World Class Science for the Marine and Freshwater Environment

Dorset and East FLAG area - Aquaculture Mapping Project

Mapping of Areas of Potential Aquaculture within the and Fisheries Local Action Group (FLAG) area - FINAL report

Authors: Simon Kershaw; Richard Heal; Paulette Posen; Ian Tew, Claire Beraud; Keith Jeffery.

Date: January 2021

Please cite this document as: Kershaw, S., Beraud, C., Heal, R., Posen, P., Tew, I. and Jeffery, K. (2021). Mapping of Areas of Potential Aquaculture within the Dorset and East Devon Fisheries Local Action Group (FLAG) area. Cefas Contract C7731, European Maritime and Fisheries Fund (EMFF) grant number ENG3016, pp. 1-205.

Page i

Executive Summary

The global human population looks set to reach 9.8 billion people by 2050 (United Nations, 2017) and our food systems will therefore be under intense pressure to produce animal protein for an increasing population. Food security is an objective of the UK administrations and aquaculture makes an important and growing contribution to this (HM Government, 2011). The UK multi-annual plan (Defra, 2015) highlights the Government’s commitment to encouraging the development of aquaculture within the new Marine Plan areas. This included emphasising to marine planners the importance of aquaculture to the environment and society and strengthening data and evidence locally and nationally for the potential for aquaculture development (marine finfish, shellfish, algae and other organisms). Locally, the Dorset & East Devon Fisheries Local Action Group (FLAG) area was selected because of the importance of fishing and seafood to coastal communities. Developing an increased production of quality local product will help bolster and supply successful local business events such as the Weymouth Seafood Festival.

The purpose of this project is to identify and map areas best suited (and with least conflict) to specific types of sustainable marine aquaculture, within the boundaries of the Dorset and East Devon FLAG area. A detailed literature review has been conducted to understand previous work on identifying and mapping areas with potential for future aquaculture production. A base map was created for the FLAG area extending from Beer Head to Poole Entrance, out to the six nautical mile (6 nm) limit from the UK coast. Baseline data and a comprehensive set of spatial data, from several different sources, was compiled for analysis, processing and mapping of aquaculture suitability extents within the area. Environmental variables underpinning aquaculture suitability were classified in optimal, suboptimal and unsuitable ranges based on published literature. Aquaculture potential was then determined for each species based on an appropriate culture method for the area. Stakeholder consultation workshops (additional to the original project scope) were undertaken to review draft maps of areas with potential for aquaculture within the FLAG district. Feedback from these has been incorporated into both the final maps of aquaculture potential and the report.

Outputs from the project comprise: • this report on the background, approach, methods, limitations, and utility of the mapping resources produced; • the creation of downloadable GIS shapefiles; • an interactive mapping tool resource and • the facility for ongoing user feedback on-line.

Page ii

Building on previous work (Heal & Capuzzo, 2019) in support of national aquaculture modelling, the project has delivered an improved level of regional mapping of aquaculture potential at a larger scale covering the inshore region of the east Devon and coast. It has improved utilisation of information on water quality underpinned by the development of a hydrodynamic computer model for the coastal waters within this FLAG region, with an emphasis on suitability of specific species and appropriate aquaculture systems. Whilst not originally incorporated into the study design, this has proved invaluable for assessing the impact of long sea outfalls on microbiological water quality for bivalve shellfish, in this tidally dynamic and hydrographically complex area. Consideration of factors to support land-based aquaculture utilising natural seawater have also been included in our assessment.

Whilst the project outputs should help target aquaculture opportunities and inform the development of new aquaculture business plans in the event of future planning applications, more detailed site assessment and potential impact assessment would still need to be undertaken. Should funds become available, the next stages for regional or community led investment could include exploring whether permissions and licences could be obtained for certain areas or zones e.g. as part of a fisheries fund project administered by a body such as the Dorset Coast Forum or the Crown Estate. This would enable easier start-up opportunities particularly for small to medium scale businesses which may best support the local economy.

Consideration of carrying capacity, economic potential or zoning for diseases were not included as part of this assessment. However, if pre-licensed areas or zones were to be created, these would need to be assessed and developed along with economic and socio-cultural considerations.

The project has developed a model and tools for detailed mapping of potential aquaculture areas which makes identification of biologically and physically suitable areas possible particularly for new entrants. Additionally, it identifies exclusions with buffer zones for various sorts of aquaculture system due to other human activity or hard constraints. This approach combines the traditional approach of looking for areas of least conflict with an ecosystem and integrated coastal management approach. The underlying data and assumptions, along with the tables and maps generated, can be built on as knowledge and experience develops further.

Currently less than 0.3 % of the sea area in the East Devon and West Dorset FLAG area is used for aquaculture production. Notwithstanding other competing resource and planning constraints and within the resource limitations of the project, on the basis of detailed research from best available information and modelling work reported on here, 68 % of the total sea area within this

Page iii

FLAG sea area has optimal aquaculture potential for aquaculture of one or more species [and ~40 % occurs outside of Marine Protected Areas (MPAs)].

Page iv

Table of Contents

Executive Summary ...... ii Table of Contents ...... v Tables ...... ix Figures ...... x Glossary ...... xiii Abbreviations ...... xv 1 Introduction ...... 1 1.1 Project rationale ...... 1 1.2 Background ...... 1 1.3 Project Aim ...... 5 1.4 Project Scope ...... 5 1.5 Project Benefits ...... 6 2 Marine aquaculture systems by culture groups ...... 7 2.1 Marine Bivalve shellfish ...... 7 Marine Bivalve Shellfish – Bottom Culture ...... 7 Marine Bivalve Shellfish – Bottom Secured Culture ...... 8 Marine Bivalve Shellfish – Suspended Culture ...... 9 Marine Bivalve Shellfish – Land-based Hatchery ...... 10 Marine Bivalve Shellfish – Depuration ...... 11 Marine Bivalve Shellfish – Classification ...... 11 2.2 Marine Crustaceans ...... 12 Marine Crustacean – Hatchery ...... 12 Marine Crustacean – Bottom culture ...... 12 Marine Crustaceans – Suspended Culture ...... 12 2.3 Marine Macro Algae ...... 13 Marine Macro Algae – Bottom Secured ...... 13 Marine Macro Algae – Suspended Culture ...... 14 Marine Macro Algae – Land-based Hatchery ...... 14 2.4 Marine Finfish ...... 15 Floating open pen cages or submersible cages ...... 15 Marine Finfish – Pump-ashore Land-based ...... 15 Marine Finfish – Recirculation System ...... 16 2.5 Marine Multi-Trophic Aquaculture ...... 16 3 Overview of optimal physiological conditions for aquaculture ...... 18 3.1 Sea surface temperature ...... 18 3.2 Salinity of the water ...... 19 3.3 Light availability...... 20 3.4 Phytoplankton (chlorophyll) levels ...... 20 3.5 Nutrient concentration ...... 21 3.6 Dissolved oxygen concentration ...... 21

Page v

3.7 General comments ...... 23 4 Approaches to mapping potential aquaculture areas ...... 24 4.1 Previous approaches ...... 24 4.2 Current approaches ...... 25 5 Data sets, processing and mapping ...... 26 5.1 Mapping existing features ...... 27 Base map ...... 27 Natural habitats ...... 28 Human activities and infrastructure features ...... 29 Regulatory and conservation areas ...... 31 5.2 Other known constraints and data limitations ...... 31 Fishing activity ...... 31 Contaminants ...... 32 Wastewater discharges and long sea outfalls...... 33 Biotoxins ...... 34 Disease ...... 35 5.3 Aesthetic considerations...... 36 6 Deriving aquaculture suitability layers...... 37 6.1 Overview of process to generate aquaculture suitability layers ...... 37 6.2 Provenance of GIS data used to determine aquaculture suitability ...... 39 Sea temperature ...... 39 Salinity ...... 39 Light penetration layer ...... 40 Phytoplankton (chlorophyll) levels ...... 40 Dissolved oxygen concentration ...... 41 Suspended Inorganic Particulate Matter (SPIM) ...... 42 Nutrient concentration (Total oxidised nitrogen) ...... 42 Significant wave height ...... 43 Peak water current ...... 43 Substrate type ...... 43 Site Exposure ...... 43 Water depth (bathymetry) ...... 44 Tides ...... 44 6.3 Aquaculture layers used for the consultation exercise...... 44 7 Stakeholder consultation ...... 45 7.1 Stakeholder consultation – outputs...... 48 7.2 Mapping refinement and revision - post consultation ...... 48 7.3 Application of other constraints to determine potential aquaculture areas ...... 48 8 Results ...... 53 8.1 Potential areas for aquaculture ...... 53 Marine near shore and offshore areas ...... 53 ...... 61

Page vi

Portland Mussel Seed Fishery ...... 62 ...... 63 The Fleet ...... 64 Active management of bivalve mollusc production areas ...... 64 Fishery Orders ...... 64 Cleaner fish ...... 64 Oysters, Mussels and bioremediation ...... 65 Land-based mariculture ...... 65 Aquaculture Park concept ...... 68 8.2 FLAG aquaculture areas, footprints and capacity ...... 68 Current aquaculture within the FLAG area ...... 68 Footprints for example aquaculture installations ...... 69 Statistics on aquaculture and its potential in the FLAG area...... 69 9 Discussion ...... 71 9.1 Areas identified with potential ...... 71 9.2 Management & co-location opportunities ...... 75 Co-location with MPA’s ...... 75 Co-location within military training areas or firing ranges...... 75 Co-location within Statutory instrument ...... 75 Co-location within no take zones...... 75 Integration of fishing and aquaculture ...... 76 9.3 Strategic development opportunities ...... 76 10 Conclusions & Recommendations ...... 77 10.1 Conclusions ...... 77 10.2 Recommendations ...... 78 11 Final Outputs from Project ...... 79 11.1 Project report...... 79 11.2 GIS Shapefiles ...... 79 11.3 Interactive mapping tool ...... 79 11.4 Stakeholder feedback on potential aquaculture areas maps ...... 79 References ...... 80 Acknowledgements ...... 92 Appendix 1: Literature Review ...... 93 A1.1 Data collection ...... 93 A1.2 Data synthesis and mapping ...... 94 A1.3 Natural and anthropogenic environmental considerations and resource pressures 95 A1.4 Spatial modelling and defining potential areas for aquaculture ...... 96 A1.5 Management requirements ...... 101 A1.6 Model and data limitations ...... 102 A1.7 Recommendations ...... 104 Appendix 2: Overview of principal aquaculture production systems ...... 106 A2.1 Bottom culture ...... 106

Page vii

A2.2 Bottom secured culture ...... 106 A2.3 Suspended longline – rope/textile/raft ...... 107 A2.4 Suspended longline – container culture ...... 107 A2.5 Floating/submersible cage culture...... 108 A2.6 Submerged cage culture ...... 108 Appendix 3: Species and culture system threshold values ...... 109 A3.1 Threshold tables...... 110 A3.2 Scientific literature used to determine threshold levels...... 133 Appendix 4: Stakeholder consultation workshops ...... 147 A4.1 Introduction ...... 147 A4.2 Workshop aim ...... 147 A4.3 Workshop tasks...... 148 A4.4 Next steps ...... 161 Appendix 5: Sea outfall modelling ...... 162 A5.1 Overview ...... 162 A5.2 Mesh construction ...... 162 A5.3 Steering file parameters ...... 162 A5.4 Tidal calibration ...... 163 A5.5 Validated model tidal peak velocity and elevation ...... 168 A5.6 Water quality results ...... 170 Appendix 6: Aquaculture Potential Maps...... 176 A6.1 Overview ...... 176 A6.2 Aquaculture Potential: Bottom-secured seaweed ...... 177 A6.3 Aquaculture Potential: Suspended seaweed ...... 178 A6.4 Aquaculture Potential: Open pen finfish ...... 179 A6.5 Aquaculture Potential: Bottom (ranching) and Suspended lobster ...... 180 A6.6 Aquaculture Potential: Bottom culture of bivalves ...... 181 A6.7 Aquaculture Potential: Bottom-secured culture of bivalves ...... 182 A6.8 Aquaculture Potential: Suspended culture of bivalves ...... 183 Appendix 7: Data layers ...... 184 Appendix 8: Interactive mapping tool ...... 185

Page viii

Tables

Table 1: Numbers of bivalve samples for biotoxin analysis by year...... 35 Table 2: Bivalve mollusc production areas - Biotoxin Closures Nos. of days by year...... 35 Table 3: Tide levels and ranges...... 44 Table 4: Potential non-environmental constraints on aquaculture development...... 49 Table 5: List of the constraining factors used to produce the exclusion layer for each species group-culture system...... 52 Table 6: List of Aquaculture potential for FLAG region for Macroalgae...... 53 Table 7: List of Aquaculture potential for FLAG region for Finfish...... 56 Table 8: List of Aquaculture potential for FLAG region for Crustacean...... 57 Table 9: List of Aquaculture potential for FLAG region for Bivalves...... 59 Table 10: Spatial area statistics for the sea area in the E. Devon & W. Dorset FLAG region...... 69 Table A 1: Threshold levels used for Saccharina latissima aquaculture. 110

