers of change, including land use and pollution.

The BICCO-Net project assessed the impact of climate change on UK biodiversity by analysing long-term monitoring data on eight terrestrial taxa. The preliminary results show complex responses across different species populations. The project included a broad range of research on the influence of climate on birds and the findings highlight that changes in climate can bring about significant constraints on bird populations as well as benefits for some selected species. Populations of some beetles, butterflies and moths seem to benefit from warmer summers but decline after warmer winters. Different species have varying sensitivities to weather patterns, but the impacts of temperature changes can be distinguished whilst precipitation changes are often more complex.

A series of projects have assessed the resilience of the protected site network (Defra, 2011b). Climate change modelling of priority habitats at protected sites has indicated a wide range of changes in vegetation and community composition can be expected. The modelling in some cases shows important differences compared to the expert review of Mitchell et al., (2007), indicating discrepancies worthy of further investigation. Priority habitats in the mid- to north of England are predicted to change the least whilst those in SE England, Northern Isles and eastern Scottish Highlands are predicted to change the most.

The CHAINSPAN project employed climate change modelling of common and widespread bird species to assess the resilience of the UK’s network of

Page 194: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

166 Biodiversity Sector Report

Special Protected Areas (SPAs) in the context of management and potential impacts from changes in habitat. Of the species considered within the research, more were considered to be likely to benefit from climate change in the short (2020) to medium (2050) term, but with increasing severity of climate change in the long term, a greater proportion of species were projected to decline in abundance. Populations of northern breeding species, particularly seabirds, were projected to be most vulnerable to future climate change. Milder winters were projected to benefit wintering populations of many waterbirds. Southerly distributed heathland species were also projected to benefit from climate change. The focus on common or widespread species means that the findings cannot yet be generalised to present an overall picture for UK avifauna.

The analysis in CHAINSPAN showed that large SPA sites were projected to continue supporting many birds into the future and therefore they are likely to remain key sites irrespective of climate change. Hence, large sites with good quality habitat will be more resilient to change and better able to accommodate colonisation by new species.This is particularly relevant for intertidal habitats, highlighting the importance of a coherent network of sites capable of hosting internationally important waterbird populations. For lowland heathland, grassland and arable habitats, modelling implies they will become increasingly suitable for current breeding SPA features associated with these habitats, if human disturbance can be minimised. The potential colonisation of sites in S. England by continental Annex I species is likely to be limited by a lack of colonists as they are often in decline in mainland Europe. For upland and montane habitats, reductions in the abundance of upland SPA species were projected with relatively small number of potential Annex I colonists, highlighting the importance of reducing existing pressures to maintain viable populations. In the marine environment, vulnerable seabird populations also require measures to reduce existing pressures including those on their food supply from commercial fishing, and potential disturbance from offshore renewable energy installations.

Defra project CR0422 investigated impacts of sea-level rise on selected terrestrial and wetland habitats, with results included within the analysis of Tier 2 risks BD2 and BD7 (Section 4.3). These results highlighted the level of risk from coastal flooding and erosion that is present now, with potential loss of important habitat in SPAs including coastal grazing marsh and saline lagoons.

The benefits of applying an ecosystem approach have been strongly emphasised by the UK National Ecosystem Assessment. Impacts on species, community composition, and fundamental ecosystem processes such as nutrient cycling, primary production, and the production/decomposition of soil organic matter, all collectively contribute to a healthy, balanced ecosystem. Changes to the biodiversity of the ecosystem, either directly or indirectly due to climate change, have the potential to change this balance, for better or worse. In addition, changes in temperature and moisture availability will impact upon the speed of biological, chemical and physical processes within the ecosystem. As we rely on the integrity of ecosystem functioning for many goods and services that humans obtain from the natural environment, this should be a major source of concern. The complexity of processes mean that our knowledge is limited but following the UK NEA, it is a major area of ongoing research, and further advances in knowledge should be expected in the next few years.

Page 195: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 167

Risk assessment approaches are relatively new to the biodiversity sector, which traditionally has assessed the vulnerability of species and habitats using approaches based upon sensitivity and exposure to climate change, combined with other stresses. However, at country level a series of initiatives have begun to implement systematic risk assessments, with information collated through the UK Inter-Agency Climate Change Forum. A vulnerability assessment of the site network in Wales is discussed in Appendix 4. Novel approaches to species-level risk assessment are also being developed and trialled (e.g. Thomas et al., 2010). These should provide a firmer basis for analysis during the next cycle of the CCRA.

8.4 Limitations of analysis and knowledge gaps

A challenge for the Biodiversity and Ecosystem Services sector has been to adjust the generic CCRA methodology to better represent the complexity of the natural environment, whilst maintaining the overarching structure. The sole use of response-functions as risk metrics was considered to be potentially misleading for the sector because multiple variables are involved (climate and non-climate), and because species, habitats and ecosystems have variable responses that are often non-linear through time. These responses are characterised by natural adaptation processes, in addition to the influence of other pressures and the role of management interventions. However, risk metrics can be useful indicators for the change in magnitude and frequency of risk that can be related to wider responses by interpretation of the broader evidence base.

Much of the climate change research conducted to date has, of necessity, been focused either upon individual species or on specific locations or habitat types. Development of systems-based approaches that can improve understanding of the multitude of interactions within the natural environment, and their links to the human environment, remain in the early stages. Continuous data records that are comparable across years or standardized across the UK are also limited, although some very good examples do exist. Use of UKCP09 data has been restricted until very recently by the lack of spatially-coherent data in the future projections; this remains a major challenge for the further development of probabilistic data in climate change risk assessments.

The basic knowledge gap is our understanding of change and ecosystem dynamics, including the interaction of people within ecosystems. This is essentially due to the complexity of responses and feedbacks involved, but also because this has often been a neglected topic in research. As a consequence, key functions and services provided by ecosystems have fundamental uncertainty in terms of how they will respond to change.

The following topics have been identified as important evidence gaps:

(a) Impacts i. More systematic UK-level collation and interpretation of site

monitoring and other available data (e.g. phenology), against inter-annual patterns of climate variability and trends of long-tem change.

ii. Upscaling of sensitivity data from freshwater ecosystems for regional and national-scale assessments to understand the interactions between water temperature, water quality and water quantity on priority habitats and species. Most information is currently only available at site level.

iii. Improved understanding of the combined impact of climate change and atmospheric pollution on ecosystems. This would facilitate further

Page 196: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

168 Biodiversity Sector Report

measures to limit pollution below critical levels to improve ecosystem resilience.

iv. Integrated modelling of species distributions and interactions, habitat shifts and landscape structure based upon a linked framework that combines bioclimate and ecological factors. This would provide a more robust evidence base than the current reliance on bioclimate envelope models to project future changes.

v. Vulnerability assessment of key locations and pathways for migratory routes (e.g. using space-for-time substitutes based upon current climate variability).

vi. Integration of ecological and geomorphological assessments for evaluation of habitat change in highly dynamic environments, particularly the coastal zone and river floodplains. This would also inform adaptation management based upon improved knowledge of the interaction of current biological and physical processes.

vii. Improved assessments of climate impacts on ecosystem functions. For example, this is particularly necessary for understanding the key role of soil biodiversity in changing ecosystem processes (such as nutrient cycling) and the implications for habitats and species biodiversity, soil organic matter and carbon storage, with implications for flood risk, water quality, soil erosion risk, habitat conservation and crop production capacities.

viii. Further work is required on ecosystem response to increased CO2 and associated feedbacks in conjunction with changing climate variables (e.g. temperature, soil moisture), based upon both modelling and experimental evidence. This has not been incorporated into most climate change impact assessments.

ix. Critical thresholds (‘tipping points) in the interactions between climate and ecosystem responses beyond which the system may undergo a major non-linear change or shift to a new ecological regime (e.g. coastal systems in response to a major storm surge event). Some recent advances have been made with regard to identifying key thresholds for animal population declines (e.g. Drake and Griffin, 2010).

(b) Adaptation i. Improved understanding of the implications of the rate of climate

change for natural adaptive responses in different ecosystems (including across different species), including the role of extreme events, and hence the limits to and thresholds for maintaining adaptive capacity.

ii. Evaluation of the effectiveness of landscape-scale initiatives, including measures to improve landscape permeability, in both short-term and long-term. This would act as a bridge between broad-scale modelling studies and site-specific monitoring/modelling to identify strategic maps of habitat creation opportunities that are robust against climate change.

iii. Evaluation of options for further development of protected areas. How will existing areas cope with change and how might they be further strengthened and better integrated into the wider landscape for maximum ecological benefits?

Page 197: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 169

iv. Detailed analysis of the scope for increasing the resilience of species and communities within their existing range, including measures to increase heterogeneity of habitat patches and increase population size of vulnerable species (e.g. based upon variation of microclimate with aspect and vegetation structure). This should include area issues: larger sites are generally considered less vulnerable to edge effects such as desiccation during droughts, but can general rules be derived for minimum site extent.

v. Exploration of biodiversity strategies that better integrate adaptation with opportunities from climate mitigation schemes. For example, the development of biomass energy (including short rotation coppice and short rotation forestry) could add to landscape and habitat diversity. Initiatives to enhance carbon storage in soils and biomass could also have significant benefits for biodiversity.

vi. Further information on the role of genetic diversity within species. This is mainly based on cultivated crops and domesticated animals at present. Although this is time-consuming and expensive to gather, it would help to identify and monitor genetic constraints on adaptation and may be used to evaluate the viability of translocation as an option for some threatened species.

vii. Improved valuation techniques for biodiversity (both qualitative and quantitative) that can incorporate dynamic system processes and interactions. Existing work has shown that results can vary across scales (relative to the level of generalization) therefore multi-level assessments would be particularly instructive in understanding trade-offs for decision-making. The prevailing approach to quantifying ecosystem services is still based on static analyses and single services, typically interactions between ecological and social systems.

viii. Better understanding of the role of culture and social capital (i.e. non-monetary benefits) in ecosystem-based management and the wider benefits of ecosystem services for human well-being.

ix. Development and implementation of a cross-sectoral approach to assess trade-offs in adaptive capacity between different sectors, with particular emphasis on reducing conflicts and enhancing complementarities between biodiversity and other sectors.

Page 198: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

170 Biodiversity Sector Report

9 Conclusion At a basic level, the CCRA has identified a range of climate-related risks to biodiversity that are highly likely to result in loss of important habitats and species. Most of the impacts are already becoming apparent in vulnerable locations. These are most pronounced in coastal habitats, some upland habitats, and in aquatic and wetland habitats. The direct risks are often compounded by indirect risks, notably through land use change and interactions with air and water pollution. These indirect risks also mean that many lowland habitats, including enclosed farmland, grassland and heathland, are also vulnerable to change and unable to take advantage of the opportunities for new biodiversity that a warming climate might bring. The increasing threat from invasive non-native species due to climate change provides additional risk for many habitats.

There also appear to be substantial risks of change in soil and ecosystem processes, but our knowledge of these remains rather limited because of the complexity of change. These issues should be of fundamental concern because biodiversity and healthy ecosystems provide many benefits to human welfare, either acknowledged or unacknowledged. Such benefits include water purification, regulation of flooding, pollination of crops, landscape amenity, food and fibre. The value of biodiversity is therefore not just provided by its intrinsic value but also increasingly recognised in terms of its economic and social value. Conventional economic assessments have previously failed to include these ‘non-market’ benefits and therefore considerably undervalued the assets that the UK holds in its natural environment. Following the National Ecosystem Assessment, and related initiatives to assess the status and value of ‘ecosystem services’, a more balanced approach is now being sought.

The importance of the network of designated sites in the UK is recognised. These protect against some of the other stresses imposed upon biodiversity, including land use intensification and pollution. They can therefore allow natural adaptation to occur and for ecosystems to adjust to the changing climate whilst maintaining their key functions and services. Unfortunately, as highlighted by the Lawton Review, sites are usually too small and the network too fragmented to be able to accommodate the large-scale ecological changes that are now inevitable due to climate change. Furthermore the distribution of sites will need to be modified to allow for species movement as a natural response to change. Natural adaptive capacity needs to be facilitated by human organisational capacity, especially as many habitats are currently maintained by human interventions.

The sector report has highlighted many uncertainties with regard to future change for Biodiversity and Ecosystem services. The complexity of issues has also meant that it has been a challenge to implement the generic CCRA methodology. However, there is compelling evidence that change is happening now and that without appropriate action the consequences will be overwhelmingly negative from an economic and social perspective, as well as due to environmental damage or failure to meet international obligations. Action therefore needs to incorporate these uncertainties and to be cross-sectoral in order that it can be transferred into the wider landscape. This will require adaptive management strategies that are flexible and resilient in the face of this uncertainty and which can be modified when new knowledge becomes available. Ultimately, if successful, these strategies will also act to buffer human society from the worst excesses of climate change.

Page 199: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 171

References

CCRA Sector Reports

Baglee, A., Haworth, A. and Anastasi, S. (2012) CCRA Risk Assessment for the Business, Industry and Services Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

McColl, L., Betts, R. and Angelini, T. (2012) CCRA Risk Assessment for the Energy Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Moffat, A.J., Morison, J.I.L., Nicoll, B., and Bain, V. (2012) CCRA Risk Assessment for the Forestry Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Brown, I., Ridder, B., Alumbaugh, P., Barnett, C., Brooks, A., Duffy, L., Webbon, C., Nash, E., Townend, I., Black, H. and Hough, R. (2012) CCRA Risk Assessment for the Biodiversity and Ecosystem Services Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Hames, D. and Vardoulakis, S. (2012) CCRA Risk Assessment for the Health Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Knox, J.W., Hurford, A., Hargreaves, L. and Wall, E. (2012) CCRA Risk Assessment for the Agriculture Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Capon, R. and Oakley, G. (2012) CCRA Risk Assessment for the Built Environment Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Ramsbottom, D., Sayers, P. and Panzeri, M. (2012) CCRA Risk Assessment for the Floods and Coastal Erosion Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Rance, J., Wade, S.D., Hurford, A.P., Bottius, E. and Reynard, N.S. (2012) CCRA Risk Assessment for the Water Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Thornes, J., Rennie, M., Marsden, H. and Chapman, L. (2012) CCRA Risk Assessment for the Transport Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Pinnegar, J., Watt, T., and Kennedy, K. (2012) CCRA Risk Assessment for the Marine and Fisheries Sector. UK 2012 Climate Change Risk Assessment, Defra, London.

Other CCRA reports

Ballard, D. (2009) Working paper on adaptive capacity. Working paper 1 for the UK's climate change risk assessment 2012. Alexander Ballard Ltd for HR Wallingford Ltd.

CCRA (2011) Systematic mapping report. UK 2012 Climate Change Risk Assessment, Defra, London.f

Defra (2010a) Method for undertaking the CCRA, UK 2012 Climate Change Risk Assessment, Defra, London.

Defra (2010b) Method for undertaking the CCRA Part II - Detailed Method for Stage 3: Assess Risk, UK 2012 Climate Change Risk Assessment, Defra, London.

CCRA (2012) CCRA Evidence Report. UK 2012 Climate Change Risk Assessment, Defra, London.

Page 200: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

172 Biodiversity Sector Report

Other References

ADAS (2010) Climate Vulnerability Assessment of Designated Sites in Wales (Final Report Ref: 209 TSG 09). ADAS, UK.

Adrian, R., O’Reilly, M., Zagarese, H., Baines, S., Hessen, D., Keller, W., Livestrong, D., Sommaruga, R., Straile, D., Van Donk, E., Weyhenmeyer, G., and Winder, M. (2009) Lakes as sentinels of climate change. Limnology Oceanography. November 2009; 54 (6): 2283-2297.

Albertson, K., Aylen, J., Cavan, G, and McMorrow, J. (2010) Climate change and the future occurrence of moorland wildfires in the Peak District of the UK. Climate Research 45, no. 24: 105-118.

Anderson, P.K., Cunningham, A.A., Patel, N.G., Morales, F.J., Epstein, P.R., and Daszak P. (2004). Emerging infectious diseases of plants: pathogen pollution, climate change and agrotechnology drivers. Trends in Ecology and Evolution.19, 535-44.

Angus, S. (2009). Dé tha cearr air a’mhachaire? Biodiversity issues for Scottish machair: an initial appraisal. Glasgow Naturalist, 25 (Supplement), 53-62.

Angus, S., Hansom, J., and Rennie, A. (2010) Oir-thìrean na h-Alba ag atharrachadh: habitat change on Scotland’s coasts – The changing nature of Scotland. Marrs, S.J., Foster, S., Hendrie, C., Mackey, E.C. and Thompson, D.B.A. (eds). TSO, Edinburgh.

Araújo, M.B., Pearson, R.G., Thuiller, W., and Erhard, M. (2005) Validation of species-climate impact models under climate change. Global Change Biology 11: 1504-1513.

Araújo, M.B., Thuiller, W., and Pearson, R.G. (2006) Climate warming and the decline of amphibians and reptiles in Europe. Journal of Biogeography 33: 1712-1728.

Arnell, N. (1998) Climate change and water resources in Britain. Climate Change 39: 83-110. Kluwer Acedemic Publishers; Netherlands.

Austin, G. and Rehfisch, M.M. (2003) The likely impact of sea level rise on waders (Charadrii) wintering on estuaries. Journal for Nature Conservation, 11, 43-58.

Austin, G., and Rehfisch, M.M. (2005) Shifting non-breeding distributions of migratory fauna in relation to climatic change. Global Change Biology, 11, 31-38.

Baillie, S.R., Marchant, J.H., Crick, H.Q.P., Noble, D.G., Balmer, D.E., Beaven, L.P., Coombes, R.H., Downie, I.S., Freeman, S.N., Joys, A.C., Leech, D.I., Raven, M.J., Robinson, R.A. and Thewlis, R.M. (2005) Breeding birds in the wider countryside: their conservation status 2004. BTO Research Report No. 385. BTO, Thetford.

Bain, C; Bonn, A.; Stoneman, R.; Bruneau, P.; Chapman, S.J.; Coupar, A.; Evans, M.; Geary, B.; Howat, M.; Hjoosten, H.; Keenleyside, C.; Lindsay, R.; Labadz, J.; Littlewood, N.; Lunt, P.; Miller, C.; Moxey, A.; Orr, H.; Reed, M.; Shepherd, M.; Smith, P.; Swales, V.; Thompson, .B.A.; van de Noort, R.; Wilson, J.D.; Worrall, F. (2011) Commission of Inquiry on UK Peatlands. Report to IUCN UK Peatland Programme, Edinburgh.

Bateman, I.J. and Georgiou, S. (2010) The socioeconomic consequences of climate change for the management of water resources: the impact of climate change on European lakes (ed D.G. George), pp. 437–452. Springer, Amsterdam

Bateman, I.J. (ed.) (2011a) Economic values from ecosystems. UK National Ecosystem Assessment Technical Report. Chapter 22. UNEP-WCMC.

Bateman, I, (ed.) (2011b) Valuing changes in ecosystem services: scenario analyses. UK National Ecosystem Assessment Technical Report. Chapter 26. UNEP-WCMC

Page 201: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 173

Bayliss, M. (2006) T7.1: Climate change and diseases of plants, animal and humans: an overview. Foresight: Infectious Diseases: preparing for the future. Office of Science and Innovation.

Beaumont, N., Hattam, C., Mangi, S., Moran, D. van Soest, D., Jones, L. and Tobermann, M. (2010) Economic analysis of ecosystem services provided by UK Coastal Margin and Marine Habitats, Final Report. The Economics Team of the UK National Ecosystem Assessment. Plymouth Marine Laboratory, Plymouth.

Beaumont, N, Townsend, M, Mangi, S, Austen, MC (2006) Marine Biodiversity: an economic valuation. Defra, London

Beaumont, N.J., Austin, M.C., Mangi, S.C., Townsend, M. (2008) Economic valuation for the conservation of marine biodiversity. Marine Pollution Bulletin 56, 386–396

Bellamy, P.H., Loveland, P.J., Bradley, R.A., Lark ,R.M., and Kirk, G.J.D. (2005) Carbon losses from all soils across England and Wales. Nature 437:245–248.

Berry, P. (2007) The potential impacts of climate change on species in England. Unpublished. Environmental Change Institute, University of Oxford, Oxford.

Berry, P.M., O'Hanley, J.R., Thomson, C.L., Harrison, P.A, Masters, G.J., and Dawson, T.P. (eds.) (2007) Modelling natural resource responses to climate change (MONARCH): MONARCH 3 Contract Report. UKCIP Technical Report, Oxford.

Berry, P.M., Harrison, P.A., Dawson, T.P. and Walmsley, C.A. (2005) Monarch 2: modelling natural resource responses to climate change. UKCIP, Oxford.

Berry, P., Burch, S., and Sanders, M. (2011) EMBEDS: Embedding biodiversity adaptation principles. Final project report to Defra.

Bianchi, F.J.J.A., Booij, C.J.H., Tscharntke, T. (2006) Sustainable pest regulation in agricultural landscapes: a review on landscape composition, biodiversity and natural pest control. Proceedings of the Royal Society, Series B, Biological Sciences, 273(1595), 1715-1727.

Billett, M.F., Charman, D.J., Clark, J.M., Evans, C.D., Evans, M.G., Ostle, N.J., Worrall, F., Burden, A., Dinsmore, K.J., Jones, T., McNamara, N.P., Parry, L., Rowson, J.G. and Rose, R. (2010) Carbon balance of UK peatlands: current state of knowledge and future research challenges. Climate Research 45: 13–29.

Black and Veatch, (2007) Tidal power in the UK. Research report 3 – Review of Severn barrage proposals, Sustainable Development Commission. Online: http://www.sd-commission.org.uk/publications/downloads/TidalPowerUK3-Severn_barrage_proposals.pdf. [Accessed: 05/01/2012]

Booth, P., Brown, I., Guillaume, P., Towers, W. (2010) Impacts of biomass and bioenergy crops on landscape, land use and the wider environment in Northern Ireland and Scotland - Final Report. Scotland and Northern Ireland Forum for Environmental Research (SNIFFER). www.sniffer.org.uk

Bosch, J., Carrascal, LM., Duran, L., Walker, S., Fisher, M.C. (2006). Climate change and outbreaks of amphibian chytridiomycosis in a montane area of Central Spain; is there a link? Proceedings of the Royal Society of London. Series B. 274:253-260.

Both, C., Bouwhuis, S., Lessels, C.M., Visser, M.E. (2006) Climate change and population declines in a long-distance migratory bird. Nature 441:81–83.

Bradley RI, Milne R, Bell J, Lilly A, Jordan C, Higgins A (2005) A soil carbon and land use database for the United Kingdom. Soil Use Management 21:363–369

Page 202: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

174 Biodiversity Sector Report

Braschler, B., Hill, J. (2007) Role of larval host plants in the climate-driven range expansion of the butterfly Polygonia c-album. Journal of Animal Ecology 76: 415-423.

Bright, J.A., Langston, R.H.W., Bullman, R., Evans, R.J., Gardner, S., Pearce-Higgins, J., and Wilson, E., 2006, Bird sensitivity map to provide locational guidance for onshore wind farms in Scotland RSPB Research Report No.20.

Britton, A., Beale, C.M., Towers, W., and Hewison, R.L. (2009) Biodiversity gains and losses: Evidence for homogenisation of Scottish alpine vegetation. Biological Conservation. Volume 142, Issue 8, Pages 1728-1739.

Broadmeadow, M.S.J, Ray, D., and Samuel, C.J.A, (2005) Climate change and the future for broadleaved tree species in Britain. Forestry 78: 145-161.

Broadmeadow, M.S.J. Morecroft, M.D. and Morison, J.I.L. (2009) Observed Impacts of climate change on UK forests to date In: Read, D.J., Freer-Smith, P.H., Morison, J.I.L., Hanley, N., West, C.C. and Snowdon, P. (eds). (2009) Combating climate change – a role for UK forests. An assessment of the potential of the UK’s trees and woodlands to mitigate and adapt to climate change. The Stationery Office, Edinburgh.

Brooker, R.W. (2006) Plant-plant interactions and environmental change. New Phytologist, 171, 273-275.

Brondízio, E.S., Gatzweiler, F.W., Zografos, C. and Kumar, M. (2011) .Chapter 4 Socio-cultural context of ecosystem and biodiversity valuation. The Economics of Ecosystems and Biodiversity (TEEB). Chapter 4: Final Report.

Brown, I., Towers, W., Rivington, M., and Black H.I.J. (2008) Influence of climate change on agricultural land-use potential: adapting and updating the land capability system for Scotland. Climate Research 37, 43–57.

Brown, I., Poggio, L., Gimona, A., and Castellazzi, M. (2011). Climate change, drought risk and land capability for agriculture: implications for land use in Scotland. Regional Environmental Change, 11, 503-518.

Brussaard, L., Behan-Pelletier, V.M., Bignell, D.E., Brown, V.K., Didden, W., Folgarait, P., Fragoso, C., Freckman, D.W., Vadakattu, V.S.R., Hattori, T., Hawksworth, D.L., Klopatek, C., Lavelle, P., Malloch, D.W., Rusek, J., Soderstrom, B., Tiedje, J.M., and Virginia, R.A., (1997) Biodiversity and ecosystem functioning in soil. AMBIO, 26(8): 563-570.

Burgess, K.A., Orford, J., Dyer, K.R., Townend, I.H., Balson, P. (2002) Futurecoast - the integration of knowledge to assess future coastal evolution at a national scale, In: 28th International Conference on Coastal Engineering, ASCE, New York, 3, pp. 3221-3233.

CABI (2010) The economic costs of invasive non-native species on Great Britain. Report CAB/001/09 to Scottish Government, Defra and National Assembly of Wales.

Caissie, D. (2006) The thermal regime of rivers: a review. Freshwater Biology, 51, 1389-1406.

Carey, P.D., Wallis, S., Chamberlain, P.M., Cooper, A., Emmett, B.A., Maskell, L.C., McCann, T., Murphy, J., Norton, L.R., Reynolds, B., Scott, W.A., Simpson, I.C., Smart, S.M., Ullyett, J.M. (2008) Countryside survey: UK results from 2007. NERC/Centre for Ecology and Hydrology, 105pp.

Carter, R.W.G. and Woodroffe, C.D. (editors) (1994) Coastal evolution: Late quaternary shoreline morphodynamics. Cambridge University Press, Cambridge, 517 pp.

Page 203: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 175

Carvalho, L., and Kirika, A. (2003) Changes in shallow lake functioning: Response to climate change and nutrient reduction. Hydrobiology, 506/509, 789–796.

Cazenave, A., and Llovel, W. (2010) Contemporary sea level rise. Annual review of marine science, 2, 145-173.

Christie M, Hyde T, Cooper R, Fazey I, Dennis P, Warren J, Colombo S and Hanley N (2011) Economic Valuation of the Benefits of Ecosystem Services delivered by the UK Biodiversity Action Plan. (Defra Project SFFSD 0702) Final Report to Defra.

Ciais, P., Reichstein, M., Viovy. N., Granier, A., Ogée, J., Allard, V., Aubinet, M., Buchmann, N., Bernhofer, C., Carrara, A., Chevallier, F., De Noblet, N., Friend, A.D., Friedlingstein, P., Grünwald, T., Heinesch, B., Keronen, P., Knohl, A., Krinner, G., Loustau, D., Manca, G., Matteucci, G., Miglietta, F., Ourcival, J.M., Papale, D., Pilegaard, K., Rambal, S., Seufert, G., Soussana, J.F., Sanz, M.J., Schulze, E.D., Vesala, T. and Valentini, R. (2005). Europe-wide reduction in primary productivity caused by heat and drought in 2003, Nature 437, 529-533.

CJC Consulting (2004) Cost Effectiveness Study of Approaches for Delivery of PSA Target Relating to SSSIs. Final report to Defra.

Clark, J.M., Bottrell, S.H., Evans, C.D., Monteith, D.T., Bartlett, R., Rose, R., Newton, R.J., and Chapman, P.J. (2010a) The importance of the relationship between scale and process in understanding long-term DOC dynamics. Science of the Total Environment. 408 (13), 2768-2775

Clark, J.M., Gallego-Sala, A.V., Allott, T.E.H., Chapman, S.J., Farewell, T., Freeman, C., House, J.I., Orr, H.G., Prentice, C. and Smith, P. (2010b) Assessing the vulnerability of blanket peat to climate change using an ensemble of statistical bioclimatic envelope models. Climate Research 45:131–150.

Clarke, D., Sanitwong Na Ayuthaya, S. (2010) Predicted effects of climate change, vegetation and tree cover on dune slack habitats at Ainsdale on the Sefton Coast, UK. Journal of Coastal Conservation, 14/2, 115-126.