Table A 2: Threshold levels used for Laminaria digitata aquaculture...... 111 Table A 3: Threshold levels used for Alaria esculenta aquaculture...... 112 Table A 4: Threshold levels used for Palmaria palmata aquaculture...... 113 Table A 5: Threshold levels used for Ulva lactuca aquaculture...... 114 Table A 6: Threshold levels used for Salmo salar aquaculture...... 115 Table A 7: Threshold levels used for Oncorhynchus mykiss aquaculture...... 116 Table A 8: Threshold levels used for Salmo trutta aquaculture...... 117 Table A 9: Threshold levels used for Gadus morhua aquaculture...... 119 Table A 10: Threshold levels used for Dicentrarchus labrax aquaculture...... 120 Table A 11: Threshold levels used for Sparus aurata aquaculture...... 121 Table A 12: Threshold levels used for Homarus gammarus aquaculture...... 123 Table A 13: Threshold levels used for Mytilus edulis aquaculture...... 124 Table A 14: Threshold levels used for Crassostrea gigas aquaculture...... 125 Table A 15: Threshold levels used for Ostrea edulis aquaculture...... 127 Table A 16: Threshold levels used for Ruditapes philippinarum aquaculture. 129 Table A 17: Threshold levels used for Pecten maximus aquaculture...... 131 Page ix

Table A 18: Table of tidal coefficient calibration ...... 164 Table A-19: Time of High and Low water tidal state, and peak velocity over spring and neap tide...... 170 Table A-20: Details of discharges included in the validated model...... 170

Figures

Figure 1: Boundary of the East Devon & Dorset FLAG area of interest...... 5 Figure 2: Example of the curvilinear effect of temperature on the growth rate of fish. Shown here is the effect on the growth rate in Atlantic cod (from Björnsson et al., 2001)...... 19 Figure 3: Graph showing the effect of temperature and pressure on the concentration of dissolved oxygen. (Taken from Engineering ToolBox, (2005). Oxygen - Solubility in Freshwater and Sea water. Available online at: www.engineeringtoolbox.com/oxygen-solubility-water-d_841.html, accessed 28/11/2018.) ...... 23 Figure 4: Base map for the East Devon & Dorset FLAG area of interest (The bathymetry layer indicates the FLAG extent out to the UK 6 nm limit)...... 28 Figure 5: Natural seabed habitats in the East Devon & Dorset FLAG area of interest, based on the JNCC UKSeaMap 2018 broad-scale physical habitat vector layer, aggregated to broad substrate types...... 29 Figure 6: Human infrastructure, water quality monitoring and other obstructions within the East Devon & Dorset FLAG area of interest...... 30 Figure 7: Recreation (including sailing) areas, aquaculture sites (existing and provisional), energy licensing areas and modelled vessel density (from MMO AIS data, 2015) within the East Devon & Dorset FLAG area of interest...... 30 Figure 8: Administrative, regulatory and conservation boundaries within the East Devon & Dorset FLAG area of interest...... 31 Figure 9: MMO boat angling model for the East Devon & Dorset FLAG area of interest...... 32 Figure 10: Locations and dry weather flow discharge rates from coastal long sea outfalls in the FLAG area...... 33 Figure 11: Biotoxin monitoring programme sampling locations SW . .. 34 Figure 12: MMO modelled land with sea view in the East Devon & Dorset FLAG area of interest...... 36 Figure 13: Flow chart summarising the approach used for identifying areas suitable for aquaculture...... 38 Figure 14: Map 1 produced for the stakeholder event...... 46

Page x

Figure 15: Map 2 produced for the stakeholder event...... 47 Figure 16: Map 3 produced for the stakeholder event...... 47 Figure 17: Maps of the desirability of aquaculture within in the FLAG region. (a) Seaweed aquaculture; (b) Finfish aquaculture; (c) Lobster aquaculture; (d) Bivalve aquaculture. Points have been georeferenced from the locations indicated by participants in the 3 consultation meetings...... 48 Figure 18: Map showing the location and extent of the constraining factors used to exclude regions in the FLAG area...... 51 Figure 19: Areas of potential for the aquaculture of macroalgae in the FLAG region by bottom-secured, suspended and floating-container systems...... 55 Figure 20: Areas of potential for the aquaculture of finfish in the FLAG region by open pen systems...... 57 Figure 21: Areas of potential for the aquaculture of crustaceans in the FLAG region by bottom (ranching) and suspended systems...... 58 Figure 22: Areas of potential for the aquaculture of bivalves in the FLAG region by bottom, bottom-secured and suspended systems...... 60 Figure 23: Indicative footprints for potential commercial scale aquaculture (finfish, bivalves & macroalgae)...... 69

Figure A-1: Telemac mesh coloured following water depth information...... 163 Figure A-2 Predicted water depth over tidal cycle at the BODC elevation point ...... 164 Figure A-3: Velocity from model prediction and observation in cm/s format X axis ...... 165 Figure A-4: Predicted and observed water elevation over the ebbing phase of the tidal cycle...... 166 Figure A-5: Comparison ABPmer predictions with Telemac predictions...... 166 Figure A-6: Comparison of current pattern in the vicinity of . . 168 Figure A-7: Model mesh with the location where the tidal state is extracted indicated with a red square...... 169 Figure A-8: Predicted water elevation over a tidal cycle, with low water indicated with the blue arrow, and high water with the green arrow...... 169 Figure A-9: Predicted velocity at the node - over tidal cycle, including the spring (green arrow) and neap (blue arrow) tide...... 170 Figure A-10: Prediction of E. coli plume for several state of tide, assuming a constant decay rate of 19.48 hours. The discharges locations are indicated with black octagon...... 172 Page xi

Figure A-11: Prediction of E. coli spreading for four tidal states. The locations of the discharge are indicated with a black dot with the associated name above for , , Bridport and Weymouth & Portland...... 173 Figure A-12: Extent of the E. coli 5 cfu/100 ml isoline off Charmouth, Bridport and Weymouth & Portland...... 174

Page xii

Glossary

Aquaculture Aquaculture is the farming of aquatic organisms, including fish, molluscs, crustaceans and aquatic plants. Bivalve Bivalve - means any marine molluscs of the class Bivalvia having a shell consisting of two hinged valves, typically connected by a ligament. Dry weather The average daily flow to the sewage treatment works flow during seven consecutive days without rain following seven days during which rainfall did not exceed 0.25 mm on any one day (excludes public or local holidays). With a significant industrial input, the dry weather flow is based on the flows during five working days if production is limited to that period. E. coli Escherichia coli (E. coli) is the most common example of faecal coliform bacteria. Coliforms can produce their characteristic reactions (e.g. production of acid from lactose) at 44°C as well as 37°C. Usually, but not exclusively, associated with the intestines of warm-blooded animals and birds and used as an indicator species for identifying potential faecal contamination of bivalve mollusc shellfish (and bathing waters). HACCP Hazard Analysis and Critical Control Point (HACCP). Mariculture Marine based aquaculture. Marine A Marine Conservation Zone (MCZ) is a type of marine Conservation nature reserve in UK waters. They are established under the Zones Marine and Coastal Access Act (2009) and are areas designated with the aim to protect nationally important, rare or threatened habitats and species. Nikuradse Relative roughness based on the diameter of uniform sand roughness grain particles. RAMSAR Ramsar Convention on Wetlands of International Importance (1971), especially as Waterfowl Habitat is an international treaty for the conservation and sustainable use of wetlands. SeaWiFS SeaWIFS (Sea-Viewing Wide Field-of-View Sensor) was a satellite-borne sensor designed to collect global ocean biological data. Active from September 1997 to December 2010, its primary mission was to quantify chlorophyll produced by marine phytoplankton (microscopic plants). Sewage Sewage can be defined as liquid, of whatever quality that is or has been in a sewer. It consists of waterborne waste from domestic, trade and industrial sources together with rainfall from subsoil and surface water. SIFCA District - means the area defined in Articles 2 and 3 of the District Southern Inshore Fisheries and Conservation Order 2010 (SI 2010 No 2198). Shapefile A shapefile is a simple, non-topological format for storing the geometric location and attribute information of geographic features. Geographic features in a shapefile can be represented by points, lines, or polygons (areas).

Page xiii

Shellfish Bed Article 22 (2) Sea Fisheries (Shellfish) Act 1967 “shellfish bed " means any bed or ground in which shellfish are usually found or which is used for the propagation or cultivation of shellfish. Spat Spat - means a larva of an oyster or similar bivalve that has settled by attaching to a surface. T90 The time necessary for 90% of the microbes to die Wiggling A hydrodynamic modelling term referring to oscillations in space at a given time or oscillations in time at a given position, a function of two-dimensional spatial resolution.

Page xiv

Abbreviations

ADO Aquaculture Development Officer AIS Automatic Identification System AMA Aquaculture Management Areas BMPA Bivalve Mollusc Production Area BNG British National Grid BODC British Oceanographic Data Centre C-SCOPE Combining Sea and Coastal Planning in Europe project CD Chart Datum CDOM Coloured Dissolved Organic Matter Cefas Centre for Environment, Fisheries & Aquaculture Science CSO Combined Sewer Overflow DCF Dorset Coast Forum Defra Department for Environment, Food and Rural Affairs DEM Digital Elevation Model DO Dissolved Oxygen DTM Digital Terrain Models DTX DTX DWF Dry Weather Flow DRN Detailed River Network EA Environment Agency EATiP European Aquaculture Technology and Innovation Platform E. coli Escherichia coli EC European Community EEC European Economic Community EFARO European Fisheries Aquaculture Research Organisations ELWM Extreme Low Water Mark EMODnet European Marine Observation and Data Network EO Emergency Overflow ETP Endangered, Threatened and Protected species; FAO Food and Agriculture Organisation FHI Cefas Fish Health Inspectorate FLAG Fisheries Local Action Group FSA Food Standards Agency GIS Geographic Information System GMA General Management Approach HPLC High Performance Liquid Chromatography HRA Habitats Regulations Assessment IFCA Inshore Fisheries and Conservation Authority

Page xv

IMTA Integrated Multi-Trophic Aquaculture JNCC Joint Nature Conservation Committee LAT Lowest Astronomical Tide MACAA Marine and Coastal Access Act 2009 MarLIN Marine Life Information Network MBA Marine Biological Association MCZ Marine Conservation Zone MHWN Mean High Water Neaps MHWS Mean High Water Springs MLWN Mean Low Water Neaps MLWS Mean Low Water Springs MMO Marine Management Organisation MODIS Moderate Resolution Imaging Spectroradiometer MPA Marine Protected Area MPN Most Probable Number MPS Marne Policy Statement n.d. No date OA Okadaic Acid pMCZ provisional Marine Conservation Zone PAR Photosynthetically Available Radiation PSU Practical Salinity Units PTX Pectenotoxins RAS Recirculating Aquaculture System RMP Representative Monitoring Point RYA Royal Yachting Association SAC Special Area of Conservation SAMS Scottish Association of Marine Science SBCC Sea Based Container Systems SCOPAC Standing Conference On Problems Associated with the Coastline SeaWiFS Sea-Viewing Wide Field-of-View Sensor SPA Special Protection Area SO Storm overflow SPIM Suspended Inorganic Particulate Matter SPM Suspended Particulate Matter SQL Structured Query Language SSSI Site of Special Scientific Interest STW Sewage Treatment Works SWOT Strengths Weaknesses Opportunities Threats analysis SWPA Shellfish Water Protected Area TCE The Crown Estate TSLE Test of Significant Likely Effect

Page xvi

UNESCO United Nations Educational, Scientific and Cultural Organization. UV Ultraviolet vMNR voluntary Marine Nature Reserve VMS Vessel Management System WGS84 World Geodetic System 1984 WPNSA Weymouth & Portland National Sailing Academy WwTW Wastewater Treatment Works

Page xvii

1 Introduction 1.1 Project rationale

The global human population looks set to reach 9.8 billion people by 2050 (United Nations, 2017), and our food systems will therefore be under intense pressure to produce animal protein for an increasing population. Food security is an objective of the UK administrations and aquaculture makes an important and growing contribution to this (HM Government, 2011)

Aquaculture is the process of farming or culturing aquatic organisms. Marine aquaculture presents an opportunity for increasing seafood production in the face of growing demand for marine protein and limited scope for expanding wild fishery harvests. As the global human population grows and average incomes rise, the demand for ocean-derived food will continue to increase (Costello et al, 2019). Given the relative sustainability of marine aquaculture compared with land-based meat production (Hall et al, 2011), the human health benefits of diets rich in fish (Tacon et al, 2016) and the findings that current total landings of all wild-capture fisheries could be produced using less than 0.015% of the global ocean area, it makes it even more pressing that we consider aquaculture’s potential (Gentry et al, 2017). New analysis has found that under optimistic projections regarding alternative marine aquaculture (mariculture) feed innovations and uptake, the ocean could provide over six times more food than it does today—more than two-thirds of the animal protein needed to feed the future global population (Costello et al., 2019).

1.2 Background

In a response to recent stagnation in the growth of many sectors of aquaculture across Europe, the EU developed a set of strategic guidelines (COM, 2013) aimed at removing barriers and facilitating sustainable aquaculture. Two key areas identified within the guidelines were securing sustainable development and growth of aquaculture through coordinated spatial planning and the development of a multi-annual national plans for aquaculture.

The UK Marine Policy Statement (MPS) 2011 provided the framework for preparing Marine Plans which are now a legal requirement under the Marine and Coastal Access Act (MCAA) 2009, sections 54 and 61 of which require these are to be kept under review at regular intervals. Adoption of Marine Plans is required by 2021. The South Marine Plans are informed by 53 policies including those supportive and protective of aquaculture.