Chapman SJ, Bell J, Donnelly D, Lilly A (2009) Carbon stocks in Scottish peatlands. Soil Use Management 25:105–112

Comiso, J.C., Parkinson, C.L., Gernsten, L., and Stock, L. (2008) Accelerated decline in the Arctic sea ice cover. Geophysical Research Letters, 35, L01703.

Communities and Local Government (CLG), (2007) Planning policy statement 1 - supplement to planning and climate change.

Communities and Local Government (CLG), (2008) IMP Steering Group Integrated Risk Management Planning: Policy Guidance – Wildfire.

Communities and Local Government (CLG), (2009) Incident recording system – questions and lists, Version 1.4 – (XML Schemas v1-0n).

Conlan, K., Wade, T., Ormerod, S., Lane, S., Durance, I., and Yu, D. (2007). Preparing for climate change impacts on freshwater ecosystems (PRINCE). Environment Agency Science Report: SC030300/SR.

Cornell, S. (2011) Valuing ecosystem benefits in a dynamic world. Climate Research, 45, 261-272.

Cox, B.A., and Whitehead, P.G. (2009) . Hydrology Research, 40(2–3): 138–152.

Crick, H.Q.P. (2004) The impacts of climate change on birds. Ibis 146 (suppl.1): 48-56.

Page 204: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

176 Biodiversity Sector Report

Crimmins, S. M., Dobrowski, S. Z., Greenberg, J. A., Abatzoglou, J. T., Mynsberge, A.R., (2011) Changes in climatic water balance driver downhill shifts in plant species’ optimum elevations. Science 21 Vol 331 No 6015 pp324-327.

Cruickshank, M.M., Tomlinson, R.W., and Milne, R. (2000). Application of CORINE land cover mapping to estimate surface carbon pools in Northern Ireland. In: Vegetation mapping: from patch to planet. Edited by Alexander, R., and Millington, A.C., pp 259 - 272. Wiley, London.

Cruickshank, M.M., Tomlinson, R.W., Devine, P.M., and Milne, R. (1998) Carbon in the vegetation and soils of Northern Ireland. Biology and Environment: Proceedings of the Royal Irish Academy, 98B, 9 - 21.

Davidson, I.C., and Hazlewood, M.S. (2005) Effect of climate change on salmon fisheries. Environment Agency Science Report: W2-047/SR.

Davies, Z.G., Wilson, R.J., Coles, S. and Thomas, C.D. (2006) Changing habitat associations of a thermally constrained species, the silver spotted skipper butterfly, in response to climate warming. Journal of Animal Ecology, 75: 247-256.

Dawson, J.J.C., and Smith, P. (2007) Carbon losses from soil and its consequences for land management. Science of the Total Environment 382, 165-190.

DECC (2010) Severn Tidal Power: Feasibility Study Conclusions and Summary Report. Online: http://www.decc.gov.uk/en/content/cms/meeting_energy/wave_tidal/severn_tidal_power/severn_tidal_power.aspx [Accessed: 04/01/2012]

Defra (2002) The Futurecoast project. Project code FD2002. Online: http://www.coastalwiki.org/coastalwiki/FUTURECOAST_project,_UK [Accessed: 05/01/2012]

Defra (2003) Review of non-native species policy. Report number PB8072. Defra. Online: https://secure.fera.defra.gov.uk/nonnativespecies/search.cfm. [Accessed: 30/03/2011].

Defra (2004) Phytophthora ramorum epidemiology: sporoulation potential, dispersal, infection, latency and survival – PH0194. http://randd.defra.gov.uk/: Defra Cromwell House, London.

Defra (2006a) R&D technical report FD2017/TR: National evaluation of the cost of meeting coastal environmental requirements. Online: http://sciencesearch.defra.gov.uk/Document.aspx?Document=FD2017_5200_TRP.pdf. [Accessed: 05/01/2012]

Defra (2006b) The UK biodiversity action plan: Highlights from the 2005 reporting round. Defra, on behalf of the UK Biodiversity Partnership.

Defra (2007a) Management and containment of Phytophthora ramorum infections in the UK – Defra, 2007. Online: http://randd.defra.gov.uk/: Defra. [Accessed: 05/12/2012]

Defra (2007b) An introductory guide to valuing ecosystem services. Online: http://www.defra.gov.uk/publications/files/pb12852-eco-valuing-071205.pdf. [Accessed: 05/12/2012]

Defra, (2009) Development of disease management approaches for the quarantine pathogens Phytophthora ramorum and P. kernoviae on Vaccinium and other heathland species in important habitats in England and Wales – PH0601. Online: http://randd.defra.gov.uk/: Defra. [Accessed: 05/12/2012]

Page 205: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 177

Defra, (2010c) Biodiversity indicators in your pocket 2010: Measuring progress towards halting biodiversity loss. Defra product code PB13400. Online: http://www.defra.gov.uk/evidence/statistics/environment/wildlife/download/pdf/biyp2010.pdf [Accessed 13/02/2011].

Defra, (2010d) Charting progress 2 feeder report: Healthy and biologically diverse seas. UK Marine Monitoring and Assessment Strategy (UKMMAS), Defra. Online: http://chartingprogress.defra.gov.uk/chapter-3-healthy-and-biologicaly-diverse-seas [Accessed 26.01.11].

Defra (2011a) Developing Tools to Evaluate the Consequences for Biodiversity of Options for Coastal Zone Adaptation to Climate Change (project CR0422). Defra, London.

Defra (2011b) An investigation into three climate change issues for Priority Habitats and protected sites. Project WC0751. Unpublished. Defra, London.

Defra (2011c) Observatory monitoring framework – indicator sheet. Online: http://archive.defra.gov.uk/evidence/statistics/foodfarm/enviro/observatory/indicators/d/de5_data.htm [Accessed: 12/01/2012]

d’Erlanger, J. (2001) Priority habitat definition statement: Lowland beech and yew woodland - v1.3. English Nature, Peterborough.

Dieleman, W., Luyssaert, S., Rey S.A., de Angelis, P., Barton, C., Broadmeadow, M., Chigwerewe, K., Crookshanks, M., Dufrêne, E., Broadmeadow, S., Jarvis, P.G., Kasurinen, A., Kellomaki, S., Le Dantec, V., Liberloo, M., Marek, M.V., Medlyn, B., Pokorný, R., Scarascia-Mugnozza, G., Temperton, V., Tingey, D., Urban, O., Ceulemans, R. and Janssens, I. (2010) Soil fertility affects tree root biomass responses to elevated CO2: A meta-analysis. Plant, Cell and Environment, 33, 2001-2011.

Doody, P., Lee, M., and Pethick, J. (2006) Coastal evolution in Suffolk: an evaluation of geomorphological and habitat change. English Nature Research Report No. 647.

Dowdy, A.J., Mills, G.A., Finkele, K., and de Groot, W. (2009) Australian fire weather as represented by the McArthur forest fire danger index and the Canadian forest fire weather index, CAWCR technical report No. 10. Online: http://www.cawcr.gov.au/publications/technicalreports/CTR_010.pdf. [Accessed: 05/12/2012]

Downing, T. E., Butterfield, R.E., Edmonds, B., Knox, J.W., Moss, S., Piper B.S., Weatherhead, E.K. (and the CCDeW project. team) (2003) Climate change and demand for water. Research report. Stockholm Environment Institute Oxford Office, Oxford.

Drake, J.M. and Griffin, B.D. (2010) Early warning signals of extinction in deteriorating environments. Nature 467,456–459.

Drewitt, A.L., and Langston, R.H.W. (2006) Assessing the impacts of wind farms on birds. Ibis 148: 29-42.

Duckworth, J.C., Bunce, R.G.H. and Malloch, A.J.C. (2000) Modelling the potential effects of climate change on calcareous grasslands in Atlantic Europe. Journal of Biogeography 27: 347-358.

Durance, I., and Ormerod, S.J. (2007) Climate change effects on upland stream invertebrates over a 25 year period. Global Change Biology, 13, 942-957.

Eaton, M.A., Appleton, G.F., Ausden, M.A., Balmer, D.E., Grantham, M.J., Grice, P.V., Hearn, R.D., Holt, C.A., Musgrove, A.J., Noble, D.G., Parsons, M., Risely, K., Stroud,

Page 206: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

178 Biodiversity Sector Report

D.A. and Wotton, S. (2010) The state of the UK’s birds 2010. RSPB, BTO, WWT, CCW, JNCC, NE, NIEA and SNH, Sandy, Bedfordshire.

ECOHARM (2001) The socio-economic impact of harmful algal blooms in European

marine waters. Online: www.bom.hik.se/ecoharm/deliverables/ecoharm.pdf. [Accessed: 05/12/2012] EftEC (2006) Valuing our Natural Environment. Final report to Defra. Research Project NR0103

Eigenbrod, F., Anderson, B.J., Armsworth, P.R., Heinemeyer, A., Jackson, S.F., Parnell, M., Thomas, C.D. and Gaston, K.J. (2009) Ecosystem service benefits of contrasting conservation strategies in a human-dominated region. Proceedings of the Royal Society B: Biological Sciences, 276, 2903-2911.

Elliott, J.A., Jones, I.D., and Thackeray, S.J. (2006). Testing the sensitivity of phytoplankton communities to changes in water temperature and nutrient load, in a temperate lake. Hydrobiologia, 559, 401-411.

Elliott, J.M., and Hurley, M.A. (1997) A functional model for maximum growth of Atlantic salmon parr, Salmo salar, from two populations in Northwest England. Functional Ecology 11, 592-603.

Elliott, J.M., and Elliott, J.A. (2010). Temperature requirements of Atlantic salmon Salmo salar, brown trout Salmo trutta and Arctic charr Salvelinus alpinus: predicting the effects of climate change. Journal of Fish Biology. 77,8 1793-1817.

Emmett, B.A., Reynolds, B., Chamberlain, P.M., Rowe, E., Spurgeon, D., Brittain, S.A., Frogbrook, Z., Hughes, S., Lawlor, A.J., Poskitt, J., Potter, E., Robinson, D.A., Scott, A., Wood, C.,Woods, C. (2010) Countryside survey: soils report from 2007. NERC/Centre for Ecology and Hydrology, 192pp. (CS Technical Report No. 9/07, CEH Project Number: C03259).

English Nature, (2005) Audit of non-native species in England. English Nature: Peterborough.

English Nature, (2004) Our coasts and seas: A 21st century agenda for their recovery conservation and sustainable use. English Nature, Peterborough, UK.

Eno, N. C., Clark, R.A., and Sanderson, W.G. (1997) Non-native marine species in British waters: A review and directory, 152 pages A4 softback, ISBN 1 86107 442 5

Environment Agency (2008) Climate change and river flows in the 2050s. Environment Agency Science Summary SC070079/SS1

Eurosion (2004) Living with coastal erosion in Europe: Sediment and space for sustainability, PART II – Maps and statistics. Online: http://www.eurosion.org/index.html. [Accessed: 05/01/2012]

Evans, C.D., Chapman, P.J., Clark, J.M., Monteith, D.T., Cresser, M.S. (2006) Alternative explanations for rising dissolved organic carbon from organic soils. Global Change Biology 12:2044–2053.

Evans, E., Ashley, R., Hall, J., Penning-Rowsell, E., Saul, A., Sayers, P., Thorne, C., and Watkinson, A. (2004) Foresight: Future flooding. Scientific summary: Volume I - Future risks and their drivers. Office of Science and Technology, London.

Fera and Newcastle University (2010). Estimating the wildlife and landscape benefits of the environmental stewardship scheme. Final report to Defra. Online:

Page 207: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 179

http://archive.defra.gov.uk/evidence/economics/foodfarm/reports/documents/estimatingthewildlife.pdf [Accessed: 05/01/2012]

Fezzi, C., Hutchins, M., Rigby, D., Bateman, I.J., Posen, P. and Hadley, D. (2010) Integrated assessment of Water Framework Directive Nitrate Reduction Measures. Agricultural Economics, 41, 123–134.

Fish, R., Burgess, J., Chilvers, J., Footitt, A., Haines‐Young, R., Russel, D., Turner, K.

and Winter, D.M. (2011) Participatory and Deliberative Techniques for Embedding an Ecosystems Approach into Decision Making. Full Technical Report. (NR0124). Defra.

Foresight, (2010) Land use futures: making the most of land in the 21st century. Online: http://bis.ecgroup.net/Publications/Foresight/LandUseFutures.aspx [Accessed: 30/03/2011].

Franco, A.M.A. (2006) Impacts of climate warming and habitat loss on extinction of species’ low-latitude range boundaries. Global Change Biology, 12: 1545-1553.

Freeman, S. (2009) Towards a method for the estimation and use of averaged multispecies trends, as indicators of patterns of change in butterfly populations. Centre for Ecology and Hydrology. Online: http://www.ukbms.org/Downloads/UKBMS_Milestone4_IndicatorTrendAnalysis.pdf [Accessed: 05/01/2012]

Friedlingstein, P., Houghten, R.A., Marland, G., Hacklar, J., Boden, T.A., Conway, T.J., Canadell, J.G., Raupach, M.R., Ciais, P., and Le Quéré, C. (2010) Update on CO2. Nature Geoscience 3: 811-812

Garner, T.W.J., Walker, S., Bosch, J., Hyatt, A.D., Cunningham, A.A., Fisher, M.J. (2005) Chytrid fungus in Europe. Emerging Infectious Diseases, 11, 1639-1641.

Garten, Jr C.T., Classen, A.T., Norby, R.J. (2009) Watering treatment surpasses elevated CO2 and temperature in importance as a determinant of soil carbon dynamics in a multi-factor climate change experiment. Plant Soil 319:85–94.

Gazzard, R. (2010) Development of wildfire statistics and risk impacts in the United Kingdom, Fire and Rescue Statistics User Group (FRSUG). Online: http://www.frsug.org/reports/ [Accessed 13/02/2011].

Gazzard, R. (2009) United Kingdom vegetation fire standard (UKVFS), Forestry Commission.

George, D. G., Maberly, S.C., and Hewitt, D.P. (2004) The influence of the North Atlantic Oscillation on the physical, chemical and biological characteristics of four lakes in the English Lake District. Freshwater Biology. 49: 760–774.