The UK multi-annual plan (Defra, 2015) highlights the Government’s commitment to encouraging the development of aquaculture within the new Marine Plan Areas. This included emphasising to marine planners the importance of Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 1 of 185

aquaculture to the environment and society and strengthening data and evidence locally and nationally for the potential for aquaculture development (marine finfish, shellfish, algae and other organisms) e.g.:

• identify the environmental conditions (physical, chemical and biological) required for different forms of marine aquaculture • assess the suitability of different forms of aquaculture within each Marine Plan area, identifying areas with the greatest potential • identify factors that may limit aquaculture development in these areas, such as water quality and other marine activities and recommend solutions • assess the legislative environmental designations in those areas (SSSIs, EMS, MCZs, etc.) and likely constraints and opportunities associated with them • determine how the aquaculture industry could be best represented in each of the Marine Plan areas and at each stage of the process of developing the Marine Plans, for example by setting up regional hubs.

Spatial planning for marine aquaculture is receiving increasing attention globally in response to interest in maximising sustainable aquatic food production (Pastres et al., 2018). In recognition of the difficulties that aquaculture has faced in obtaining access to space several initiatives at both UK and European levels have set targets or visions for addressing these problems i.e.:

• The ‘creation of prioritised aquaculture areas’ in the Seafood 2040 A Strategic Framework for England (Austin, 2017) • The European Aquaculture Technology and innovation Platform (EATiP) vision for the future states ‘…The development and application of planned zoning for aquaculture production will enable social and economic benefits for coastal and rural communities.’ (EATiP, 2017). • The European Fisheries and Aquaculture Research Organisations (EFARO) vision of where European aquaculture will be in 2030, 2050 and beyond, outlining the main drivers and game changers recommends that ‘By 2030 Suitable zones have been allocated to aquaculture, so that coastal and rural communities benefit socially and economically.’

As highlighted by the FAO & World Bank (2015), to ensure success and sustainability of future aquaculture development, a holistic approach should be taken to the designation of areas, in accordance with the Code of Conduct for Responsible Fisheries (FAO, 2011) and the Ecosystem Approach to Aquaculture (FAO, 2010). The three key guiding principles of an ecosystem approach for aquaculture are:

1. aquaculture development and management should take account of the full range of ecosystem functions and services, and should not threaten the sustained delivery of these to society 2. aquaculture should improve human well-being and equity for all relevant stakeholders Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 2 of 185

3. aquaculture should be developed in the context of other sectors, policies and goals.

Once scoping has identified aquaculture as a priority at the national level, detailed plans are elaborated for progressively smaller geographical units at the regional and local levels, as appropriate (FAO & World Bank, 2015).

The process of spatial planning usually consists of the following steps:

• Aquaculture zoning: bringing together the criteria for locating aquaculture and other activities to define broad zones suitable for different activities or mixes of activities; • Site selection: identifying the most appropriate locations for individual farm development within zones.

Furthermore, as aquaculture develops within a plan area further steps are required for:

• Aquaculture management areas (AMAs): within zones, AMAs contain several individual farms that share a common water supply and/or are in such proximity that disease and water quality are best managed collectively rather than by individual farms; • Carrying capacity estimation via a range of modelling approaches suitable for the species or location; • Implementation and compliance monitoring for the farms in relation to site position, stock biomass, adherence to shared management agreements such as biosecurity.

The Dorset and East Devon FLAG area is covered by the South Inshore Marine Plan that was adopted in July 2018 and extends up through the inshore area up to mean high water springs (MHWS) or the tidal limit. Within an earlier sustainability Appraisal Report (MMO, 2016) the MMO summarised information on relevant parts of the environment, economy and society that could be affected (negatively or positively) by the plans. It was intended for the plans to comply with the goals of the UK 2005 Sustainable Development Strategy, by using sound science responsibly to promote and achieve a healthy and sustainable society and economy, while living within environmental limits. The draft and final Plan considers a wide range of issues, ranging from community, culture and economy, through environmental, ecological and geophysical constraints, to marine activities including ports, fishing and offshore energy production. In summary, the objectives are:

1. Promote effective use of space to support existing, and facilitate future, sustainable economic activity through the encouragement of coexistence, mitigation of conflicts and minimisation of development footprints; 2. Manage existing, and facilitate the provision of new, infrastructure supporting marine and terrestrial activity; Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 3 of 185

3. Support diversification of activities which improve socioeconomic conditions in coastal communities; 4. Support marine activities that increase or enhance employment opportunities at all skills levels among the workforce of coastal communities, particularly where they support existing or developing industries within the South Marine Plan areas; 5. Avoid, minimise, or mitigate displacement of marine activities, particularly where of importance to adjacent coastal communities, and where this is not practical to ensure adverse effects on social benefits are avoided; 6. Maintain and enhance inclusive access to, and within, the South Marine Plan areas appropriate to its setting and in a way that is equitable to users; 7. Support the reduction of the environmental, social and economic impacts of climate change, through encouraging the implementation of mitigation and adaptation measures that • avoid proposals’ indirect contributions to greenhouse gas emissions and reduce vulnerability • improve resilience to climate and coastal change • consider habitats that provide related ecosystem services; 8. Identify and conserve heritage assets that are significant to the historic environment of the South Marine Plan areas; 9. Consider the seascape and its constituent marine character and visual resource and the landscape of the South Marine Plan areas; 10. Support marine protected area objectives and the delivery of a well- managed, ecologically coherent, network with enhanced resilience and capability to adapt to change; 11. Complement and contribute to the achievement or maintenance of • Good Environmental Status under the Marine Strategy Framework Directive, and • Good Ecological Status or Potential under the Water Framework Directive with respect to descriptors for marine litter, non-indigenous species and underwater noise; 12. Safeguard space for, and improve the quality of, the natural marine environment, including to enable continued provision of ecosystem goods and services, particularly in relation to coastal and seabed habitats, fisheries and cumulative impacts on highly mobile species.

This set the background requirements and factors for consideration in identification of, and planning for, potential future extension and establishment of aquaculture and associated activities within the South Marine plan, underpinned by evidence projects (MMO,2013a and MMO 2013b).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 4 of 185

1.3 Project Aim

The purpose of this project was to build on the previous small-scale mapping of potential aquaculture areas supplied for the English marine plans (Heal & Capuzzo, 2019) and develop a more detailed regional resource as per FAO & World Bank (2015) which identifies, refines and improves mapped areas best suited (and with least conflict) to specific types of sustainable marine aquaculture, within the boundaries of the Dorset and East Devon Fisheries Local Action Group (FLAG) area. The area of interest (Figure 1) was delineated by river catchment boundaries on the landward side and the UK 6 nautical mile limit at the southernmost seaward extent, as described in detail in Section 5.

Figure 1: Boundary of the East Devon & Dorset FLAG area of interest.

Following consultation with the industry, produce an in-depth analysis which should benefit aquaculture interests in the FLAG area and act as an exemplar approach for others to adopt.

In consultation with key stakeholders the project will also seek to identify potential management and mitigation measures that may facilitate sustainable aquaculture development in the FLAG area, without unduly restricting opportunities for fishing.

1.4 Project Scope

The scope of the project was limited to marine aquaculture within the boundaries of the Dorset & East Devon FLAG area and the species chosen were restricted to those considered to be suitable under current or future environmental conditions.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 5 of 185

For bivalve shellfish, these species included the blue mussel (Mytilus edulis), the Pacific Oyster (Crassostrea gigas) and native oysters (Ostrea edulis), Manila clam (Ruditapes philippinarum) and Atlantic king scallops (Pecten maximus).

For macroalgae currently exploited species include Kelp (Saccharina latissima), Oarweed (Laminaria digitata), Bladderlock or Winged Kelp (Alaria esculenta) and Dulse (Palmaria palmata) were selected, along with the sea lettuce (Ulva lactuca) which is fast-growing and cultivated in parts of Europe and Asia.

The finfish species selected included two commonly farmed species in Northern Europe the Atlantic salmon (Salmo salar) and the rainbow trout (Oncorhynchus mykiss), two potential species for colder marine waters the sea trout (Salmo trutta) and the Atlantic cod (Gadus morhua) with potential for farming in submerged cages and two potential warmer waters species the gilthead bream (Sparus aurata) and the sea bass (Dicentrarchus labrax) whose range is moving further north.

For Crustaceans in Northern Europe, the European lobster and crab, are still almost exclusively obtained by capture fisheries. Aquaculture has been used for restocking and farming of the European lobster (Homarus gammarus) and therefore this was the only species considered.

Based on the potential aquaculture species selected the relevant culture systems that are either currently in use or under development were assessed and included within the scope of the project. These systems included bottom and bottom secured culture, suspended longline (rope, textile, rafts & containers) and both floating and submerged cage culture. An overview of principle aquaculture systems is provided in Appendix 2.

Land based mariculture (including pump ashore and land based Recirculating Aquaculture Systems [RAS]) were covered within the scope of the project. However, freshwater aquaculture of any sort was not considered within the scope of this project.

Whilst the project aimed to be as comprehensive as possible and utilise best available data and approaches this did not extend to modelling of carrying capacity or assigning farm management areas. This is because at this stage these would be purely speculative and theoretical. However, these should be developed and implemented by regulators as any potential aquaculture areas begin to be developed.

1.5 Project Benefits

The outputs and benefits of the project include:

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 6 of 185

• A detailed appraisal of aquaculture potential in the FLAG region based on up-to-date data and stakeholder consultation. • Access to a completed report and its outputs to support business development plans and diversification opportunities. • GIS map access via Cefas data Hub & FLAG Aquaculture Hub • A mapping resource that is available to feed into future revision of the Marine Management Organisation (MMO) South Marine Plan. • Improved knowledge and advancement of the processes for carrying out this type of mapping exercise.

2 Marine aquaculture systems by culture groups

Information is provided below on the systems requirements for all the differing culture groups. Additionally, an overview of the principal marine aquaculture production systems which was developed for illustrative use in the stakeholder workshops is provided in Appendix 2.

2.1 Marine Bivalve shellfish

Marine Bivalve Shellfish – Bottom Culture

Bottom culture involves a chosen nursery area that is sown with bivalve seed, called spat and allowed to grow to harvestable size. Maintenance of the stock is minimal and limited to occasional monitoring and predator removal (Laing and Spencer, 2006). In the UK, the following species are regularly cultivated in this way:

• Native Oyster, Ostrea edulis (Seafish 2019c) • Pacific Oyster, Crassostrea gigas (Seafish 2019d) • Blue Mussel, Mytilus edulis (Seafish 2019g) • Manilla Clam, Ruditapes philippinarum (Seafish 2019a) • King Scallop, Pecten maximus (Seafish 2019f)

For example, the cultivation of mussels directly on the seabed occurs in intertidal and subtidal areas in sheltered bays where there is typically an abundant natural spatfall. Spat are collected at an age of approximately 1 year old, when they measure between 10-30 mm shell length (SL) (Spencer, 2002). Spat are generally collected locally from less suitable grounds and re-laid in higher densities in more productive areas. (SIDS, 2008). Mussels take from 18 – 36 months to reach harvest size of 45 mm. The time is dependent on the location and productivity of the cultivation site and the initial size of seed relayed.

Harvesting bottom cultured mussels can vary from mechanised dredging and power dredging to simple hand-collection or hand-raking on lower yield operations (Spencer, 2002).

Scallops can also be re-laid, or sown, on the bottom (often from suspended longline culture in pearl/lantern nets) for on growing after dredging has occurred to remove Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 7 of 185

starfish predators from the culture plots. Scallops tend to be stocked up to 5 individuals per m2, will disperse naturally and can reach market size in 2.5-3.5 years. Harvesting is usually performed by dredges or by hand diving and can result in up to 80% of the stock being recaptured. As there is no containment of the scallops, it is basically a restocking exercise (SIDS, 2008).

Bottom culture can be undertaken in deeper waters (up to 20 m) using boats and dredges (Laing, 2002; Laing and Spencer, 2006).

Marine Bivalve Shellfish – Bottom Secured Culture

These include trestles, trays & secured mesh bags installed to posts/poles driven into the seabed, so they are accessible at mean spring low water mark. They need regular attention so must have good access and they need to be in relatively protected sites safe from excessive wind/wave action. These systems are most suited to the cultivation of mussels and oysters which can tolerate short periods of air exposure at low tides.

Pole Culture

Mussels can be cultured on a series of poles driven into the seabed. Known as the bouchot method (which originated in France in the 13th century) it consists of vertical poles driven into the seabed and is essentially a shallow water technique, with access to the poles at low tide or by diving. The poles are typically 4-7 m long (with 2-3 m rising above the seabed) and 25-30 cm in diameter and are made from hardwoods such as oak and pine, or more recently aluminium. The poles are placed in parallel rows at right angles to the shoreline and are usually positioned around the ELWM (Extreme Low Water Mark). For this method, the spat is collected on poles or coir rope attached to metal frames placed in deeper water. The seed are then transferred to tubular nets, which are then wound around the bouchot poles and nailed into place. From these nets the seed will migrate out onto the surface of the poles and settle (SIDS, 2008).

Trestle Culture

Both Pacific and Native Oysters can be cultivated in meshed bags (typically 1 m long and 0.5 m wide) secured onto intertidal trestles (typically 3 m long and 0.5 m high) driven into the seabed. Seed or ‘spat’ oysters may be purchased from a few dedicated hatcheries and are available in a variety of size grades, usually from 4 – 30 mm shell length and are placed in bags with an appropriate mesh size. Bags are regularly attended to be turned and for fouling and predators to be removed turned. Stock is occasionally graded into larger meshed bags and they can be harvested from 75 g upwards and it can take 2.5 – 3 years cultivation before the first harvest (Seafish 2019d).