GHK (2011a) Benefits of Sites of Special Scientific Interest. Final Report to Defra. Online: http://randd.defra.gov.uk/ [Accessed: 05/01/2012]

GHK (2011b) Costing potential actions to offset the impact of development on biodiversity. Final Report to Defra. Online: http://randd.defra.gov.uk/ [Accessed: 05/01/2012]

Gibson, C.E. (2011) Northern Ireland state of the seas report. Online: http://www.doeni.gov.uk/niea/water-home/state_of_the_seas_ni_report.htm [Accessed: 05/01/2012]

Golding, N., and Betts, R. (2008) Fire risk in Amazonia due to climate change in the HadCM3 climate model: Potential interactions with deforestation, Global Biogeochemical Cycles, 22, GB4007.

Page 208: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

180 Biodiversity Sector Report

Graham, C.T., and Harrod, C. (2009) Implications of climate change for the fishes of the British Isles. Journal of Fish Biology 74, pp. 1143–1205.

Grant, J.W.A., Steingrimsson, S.O., Keeley, E.R., and Cunjak, R.A. (1998) Implications of territory size for the measurement and prediction of salmonid abundance in streams. Canadian Journal of Fisheries and Aquatic Science 55 (Suppl. 1): 181–190.

Green, R.E., Collingham, Y,C. Willis, S.G., Gregory, R.D., and Smith, K.W. (2008) Performance of climate envelope models in retrodicting recent changes in bird population size from observed climatic change. Biology Letters 4, 599-602.

Green, S., and Ray, D. (2009) Potential impacts of drought and disease on forestry in Scotland, Forestry Commission Scotland.

Gregory, R.D., Wilkinson, N.I., Noble, D.G., Robinson, J.A., Brown, A.F, Hughes, J., Procter, D.A., Gibbons, D.W., and Galbraith, C.A. (2002) The population status of birds in the United Kingdom, Channel Islands and Isle of Man: an analysis of conservation concern 2002-2007. British Birds 95: 410-450.

Gurney, W.S.C., Bacon, P.J., Tyldesley, G., and Youngson, A.F. (2008) Process-based modelling of decadal trends in growth, survival, and smolting of wild salmon (Salmo salar) parr in a Scottish upland stream. Canadian Journal of Fisheries and Aquatic Sciences, 65, 2606-2622.

Guttal, V., and Couzin, I. D. (2010). Social interactions, information use and the evolution of collective migration. Proceedings of the National Academy of Sciences USA, 107(37), 16172-16177.

Hammond, D. and Pryce, A. (2007) Climate change impacts and water temperature. Science Report SC060017/ SR. Bristol: Environment Agency. Online: http://publications.environment-agency.gov.uk/pdf/SCHO0707BNAG-e-e.pdf [Accessed: 05/01/2012]

Hannaford, J., and Marsh, T.J. (2006) An assessment of trends in UK runoff and low flows using a network of undisturbed catchments. International Journal of Climatology., 26, 1237–1253.

Hannah, L., Midgley, G.F., Andelmand, S., Araujo, M.B., Hughes, G., Martinez-Meyer, E., Pearson, R. and Williams, P.H. (2007) Protected area needs in a changing climate. – Frontiers in Ecology and Environment 5: 131-138.

Harries, T., and Penning-Rowsell, E. (2011) Victim pressure, institutional inertia and climate change adaptation: The case of flood risk. Global Environmental Change 21, 188–197.

Harris, D., Crabtree, B., King, J., Newell-Price, P. (2006) Economic Valuation of Soil Functions Phase 1: Literature Review and Method Development. Report to Defra. Online: http://www.cjcconsulting.co.uk/pdfs/FINALEcon%20Valuation%20of%20Soil%20Functions%20Rept%20Phase%20I_0806.pdf [Accessed: 05/01/2012]

Harrison, P.A., Berry, P.M. and Dawson, T.P. (Eds.) (2001) Climate change and nature conservation in Britain and Ireland: Modelling natural resource responses to climate change (the MONARCH project). UKCIP Technical Report, Oxford.

Harvell, C.D., Kim, K., Burkholder, J.M., Colwell, R.R., Epstein, P.R., Grimes, D.J., Hofmann, E.E., Lipp, E.K., Osterhaus, A.D.M.E., Overstreet, R.M., Porter, J.W., Smith, G.W., Vasta, G.R. (1999) Emerging marine diseases - climate links and anthropogenic factors. Science 285, 1505–1510.

Page 209: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 181

Hassall, C., Thompson, D.J., French, G.C., and Harvey, I.F. (2007) Historical changes in the phenology of British Odonata are related to climate. Global Change Biology,13, 933-941.

Haughton, A.J., Bond, A.J., Lovett, A.A., Dockerty, T., Sunnenberg, G., Clark, S.J., Bohan, D.A., Sage, R.B., Mallott, M.D., Mallott, V.E., Cunningham, M.D., Riche, A.B., Shield, I.F., Finch, J.W., Turner, M.M., and Karp, A. (2009) A novel, integrated approach to assessing social, economic and environmental implications of changing rural land-use: a case study of perennial biomass crops. Journal of Applied Ecology, 46, 315-322.

Heath, J,, Ayres, E., Possell, M., Possell, M., Bardgett, R.D., Black, H.I.J., Grant, H., Ineson, P., and Kerstiens, G. (2005) Rising atmospheric CO2 reduces sequestration of root-derived soil carbon. Science, 309, 1711–1713.

Heikkinen, R.K., Luoto, M., Leikola, N., Pöyry, J., Settele, J., Kudrna, O., Marmion, M., Fronzek, S., Thuiller, W. (2010) Assessing the vulnerability of European butterflies to climate change using multiple criteria. Biodiversity Conservation. 19:695–723.

Hewson, C.M., and Noble, D.G. (2009) Population trends of breeding birds in British woodlands over a 32-year period: relationships with food, habitat use and migratory behaviour, Ibis 151, pp. 464–486.

Hickling, R., Roy, D. B., Hill, J. K., Fox, R., and Thomas, C. D. (2006). The distributions of a wide range of taxonomic groups are expanding polewards. Global Change Biology, 12, 1–6.

Hiederer, R., and Durrant, T. (2010). Evaluation of BioSoil demonstration project. Preliminary data analysis. JRC scientific and technical reports. European Union, Luxembourg. Online: http:// eusoils.jrc.ec.europa.eu/ESDB_Archive/eusoils_docs/ other/EUR24258.pdf. [Accessed 13/02/2011].

Hill, J.K., Thomas, C.D., Fox, R., Telfer, M.G., Willis, S. G., Asher. J., and Huntley B. (2002) Responses of butterflies to twentieth century climate warming: implications for future ranges. Proceedings of the Royal Society B, London. 269, 2163-2171.

HM Treasury (2003) The Treasury Green Book – Appraisal and Evaluation in Central Government. London.

Holden, J., Chapman, P., Evans, M., Hubacek, K., Kay, P., and Warburton, J. (2006) Vulnerability of organic soils. Defra Project SP0532.

Holden, J., Shotbolt, L., Bonn, A., Burt, T.P.L., Chapmana, P.J., Dougille, A.J., Frasere, E.D.G., Hubaceke, K., Irvinea, B., Kirkbya, M.J., Reede, M.S., Prellf, C., Staglg, S., Stringerh, L.C., Turnera, A., and Worralli, F. (2007) Environmental change in moorland landscapes. Earth Science Review 82:75–100.

Hopkins, J.J., Allison, H.M., Walmsley, C.A., Gaywood, M., Thurgate, G. (2007) Conserving biodiversity in a changing climate: guidance on building capacity to adapt. Published by Defra on behalf of the UK Biodiversity Partnership.

Hossell, J.E., Briggs, B., and Hepburn, I.R. (2000) Climate change and UK nature conservation: a review of the impact of climate change on UK species and habitat conservation policy. DETR, HMSO, London.

Huin, N., and Sparks, T.H. (1998) Arrival and progression of the Swallow Hirundo rustica through Britain. Bird Study, 45:361-370.

Hulme, P.E. (2007) Biological invasions in Europe: drivers, pressures, states, impacts and responses. In: Hester R, Harrison RM (eds) Biodiversity under threat issues in

Page 210: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

182 Biodiversity Sector Report

environmental science and technology. Royal Society of Chemistry, Cambridge. 25:56–80.

Huntley, B, Collingham, Y.C., Willis, S.G., and Green, R.E. (2008) Potential impacts of climatic change on European breeding birds. Public Library of Science ONE 3(1): e1439.

Huntley, B., Green, R.E., Collingham, Y.C., and Willis, S.G. (2007) A climatic atlas of European breeding birds. Barcelona, Spain: Lynx Edicions.

IPCC (2007) Climate Change 2007: The physical science basis. Contribution of working group I to the fourth assessment report of the Intergovernmental Panel on Climate Change [Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., and Miller, H.L. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 996 pp.

ISCCC (2009) International Scientific Congress Climate Change: Global risks, challenges and decisions – Synthesis report. ISCCC, Copenhagen.

Ise, T., Dunn, A.L., Wofsy, S.C., and Moorcroft P.R. (2008) High sensitivity of peat decomposition to climate change through water-table feedback. Nature Geoscience 1: 763 – 766.

Jacobs (2004) An Economic Assessment of the Costs and Benefits of Natura 2000 Sites in Scotland (Environment Group Research Report 2004/05). Scottish Executive

Janssens, I.A., Freibauer, A., Ciais, P., Smith, P., Nabuurs, G.J., Folberth, G., Schlamadinger, B., Hutjes, R.W.A., Ceulemans, R., Schulze, E.D., Valentini, R., and Dolman, A.J. (2003) Europe’s terrestrial Biosphere Absorbs 7 to 12 % of European anthropogenic CO2 emissions. Science 300: 1438-1541.

Jenkins, G.J., Perry, M.C., and Prior, M.J. (2009) The climate of the United Kingdom and recent trends. Met Office Hadley Centre, Exeter, UK.

JNCC (Joint Nature Conservation Committee) (2011) Treating Alzheimer’s disease with daffodils. Online: <http://www.jncc.gov.uk/page-5721> [Accessed: 05/01/2012]

Johnson, P.T.J., and Thieltges, D.W. (2010) Diversity, decoys and the dilution effect: how ecological communities affect disease risk. The Journal of Experimental Biology, 213: 961-970.

Keesing, F., Belden, L.K., Daszak, P., Dobson, A., Harvell, C.D., Holt., R.D., Hudson, P., Jolles, A., Jones, K.E., Mitchell, C.E., Myers, S.S., Bogich. T., and Ostfeld, R.S. (2010) Impacts of biodiversity on the emergence and transmission of infectious diseases. Nature, 468,647–652.

King, S.E., and Lester, J.N. (1995) The value of salt marsh as a sea defence. Marine Pollution Bulletin, 30(3), 180-189.

Kirby, K.J., Smart, S.M., Black, H.I.J., Bunce, R.G.H., Corney, P.M., and Smithers, R.J. (2005) Long term ecological change in British woodlands (1971–2001), English Nature, Peterborough.

Kirk, G.J.D., and Bellamy P.B. (2010) Analysis of changes in organic carbon in mineral soils across England and Wales using a simple single-pool model. European Journal of Soil Science, 61, 401–411.

Knox, J.W., Morris, J., Weatherhead, E.K., and Turner, A.P. (1999) Mapping the financial benefits of spray irrigation and potential financial impact of restrictions on abstraction: a case study in Anglian Region. Journal of Environmental Management, 58, 45–59.

Page 211: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 183

Kuhn, K., Campbell-Lendrum, D., Haines, A., and Cox, J. (2005) Using climate to predict infectious disease epidemics. Geneva: World Health Organization. Online: www.who.int/globalchange/publications/infectdiseases/en/index.html. [Accessed: 05/01/2012]

Landsberg, J. (2002) The effects of harmful algal blooms on aquatic organisms. Reviews in Fisheries Science, 10, 113-190.

Langan, S.J., Johnston, L., Donaghy, M.J., Youngson, A.F., Hay, D.W., and Soulsby, C. (2001) Variation in river water temperatures in an upland stream over a 30-year period. Science of the Total Environment, 265:195–207.

Lawton, J.H., Brotherton, P.N.M., Brown, V.K., Elphick, C., Fitter, A.H., Forshaw, J., Haddow, R.W., Hilborne, S., Leafe, R.N., Mace, G.M., Southgate, M.P., Sutherland, W.J., Tew, T.E., Varley, J., and Wynne, G.R. (2010) Making space for nature: a review of England’s wildlife sites and ecological network. Report to Defra.

Lee, M. (2001) Coastal defence and the Habitats Directive: predictions of habitat change in England and Wales. The Geographical Journal 167:139–56.

Lehikoinen, E., Sparks, T.H., Zalakevicius, M. (2004) Arrival and departure dates. Advances in Ecological Research 35:1–31.

Le Quéré, C., Raupach, M.R., Canadell, J.G., and Marland, G.M. (2009) Nature Geosciences, 2, 831-836.

Lichter, J., Billings, S.A., Ziegler, S.E., Gaindh, D., Ryals, R., Finzi, A.C., Jackson, R.B., Stemmler, E.A., and Schlesinger, W.H. (2008) Soil carbon sequestration in a pine forest after nine years of atmospheric CO2 enrichment. Global Change Biology 14:2910-2922 Nature Reports-Climate Change 3:2.

LIFE Peatlands Project, (2005) The Peatlands of Caithness and Sutherland: Management Strategy 2005-2015. Scottish Natural Heritage.

Lilly, A., Birnie, R.V.B., Futter, M.N., Grieve, I.C., Higgins, A., Hough, R., Jones, M.A., Jordan, C., Nolan, A.J., Stutter, M.I., Towers, W., and Baggaley, N.J. (2009) Climate change, land management and erosion in the organic and organo-mineral soils in Scotland and Northern Ireland. SNH Commissioned Report, No. 325 (ROAME No. F06AC104 - SNIFFER UKCC21).

Lindner, M., Maroschek, M., Netherer, S., Kremer, A., Barbati, A., Garcia-Gonzalo, J., Seidl, R., Delzon, S., Corona, P., Kolstrom, M., Lexer, M. J. and Marchetti, M. (2010) Climate change impacts, adaptive capacity and vulnerability of European forest ecosystems. Forest Ecology and Management, 259, 698 – 709.

Littlewood, N., Anderson, P., Artz, R., Bragg, O., Lunt, P., and Marrs, R. (2010) Peatland biodiversity: Technical Review. IUCN UK Peatland Programme.

Lorenz, M., Fischer, R., Becher, G., Granke, O., Seidling, W., Ferretti, M., Schaub, M., Calatayud, V., Bacaro, G., Gerosa, G., Rocchini, D. and Sanz, M. (2008) Forest condition in Europe: 2007. Technical Report of ICP Forests. Federal Research Centre for Forestry and Forest Products and University of Hamburg.