Both these techniques favour firmer substrates to provide better support for the aquaculture containers and so that they are less likely to result in high silt loads in the water. This means less maintenance is required to ensure clogging of mesh does not occur (Laing and Spencer, 2006).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 8 of 185

Marine Bivalve Shellfish – Suspended Culture

The methods of suspended culture utilised in different regions originate from a combination of traditional methods, local availability of materials and site. The advantages of suspended culture methods are the increased exposure to water currents and the reduction of predation on the cultured population from benthic predators.

There are two principle methods of suspended culture and a variety of different designs and methods have arisen to take advantage of local conditions:

• Raft culture • Longline culture which includes: o Rope cultures o Suspended oyster bags o Suspended pearl/lantern nets o Ear hanging for King Scallops.

The continuity of suspended mussel cultivation is greatly dependant on the acquisition of spat (mussel seed) which is predominantly obtained from natural populations. Consequently, most rope-grown mussel cultivators collect their own stock from the wild spat-fall. The rope droppers are coiled so that they remain in the top 2 – 3 m of the water column and they are placed on the line or raft in time to collect the natural spat. The exact timing of settlement varies around the country, but the main spat-fall is generally in early summer, although later and earlier settlements also take place. Mussels can spawn several times during the year (Seafish 2019e). Hatchery produced spat are available - however, for most markets the use of these spat is uneconomical (SIDS, 2008).

Raft culture

The use of rafts to cultivate mussels is becoming increasingly popular and is utilised by many countries including Ireland and Scotland (SIDS, 2008). Raft culture is based around a framework of timber supported by floats, with a series of ropes attached within the framework of the raft. The rafts require a water depth several metres deeper than the length of the ropes to prevent the grounding of the mussels during low tides, which would lead to damaging of the mussels and increased predation.

Spat collection occurs at the farm on ropes attached to the raft. The seed are then stripped from the ropes and placed into cotton socking which is then reattached to the ropes which are then replaced onto the rafts. Thinning of the mussels occurs throughout the grow out process to manage the size at which the mussels can be harvested.

Longline culture

Longline culture of bivalve shellfish is the method of choice in more exposed waters due to the flexibility of the structures. A typical arrangement will consist of a series of horizontal ropes that are held in the top 3 m of the water column by buoys and anchored to the seabed. From the headlines are a series of dropper ropes, which can carry the shellfish directly (mussels, ear hung scallops) or containers to hold the shellfish (oyster

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 9 of 185

bags, pearl/lantern nets). Rope/line-grown shellfish aquaculture in the UK is dominated by Blue Mussels (Mytilus edulis) and King Scallops (Pecten maximus).

Blue Mussels

Mussels are commonly cultivated by longlines connected to large floats and anchored at both ends, which support many “droppers” (ropes). The dropper ropes are typically 6- 10 m in length and spaced at 50 cm intervals (Spencer, 2002). The coiling “droppers”, on which the mussels are grown, are suspended in the top 2–3 m of the water column in time for the wild spat fall. Once seed mussels have settled the droppers are uncoiled to their full length. Little further management is required in this form of aquaculture until the bivalves have reached the desired size for harvesting. The carrying capacity of the longlines is dependent on the volumetric capacity of the floats, and as the mussels grow, additional floats are required to support the crop.

King Scallops

Cultivation methods using pearl and lantern nets suspended from long-lines are typically used to raise scallop seed and for on growing towards market size. Pearl nets are generally used for small (10–30 mm shell length) scallops and lantern nets for larger animals (Laing, 2002). As with mussel suspended culture, the main lines are held afloat by buoys and anchored by mooring blocks. Farmers use spacing and orientation of the culture ropes to control growth and condition of the crop (Spencer, 2002). Grow out methods used in the UK are similar to those in Japan, adjusted for the local conditions, and market sizes are reached in 4-5 years for King Scallops (100-150 mm shell length) (SIDS, 2008). It is common for scallops to be relayed to the seabed once they have reached ~ 60mm shell length as evidence suggests that scallops cultured for more prolonged periods in suspension have weaker shells and more prone to predation (Goulden, 2018).

Marine Bivalve Shellfish – Land-based Hatchery

Bivalve shellfish aquaculture is comprised of three stages: hatchery, nursery, and grow- out. Farmers rely upon either hatchery seed production or procurement of wild seed. In some geographic locations, large natural beds exist solely for the latter purpose (Getchis, 2014).

Hatcheries are typically land-based facilities that draw seawater from a local water body or a saltwater well. Wild or cultivated adult bivalve shellfish are brought into the facility and placed in tanks for broodstock conditioning. Gametes are collected, fertilized, and then maintained in static tanks or recirculating systems and reared through metamorphosis. The production of the correct species of algae or plankton to feed the developing shellfish larvae and juveniles is key to a successful hatchery. This requires considerable skill and experience and investment in live food production technology. Some hatcheries are experimenting with flow-through production methods that allow high-density larval culture. Post-set shellfish may remain in the hatchery for a period to

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 10 of 185

achieve a larger size or may be moved into the nursery and growout stages (Getchis, 2014).

Nursery culture can continue on land or can involve seeding the bivalve shellfish in floating or submerged containers or gear (e.g. upwellers, bags, trays, cages, longlines, etc.) or directly to the bottom. While many farmers choose to rear their product in containers, a large proportion of the industry plants significantly larger seed (which are less vulnerable to predation) directly on the bottom or use some combination of both gear and bottom culture.

Marine Bivalve Shellfish – Depuration

Depuration is a process in which harvested bivalve shellfish are held in closed water recirculating units for the purpose of reducing the level of potentially harmful bacterial/viral loading in the flesh of the shellfish1. The requirement to depurate harvested shellfish is a function of the classification of the water within the region of cultivation.

Marine Bivalve Shellfish – Classification

Under food hygiene legislation commercial bivalve mollusc production areas used for harvesting wild or cultured bivalve molluscs must be sampled and classified. The classification system is determined by the Food Standards Agency (FSA) using a standard testing process and currently provides the following classification levels based on time-series analysis of the levels of the indicator bacteria Escherichia coli (E. coli) in shellfish flesh:

Class A – Shellfish samples must not exceed 230 E. coli per 100 g in 80% samples (No sample may exceed 700). No post-harvesting treatment required. Can go direct for human consumption. Class A product can attract a premium price and may be required by some food retail customers along with evidence of an effective Hazard Analysis and Critical Control Point (HACCP) system.

Class B – Shellfish samples must not exceed 4,600 E. coli per 100 g in 90% of samples (upper limit 46,000). Can go for human consumption after purification (depuration1), relaying in an approved Class A relay area or following application of an EC approved heat treatment process.

Class – C Shellfish samples must not exceed 46,000 E. coli per 100 g. Can go for human consumption only after relaying for at least 2 months in an approved relaying area, followed (if required) by treatment in a depuration system or following application of approved heat treatment process.

Prohibited - >46,000. No harvesting permitted

1 The depuration unit is usually land based and comprises of large holding tanks through which seawater is recirculated. An in-line water disinfection system (usually UV) is installed in the recirculated flow which kills bacteria and viruses which are purged from the gut and tissues of the shellfish usually over a period of 42 hours Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 11 of 185

2.2 Marine Crustaceans

For this report we will be concentrating on the European lobster (Homarus gammarus) which is currently the only marine crustacean whose culture is developed enough to attempt commercial production.

Marine Crustacean – Hatchery

Lobster hatcheries operate at elevated seawater temperatures and consequently tend to make use of recirculation technologies. Lobster hatcheries do not tend to be fully commercial enterprises. In general, they are operated to provide juvenile lobsters for release in to the wild as a stocking or fishery management measure. (Seafish 2019b)

Eggs are collected from naturally gravid females known as berried females (the collection of which require special permissions). The berried females are held and conditioned in the hatchery and as the larvae hatch, they are collected by using low powered light which they are attracted to. The larvae are then reared in upwelling containers until the reach metamorphosis to the post larval (stage 4) in 2-3 weeks. Metamorphosed larvae have the appearance of the adult lobster and swim against the current in the tub with their claws extended in front of them. The stage 4 larvae are reared for at least another 2 months in individual compartments (due to their cannibalistic nature) when they can be considered for release into the wild. (Seafish 2019b)

Marine Crustacean – Bottom culture

This is where hatchery reared juvenile lobsters are released into the wild. This is more akin to ranching as opposed to a more recognisable form of contained aquaculture. The principle is that if they are released in suitable grounds the juvenile lobsters will remain in the area. Depending upon the size of juvenile released, it can take from 4 – 7 years before the first animals above the minimum landing size are recovered. The animals can be recaptured in conventional fishing gear (pots or creels) set in the normal way. It has been estimated that, depending upon the size of juvenile released, up to 40% of the stock may be surviving to marketable size. It has been observed that animals of hatchery origin can contribute between 5 – 10% of the gross fishery landings of an area. (Seafish 2019b)

Marine Crustaceans – Suspended Culture

In the past 20 years there have been a handful of trials on the culture of European Lobster (Homarus gammarus) in containers suspended in the sea (Bignell et al, 2016). This is known as sea-based container culture or SBCC. The most recent of these involved the rearing of Stage 6 hatchery reared juvenile European Lobster in purpose designed modular container systems attached to a Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 12 of 185

more standard longline system – similar to that used for cultivating rope grown mussels.

Stage 6 juvenile lobsters were transferred to a stage 1 container capable of holding 20 individuals in isolation (to prevent losses due to the natural cannibalistic nature of European lobster). The lobsters were not provided with any additional feed and relied on suspension and raptorial feeding of plankton, zooplankton and fouling organisms (Beal et al., 2012). The lobsters were placed into to Stage 2 containers after 12 months at sea, which provided more space for the individual lobsters and a year later the lobsters were temporarily removed from the Stage 2 containers, so they could be cleaned. In this way it was possible to grow lobster to 150-200 g in 2-3 years.

Although this is not yet currently a marketable size for lobster and the economic viability of the culture to this stage was challenging; the concept of growing European lobster in sea-based containers without supplemental feeding with a survival rate of over 60% from juvenile stocking was proven. Future developments may well establish this as a viable culture system (particularly in integrated multitrophic aquaculture applications) where the production of Lobster is seen as one of the outputs of the production system.

2.3 Marine Macro Algae

Seaweed farming is a newly emerging technology in the UK which has many potential benefits:

• Water remediation and habitat remediation • Reduction in the frequency of algal blooms • Production of pseudo habitats - providing nursery grounds and shelter • Provision of Increased nutrients for shellfish • Coastal protection by wave/current attenuation

The production techniques for seaweed are wide ranging though production techniques can be split between bottom rope culture and suspended rope culture.

Marine Macro Algae – Bottom Secured

Bottom rope culture is more widely used for the more delicate bottom dwelling species (such as Ulva spp and Palmaria spp). With this technique ropes are secured along the seabed (often close to the shore and seaweed attached at set points along the ropes Depending on the cultivated species and placement of the bottom secured ropes often they are often uncovered at low tide which helps with collection. Importantly bottom rope culture needs to be established in very

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 13 of 185 sheltered areas since the culture system and the algae growing on them are very sensitive to disturbance caused by wave/tide and wind action.

Marine Macro Algae – Suspended Culture

Suspended rope culture of seaweeds follows a process very similar to that of suspended rope grown shellfish and is better suited to species with long fronds (kelps). Like with rope grown shellfish the seaweed is attached to ropes which usually run horizontally under the water at a depth which permits access to good light levels. In inshore environments where water turbidity will affect the depth to which light can penetrate, the water column the culture lines may be set just a few feet below the water surface. In more offshore environments where the water is often clearer, and light can penetrate deeper into the water column the seaweed can be also be cultured on rope droppers from rafts to provide a greater yield per unit area.

Currently there is no specific regulation for seaweed farming in the UK and the established licensing procedure for finfish and shellfish would not always be applicable to seaweed farms. The lack of information on potential environmental impacts of seaweed farms and their management (e.g. control of nuisance species) results in uncertainties for the regulator on how to interpret requirements under national and international regulations for Seaweed in the UK (Capuzzo & McKie, 2016).

Marine Macro Algae – Land-based Hatchery

Seaweeds are cultured by the production of spores from suitably ripening individuals which is aided by the manipulation of the light wavelength that they are held in prior to spore production. Macroalgal spores are collected from ripe plants and are then seeded onto polyamide strings. Here the spores can germinate to form tiny plants ~ 2 mm long, which are transferred to sea after a few months.

An important consideration regarding the production of macroalgal spores for culture in the open sea is that they need to be collected from macroalgal species that are currently local and indigenous to the area where the macroalgae production is intended. This will ensure that the environmental and genetic integrity of local macroalgae populations will be maintained in the presence of increased cultivation of selected macroalgae species.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 14 of 185

2.4 Marine Finfish

Floating open pen cages or submersible cages

Marine finfish production in the open sea typically takes place in sea cages which are open pen cages (often with covering nets to mitigate against escapees and predation) anchored securely to the seabed. These open pen cages can be square, rectangular or circular. The most commonly used open pens are circular and are typically 90-110 m in diameter and 15-20 m deep, with several cages employed at a site to create a cage farm (Scott and Muir, 2000; Cardia, F & Lovatelli, 2015). Circular cages are advantageous as they better utilise space, have a beneficial cost/volume ratio and are more resistance to dynamic stresses experience in exposed sites.

There are many different types of offshore cages structure including floating flexible, floating rigid, semi-submersible flexible, semi-submersible rigid and submersible rigid (Scott and Muir, 2000). The advantage of submersible or semi- submersible cages is that they can be positioned away from the high energy regimes of the surface water, so could potentially be cheaper and reduce risks of loss of stocks. Currently floating circular cages are most widely considered as they are the most economically proven at present.

In addition to the net pens themselves floating open pen culture systems can include, on-farm feed hoppers and top nets, feed barges, lifting gear, on-farm accommodation for staff, ancillary rafts for generators & other infrastructure, feed pipes and above & below sea-level lighting.

Choosing the correct site for the sea cages is important because it influences the mortality and growth conditions of the finfish which affect profitability but is also important for mitigating environmental impact. A key factor to consider is the exposure of the site to wind and wave action. An offshore and exposed site requires higher specification equipment and increased running costs, but fish welfare is improved and environmental impact lessened. A sheltered and protected site will have fewer damaging waves and currents but is likely to have a greater environmental impact (Cardia, F & Lovatelli, 2015).

Marine Finfish – Pump-ashore Land-based

Suitably located land-based sites with close access to the sea to allow the development of a seawater intake would be suitable for pump ashore land based marine aquaculture. These systems allow seawater to be abstracted, processed as required (filtration/disinfection) and fed directly into a series of tanks used for culturing a wide range of species. It is not normal that the temperature or the composition of the incoming water is altered significantly (cf. recirculating systems) so that the species cultured are matched with the main physical and

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 15 of 185

biological properties of the incoming water. Water passes through the rearing system and is returned to the sea with little or no re-use. Some form of settlement or waste removal is usually employed before the seawater is returned to the ocean. These systems are good for the following applications:

• Hatchery - rearing of delicate juvenile stages where good access and stock monitoring are important • Rearing of high values species • Holding and conditioning of broodstock.

Marine Finfish – Recirculation System

Recirculating Aquaculture Systems (RAS) are a development on the theme of a pump ashore land-based farm in which water is continually recirculated around containment tanks in which the fish are housed. This allows the quality of the water to be controlled and reduces the environmental impact of the farming process. These are particularly good for maintaining species at their optimum temperature or for holding species where additional biosecurity is important. A RAS is a partially closed system in which typically about 10% of the total volume of water in the culture system is replaced daily. Key to the successful operation of a RAS is the biological filter which helps to break down potentially harmful waste products (ammonia and nitrite) produced by the stock in the system. In addition to the biological filter, RAS will typically have mechanical filtration, protein skimmers, aeration/oxygenation systems, carbon dioxide degassing systems, ozone/UV sterilisation systems and water quality monitoring technology. They are usually complex systems which do not respond well to sudden changes in water quality or stock loading. In this respect they require considerable skill and expertise to manage and operate profitably backed up by sound design and technical support.

2.5 Marine Multi-Trophic Aquaculture

Integrated multi-trophic aquaculture (IMTA) is the co-cultivation of fed species (such as finfish or shrimp) together with extractive species, such as suspension- feeding (e.g., mussels and oysters) and deposit-feeding (e.g., sea-cucumbers and sea-urchins) invertebrates and macroalgae (e.g., kelps), which may feed on the organic and inorganic effluents generated by the fed species. By establishing such integrated cultivation systems, the sustainability of aquaculture may be increased (Chopin et al., 2001; Neori et al., 2004; FAO, 2006; Buchholz et al., 2012). However, for IMTA to be economically viable, each of the individual components must be marketable (Chopin et al., 2010), or adding value through accounting for the ecosystem services that extractive species provide (Buchholz et al., 2012).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 16 of 185

Integrated Multi-Trophic Aquaculture (IMTA) is put forward as a promising solution for the sustainable development of aquaculture.

IMTA farmers combine species that need supplemental feed such as fish, with “extractive” species. Extractive species use the organic and inorganic materials and by-products from the other species for their own growth. Extractive species can be primary producers (algae and plant species that transform inorganic nutrients into organic biomass) or secondary producers (those that use organic material from the water column or the seabed as food). The secondary producers can be either filter feeders (generally shellfish that sieve organic particles such as algae from the water column) or deposit feeders (organisms such as worms, sea urchins, sea cucumbers etc. that feed on organic material on or within the sediment). Extractive species act as living filters. The natural ability of these species to recycle the nutrients (or wastes) that are present in and around fish farms can help growers improve the environmental performance of their sites. In addition, the extractive species have commercial value as marketable products, providing extra economic benefits.

The IMTA concept is sometimes used in a strict sense: having the different trophic levels integrated in one farm or business, at the same site. However, trophic links in aquatic ecosystems can extend over a large spatial scale. Depending on the local hydrodynamics and the biogeochemical processes involved, spatial separation may be even beneficial. For example, as the transformation from fish waste into nutrients takes time and if there is a residual current, the best location for a seaweed farm to optimally utilise these nutrients may not be in the immediate vicinity of the fish farm, but further downstream. Conversely, cultivation of different species at the same site may occur, without a direct trophic link between the crops, due to the economic benefit of co-location and being able to use a production site in all seasons and for multiple products at the same site. This type of co-location is perhaps a bit beyond the strict definition of IMTA, but still highly relevant for developing business cases for new forms and aquaculture. As aquaculture heads towards more organised spatial planning, IMTA will have to become a reality to optimise limited space https://impaqtproject.eu/about-imta/.

Ideally, the biological and chemical processes in an IMTA system should balance. This is achieved through the appropriate selection and proportions of different species providing different ecosystem functions. The co-cultured species should be more than just biofilters; they should also be harvestable crops of commercial value (Chopin, 2006). A working IMTA system should result in greater production for the overall system, based on mutual benefits to the co- cultured species and improved ecosystem health, even if the individual

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 17 of 185

production of some of the species is lower compared to what could be reached in monoculture practices over a short-term period (Neori et al.,2004, FAO 2009).

3 Overview of optimal physiological conditions for aquaculture

For the purposes of this study the environmental factors are considered separately and appropriately for each species. However, it should be noted that there is often a complex interplay between environmental factors. Some the environmental drivers of physiological conditions used in this study are outlined in a previous report (Heal and Cappuzzo, 2019) and are repeated in this section to aid the reader.

3.1 Sea surface temperature

For all living organisms, their ambient temperature plays an important role in their survival, growth and reproduction. Endothermic organisms maintain a constant body temperature above their surroundings, but this is an energy inefficient process and is often accompanied by slow growth and larger diets.

All the organisms considered in this report are ectotherms and their body temperatures are close to that of their surroundings. One major effect of this is on their metabolic rate which is much reduced at lower temperatures thereby affecting key processes such as digestion, assimilation (growth) and movement (Schulte, 2015). These effects can be simply described using thermal performance curves which tend to be unimodal and left skewed (Dell et al., 2011). Generally, increasing the surrounding temperature, increases metabolism and therefore the growth rate of the organism increases also. This is especially true of finfish because the surrounding water is such a good conductor of heat and growth rates will rise to an optimal level, beyond which the rate decreases (Björnsson et al., 2001; Elliott and Elliott, 2010). Above the optimum temperature heat loss to the environment exceeds gains from increased metabolism. The curvilinear growth profile for Atlantic cod is shown in Figure 2: Example of the curvilinear effect of temperature on the growth rate of fish. Shown here is the effect on the growth rate in Atlantic cod (from Björnsson et al., 2001).and illustrates the presence of an optimal temperature for growth.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 18 of 185

Figure 2: Example of the curvilinear effect of temperature on the growth rate of fish. Shown here is the effect on the growth rate in Atlantic cod (from Björnsson et al., 2001).

Temperatures above that which the organism normally experiences will cause high levels of mortality. This ‘critical temperature’ is specific to an organism and reflects their adaptations to minimising the effect of temperatures beyond the optimum. It is therefore important to consider the range over which the organism survives and grows, along with an optimal range when the growth rate is at its highest (Schulte, 2015).

3.2 Salinity of the water

Changes in salinity affect the osmotic regulatory capacity of organisms, and particularly the ability of living cells of taking up water through osmosis (Rivera- Ingraham, 2017). Organisms can respond to changes in salinity with osmotic acclimation; in general, it is possible to categorise species into euryhaline organisms (more tolerant to changes in salinity) and stenohaline organisms (less tolerant to changes in salinity). Fluctuations in salinity (e.g. in estuarine, intertidal or lagoon areas) can have severe effects on less tolerant organisms, potentially leading to mortality.

Adult finfish in seawater deal with an imbalance of a high level of salts in the external environment versus a comparatively lower level inside the cells of the organism. Dealing with this by osmoregulation is estimated to take 20 to 50% of the total energy output, although this may be an over-estimate (Bœuf and Payan, 2001). However, the effect of salinity on fish growth is multifactorial influencing the metabolic rate, intake of food, assimilation and hormonal action (Bœuf and Payan, 2001). Typically, highest growth rates are observed in intermediary salinity conditions (salinities of 8-20 psu).

Marine shellfish (crustaceans and bivalve molluscs) living in coastal regions can experience large changes in the salinity of the surrounding water. Coastal salinity Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 19 of 185

can fluctuate between near freshwater (near 0 salinity) to hypersaline seawater in supra-tidal pools (>150 salinity) (Rivera-Ingraham and Lignot, 2017). To acclimatise to changes in salinity the marine shellfish must make cellular adjustments to prevent loss of water and turgor pressure (hypertonic challenge) or uptake of water or cell lysis (hypotonic challenge). This is achieved through a combination of changes to membrane permeability, osmotic effectors in the cytosol, active transport processes and the production of ammonia (Rivera- Ingraham and Lignot, 2017). These processes require an increase in energy expenditure and therefore specific species balance their energy budgets according to potential niches that can be occupied (Rivera-Ingraham and Lignot, 2017). Most bivalves and crustaceans are osmo-conformers and they keep their internal conditions isosmotic to their environment (Rivera-Ingraham and Lignot, 2017).

The salinity of water changes in response to varying water temperature. However, the wide fluctuations in salinity are observed in transitional and coastal waters, subjected to river runoff. Estuarine areas can have large fluctuations in salinity due to freshwater input from rivers which can vary seasonally. In some coastal lagoons the water movement is restricted, and salinity increases as water evaporates. This can also occur in the intertidal zone, especially in coastal rock pools.

Salinity is the amount of dissolved salts in water, it is reported in practical salinity units (psu) a unit based on the properties of sea water conductivity. It is equivalent to parts per thousand (ppt) or to g/kg.

3.3 Light availability

Without light (specifically PAR, photosynthetically available radiation), seaweeds (as well as any other plant) would not be able to carry out photosynthesis and therefore survive. However, too much light has also detrimental effects as it can lead to cellular damage and death (photo-inhibition; see review by Hanelt and Figueroa, 2012).

3.4 Phytoplankton (chlorophyll) levels

For filter feeder organisms, such as mollusc bivalves, the availability of phyto- plankton biomass is important, as micro-algae can represent the major food source for these organisms. Studies (Laing and Spencer, 2006; Strand et al., 2016) have shown that up to 85% of growth performance differences between sites culturing bivalves is due to the combined effects of water temperature and primary productivity (i.e. phytoplankton concentration).

Phyto-plankton concentrations show seasonal fluctuations in relation to changes in the environmental conditions (i.e. nutrient and light availability and

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 20 of 185

temperature) and bivalves cope by supplementation of their diet with particulate organic matter. Furthermore, phyto-plankton are difficult to characterise to species level and the relative nutritional value is not known for many species.

Phyto-plankton are autotrophic, using photosynthesis to fix carbon and generate organic biomass. This process requires chlorophyll pigments and the concentration of these pigments in the sea water can be used to assess the phytoplankton levels. This is useful as it allows remote sensing by satellite over large areas (see Snyder et al., 2017 for application of this process to aquaculture site selection along the Maine coast). However, using this as a proxy for the phytoplankton levels has been questioned (Bourlès et al., 2009) with studies showing a disparity between chlorophyll levels and phytoplankton abundance and/or biomass. It is suggested that where environmental conditions are relatively stable chlorophyll levels are a good indicator of phytoplankton (Ren and Schiel, 2008) but this association is less strong in more varied ecosystems. In general, the ratio between carbon and chlorophyll in phytoplankton cells is not fixed but changes during the year and with location (e.g. Jakobsen and Markager, 2016; Lyngsgaard et al., 2017).

3.5 Nutrient concentration

Seaweeds require nutrients (e.g. nitrogen and phosphorus) for growth and metabolic processes. The uptake rate of the dissolved nutrients from the surrounding water is affected by factors such as light availability, temperature, desiccation, water movement, and the chemical form of the nutrients (Harrison and Hurd, 2001). The rate is not constant with time but may vary depending on the level of nutrients in the nutrient-pools in the plant tissue (i.e. whether the plant is in a nutrient-limited stage or not; (Harrison and Hurd, 2001). In this study we focused on availability of total oxidised nitrogen (TOxN), nitrate and nitrite concentrations.

3.6 Dissolved oxygen concentration

The concentration of dissolved oxygen (DO) in water is influenced by many factors including water temperature, salinity, atmospheric pressure, water depth and biological activity (i.e. photosynthesis and respiration). There is an inverse relationship between the temperature of water and the concentration of oxygen dissolved in it (Benson and Krause, 1984; Garcia and Gordon, 1992). The DO concentration of seawater is typically around 8-11 mg/l.