Luisetti, T., Turner, R.K., Bateman, I.J., Morse-Jones, S., Adams, C. and Fonseca, L. (2011) Coastal and marine ecosystem services valuation for policy and management: managed realignment case studies in England. Ocean and Coastal Management, 54, 212–224.

Maclean, I.M.D., Austin, G.E., Rehfisch, M.M., Blew, J., Crowe, O., Delany, S., Devos, K., Deceuninck, B., Gunther, K., Laursen, K., Van Roomen, M., and Wahl, J. (2008)

Page 212: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

184 Biodiversity Sector Report

Climate change causes rapid changes in the distribution and site abundance of birds in winter. Global Change Biology 14: 2489-2500.

Maggs, C., Mineur, F., Bishop J., and McCollin, T. (2010) Non-natives in MCCIP annual report card 2010-11, MCCIP Science Review, 15pp. Online: www.mccip.org.uk/arc [Accessed: 05/01/2012]

Manchester, S.J., and Bullock, J. M. (2000) The impacts of non-native species on UK biodiversity and the effectiveness of control. Journal of Applied Ecology, 37, 845-864.

Marsh, T., Cole, G., and Wilby, R. (2007) Major droughts in England and Wales, 1800–2006, Weather 62(4), pp. 87-93.

McMorrow, J., Aylen, J., Albertson, K., Cavan, G., Lindley, S., Handley, J., and Karooni, R. (2005) Change and the visitor economy technical report 3: Moorland wildfires in the Peak District, University of Manchester.

MEA (Millennium Ecosystem Assessment) (2003) Ecosystems and human well-being: A framework for assessment. Online: http://www.maweb.org/en/Framework.aspx. [Accessed: 05/01/2012]

Meentemeyer, R., Rizzo, D., Mark, W., and Lotz, E. (2004) Mapping the risk of establishment and spread of sudden oak death in California. Forest Ecology and Management, 200, 195-214.

Menéndez, R., González Megías, A., Hill, J.K., Braschler, B., Willis, S.G., Collingham, Y., Fox, R., Roy, D.B., and Thomas, C.D. (2006) Species richness changes lag behind climate change. Proceedings of the Royal Society B, London, 273, 1465-1470.

Metcalfe, N.B., and Thorpe, J.E. (1990) Determinants of geographic variation in the age of seaward migrating salmon (Salmo salar). Journal of Animal Ecology 59, 131–145.

Met Office, (2005) The Met Office fire severity index for England and Wales Met Office report ref: M/BO/P87.

Metroeconomica (2006) Project E – Quantify the cost of impacts and adaptation in climate change impacts and adaptation: Cross-regional research programme. Defra, London.

Mieszkowska, N. (2010) Intertidal habitats and ecology. In: MCCIP annual report card 2010-11, MCCIP science review.

Milne R. and Brown T.A.W. (1997) Carbon in the vegetation and soils of Great Britain. Journal of Environmental Management 49,413–33.

Mitchell, R.J., Morecroft, M.D., Acreman, M., Crick, H.Q.P., Frost, M., Harley, M., Maclean, I.D.M., Mountford, O., Piper, J., Pontier, H., Rehfisch, M.M., Ross, L.C., Smithers, R.J., Stott, A., Walmsley, C.A., Watts, O., and Wilson, E. (2007) England Biodiversity Strategy - towards adapation to climate change. Final report to Defra for contract CRO327. Defra, 177pp.

Moller, I., Spencer, T., French, J.R., Leggett, D.J., and Dixon, M. (2001) The sea-defence value of salt marshes: Field evidence from north Norfolk. Journal of the Chartered Institution of Water and Environmental Management, 15, (2), pp 109-116.

Monteith, D.T., Stoddard, J.L., Evans, C.D., de Wit, H.A., Forsius, M., Hogasen, T., Wilander, A., Skjelkvale, B.L., Jeffries, D.S., Vuorenmaa, J., Keller, B., Kopacek, J., and Vesely, J. (2007) Dissolved organic carbon trends resulting from changes in atmospheric deposition chemistry. Nature 450, 537–541.

Page 213: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 185

Moran, D. and Dann, S. (2008) The economic value of water use: implications for implementing the Water Framework Directive in Scotland. Journal of Environmental Management, 87,484–496.

Morecroft, M.D., Bealey, C.E., Howells, O., Rennie, S., Woiwod, I.P. (2002) Effects of drought on contrasting insect and plant species in the UK in the mid-1990s. Global Ecology and Biogeography, 11, 7–22.

Morecroft, M.D., Bealey, C.E, Beaumont, D.A, Benham, S., Brooks, D.R., Burt, T.P, Critchley, C.N.R, Dick, J., Littlewood, N.A., Monteith, D.T., Scott, W.A., Smith, R.I., Walmsley, C., and Watson, H. (2009) The UK environmental change network: Emerging trends in the composition of plant and animal communities and the physical environment. Biological Conservation: 142: 2814-2832.

Morris, J., Weatherhead, E.K., Knox, J., Vasilieou, K., de Vries, T., Freeman, D., Leiva, F. and Twite, C. (2004) Irrigation: The Case of England and Wales. (eds J. Burbel and C.G. Martin), The Sustainability of European Irrigation under Water Framework Directive and Agenda 2000. Office for Official Publications of the European Communities, Luxembourg.

Moss, B. (1998) Ecology of fresh waters: Man and medium, Past to future. Edition 3. Blackwell Science Cambridge UK.

Natural England (2009) TIN051: Bats and onshore wind turbines. Online: http://www.naturalengland.org.uk/ourwork/regulation/wildlife/advice/advisoryleaflets.aspx. [Accessed: 05/01/2012]

National Trust (2005) Shifting shores. Online: http://www.nationaltrust.org.uk/main/w-shifting_shores.pdf [Accessed: 12/01/2012]

Nayak, D.R., Miller, D., Nolan, A., Smith, P., and Smith, J. (2008) Calculating carbon savings from wind farms on Scottish peat lands: A new approach. Scottish Government. Online: http://www.scotland.gov.uk/Resource/Doc/229725/0062213.pdf. [Accessed: 05/01/2012]

Nera and Accent (2007) Report on The Benefits of Water Framework Directive Programmes of Measures in England and Wales. Final Report to Defra.

Nicholls, R.J., Cooper, N.J., Townend, I.H. (2007) The management of coastal flooding and erosion, In: Thorne, C.R., Evans, E.P., and Penning-Rowsell, E.C. (Eds.), Future flood and coastal erosion risks, Thomas Telford, London, pp. 392-413.

Nickus, U. (2007) Water problems at the European scale – The EURO-LIMPACS project. Man and Environment, 3, The Water Balance of the Alps.

Nickus, U., Bishop, K., Erlandsson, M., Evans, C.D., Forsius, M., Laudon, H., Livingstone, D.M., Monteith, D. and Thies, H. (2009).Chapter 3: Direct impacts of climate change on freshwater ecosystems. In: Kernan, M., Battarbee, R.W., Moss, B. (Eds.) Climate change impacts on freshwater ecosystems. Wiley-Blackwell.

Nõges, T., Nõges, P., Jolma, A., and Kaitaranta, J. (2009) Impacts of climate change on physical characteristics of lakes in Europe. JRC Scientific and technical Reports. European Communities.

Orr, J.C., Fabry, V.J., Aumont, O., Bopp, L., Doney, S.C., Feely, R.A., Gnanadesikan, A., Gruber, N., Ishida, A., Joos, F., Key, R.M., Lindsay, K., Maier-Reimer, E., Matear, R., Monfray, P., Mouchet, A., Najjar, R.G., Plattner, G.K., Rodgers, K.B., Sabine, C.L., Sarmiento, J.L., Schlitzer, R., Slater, R.D., Totterdell, I.J., Weirig, M.F., Yamanaka, Y., and Yool, A. (2005) Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature 437, 681-686.

Page 214: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

186 Biodiversity Sector Report

Orr, H.G., Wilby, R.L., Hedger, M.M., and Brown, I. (2008) Climate change in the uplands: a UK perspective on safeguarding regulatory ecosystem services. Climate Research 37:77–98.

OSPAR (2010) Quality status report 2010. OSPAR Commission, London.176pp.

Parish, F., Sirin, A., Charman, D., Joosten, H., Minayeva, T. and Silvius, M. (eds.) (2008) Assessment on peatlands, biodiversity and climate change. Global Environment Centre (Kuala Lumpur) and Wetlands International (Wageningen).

Peach, W.J., Robinson, R.A., and Murray, K.A. (2004) Demographic and environmental causes of the decline of rural song thrushes Turdus philomelos in lowland Britain. Ibis 146:50–59.

Pearce-Higgins, J.W. (2010) Using diet to assess the sensitivity of northern and upland birds to climate change. Climate Research 45:119–130.

Pearce-Higgins, J.W., Stephen, L., Langston, R.H.W., Bainbridge, I.P. and Bullman, R. (2009) The distribution of breeding birds around upland wind farms. Journal of Applied Ecology 46: 1323-1331.

Perry, A.L., Low, P.J., Ellis, J.R., and. Reynolds, J.D. (2005) Climate change and distribution shifts in marine fishes. Science, 308: 1912-1915.

Peterken, G.F. and Mountford, E.P. (1996) Effects of drought on beech in Lady Park Wood, an unmanaged mixed deciduous woodland. Forestry 69 (2): 125-136.

Pethick, J. (1999) Future sea-level changes in Scotland: options for coastal management. In: Baxter, J., Duncan, K., Atkins, S. and Lees, G., Editors, 1999. Scotland's Living Coastline, HMSO, Norwich, pp. 45–62.

Porter, J.H., Parry, M.L., and Carter, T.R. (1991) The potential effects of climatic change on agricultural insect pests. Agricultural and Forest Meteorology, 57, 221-240.

Pye, K., and Saye, S. (2005) The geomorphological response of Welsh sand dunes to sea-level rise over the next 100 years and the management implications for SAC and SSSI sites. Countryside Council of Wales Contract Science Report No 670.

Read, D., Freer-Smith, P. H., Morison, J., Hanley, N., West, C., Snowdon, P. & (Eds.), M. B. (2009) Combating Climate Change - A Role for UK Forests. An assessment of the potential of the UK's trees and woodlands to mitigate and adapt to climate change. Edinburgh: The Stationery Office.

Rees, S., Angus, S., Rhind, P., Doody, J.P. (2010) Coastal margin habitats. MCCIP annual report card science review 2010-2011. Online: http://www.mccip.org.uk/arc. [Accessed 05/01/2012]

Rennie, A.F., and Hansom, J.D. (2010) Sea level trend reversal: land uplift outpaced by sea level rise on Scotland's coast. Geomorphology, 125 (1). pp. 193-202.

Robinson, R.A., Crick, H.Q.P., Learmonth, J.A., Maclean, I.M.D., Thomas, C.D., Bairlein, F., Forchhammer, M.C., Francis, C.M., Gill, J.A., Godley, B.J., Harwood, J., Hays, G.C., Huntley, B., Hutson, A.M., Pierce, G.J., Rehfisch, M.M., Sims, D.W., Santos, M.B., Sparks, T.H., Stroud, D.A., and Visser, M.E. (2009) Travelling through a warming world - climate change and migratory species. Endangered Species Research 7: 87-99.

RSPB (2010) The Local Value of Seabirds: Estimating spending by visitors to RSPB coastal reserves and associated local economic impact attributable to seabird. The RSPB, Sandy, UK.

Page 215: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 187

Ruiz, G.M., Carlton, J.T., Grosholz, E.D., and Hines, A.H. (1997) Global invasions of marine and estuarine habitats by non-indigenous species: mechanisms, extent and consequences. American Zoologist, 37, 621-632.

Ruiz, G.M., Fofonoff, P.W., Carlton J.T., Who, M.J., and Hines A.H. (2000) Invasion of coastal marine communiities in North America: apparent patterns, processes and biases. Annual Review of Ecology and Systematics, 31, 481 - 531.

Sabine, C.L., Feely, R.A., Gruber, N., Key, R.M., Lee, K., Bullister, J.L., Wanninkhof, R., Wong, C.S., Wallace, D.W.R., Tilbrook, B., Millero, F.J., Peng, T.H., Kozyr, A., Ono, T., and Rios, A.F. (2004). 'The Oceanic Sink for Anthropogenic CO2', Science, 305 (5682), 367-371.

SAC (Scottish Agricultural College) (2008) Farming’s retreat from the hills. Online: http://www.sac.ac.uk/ruralpolicycentre/publs/researchreports. [Accessed: 05/01/2012]

Sagoff, M. (2011) The quantification and valuation of ecosystem services. Ecological Economics, 70, 497–502.

Saino, N., Ambrosini, R., Rubolini, D., von Hardenberg, J., Provenzale, A., Hüppop, K., Hüppop, O., Lehikoinen, A., Lehikoinen, E., Rainio, K., Romano, M., Sokolov, L. (2010) Climate warming, ecological mismatch at arrival and population decline in migratory birds. Proceedings Of The Royal Society Series B-Biological Sciences. (published online). DOI: 10.1098/rspb.2010.1778.

Samaritani, E., Siegenthaler, A., Yli-Petäys, M., Buttler, A., Christin, P.A., Mitchell, E.A.D. (2010) Seasonal net ecosystem carbon exchange of a regenerating cutover bog in the Swiss Jura Mountains. Restoration Ecology. In press. doi: 10.1111/j.1526-100X.2010.00662.x.

Sansford, C., Inman, A., Baker, R., Brasier, C., Frankel, S., De Gruyter, J., Husson, C., Kehlenbeck, H., Kessel, G., Moralejo, E., Steeghs, M., Webber, J., and Werres, S. (2009) Report on the risk of entry, establishment, spread and socio-economic loss and environmental impact and the appropriate level of management for Phytophthora ramorum for the EU. http://rapra.csl.gov.uk/: EU Sixth Framework Project RAPRA.

Schils, R., Kuikman, P., Liski, J., van Oijen, M., Smith, P., Webb, J., Alm, J., Somogyi, Z., van den Akker, J., Billett, M., Emmett, B., Evans, C., Lindner, M., Palosuo, T., Bellamy, P., Alm, J., Jandl, R., and Hiederer, R. (2008) CLIMSOIL: Review Of Existing Information On The Interrelations Between Soil And Climate Change. Final Report to EU.

Scottish Government (2011) Scotland's marine atlas: Information for the National Marine Plan. Online: http://www.scotland.gov.uk/Publications/2011/03/16182005/0. [Accessed: 05/01/2012]

Shepherd, D., Burgess, D., Jickells, T., Andrews, J., Cave, R., Turner, R.K., Aldridge, J., Parker, E.R., and Young, E. (2007) Modelling the effects and economics of managed realignment on the cycling and storage of nutrients, carbon and sediments in the Blackwater estuary UK. Estuarine, Coastal and Shelf Science, 73, pp. 355-367.

Smith, P. (Ed.) (2011) Regulating Services. Chapter 14. UK National Ecosystem Assessment Technical Report. Chapter 14. UNEP-WCMC.

Smith, P., Chapman, S.J., Scott, W.A., Black, H.I.J., Wattenbach, M., Milne, R., Campbell, C.D., Lilly, A., Ostle, N., Levy, P.E., Lumsdon, D.G., Millard, P., Towers, W., Zaehle, S., and Smith, J.U. (2007a). Climate change cannot be entirely responsible for soil carbon loss observed in England and Wales, 1978-2003. Global Change Biology 13, pp 2605-2609.

Page 216: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

188 Biodiversity Sector Report

Smith, P., Martino, D., Cai, Z., Gwary, D., Janzen, H.H., Kumar, P., McCarl, B., Ogle, S., O'Mara, F., Rice, C., Scholes, R.J., Sirotenko, O., Howden, M., McAllister, T., Pan, G., Romanenkov, V., Schneider, U. and Towprayoon, S. (2007b) Policy and technological constraints to implementation of greenhouse gas mitigation options in agriculture. Agriculture, Ecosystems and Environment 118, pp 6-28.

Smithers, R.J., Cowan, C., Harley, M., Hopkins, J.J., Pontier, H., and Watts, O. (2008) England biodiversity strategy - climate change adaptation principles: Conserving biodiversity in a changing climate. Defra, London.

SNH (Scottish Natural Heritage) (2000) Windfarms and Birds: Calculating a theoretical collision risk assuming no avoiding action. Online: http://www.snh.org.uk/pdfs/strategy/renewable/COLLIS.pdf [Accessed: 05/01/2012]

Sparks, T., and Tryjanowski, P. (2007) Patterns of spring arrival dates differ in two hirundines, Climate Research 35:159-164.

Sparks, T.H., Collinson, N., Crick, H., Croxton, P., Edwards, M., Huber, K., Jenkins, D., Johns, D., Last, F., Maberly, S., Marquiss, M., Pickup, J., Roy, D., Sims, D., Shaw, D., Turner, A., Watson, A., Woiwod, I., and Woodbridge, K. (2006) Natural Heritage Trends of Scotland: phenological indicators of climate change. Scottish Natural Heritage Commissioned Report No. 167 (ROAME No. F01NB01).

Stafford-Smith, M., Horrocks, L., Harvey, A. and Hamilton, C. (2011) Rethinking Adaptation for a 4°C World. Philosophical Transactions Royal Society, 369, 196-216.

Steffen, W., Burbidge, A.A., Hughes, L., Kitching, R., Lindenmayer, D. Musgrave, W., Stafford Smith, M. and Werner, P.A. (2009) Australia’s biodiversity and climate change: a strategic assessment of the vulnerability of Australia’s biodiversity to climate change. A report to the Natural Resource Management Ministerial Council commissioned by the Australian Government. CSIRO Publishing.

Stott, P.A., Stone, D.A., and Allen. M.R. (2004) Human contribution to the European heatwave of 2003. Nature, 432, 610-614. Corrigem. Nature, 436, 1200.

Stribley, G.H. (2005) Decline in the health of beech (Fagus sylvatica L.) trees in southern England monitored from 1989 to 2002. Quarterly Journal of Forestry 99(3): 193-200.

Swansburg, E., Chaput, G., Moore, D., Caissie, D., and El-Jabi, N. (2002). Size variability of juvenile Atlantic salmon: links to environmental conditions. Journal of Fish Biology, 61, 661–683.

Sytsma, M.D., and Anderson, L.W.J. (1993) Biomass, nitrogen, and phosphorus allocation in parrotsfeather (Myriophyllum aquaticum). Journal of Aquatic Plant Management 31:244-248.

Tait, J., Meagher, L., Lyall, C., and Suk, J. (2006) Foresight. Infectious diseases: Preparing for the future. Risk analysis. Office of Science and Innovation, London, pp 47.

Taylor, J.A., Murdock, A.P., Pontee, N.I. (2004) A macroscale analysis of coastal steepening around the coast of England and Wales, The Geographical Journal, 170(3), 179-188.

Teasdale, P., Collins, P.E.F., Firth, C.R. and Cundy, A.B. (2011) Recent estuarine sedimentation rates from shallow inter-tidal environments in Western Scotland: implications for future sea-level trends and coastal wetland development Quaternary Science Reviews, 30 (1-2). pp. 109-129.

Page 217: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 189

TEEB (2010) The Economics of Ecosystems and Biodiversity: Ecological and Economic Foundations. Earthscan, London.

TEEB (2008) The Economics of Ecosystems and Biodiversity. Interim report. Earthscan, London.

Thackeray, S.J., Sparks, T.H., Frederiksen, M., Burthe, S., Bacon, P., Bell, J.R., Botham, M.S., Brereton, T.M., Bright, P.W., Carvalho, L., Clutton-Brock, T., Dawson, A., Edwards, M., Elliott, J.M., Harrington, R., Johns, D., Jones, I.D., Jones, J.T., Leech, D.I., Roy, D.B., Scott, W.A., Smith, M., Smithers, R.J., Winfield, I.J., and Wanless, S. (2010) Trophic level asynchrony in rates of phenological change for marine, freshwater and terrestrial environments. Global Change Biology. 16:12, 3304–3013.

Thomas, C.D., Hill, J.K., Anderson, B.J., Bailey, S., Beale, C.M., Bradbury, R.B., Bulman, C.R., Crick, H.Q.P., Eigenbrod, F., Griffiths, H.M., Kunin, W.E., Oliver, T.M., Walmsley, C.A., Watts, K., Worsfold, T., and Yardley, T. (2010) A framework for assessing threats and benefits to species responding to climate change. Methods in Ecology and Evolution, 2, 125-142.

Thomas, J.A., Bourn, N.A.D., Clarke, R.T., Stewart, K.E., Simcox, D.J., Pearman, G.S., Curtis, R., and Goodger, B. (2001) The quality and isolation of habitat patches both determine where butterflies persist in fragmented landscapes. Proceedings of the Royal Society of London. Series B, Biological Sciences, 268, 1791–1796.

Thomas, S. (2010) Horizon-scanning for invasive non-native plants in Great Britain. Natural England Commissioned Reports, Number 053.

Tooley, P.W., Browning, M., and Berner, D. (2008) Recovery of Phytophthora ramorum following exposure to temperature extremes. Plant Disease, 92, 431-437.

Tooley, P.W., Browning, M., Kyde, K.L., and Berner, D. (2009). Effect of temperature and moisture period on infection of rhododendron 'Cunningham's White' by Phytophthora ramorum. Phytopathology, 99, 1045-1052.

Towers, W., Grieve, I.C., Hudson, G., Campbell, C.D., Lilly, A., Davidson, D.A., Bacon, J.R., Langan, S.J., and Hopkins D.W, (2006) Scotland's soil resource - current state and threats. Report to Scottish Executive. Online: http://www.scotland.gov.uk/Publications/2006/09/21115639/0. [Accessed: 05/01/2012]

Townend, I.H., Scott, C.R. and Dixon, M. (2010) Managed realignment: A coastal flood management strategy, In: Pender G., and Faulkner H. (Eds.), Flood Risk Science and Management, Blackwell Publishing Ltd, Oxford, pp. 60-86.

Trivedi, M.R., Morecroft, M.D., Berry, P.M., and Dawson, T.P. (2008). Potential effects of climate change on plant communities in three montane nature reserves in Scotland, UK. Biological Conservation, 141 (6).

Turner, R.K., J. Paavola, P. Cooper, S. Farber, V. Jessamy and S. Georgiou (2003) Valuing nature: lessons learned and future research directions. Ecological Economics 46, 493–510.

Turner, R.K., Burgess, D., Hadley, D., Coombes, E.G. and Jackson, N. (2007) A cost-benefit appraisal of coastal managed realignment policy. Global Environmental Change, 17, 397–407.

UKCIP (2003) Climate adaptation: Risk, uncertainty and decision making. Willows, R., and Connell, R., UK Climate Impacts Programme, Oxford, UK.

UK National Ecosystem Assessment (2011) The UK National Ecosystem Assessment: synthesis of the key findings. UNEP-WCMC, Cambridge.

Page 218: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

190 Biodiversity Sector Report

Varnham, K. (2006) Non-native species in UK overseas territories: A review. JNCC Report No.372. Joint Nature Conservation Committee, Peterborough.

Visser, M.E. (2008) Keeping up with a warming world; assessing the rate of adaptation to climate change. Proceedings of the Royal Society Series B, 275, 649–659.

Visser, M.E., Holleman, L.J.M., and Gienapp, P. (2006) Shifts in caterpillar biomass phenology due to climate change and its impact on the breeding biology of an insectivorous bird. Oecologia 147:164–172.

Visser, M.E., Perdeck, A.C., van Balen, J.H., and Both, C. (2009) Climate change leads to decreasing bird migration distances. Global Change Biology 15, 1859–1865.

Walmsley, C.A., Smithers, R.J., Berry, P.M., Harley, M., Stevenson, M.J., and Catchpole, R. (eds.) (2007) MONARCH - Modelling Natural Resource Responses to Climate Change: A Synthesis for Biodiversity Conservation. UKCIP, Oxford.

Warren, M.S., Hill, J.K., Thomas, J.A., Asher, J., Fox, R., Huntley, B., Roy, D.B., Telfer, M.G., Jeffcoate, S., Harding, P., Jeffcoate, G., Willis, S.G., Greatorex-Davies, J.N., Moss, D., and Thomas, C.D. (2001) Rapid responses of British butterflies to opposing forces of climate and habitat change. Nature: 414: 65-69.

Webb, B.W., and Walsh, A.J. (2004) Changing UK river temperatures and their impact on fish populations. Hydrology: Science and Practice for the 21st Century. Volume II 177-191.

Webb, B.W. (1996) Trends in stream and river temperature. Hydrological Processes, 10, 205-226.

Webb, J., Drewitt, A.L., and Measures, G. (2010) Managing for species: integrating the needs of species into habitat management. Natural England Research Report 024. Natural England.

Wesche, S. (2003) The implications of climate change for the conservation of beech woodlands and associated flora in the UK. English Nature Research Reports 528. English Nature, Peterborough.

Whitehead, P., Wilby, R., Battarbee, R., Kernan, M., and Wade, A. (2009). A review of the potential impacts of climate change on surface water quality. Hydrological Sciences 54(1): 101-123.

Wilby, R.L. (2006) When and where might climate change be detectable in UK river flows? Geophysical Research Letters 33: L19407.

Wilby, R.L., Orr, H.G., Hedger, M., Forrow, D., and Blackmore, M. (2006). Risks posed by climate change to delivery of wWater Framework Directive objectives. Environment International 32:1043–1055.

Wilby, R.L., Orr, H., Watts, G., Battarbee, R.W., Berry, P.M., Chadd, R., Dugdale, S.J, Dunbar, M.J., Elliott, J.A., Extence, C., Hannah, D.M., Holmes, N., Johnson, A.C., Knights, B., Milner, N.J., Ormerod, S.J., Solomon, D., Timlett, R., Whitehead, P.J., and Wood, P.J. (2010) Evidence needed to manage freshwater ecosystems in a changing climate: turning adaptation principles into practice. Science of the Total Environment, 408, 4150-4164.

Willis, K.J., and Bhagwat, S.A. (2009) Biodiversity and climate change. Science, 326(5954): 806-807.

Wilson, S.M. (2006) The european beech (Fagus sylvatica L.) in Scotland: History, distribution and ecological potential. Scottish Forestry Vol. 60 No. 4 pp. 4-12.

Page 219: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 191

Winfield, I.J., Fletcher, J.M., and James, J.B. (2008) A review of recent research and translocation activities concerned with the gwyniad of Llyn Tegid. Draft Final Report. Centre for Ecology and Hydrology.

Winfield, I.J., Hateley, J., Fletcher, J.M., James, J.B., Bean, C.W., and Clabburn, P. (2010) Population trends of Arctic charr (Salvelinus alpinus) in the UK: assessing the evidence for a widespread decline in response to climate change. Hydrobiologia.

Wormworth, J., and Mallon, K. (2006) Bird species and climate change. Climate Risk/ World Wildlife Fund.

Worrall, F., Evans, M.G., Bonn, A., Reed, M.S., Chapman, D., and Holden, J. (2009) Can carbon offsetting pay for upland ecological restoration? Science of The Total Environment 408(1), pp. 26-36.

WRC (1999) Potential costs and benefits of implementing the proposed water resources framework directive. Final report to the Department of the Environment, Transport and the Regions (report no. DETR 4477/5)

Wright, A., and Woodhall, J. (2006) T5.11: Review: Case study of phytophthora ramorum, an emerging plant pathogen in cultivated and wild environments. Office of Science and Innovation.

Page 220: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

192 Biodiversity Sector Report

Page 221: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 193

Appendices

Page 222: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

194 Biodiversity Sector Report

Page 223: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 195

Appendix 1 Acknowledgements

This report incorporate inputs from many individuals at expert workshops. The following people have contributed to this work.