Most marine organisms are poikilotherms having a metabolic rate that is predominantly controlled by the surrounding water temperature. Any increase in temperature results in an increase in metabolic rate with a subsequent increase in oxygen uptake and growth. However, at increased temperatures oxygen

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 21 of 185 availability decreases as the amount of oxygen dissolved in the seawater reduces (Figure 3).

For example, at 1 atmosphere pressure seawater at 5°C has 9.9 mg/l DO compared to 7.2 mg/l for water at 20°C (Benson and Krause, 1984; Garcia and Gordon, 1992). Furthermore, a decrease in salinity results in an increase in DO (Figure 3). In freshwater, at 1 atmosphere pressure the DO is 12.8 mg/l at 5°C and 9.1 mg/l at 20°C (Benson and Krause, 1984; Garcia and Gordon, 1992).

Factors which can affect the level of oxygen dissolved in the water include movement of the water, mixing of the water with air and the presence of marine organisms. Bio-fouling of aquaculture apparatus can lead to an increase in micro-organism growth (decomposers) which increase the Biological Oxygen Demand (BOD) producing localised reductions in DO levels (Fitridge et al., 2012).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 22 of 185

[a] Seawater (Salinity – 35 psu)

[b] Freshwater (Salinity – 0 psu)

Figure 3: Graph showing the effect of temperature and pressure on the concentration of dissolved oxygen. (Taken from Engineering ToolBox, (2005). Oxygen - Solubility in Freshwater and Sea water. Available online at: www.engineeringtoolbox.com/oxygen-solubility-water- d_841.html, accessed 28/11/2018.)

3.7 General comments

Under the EU Horizon 2020 funded ‘AquaSpace’ project, Boogert et al., 2018, performed an extended literature review on tolerance for each of 45 farmed species and optimal ranges for thirteen environmental parameters. These were considered alongside other sources to derive optimal and sub-optimal threshold values for the potential aquaculture species identified in this project.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 23 of 185

4 Approaches to mapping potential aquaculture areas 4.1 Previous approaches

A detailed literature review was conducted to identify previous work on identifying and mapping areas with potential for future aquaculture production. Appendix 1 includes fuller detail on elements of previous studies, such as: data collection, synthesis and mapping; natural and anthropogenic environmental considerations and resource pressures; spatial modelling and defining potential areas for aquaculture; co-location opportunities, and constraints to development; model and data limitations. The main findings are summarised in this section (previous approaches) and the following section (current approaches).

Studies examined addressed a range of factors and issues surrounding the siting of aquaculture farms, including consideration of environmental conditions, sensitive features and other users of the marine environment. Attention was directed towards studies considering spatial aspects of aquaculture farming and suitability mapping.

One important aspect in this type of assessment is data availability. Location and extent of natural features are often captured in localised areas, such as through series of surveys and monitoring initiatives in the Lyme Bay area (Munro & Baldock, 2012; Wood, 2007) to record qualitative and quantitative information on seabed habitats, species assemblages, ecosystem health, etc. Of direct relevance to the current work is a report providing a detailed methodology of GIS processing applied in the Lyme Bay Habitat Risk Assessment (Marine Planning Consultants Ltd, 2014a).

A management-focused approach was taken in the Lyme Bay Integrated Fisheries Management Plan analysis (Marine Planning Consultants Ltd, 2014b), which incorporated habitat, species and fisheries assessments within a comprehensive SWOT (Strengths, Weaknesses, Opportunities & Threats) analysis. A risk matrix was compiled for nine different types of fishing activity versus the full range of habitats, benthic organisms, fish, mammals, turtles and birds in the area, along with a fisheries sustainability assessment and range of management options for five key fisheries. More topic-oriented mapping exercises have included, in 2010, the development by Cefas of a set of national data layers providing information on the spatial extent of fishing activities, including intensity, for inshore waters around England and Wales. Outputs from this were compared with those of the Finding Sanctuary initiative (Lieberknecht et al., 2011), to develop an approach which

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 24 of 185

may aid a future update of the National Inshore Fisheries Data layer. See Appendix 1 for further details of this work.

GPS, data-logger technology and dye studies have also been used to collect, analyse and visualise data in a spatial framework (Huang et al., 2009), and an integrative methodology for site selection for shellfish aquaculture (Silva et al., 2011) combined GIS and dynamic farm-scale carrying capacity to generate a spatial suitability index. Socio-economic and environmental factors have also been considered (Ferreira et al., 2007; Silva et al., 2011).

Commercial site selection software, such as ShellGIS (http://www.shellgis.com/Default.aspx), have been developed for shellfish growers, water body managers and regulators. The tool embeds state-of-the-art models: ShellSIM (Hawkins et al., 2012; 2013a, b); MIKE 21 (Warren and Bach, 1992); and FLOW 3D (Sicilian et al., 1987) that account for interactive effects of culture type, shellfish density and hydrodynamics. ShellGIS is a transferable tool which, for a choice of 14 commonly-cultured shellfish species, enables dynamic predictions of shellfish production, environmental effects and profitability, according to culture practice and site selection within a given GIS domain (Newell et al., 2012a, b; 2013).

The MMO (2013a) developed a model to identify aquaculture trends in the context of the growing importance of aquaculture in supporting food security, in line with the objectives of the UK Administrations (UK Marine Policy Statement) – https://www.gov.uk/government/publications/uk-marine-policy-statement). The model, which included a high-level cost benefit analysis of potential economic returns for each aquaculture type, was the first assessment of its type on the spatial potential for aquaculture in East and South MPAs. Final outputs identified and mapped areas of current, near future and future, potential development per species/aquaculture method, including infrastructure and planning constraint areas, investment factors and relative investment required. Detailed modelling steps, including data processing requirements, are given in MMO (2013b).

4.2 Current approaches

The current work examines several studies where existing data have been used to synthesise new information, including incorporation within a GIS for analysis and/or mapping to derive spatially explicit information. A small selection of such studies is summarised here, with further detail in Appendix 1.

ABPmer (2015) have provided a comprehensive list of data layers and sources collated as part of their spatial assessment, including identification of unused data, with reasons for their exclusion where relevant. With a focus on

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 25 of 185

sustainable aquaculture, Falconer et al. (2017) reviewed the application of GIS in aquaculture planning and management in its entirety, from data collection and database construction, through spatial analysis, model development and decision support, to capacity building and stakeholder engagement. They also highlighted the relatively recent innovation of web-based applications, opening a range of new possibilities for GIS. Further details of model applications are given in Appendix 1.

A number of interactive web tools have been developed (e.g. the C-SCOPE Coastal Atlas, Belgium, and the Dorset Explorer, both intended to provide information and support planning and decision-making in the coastal zone – see Appendix 1). A more ambitious international research project, aiming to understand spatial and socio-economic constraints on the expansion of aquaculture, led to the development of the open source AquaSpace tool (Gimpel et al., 2018) for integrated assessment within a spatial framework, of socio- economic and environmental risks and opportunities for potential aquaculture systems.

Detailed analyses of GIS-based assessments include Falconer et al. (2017) and Ross (2014), who highlight the benefits and flexibility of such systems, including the facility to incorporate stakeholder feedback. Silva et al. (2016) incorporate detailed analyses of remote sensing observations, and Falconer et al. (2017) and (Marine Planning Consultants Ltd, 2014b) stress the importance of environmental considerations in aquaculture planning assessments. Limitations were also identified (MMO, 2013a; Marine Planning Consultants Ltd., 2014b) and recommendations made (MMO, 2013a; Ross, 2014; ABPmer, 2015; Austun, 2014; Falconer et al., 2017; Cefas & FSA, 2017).

5 Data sets, processing and mapping

Constraints to current and future aquaculture production fall into various categories, so it was necessary to identify and compile a range of data relevant to viability of production within the East Devon and Dorset FLAG area of interest. Consideration was given to environmental conditions, natural and man-made features, human activities and regulatory areas, technical limitations and species-specific requirements.

A comprehensive set of spatial data, from several different sources, was compiled for analysis, processing and mapping of aquaculture suitability extents within the area. Details of data identified for inclusion within the GIS are given in the attached document, AquacultureMappingProject_GIS_Catalogue.pdf, (Appendix 7). Mapping data were processed using ArcGIS Desktop 10.5 (https://www.esri.com) or QGIS Desktop 3.2.1 (https://www.qgis.org), in vector (shapefile) or raster (gridded) format, as necessary.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 26 of 185

5.1 Mapping existing features

All data were projected to British National Grid and extracted to the area bounded by: • a line due south from Beer Head to the UK 6 nm limit, forming the western extent; • a line extending southeast to the UK 6 nm limit from, and in the same orientation as, the training wall at the entrance to Poole Harbour, forming the eastern extent; • the UK 6 nm limit forming the southern extent; • aggregated terrestrial boundaries of river catchments draining to the coast between Beer Head and the training wall at Poole Harbour forming the northern extent.

Data were subset into layers representing (i) base map; (ii) human activities and infrastructure features; (iii) natural habitats; (iv) regulatory and conservation areas; (iv) physical and biological parameters (e.g. temperature, salinity, current speed).

Base map

A base map (Figure 4) was created for the area of interest, comprising (a) terrestrial features (50 m raster digital terrain model – OS Terrain 50 DTM) overlaid with the Environment Agency Detailed River Network (DRN) vector layer clipped to the river catchments draining to the defined extent of coastline; (b) OS Meridian 2 coastline for the area; (c) marine raster bathymetry.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 27 of 185

Figure 4: Base map for the East Devon & Dorset FLAG area of interest (The bathymetry layer indicates the FLAG extent out to the UK 6 nm limit).

As consistently high-resolution raster bathymetry data were unavailable for the entire area of interest, GIS techniques were used to mosaic together data from two sources: i. high resolution (1 m) raster data of coastal bathymetry collected and compiled by the Channel Coastal Observatory (https://www.channelcoast.org/); ii. the coarser Defra Marine digital elevation model (DEM) (1 arc second) (https://data.gov.uk/) raster layer, resampled at 10 m resolution.

In this way, it was possible to derive full bathymetric coverage of the area at the highest resolution available.

Natural habitats

Seabed substrate data were extracted from the JNCC UKSeaMap 2018 broad- scale physical habitat vector layer, aggregated to the following classes: mud,

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 28 of 185 sand, coarse substrate, mixed sediment, rock or other hard substrata, unclassified (Figure 5).

Figure 5: Natural seabed habitats in the East Devon & Dorset FLAG area of interest, based on the JNCC UKSeaMap 2018 broad-scale physical habitat vector layer, aggregated to broad substrate types.

Natural coastal habitat areas, including priority habitats, available to Cefas as composite data layers under the Defra SPIRE Agreement, were clipped to the area of interest. Layers represented intertidal substrate & foreshore (mud, sand, gravel, boulders, rock); mudflats; saltmarsh; maritime cliffs & slope; coastal vegetated shingle; coastal sand dunes. Most coastal and intertidal features are too small to be visible at the scale of Figure 4.

Human activities and infrastructure features

Point and line data were obtained and processed for a variety of infrastructure features relating to industry (subsurface cabling, wastewater discharges, fishing ports, shoreline constructions); monitoring points (bathing water quality); navigation and obstruction (surface buoys & navigational aids, sailing & racing marks, mooring points & berths, transportation routes, shipwreck sites) (Figure 6). Boundaries were obtained or derived for marine recreational areas (sailing & mooring areas, wind/kitesurfing seasonal exclusion zones, dive sites); existing and provisional aquaculture areas; offshore energy license areas), and

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 29 of 185 the MMO polygon grid of AIS vessel density data (annual average for all vessels, 2015) was also obtained and clipped to the area of interest (Figure 7).

Figure 6: Human infrastructure, water quality monitoring and other obstructions within the East Devon & Dorset FLAG area of interest.

Figure 7: Recreation (including sailing) areas, aquaculture sites (existing and provisional), energy licensing areas and modelled vessel density (from MMO AIS data, 2015) within the East Devon & Dorset FLAG area of interest.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 30 of 185

Regulatory and conservation areas

Boundaries of administrative, regulatory and conservation areas compiled within the area of interest included: MCZs; Ramsar sites; Special Areas of Conservation (SACs); Special Protection Areas (SPAs); Sites of Special Scientific Interest (SSSIs); protected sites for nesting and overwintering birds; protected wreck sites; MOD exclusion zones & military practice areas; Portland Port Authority area; Southern Inshore Fisheries and Conservation Authority (IFCA) Byelaw area; Lyme Bay Statutory Instrument area were mapped and are shown in Figure 8.

Figure 8: Administrative, regulatory and conservation boundaries within the East Devon & Dorset FLAG area of interest.

5.2 Other known constraints and data limitations

Fishing activity

Commercial and recreational fishing are important to the coastal communities and the district. Defra propose to implement a statutory instrument requiring inshore Vessel Monitoring System (iVMS) on all licenced fishing vessels under 12 m. Data from such a system would allow the mapping of commercial fishing activity. Whilst iVMS has been trailed on several vessels that operate in Lyme Bay, the data from this were not available to us, to take into consideration in this study.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 31 of 185

Within marine protected areas (MPAs) restrictions are placed on certain types of fishing activity as part of management measures aimed at protecting conservation features. For areas designated as Marine Conservation Zones (MCZs) the types of fishing gear and their compatibility with different habitat types have been developed into a matrix to allow certain types of fishing to continue. However, the same exercise has not yet been carried out for aquaculture. Aquaculture will not necessarily be precluded from conservation areas, depending on the type of restriction and the method of aquaculture production under consideration.