Biodiversity and ecosystem services expert group

Name Organisation

Pam Berry Environmental Change Institute, University of Oxford

Iain Brown The Macaulay Land Use Research Institute

Mary Christie Scottish Natural Heritage

Humphrey Crick Natural England

Fiona Mulholland Department of the Environment, Northern Ireland

Deborah Proctor Joint Nature Conservation Committee

Clive Walmsley Countryside Council for Wales

Biodiversity and ecosystem services sector workshop, 11th May 2010

Name Organisation

Tim Ashelford Government Office for Yorkshire and the Humber

Claire Barnett Entec UK Ltd

Pam Berry Environmental Change Institute, University of Oxford

Iain Brown The Macaulay Land Use Research Institute

Mary Christie Scottish Natural Heritage

Humphrey Crick Natural England

Mark Diamond Environment Agency

Willie Duncan SEPA

Richard Findon Defra

Michael Graham English National Park Authorities Association

Alice Hardiman RSPB

Mike Harley AEA

Brian Huntley University of Durham

Fiona Mulholland Dept of the Environment – Northern Ireland

Paula Orr CEP

James Pearce-Higgins British Trust for Ornithology

Deborah Proctor Joint Nature Conservation Committee

Chris Spray Scottish Alliance for Geosciences, Environment and Society (SAGES)

Clive Walmsley Countryside Council for Wales

Page 224: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

196 Biodiversity Sector Report

Additional consultees

Name Organisation

Richard Betts Met Office

Julia McMorrow University of Manchester

James Pearce-Higgins British Trust of Ornithology

Rob Gazzard Forestry Research

Harriet Orr Environment Agency

Jo Clarke University of Reading

Jo House University of Bristol

Tim Sparks Independent consultant

Page 225: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 197

Appendix 2 Tier 1 impacts scores

The following tables show the scores that were applied to each of the impacts to determine the Tier 2 list.

For this sector a slight adjustment was made to the generic scoring approach set out in the methodology report (Defra, 2010a) based upon the definition of environmental, social and economic criteria. The consequences for biodiversity were used to define the environmental criteria (Table 1), whilst the consequences for ecosystem services were used to jointly define the social and economic criteria (Table 2). Therefore, collectively those impacts that were identified as having the potential to cause major changes in biodiversity AND to modify the benefits that humans obtain from the natural environment scored highest for ‘consequences’. The translation of this impact potential into actual consequences was scored by the likelihood criteria (higher likelihood equates to higher score) and then further scored based upon urgency criteria (higher urgency equates to higher score) (Table 3).

Page 226: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

198

Table 1: Biodiversity scoring Potential loss/gain of BAP habitats/species (3-High, 2-Med, 1-Low)

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins cultivated farmland semi-nat grasslands woodlands montane, heath and bog

freshwater and wetlands

urban Biodiversity Score

RANGE SHIFTS 1 species unable to track changing climate space

T,P,PET Medium 2 2 2 2 3 2 1 3

2 species unable to find suitable microclimate

T,P, PET,wind

Medium 1 2 2 2 2 2 1 2

3 opportunities for new (priority) species

T,P,PET Low 2 2 2 2 2 2 2 2

SEASONAL SHIFTS AND CHANGES IN PHENOLOGY

4

asynchrony between a species breeding cycle and its food supply

T Medium 2 2 2 2 2 2 2 2

5 change in life cycles (esp. insects)

T Medium 2 2 2 2 2 2 2 2

6 changes in species migration patterns

T Medium 3 1 2 2 2 2 1 3

CHANGES IN PESTS AND DISEASES AND INVASIVE NON-NATIVE SPECIES

7 increased risks from pests

Tmin,P Medium 3 2 2 2 2 3 2 3

8 increased risks from diseases

Tmin,P Medium 3 2 2 2 2 3 2 3

9 increased risk from novel pathogens

Tmin,P Low 3 3 3 3 3 3 3 3

CHANGES IN INTERACTIONS AND COMMUNITY STRUCTURE

10 generalists favoured over specialists (eg. ruderal spp.)

CO2, T, P Medium 2 2 2 2 2 2 2 2

11

changing competition between C3 and C4 photosynthesis plants

CO2, T, P Low 1 1 1 1 1 1 1 1

12 changing interactions due to differences in growth/survival rates

CO2, T, P Low 2 1 2 2 2 2 1 2

Page 227: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

199

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins cultivated farmland semi-nat grasslands woodlands montane, heath and bog

freshwater and wetlands

urban Biodiversity Score

13 changing interactions between trophic levels

CO2, T, P Low 3 2 2 3 2 2 2 3

14 changes in genetic diversity

CO2, T, P Medium 2 2 2 2 2 2 2 2

15 impacts of changing nutrient supply

CO2, T, P Low 3 2 2 2 3 3 1 3

GEOMORPHOLOGICAL AND HYDRO-ECOLOGICAL HABITAT CHANGE

16 coastal evolution impacts on intertidal, grazing marsh etc.

SLR, wave High 3 2 1 1 0 2 1 3

17 floodplain evolution P High 0 0 1 1 1 2 0 2

18

increased water temperature.and stratification of water bodies

T Medium 0 0 0 0 0 2 0 2

19 less snow cover T High 0 0 0 0 2 1 0 2

20 high flow impacts on spawning beds

P High 0 0 0 0 0 2 0 2

21 low flow impacts via BOD

T,P High 0 0 0 0 0 2 0 2

22 saline intrusion SLR Medium 0 1 0 0 0 1 0 1

23 increased soil moisture deficits and drying

T,P,PET High 2 1 3 2 3 2 1 3

24 increased soil erosion T,P,PET High 0 1 1 1 2 1 1 2

25 increased waterlogging

P High 0 1 2 2 1 0 1 2

HABITAT DISTURBANCE BY EXTREME EVENTS

26 windthrow during storms

Wind High 0 0 0 1 0 0 0 1

27 major coastal flood/reconfiguration

Wave, surge

Medium 3 2 1 0 0 2 0 3

Page 228: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

200

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins cultivated farmland semi-nat grasslands woodlands montane, heath and bog

freshwater and wetlands

urban Biodiversity Score

28 major fluvial flood P High 0 1 1 1 0 0 1 1

29 major drought events

P, T High 2 2 2 2 2 2 1 2

30 loss of niche space T,P, PET,Rhu

Medium 2 1 2 1 2 2 1 2

31 Increased risk from wildfire

T,P,PET Medium 0 1 2 2 2 0 1 2

CHANGES TO ECOSYSTEM PROCESSES/FUNCTIONING

32 changes in primary productivity

T,P, CO2 Low/Medium 2 1 2 1 2 2 1 2

33 changes in soil organic carbon

T,P, CO2 Low 2 1 2 2 3 2 1 3

34 faster decomposition and nutrient cycling

T,P, CO2 Medium 1 1 1 1 1 2 1 2

35 changes in soil microbial activity

T,P, CO2 Low 1 1 1 1 1 1 1 1

INDIRECT EFFECTS VIA LAND USE CHANGE

36 agricultural intensification (i.e. human use of NPP)

T,P,PET High 1 3 2 2 1 2 1 3

37 agricultural abandonment

T,P,PET Medium 0 1 2 1 2 1 0 2

38 increased water pollution risk and eutrophication

P,t High 3 0 0 0 0 3 0 3

39 impacts of atmospheric deposition (e.g. N, SO2, O3)

T,P Low 2 1 2 1 2 2 1 2

40 climate mitigation measures (positive/negative)

T Low 2 1 1 2 2 2 2 2

41 increase in flood defence structures

P High 3 0 0 0 0 2 0 3

42 increased societal water demand

T,P, PET Medium 1 1 2 2 2 3 2 3

Page 229: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

201

Table 2: Ecosystem services scoring

Ecosystem sensitivity (keystone spp. and habitat integrity most important) (3-High, 2-Med, 1-Low)

Consequences for Ecosystem services

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins

cultivated farmland

semi-nat grasslands

woodlands montane, heath and bog

freshwater and wetlands

urban Provisioning Cultural Regulating

Total score

RANGE SHIFTS 1 species unable to track changing climate space

T,P,PET Medium 2 1 2 2 2 2 1 1 1 2 1.33

2 species unable to find suitable microclimate

T,P, PET,wind

Medium 1 1 1 1 1 1 1 1 1 1 1.00

3 opportunities for new (priority) species

T,P,PET Low 2 2 2 2 2 2 2 1 1 2 1.33

SEASONAL SHIFTS AND CHANGES IN PHENOLOGY

4 asynchrony between a species breeding cycle and its food supply

T Medium 2 2 2 2 2 2 2 1 2 1 1.33

5 change in life cycles (esp. insects)

T Medium 2 2 2 2 2 2 2 1 1 2 1.33

6 changes in species migration patterns

T Medium 2 1 2 2 2 2 1 1 2 2 1.67

CHANGES IN PESTS AND DISEASES AND INVASIVE NON-NATIVE SPECIES

7 increased risks from pests

Tmin,P Medium 3 2 2 2 2 3 2 2 1 2 1.67

8 increased risks from diseases

Tmin,P Medium 3 2 2 2 2 3 2 2 1 2 1.67

9 increased risk from novel pathogens

Tmin,P Low 3 3 3 3 3 3 3 2 1 2 1.67

CHANGES IN INTERACTIONS AND COMMUNITY STRUCTURE

10 generalists favoured over specialists (eg. ruderal spp.)

CO2, T, P Medium 2 2 2 2 2 2 2 1 2 2 1.67

11 changing competition between C3 and C4 photosynthesis plants

CO2, T, P Low 1 1 1 1 1 1 1 1 1 1 1.00

12 changing interactions due to differences in growth/survival rates

CO2, T, P Low 2 1 2 2 2 2 1 1 1 2 1.33

13 changing interactions between trophic levels

CO2, T, P Low 3 2 2 2 3 3 2 1 1 2 1.33

14 changes in genetic diversity

CO2, T, P Low 2 2 2 2 2 2 2 2 1 2 1.67

Page 230: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

202

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins

cultivated farmland

semi-nat grasslands

woodlands montane, heath and bog

freshwater and wetlands

urban Provisioning Cultural Regulating

Total score

15 impacts of changing nutrient supply

CO2, T, P Low 3 2 3 2 3 3 1 2 1 2 1.67

GEOMORPHOLOGICAL AND HYDRO-ECOLOGICAL HABITAT CHANGE

16 coastal evolution impacts on intertidal, grazing marsh etc.

SLR, wave High 3 2 1 1 0 2 1 1 2 3 2.00

17 floodplain evolution P High 0 0 1 1 1 2 0 1 2 2 1.67

18 increased water temperature and stratification of water bodies

T Medium 0 0 0 0 0 2 0 2 1 2 1.67

19 less snow cover T High 0 0 0 0 2 1 0 1 1 2 1.33

20 high flow impacts on spawning beds

P High 0 0 0 0 0 2 0 2 1 1 1.33

21 low flow impacts via BOD

T,P High 0 0 0 0 0 2 0 2 1 2 1.67

22 saline intrusion SLR Medium 0 1 0 0 0 1 0 1 1 1 1.00

23 increased soil moisture deficits and drying

T,P,PET High 2 1 3 2 3 2 1 2 2 3 2.33

24 increased soil erosion T,P,PET High 0 1 1 1 2 1 1 1 2 1 1.33

25 increased waterlogging

P High 0 1 2 2 1 0 1 2 2 1 1.67

HABITAT DISTURBANCE BY EXTREME EVENTS

26 windthrow during storms

Wind High 0 0 0 1 0 0 0 1 2 2 1.67

27 major coastal flood/reconfiguration

Wave, surge Medium 3 2 1 0 0 2 0 2 2 3 2.33

28 major fluvial flood P High 0 1 1 1 0 0 1 2 2 2 2.00

29 major drought events

P, T High 2 2 2 2 2 2 1 2 2 3 2.33

30 loss of niche space T,P, PET,Rhu Medium 2 1 2 1 2 2 1 1 1 1 1.00

31 Increased risk from wildfire

T,P,PET Medium 0 1 2 2 2 0 1 1 2 2 1.67

CHANGES TO ECOSYSTEM PROCESSES/FUNCTIONING

32 changes in primary productivity

T,P, CO2 Low/Medium 2 1 2 2 2 2 1 2 1 2 1.67

33 changes in soil organic carbon

T,P, CO2 Low 2 1 2 2 3 2 1 2 2 3 2.33

34 faster decomposition and nutrient cycling

T,P, CO2 Medium 1 2 2 2 1 2 1 2 1 3 2.00

Page 231: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

203

Risk Type # Specific Risks Climate parameter

Evidence Base

coastal margins

cultivated farmland

semi-nat grasslands

woodlands montane, heath and bog

freshwater and wetlands

urban Provisioning Cultural Regulating

Total score

35 changes in soil microbial activity

T,P, CO2 Low 1 1 1 1 1 1 1 1 1 2 1.33

INDIRECT EFFECTS VIA LAND USE CHANGE

36 agricultural intensification (i.e. human use of NPP)

T,P,PET High 1 3 2 2 1 2 1 2 1 3 2.00

37 agricultural abandonment

T,P,PET Medium 0 1 2 1 2 1 0 1 2 1 1.33

38 increased water pollution risk and eutrophication

P,t High 3 0 0 0 0 3 0 2 2 2 2.00

39

impacts of atmospheric deposition (e.g. N, SO2, O3)

T,P Low 2 1 2 1 2 2 1 1 1 2 1.33

40 climate mitigation measures (positive/negative)

T Low 2 1 1 2 2 2 2 1 2 3 2.00

41 increase in flood defence structures

P High 3 0 0 0 0 2 0 1 2 2 1.67

42 increased societal water demand

T,P, PET Medium 1 1 2 2 2 3 2 2 2 2 2.00

Page 232: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

204

Table 3: Overall scores and ranking

# Risks Biodiversity Score

ESS Score Likelihood Score

Urgency Score Total Score Ranking

23 increased soil moisture deficits and drying 3 2.33 3 3 85.2 1

16 coastal evolution impacts on intertidal, grazing marsh etc.

3 2.00 3 3 77.8 2

7 increased risks from pests 3 1.67 3 3 70.4 3

8 increased risks from diseases 3 1.67 3 3 70.4 3

40 climate mitigation measures (positive/negative) 2 2.00 3 3 66.7 5

1 species unable to track changing climate space 3 1.33 3 3 63.0 6

27 major coastal flood/reconfiguration 3 2.33 2 3 56.8 7

33 changes in soil organic carbon 3 2.33 3 2 56.8 7

38 increased water pollution risk and eutrophication 3 2.00 2 3 51.9 9

42 increased societal water demand 3 2.00 2 3 51.9 9

36 agricultural intensification (i.e. human use of NPP) 3 2.00 3 2 51.9 9

29 major drought events 2 2.33 2 3 49.4 12

41 increase in flood defence structures 3 1.67 2 3 46.9 13

6 changes in species migration patterns 3 1.67 3 2 46.9 14

10 generalists favoured over specialists (eg. ruderal spp.) 2 1.67 3 2 39.5 15

18 increased water temperature.and stratification of water bodies

2 1.67 3 2 39.5 15

4 asynchrony between a species breeding cycle and its food supply

2 1.33 3 2 34.6 17

12 changing interactions due to differences in growth/survival rates

2 1.33 3 2 34.6 17

15 impacts of changing nutrient supply 3 1.67 2 2 31.3 19

2 species unable to find suitable microclimate 2 1.00 3 2 29.6 20

21 low flow impacts via BOD 2 1.67 2 2 26.3 21

Page 233: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

205

# Risks Biodiversity Score

ESS Score Likelihood Score

Urgency Score Total Score Ranking

25 increased waterlogging 2 1.67 2 2 26.3 21

31 Increased risk from wildfire 2 1.67 2 2 26.3 21

28 major fluvial flood 1 2.00 2 2 24.7 24

20 high flow impacts on spawning beds 2 1.33 2 2 23.0 25

24 increased soil erosion 2 1.33 2 2 23.0 25

37 agricultural abandonment 2 1.33 2 2 23.0 25

39 impacts of atmospheric deposition (e.g. N, SO2, O3) 2 1.33 2 2 23.0 25

34 faster decomposition and nutrient cycling 2 2.00 3 1 22.2 29

22 saline intrusion 1 1.00 3 2 22.2 29

30 loss of niche space 2 1.00 2 2 19.8 31

3 opportunities for new (priority) species 2 1.33 3 1 17.3 32

5 change in life cycles (esp. insects) 2 1.33 3 1 17.3 32

19 less snow cover 2 1.33 3 1 17.3 32

13 changing interactions between trophic levels 3 1.33 2 1 14.0 35

14 changes in genetic diversity 2 1.67 2 1 13.2 36

17 floodplain evolution 2 1.67 2 1 13.2 36

35 changes in soil microbial activity 1 1.33 2 1 9.1 38

9 increased risk from novel pathogens 3 1.67 1 1 7.8 39

32 changes in primary productivity 2 0.00 3 1 7.4 40

11 changing competition between C3 and C4 photosynthesis plants

1 1.00 2 1 7.4 40

26 windthrow during storms 1 1.67 1 1 5.3 42

Page 234: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

206

Page 235: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 207

Appendix 3 Application of climate change projections

The full results for the BD2 risk metric with respect to UKCP09 data are presented below.

BD2 – Loss of BAP habitats caused by coastal erosion (to the nearest hectare)

The following table provides summary data on the area (in hectares) of BAP habitat that may be lost due to coastal erosion in England.

The table shows overall totals for the P50 Medium Emissions climate change scenario.

Data have not been calculated for socio economic scenarios, as these focus on growth in population and property numbers.

BAP habitat type 2020s 2050s 2080s

Coastal Floodplain and

Grazing Marsh 7 36 66

Deciduous Woodland 24 115 239

Fen 2 6 7

Purple Moor Grass and

Rush Pasture 1 2 3

Reedbed 2 11 21Saline Lagoon 6 23 41

Total 41 193 377

The following tables provide more detailed data by Area and climate change scenario.

Coastal Floodplain and Grazing Marsh

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 4 4 5 16 20 23 27 38 16 25 31 38 58

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 0 0 0 0 0 0 0 0 0 0

England North West 0 0 0 0 0 0 0 0 0 1 1 1 2

England South East 1 1 1 2 2 3 3 4 3 4 5 6 8

England South West 2 2 2 8 9 10 11 14 25 28 30 32 38

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 0 0 0 0 0 0 0 0 0 0 0 0 0

Wales Wales

Total England 6 7 8 26 32 36 41 57 45 58 66 76 106

Deciduous Woodland

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 2 2 3 7 8 9 10 12 7 11 12 14 19

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 0 0 0 0 0 0 0 0 0 0

England North West 0 0 0 2 3 4 4 6 7 9 9 10 14

England South East 11 12 13 38 50 57 62 76 82 96 103 113 140

England South West 8 9 9 37 40 44 47 55 98 106 110 115 130

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 1 1 1 2 2 2 2 3 4 4 5 5 6

Wales Wales

Total England 22 24 26 87 103 115 125 152 198 225 239 256 309

Fen

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 1 1 1 3 4 4 4 5 3 4 4 5 6

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 0 0 0 0 0 0 0 0 0 0

England North West 0 0 0 0 0 0 0 0 0 0 1 1 1

England South East 0 0 0 0 0 0 0 0 0 0 0 0 0

England South West 0 0 1 1 1 1 2 2 2 3 3 3 3

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 0 0 0 0 0 0 0 0 0 0 0 0 0

Wales Wales

Total England 2 2 2 5 5 6 6 7 6 7 7 8 10

2020s 2050s 2080s

Nation UKCP09 Region

2020s 2050s 2080s

Nation UKCP09 Region

2020s 2050s 2080s

Nation UKCP09 Region

Page 236: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

208 Biodiversity Sector Report

Purple Moor Grass and Rush Pasture

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 0 0 0 1 1 1 1 1 1 1 1 1 1

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 1 1 1 1 1 1 1 1 1 2

England North West 0 0 0 0 0 0 0 0 0 0 0 0 0

England South East 0 0 0 0 0 0 0 0 0 0 0 0 0

England South West 0 0 0 0 0 0 0 0 0 0 0 0 0

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 0 0 0 0 0 0 0 0 0 0 0 0 0

Wales Wales

Total England 1 1 1 2 2 2 2 2 2 2 3 3 3

Reedbed

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 0 0 0 2 3 5 6 9 4 8 10 12 19

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 2 2 2 2 3 3 4 4 4 5

England North West 0 0 0 0 0 0 0 0 0 0 0 0 0

England South East 0 0 0 1 1 1 1 1 1 1 1 1 1

England South West 1 1 1 4 4 4 4 5 7 7 7 7 8

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 0 0 0 0 0 0 0 0 0 0 0 0 0

Wales Wales

Total England 2 2 2 8 10 11 13 17 14 19 21 24 33

Saline Lagoon

Medium Medium Medium Low Low Medium High High Low Low Medium High Highp10 p50 p90 p10 p50 p50 p50 p90 p10 p50 p50 p50 p90

England East Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0

England East of England 2 2 2 5 8 11 16 28 6 15 22 30 51

England London 0 0 0 0 0 0 0 0 0 0 0 0 0

England North East 0 0 0 0 1 1 1 1 1 1 1 1 2

England North West 0 0 0 0 0 0 0 0 0 0 0 0 0

England South East 0 0 0 2 2 2 2 3 5 5 5 5 6

England South West 0 0 0 0 0 0 0 0 1 1 1 1 1

England West Midlands 0 0 0 0 0 0 0 0 0 0 0 0 0England Yorkshire and The Humber 3 3 3 6 8 9 9 10 10 11 12 12 14

Wales Wales

Total England 5 6 6 14 19 23 28 43 22 33 41 50 74

37 41 45 141 171 193 215 277 288 345 377 417 534Overall total

2020s 2050s 2080s

Nation UKCP09 Region

2020s 2050s 2080s

2080s

Nation UKCP09 Region

Nation UKCP09 Region

2020s 2050s

Page 237: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 209

Appendix 4 Climate vulnerability assessment of designated sites in Wales – a summary

A vital aspect of understanding the impact of climate change on the natural environment is identifying where particular vulnerabilities lie. In doing so, actions to reduce vulnerability by increasing the resilience of species and habitats can be prioritised to ecosystems and populations that are most sensitive to change (ADAS, 2010). Calculating where vulnerable areas are is a very complex thing to do, particularly at a national scale, as many different factors drive an area’s vulnerability and these may differ from one area to another.

The Countryside Council for Wales (CCW) commissioned a piece of work to develop the tools for identifying, at a national scale, vulnerabilities to climate change in the protected conservation network71 in order to identify those sites where adapting to and mitigating climate change will be of highest priority (ADAS, 2010). They developed a Climate Vulnerability Index (CVI), based upon sites’ inherent sensitivity to climate change using the vulnerability of the component habitats and species for which sites were notified, along with site factors that affect adaptive capacity and resilience including condition of site features, the extent of other stressors/management issues, and connectivity between the broad habitats within the site and the wider landscape71.

The project covers terrestrial and freshwater Sites of Special Scientific Interest (SSSI) and Natura 2000 protected areas (Special Areas for Conservation (SAC) and Special Protection Areas (SPA))71. The project utilised the following key datasets:

Climate impact-risk assessments for individual habitats and species

CCW ‘Special Sites’ database: features on protected sites, condition assessments and current management issues.

Each site was given a score in accordance with the magnitude of each of the components of vulnerability and these were compiled into the CVI, which was used to rank sites. New information on species and habitat impacts and an increase in the number of sites that are to be assessed means that the results should be considered provisional, but are nevertheless useful in demonstrating the applicability and robustness of the method.

Some of the sites with the highest CVI include those for which climate vulnerability has already been identified by local assessments; Cemlyn Bay tidal lagoon and Berwyn are two such instances. Other sites with a high CVI haven’t previously been thought of as being particularly vulnerable to climate change. For example, Cors Hirdre has a number of medium impact-risk habitat features and several management issues that may be impacted by climate change. Its peatlands also have poor habitat connectivity. These factors contribute to a high CVI score, which is informative for driving further research and management action. A large proportion of the high-ranking CVI sites contain aquatic habitats. This reflects the importance of water management issues within these sites with regard to climate change, and the influence that hydrology has on so many habitats. Overall, the provisional results indicate that more than two-thirds of habitats and species in the conservation network of Wales are at a medium or high risk from climate change (ADAS, 2010).

Figures 1 and 2 show the assessed SSSI and Natura 2000 sites in Wales categorised by their CVI values and Figure 3 shows the Natura 2000 sites whose habitat condition is high risk. This indicates the wide distribution of sites of conservation importance that

71

McCall, R. (2011) Email “Assessing vulnerability - CVI index use in the CCRA” 27th January

Page 238: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

210 Biodiversity Sector Report

are potentially vulnerable to climate change. It should be noted that whilst this type of approach is very useful at the national scale as preliminary overview, it is not a substitute for assessing vulnerability of special sites at the local scale71.

Figure 1. SSSIs in Wales categorised by the CVI (Climate Vulnerability Index), calculated as a simple unweighted geometric mean of the individual vulnerability components. Sites shown as having no data have not yet been assessed.

Page 239: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 211

Figure 2 Natura 2000 sites in Wales categorised by the CVI (Climate Vulnerability Index), calculated as a simple unweighted geometric mean of the individual vulnerability components. Sites shown as having no data have not yet been assessed.

Page 240: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

212 Biodiversity Sector Report

Figure 3 Natura 2000 sites in Wales categorised by the proportion of the site area having at least one feature with a high risk condition status. Sites shown as having no data have not yet been assessed.

Page 241: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 213

Appendix 5 Magnitude, confidence and presentation of results

Table A5.1 defines the magnitude classes used in the assessment. These were used for scoring impacts in the Tier 2 selection process as well as for scoring risk levels for the scorecards presented for each metric in Chapter 4. For the scorecard, the risk/opportunity level relates to the most relevant of the economic/environmental/social criteria.

Page 242: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

214 Biodiversity Sector Report

Table A5.1 Guidance on classification of relative magnitude: qualitative descriptions of high, medium and low classes

Class Economic Environmental Social

Hig

h

Major and recurrent damage to property and infrastructure

Major consequence on regional and national economy

Major cross-sector consequences

Major disruption or loss of national or international transport links

Major loss/gain of employment opportunities

~ £100 million for a single event or per year

Major loss or decline in long-term quality of valued species/habitat/landscape

Major or long-term decline in status/condition of sites of international/national significance

Widespread Failure of ecosystem function or services

Widespread decline in land/water/air quality

Major cross-sector consequences

~ 5000 ha lost/gained

~ 10000 km river water quality affected

Potential for many fatalities or serious harm

Loss or major disruption to utilities (water/gas/electricity)

Major consequences on vulnerable groups

Increase in national health burden

Large reduction in community services

Major damage or loss of cultural assets/high symbolic value

Major role for emergency services

Major impacts on personal security e.g. increased crime

~million affected

~1000s harmed

~100 fatalities

Med

ium

Widespread damage to property and infrastructure

Influence on regional economy

Consequences on operations & service provision initiating contingency plans

Minor disruption of national transport links

Moderate cross-sector consequences

Moderate loss/gain of employment opportunities

~ £10 million per event or year

Important/medium-term consequences on species/habitat/landscape

Medium-term or moderate loss of quality/status of sites of national importance

Regional decline in land/water/air quality

Medium-term or Regional loss/decline in ecosystem services

Moderate cross-sector consequences

~ 500 ha lost/gained

~ 1000 km river water quality affected

Significant numbers affected

Minor disruption to utilities (water/gas/electricity)

Increased inequality, e.g. through rising costs of service provision

Consequence on health burden

Moderate reduction in community services

Moderate increased role for emergency services

Minor impacts on personal security

~ 100s thousands affected, ~100s harmed, ~10 fatalities

Lo

w

Minor or very local consequences

No consequence on national or regional economy

Localised disruption of transport

~ £1 million per event or year

Short-term/reversible effects on species/habitat/landscape or ecosystem services

Localised decline in land/water/air quality

Short-term loss/minor decline in quality/status of designated sites

~ 50 ha of valued habitats damaged/improved

~ 100 km river quality affected

Small numbers affected

Small reduction in community services

Within ‘coping range’

~10s thousands affected

The levels of confidence used by the CCRA can be broadly summarised as follows:

Low - Expert view based on limited information, e.g. anecdotal evidence.

Medium - Estimation of potential impacts or consequences, grounded in theory, using accepted methods and with some agreement across the sector.

Page 243: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

Biodiversity Sector Report 215

High - Reliable analysis and methods, with a strong theoretical basis, subject to peer review and accepted within a sector as 'fit for purpose'.

The lower, central and upper estimates provided in the scorecards relate to the range of the estimated risk or opportunity level. For risk metrics that have been quantified against UKCP09, this range relates to the results that are given for the low emissions, 10% probability level (lower); medium emissions, 50% probability level (central); and high emissions, 90% probability level (upper). For the risk metrics that have been estimated with a more qualitative approach, these estimates cover the range of potential outcomes given the evidence provided.

The CCRA analysis uses three discrete time periods to estimate future risks up to the year 2100: the 2020s (2010 to 2039), 2050s (2040 to 2069) and the 2080s (2070 to 2099).This is consistent with the UKCP09 projections.

Page 244: Climate Change Risk Assessment for the Biodiversity and …randd.defra.gov.uk/Document.aspx?Document=10071_CCRAfortheBi… · climate change risks and these interventions need to

216 Biodiversity Sector Report