An initial review of the data on fish spawning areas did not provide enough spatially detailed information to be used within the study. However, the MMO advised that improved information was currently being developed.

Spatial distribution of recreational boat angling (Figure 9) was included for stakeholder consultation reference purposes, although these were not used in the suitability assessment.

Figure 9: MMO boat angling model for the East Devon & Dorset FLAG area of interest.

Contaminants

Natural and anthropogenic contaminants, including those from terrestrial areas adjacent to coastal waters, have the potential to impact on growth, health and survival of aquatic organisms (Davies and Vethaak, 2012; Bignell et al., 2016) Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 32 of 185 and should be taken into consideration in site assessment for aquaculture production from the perspective of both the organism and the end user.

Wastewater discharges and long sea outfalls

In the context of coastal water quality, pipeline discharges of wastewater direct to the sea represent a potentially significant point source of localised contamination. The impact of these sea outfalls depends largely on: the level of sewage treatment received prior to discharge, their relative location to sensitive receptors, including their proximity to, filter feeding, bivalve mollusc production areas (both wild and cultured), the volume of wastewater discharged and the dispersive regime at the point of discharge.

Several long sea outfalls occur within and close to the FLAG area (Figure 10).

Figure 10: Locations and dry weather flow discharge rates from coastal long sea outfalls in the FLAG area.

Siting of new ‘offshore’ bivalve mollusc aquaculture sites would ideally be in waters that reflect hygiene classification ‘Class A’2 in bivalve shellfish flesh, and may be used for direct human consumption, rather than be required to undergo relaying in better quality waters, or subject to purification (depuration) or heat treatment before being released to market.

Under the original project scope of the intention was to make a first assessment of this using a simple box model dilution calculation based on discharge bacterial loading, water depth at outfall location and mean ebb and flood tide current

2 Also see section 2.1.6 in relation to shellfish hygiene classifications Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 33 of 185 speed. Preferably, sea outfall dispersion would be undertaken using a calibrated hydrodynamic model of the area. This method allows improved prediction of effluent dispersion. Fishmongers’ Company was approached and was able to provide some additional limited funding which has enabled Cefas to undertake basic hydrodynamic modelling of dispersion for these outfalls.

Biotoxins

Within this Flag region there is a limited amount of data available on algal counts in seawater and bivalve flesh monitoring for biotoxins. This is collected from bivalve mollusc production areas (BMPAs) as follows. In the Fleet (the only current BMPA site in FLAG area), water sampling is undertaken fortnightly from April to September and four-weekly from October to March and monitoring of flesh for biotoxin in Pacific oysters is undertaken four-weekly. In Poole Harbour and Lyme Bay (nearest sites outside the FLAG area). Water sampling for algal counts and monitoring of flesh for biotoxin in mussels is undertaken fortnightly from April to September and four-weekly from October to March.

Biotoxins risk represents a significant planning issue for new bivalve mollusc farm sites. One ‘open coast’ mussel producer in Cornwall has now purchased another site in South Wales to be able to guarantee supply of product when biotoxins close their Cornish site. One or more sites off the Dorset coast may also be of interest to producers seeking to spread the risk of biotoxin contamination over multiple locations. Identification of additional offshore sites could therefore help to attract, potentially significant, inward investment into the FLAG region (M. Syvret pers. comm).

Figure 11 shows the locations for biotoxin monitoring in bivalve mollusc shellfish flesh in South West England.

Table 1 summarises the numbers of bivalve mollusc shellfish flesh samples taken for biotoxin monitoring by year within the FLAG district (The Fleet and Portland) and in the nearest production areas that flank the district (Lyme Bay and Poole) for 2013-2019. Table 2 provides an approximation of the number of days each year production was shut during 2013-2019.

a) Flesh sampling locations b) Water sampling locations Source: (Cefas, 2018). Figure 11: Biotoxin monitoring programme sampling locations SW England.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 34 of 185

Table 1: Numbers of bivalve samples for biotoxin analysis by year. Year 2013 2014 2015 2016 2017 2018 2019 Lyme Bay Site 1 West mussels* n/s n/s 9 25 23 24 18 Lyme Bay Ensis West razors* n/s n/s n/s n/s n/s n/s 5 Fleet Oysters Pacific oysters 13 17 12 15 14 13 13 Portland N.E. Breakwater mussels 13 15 n/s n/s n/s n/s n/s Portland N.E. Breakwater Pacific oysters n/s 1 n/s n/s n/s n/s n/s Poole Coastal mussels n/s n/s n/s n/s n/s n/s 8 Poole W. Brownsea 1 mussels* 6 3 1 n/s n/s 1 12 Poole W. Brownsea 1 Pacific oysters* 5 3 1 n/s n/s 1 n/s n/s = not sampled. Monthly sampling unless indicated (* fortnightly sampling Apr-Sep). Source: Food Standards Agency - Cefas Annual Reports (Cefas, 2013, 2014, 2015, 2016, 2017, 2018).

Table 2: Bivalve mollusc production areas - Biotoxin Closures Nos. of days by year. Year 2013 2014 2015 2016 2017 2018 2019 Lyme Bay Site 1 West mussels* n/s n/s 42 87 37 16 0 Lyme Bay Ensis West razors* n/s n/s n/s n/s n/s n/s 0 Fleet Oyster Farm Pacific oysters 0 0 n/s n/s n/s n/s n/s Fleet Oysters B025AI Pacific oysters n/s 0 0 0 0 0 0 Portland N.E. Breakwater mussels 0 8 n/s n/s n/s n/s n/s Portland N.E. Breakwater Pacific oysters n/s 0 n/s n/s n/s n/s n/s Poole Coastal mussels n/s n/s n/s n/s n/s n/s 0 Poole W. Brownsea 1 mussels* 0 0 0 n/s n/s 0 0 Poole W. Brownsea 1 B54BL Pacific oysters 0 0 0 n/s n/s 0 n/s n/s = not sampled. N.B. Days closed are approximate based on date sampled and date of second clear result reported following exceedance of regulatory toxin limit. Monthly sampling unless indicated (* fortnightly sampling Apr-Sep). Source: Food Standards Agency – Annual reports on the results of the Biotoxin and Phytoplankton Official Control Monitoring Programmes for England & Wales (Cefas, 2013-18).

In relation to prospective new aquaculture business ventures a check of recent biotoxin history of adjacent areas and baseline sampling at high-risk times of the year could form a basis for risk analysis. In-situ testing would provide the most case specific information, but several environmental variables would need considering when examining the data.

Disease

Disease outbreaks can have significant consequences for aquaculture businesses. Under conditions of authorisation all aquaculture production businesses must have a bio-security measures plan in place that documents what they do to minimise risks of introduction or spread from their farm sites.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 35 of 185

Within this FLAG area there is an historical designation for bonamia, which is a parasite of native oysters. The zone is very broad and bisects the FLAG area along a line due south of , the area to the east of this line being declared a bonamia positive zone while the area to west is classified as bonamia free. This can have implications for permitting the movement of shellfish from one zone to another.

https://assets.publishing.service.gov.uk/government/uploads/system/uploads/atta chment_data/file/545851/Dorset_Hampshire_Bonamia_Ostreae_CD02.pdf

Proximity to other farms, and farm management agreements, will need to be established in future should a significant aquaculture industry begin to develop.

5.3 Aesthetic considerations

This part of West Dorset and East Devon is a World Heritage site and, as such, visual amenity may form part of any detailed impact assessment should applications for licences be submitted in the future. It is not only possible to model the extent of visibility of the sea from different coastal positions (e.g. as previously estimated by the MMO – see Figure 12, but also of different aquaculture structures and their heights from the coastline. However, modelling of any visual impact was beyond the scope of this project.

Figure 12: MMO modelled land with sea view in the East Devon & Dorset FLAG area of interest.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 36 of 185

A guidance document has been produced for Scottish aquaculture (Scottish Natural Heritage, 2011) that helps with positioning and minimising of visual impacts of aquaculture businesses.

6 Deriving aquaculture suitability layers 6.1 Overview of process to generate aquaculture suitability layers

Environmental variables were classified in optimal, suboptimal and unsuitable ranges based on published literature. Optimal reflected that the species survives and grows best under these conditions; suboptimal that the species can survive under these conditions, but growth is reduced; and unsuitable that the species cannot survive or grow when exposed to these conditions in the medium to long term (not acute exposure).

Comparison of these threshold levels with spatial data provided a suitability map showing the sites where the species can survive or thrive. For each factor, the threshold levels were used to reclassify the data as being either optimal (2), sub- optimal (1) or unsuitable (0). Any area in which one or more of these factors was unsuitable for culture was deemed unsuitable. The suitability of the remaining areas was determined by calculating the average score for all the scores in that area and the score was normalised to give a final suitability score of scale of 0 to 1. Using this method an area which was suboptimal for all considered factors scored 0 and was deemed unsuitable. Therefore, any suitability score above 0 indicates that at least 1 factor is optimal for that region.

The following environmental layers were used to determine aquaculture potential:

• Constraints on growth o Sea surface temperature o Salinity o Light penetration (for macroalgal photosynthesis) o Phytoplankton (chlorophyll) levels (for filter feeders) • Water Quality Parameters o Dissolved oxygen concentration o Suspended inorganic particulate matter • Nutrient availability (for macroalgae) o Total oxidised nitrogen concentration • Physical constraints o Significant wave height o Peak current speed o Substrate type o Site exposure level o Water depth

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 37 of 185

The source of the data layers is covered in detail in Section 6.2 Aquaculture potential was determined for each species based on the culture method employed and the threshold levels used can be found in Appendix 3. The overall approach used for identifying areas suitable for aquaculture is summarised in Figure 13 and more detail can be found in Appendix 6.

1. Obtain environmental

data layers for 2. FLAG region Reclassify each data layer

according to 3. threshold Combine reclassified

layers to 4. generate a Exclude areas suitability layer where aquaculture is restricted

Figure 13: Flow chart summarising the approach used for identifying areas suitable for aquaculture.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 38 of 185

6.2 Provenance of GIS data used to determine aquaculture suitability

The source of the GIS data layers used to create the aquaculture potential layers is outlined in detail in this section. Several of the data layers, such as sea surface temperature, were similar to those used to determine biological suitability in the MMO report (Heal and Cappuzzo, 2019). To aid the reader the lineage of this data is presented below. Additional data layers that are more suited to the FLAG region have been selected for other parameters, such as water currents, and some are specific to this study, such as the suspended particulate inorganic matter (SPIM). Data lineage for these data sets is also provided. Overall, data sources include in situ measurements, model output and satellite data which vary in the level of resolution and reach into the littoral areas.

Sea temperature

Daily sea surface temperature data for a 9-year period (2006-2014) was obtained from the Ocean Sea Temperature and Ice Analysis Product (OSTIA); see description in Donlon, C.J. et al., 2012) of the UK Met Office with delivery via the Copernicus Marine data portal (http://marine.copernicus.eu/). Particularly, the following products were used: • SST_GLO_SST_L4_REP_OBSERVATIONS_010_011; • SST_GLO_SST_L4_NRT_OBSERVATIONS_010_001. The minimum and maximum temperature values (calculated as the 2.5th and 97.5th percentile, respectively) were derived on an annual basis and used to set a temperature range. In this way it can be expected that at least 95% of daily temperature values fell within this range. Gaps in the data layers were filled in using an Inverse Distance Weighted (IDW) algorithm using a search radius of 12 points in ArcGIS 10.5. This raster had a spatial resolution of 0.03° (WGS84) and, for use in this study, it was projected to British National Grid (BNG), clipped to the FLAG region, and resampled to a resolution of 500 m.

Salinity

A dataset of all available measurements of salinity for the upper 25 m of the water column was prepared, using data compiled for the UK Eutrophication Comprehensive Assessment (from 1990-2014; https://oap.ospar.org/en/ospar- assessments/intermediate-assessment-2017/), with additional data from the UK Environment Agency (2007-2012).

The extreme values of the 95% data interval (2.5th and 97.5th, percentile values) were calculated on an annual basis and used to set a minimum and maximum salinity range, using a 1/12th degree grid. The dataset was cropped to include only stations within the UK Economic Exclusion Zone. Sampling stations were not equally distributed across the study area, therefore inverse distance weighted Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 39 of 185 interpolation was used to interpolate onto a 1/12th degree grid, using at least 3 data points and a search radius of 1/3rd degree to produce a suitable layer of salinity. The raster was masked to the FLAG region with a resolution of 500 x 500 m with the British National Grid (BNG) spatial projection.

Light penetration layer

Daily maps of the light attenuation coefficient from MODIS (Kd(PAR); Gohin, Druon and Lampert, 2002)), generated by Ifremer (Brest, France; https://archimer.ifremer.fr/doc/0017 8/28883/27369.pdf were averaged by month across 10-years (from 2002 to 2012). The estimated Kd (PAR) for spring (March/April/May) was used to calculate the depth at which the available light is equivalent to 10% of surface light, using Equation 1.

Kd(PAR) * z10% Ez10% = 0.1 * E0 = E0 * e- (1)

Where Ez10% is the irradiance (Photosynthetic Available Radiation, PAR) at depth z10%; E0 is the irradiance (PAR) just below the surface; and z10% is the depth at which the irradiance is 10% of the surface irradiance.

The 10% surface light depth was chosen as descriptor of the light climate to ensure that a portion of the water column fell in an optimal range of irradiance for photosynthetic activity (i.e. not light limited). The raster was masked to the FLAG region with a resolution of 500 x 500 m with the British National Grid (BNG) spatial projection.

Phytoplankton (chlorophyll) levels

In situ measurements of chlorophyll were preferred over remotely acquired satellite data because they are inherently more reliable in complex coastal waters. Estimates of chlorophyll from satellite data can be influenced by variable levels of Suspended Particulate Matter (SPM) and Coloured Dissolved Organic Materials (CDOM) often experienced in case-2 waters (coastal waters).

All available measurements of water column chlorophyll for the upper 25 m of the water column during the months March to October inclusive (coincident with the phytoplankton growing season) were prepared, using data compiled for the UK Eutrophication Comprehensive Assessment (2000-2014); with additional data from the UK Environment Agency (2007-2012). Chlorophyll samples were analysed by different techniques such as fluorescence, spectrophotometry and HPLC; only HPLC provides a measurement of chlorophyll-a, while fluorescence, for example, cannot distinguish between chlorophyll-a and chlorophyllide. Therefore, for this study we simply refer to chlorophyll.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 40 of 185

As for salinity, the dataset was cropped to include only stations within the UK Economic Exclusion Zone. Sampling stations were not equally distributed across the study area, therefore inverse distance weighted interpolation was used to interpolate onto a 1/12th degree grid, using at least 3 data points and a search radius of 1/3rd degree to produce a suitable layer of nutrient concentration.

To fill in any gaps in the data layer, satellite data from the Ocean Colour Climate Change Initiative dataset was used (downloaded from https://climate.esa.int/en/projects/ocean-colour/ on 28/02/2019). Data from 2000 to 2018 was extracted for the months March to October inclusive and a spatial mean was generated in R using the Raster package. Using ArcGIS 10.5 this data layer was used to fill in gaps from the in-situ layer and resulted in a raster layer with a resolution of 1/12 degree. The raster was masked to the FLAG region with a resolution of 500 x 500 m with the British National Grid (BNG) spatial projection.

Dissolved oxygen concentration

All available measurements of water column dissolved oxygen concentration during the annual period of March to October were prepared, using data compiled for the UK Eutrophication Comprehensive Assessment (1990-2014), with additional data from the UK Environment Agency (2007-2012) – (available at http://data.Cefas.co.uk/). As for salinity, the dataset was cropped to include only stations within the UK Economic Exclusion Zone. Sampling stations were not equally distributed across the study area, therefore inverse distance weighted interpolation was used to interpolate onto a 1/12th degree grid, using at least 3 data points and a search radius of 1/3rd degree to produce a suitable layer of dissolved oxygen concentration.

Due to limited data availability, there were some small gaps in coverage of the English coast of this layer, notably in Lyme Bay. Data from the SWEEP Model from Plymouth Marine Laboratory (Lessin, Artioli and Wakelin, 2018) was used to fill these data gaps. The model provides monthly estimates of various marine parameters and the data made available under the Creative Commons Attribution 4.0 International Licence (http://creativecomm ons.org/licenses/by/4.0/legalcode). The dissolved oxygen concentration output from the model was extracted between 1st Jan 1990 and 31st December 2018 for the depths of 25 m or above and a mean value was taken. Using ArcGIS 10.5 this data was used to fill data gaps in the SW extent of the dissolved oxygen layer.

Any remaining gaps were filled by using an Inverse Distance Weighted algorithm with a 12-point search radius to produce the completed dissolved oxygen layer. The raster was masked to the FLAG region with a resolution of 500 x 500 m with the British National Grid (BNG) spatial projection. Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 41 of 185

Suspended Inorganic Particulate Matter (SPIM)

The Suspended Inorganic Particulate Matter (SPIM) layer was generated from the SPIM layer from the Cefas data hub (http://data.Cefas.co.uk/#/View/18133). This data is generated from satellite data and therefore is less accurate in close coastal areas. Data was taken from 2002 onwards due to the use of older SeaWiFS data prior to 2002. The spatial resolution of the data was 1100m and the extent was from 13°W to 12°E and 36°N to 60°N.

The NetCDF file was processed in R and for each year a monthly average of SPIM was calculated. The 95% percentile was then used as the SPIM. This allows a degree of coping mechanisms for the species in the face of higher, but transient, SPIM levels.

The output raster was clipped to the FLAG region, resampled to a 500 x 500 m grid size and projected to British National Grid.

Nutrient concentration (Total oxidised nitrogen)

All available measurements of water column Total Oxidised Nitrogen (TOxN) for the upper 25 m of the water column during the winter months (November to February) were prepared, using data compiled for the UK Eutrophication Comprehensive Assessment (1990-2014), with additional data from the UK Environment Agency (2007-2012) – (available at http://data.Cefas.co.uk/) As for salinity, the dataset was cropped to include only stations within the UK Economic Exclusion Zone. Sampling stations were not equally distributed across the study area, therefore inverse distance weighted interpolation was used to interpolate onto a 1/12th degree grid, using at least 3 data points and a search radius of 1/3rd degree to produce a suitable layer of nutrient concentration.

TOxN concentration was estimated from measurements in the upper 25 m of the water column, therefore it was assumed that TOxN concentration was uniform throughout this part of the water column.

Due to limited data availability, there were some small gaps in coverage of the English coast of this layer, notably in Lyme Bay. Data from ERSEM, Plymouth Marine Laboratory (Lessin et al., 2018) was used to fill these data gaps. The model provides monthly estimates of various marine parameters and the data made available under the Creative Commons Attribution 4.0 International Licence (http://creativecommons.org/licenses/by/4.0/legalcode). The total nitrate concentration output from the model was extracted between 1st Jan 1990 and 31st December 2018 for the depths of 25 m or above and a mean value was taken. Using ArcGIS 10.5 this data was used to fill data gaps in the south west extent of the total oxidised nitrogen layer.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 42 of 185

The final TOxN layer was completed using ArcGIS 10.5 by using an Inverse Distance Weighted algorithm with a 12-point search radius to fill in data gaps using interpolation.

The raster was masked to the FLAG region with a resolution of 500 x 500 m with the British National Grid (BNG) spatial projection.

Significant wave height

The significant wave height layer was generated from wave data taken from the ABPmer Renewables Atlas. The wave shapefile was downloaded from ABPmer Renewables Atlas (ABPmer, 2008). The polygon layer was projected to British National Grid and a 500 x 500 m grid size raster generated for the FLAG region in ArcGIS using data from the Annual Significant Wave Height.

Peak water current

The peak current speed was extracted from the TELEMAC model used to model the sea outfalls (see Appendix 5). The temporal maxima across the grid were extracted as a point shapefile which was projected to British National Grid coordinate reference system. Using an Inverse Distance Weighting algorithm, the maximum peak currents were interpolated as a raster layer with a grid size of 500 x 500 m grid size. This raster was clipped to the FLAG region.

Substrate type

The type of substrate present on the seabed Substrate Value was taken from the 2018 UK Sea map which Rock and Hard (RH) 1 was downloaded from the JNCC website Coarse (CS) 2 (www.jncc.gov.uk) on 26 April 2019. The vector Mixed Sediment (XS) 3 shapefile was converted to a raster layer which Sandy mud (SM) 4 was masked to the FLAG region with a spatial Sand (S) 4 resolution of 500 x 500 m, using the ‘Polygon to Muddy sand (MS) 4 Fine mud (FM) 5 raster’ tool in ArcGIS and the “Substrate” Seabed (SB) 6 column from the data attribute table. The spatial reference used the British National Grid coordinate system. The raster was reclassified using the Folk 5 classification system (Folk, 1954), as shown on the right.

Site Exposure

The site exposure data was taken from the 2018 UK Sea Substrate Value map which was downloaded from the JNCC website Low 1 (www.jncc.gov.uk) on 26 April 2019. Classification of site Moderate 2 exposure is based on the predicted kinetic energy due to High 3 wave and tidal action. The vector shapefile was converted to Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 43 of 185

a raster layer which was masked to the FLAG region with a spatial resolution of 500 x 500 m, using the ‘Polygon to raster’ tool in ArcGIS and the “Energy” column from the data attribute table. This represents the kinetic energy at the seabed due to wave action and is ranked Low, Moderate or High. The spatial reference used the British National Grid coordinate system. The raster was reclassified as shown on the left.

Water depth (bathymetry)

The high-resolution bathymetry layer used to determine aquaculture suitability was derived as described in Section 3.1. The raster was resampled at 500 x 500 m in ArcGIS with the origin matching the other raster layers.

Tides

Tide levels for selected ports within the FLAG area are shown in Table 3 below.

Table 3: Tide levels and ranges. Tide Range Predicted heights in metres above chart datum (mAOD) Port (m) HAT MHWS MHWN MSL MLWN MLWS LAT Spring Neap

Lyme Regis 4.8 4.3 3.1 2.4 1.7 0.6 0.2 3.7 1.4

Portland 2.5 2.1 1.4 1.04 0.8 0.1 -0.2 2.0 0.6

Swanage 2.5 2.1 1.8 1.63 1.4 0.8 0.2 1.3 0.4

Data from Admiralty TotalTide©

6.3 Aquaculture layers used for the consultation exercise.

To simplify the presentation of mapping for the consultation phase, maps of potential aquaculture suitability for individual species (shown in Appendix 6) were combined based on aquaculture groups. The groups used were:

• Finfish o Atlantic salmon (Salmo salar) o Rainbow trout (Oncorhynchus mykiss) o Sea trout (Salmo trutta trutta) o Atlantic cod (Gadus morhua) o Sea bass (Dicentrarchus labrax) o Sea bream (Sparus aurata) • Seaweed o Sugar kelp (Saccharina latissima) o Oarweed (Laminaria digitata) o Winged kelp (Alaria esculenta) o Dulse (Palmaria palmata) Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 44 of 185

o Sea lettuce (Ulva lactuca) • Crustacean o Lobster (Homarus gammarus) • Bivalve o Blue mussel (Mytilus edulis) o Pacific oyster (Crassostrea gigas) o Native (European) oyster (Ostrea edulis) o Manila clam (Ruditapes philippinarum) o King scallop (Pecten maximus) For each aquaculture group, a suitability map was generated for each potential species and culture method combination. These suitability maps were combined for each culture method available for that species group. For example, the seaweed group had 3 possible culture methods; bottom-secured, suspended and floating-container but a suitability map for sea lettuce was only produce for floating-containers. To combine the suitability maps the maximum value was taken for each location. This was achieved by stacking the appropriate raster layers and extracting the maximum value of each cell. Therefore, each resulting ‘aquapotential’ map shows the potential of culturing one or more species within the species group at that location using the specified method. Aquapotential maps were produced as follows:

• Finfish o Open pens • Seaweed o Bottom-secured o Suspended o Floating containers • Lobster o Bottom culture o Suspended • Bivalve o Bottom culture o Bottom-secured o Suspended To generate the aquaculture potential layers for the FLAG region, areas showing potential were removed where there was potential conflict between the aquaculture system and current activities or structures present. The final aquaculture potential maps are shown in Section 8.1.

7 Stakeholder consultation

In July 2019, the Dorset Coast Forum facilitated delivery of three Cefas-led stakeholder consultation workshops to review draft maps of areas with potential Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 45 of 185 for aquaculture within the FLAG district. These were held on separates dates in Weymouth, and Lyme Regis to try and maximise the opportunity for stakeholder engagement across the FLAG.

Three sets of maps (Figure 14 to Figure 16 below) were produced for perusal by participants and to invite comment on suitability for possible future location of aquaculture sites, based on availability of space, location in relation to marine protected areas, biological and physical suitability, and potential conflict with other users of the marine environment. Each set of maps was divided into west and east areas of the Dorset and East Devon FLAG area, to allow adequate space for comment, stickers, post-it notes to be applied directly to the maps.

No consideration of tidal current as an environmental variable, influencing potential aquaculture species/system suitability was presented in the mapping at the time of the workshops.

Showing human infrastructure and activities in the E. Devon and Dorset FLAG area (left-hand map shows western extent from Beer to Portland; right-hand map shows eastern extent from Portland to Purbeck). Pairs of inset maps at the bottom of each panel are indicative of MMO modelled AIS vessel density, and boat angling activity, respectively. Figure 14: Map 1 produced for the stakeholder event.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 46 of 185

Showing conservation designations in the E. Devon and Dorset FLAG area (left-hand map shows western extent from Beer to Portland; right-hand map shows eastern extent from Portland to Purbeck). Inset maps on the extreme right-hand and left-hand panels indicate areas of biological and physical suitability for 4 species groups: bivalves, finfish, lobster and seaweed, respectively. Figure 15: Map 2 produced for the stakeholder event.

Suitability maps inset in Figure 15 were derived by overlaying all the species suitability scores for a given species grouping, then mapping the maximum score for any species within that group. Therefore, the maps indicate where optimal suitability occurs for one or more species within a species grouping.

Showing biological and physical suitability for 4 species groups (bivalves, finfish, lobster and seaweed, respectively) in the E. Devon and Dorset FLAG area, excluding those areas precluded from aquaculture development due to other restrictions (left-hand maps show western extent from Beer to Portland; right-hand maps show eastern extent from Portland to Purbeck). Figure 16: Map 3 produced for the stakeholder event.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 47 of 185

7.1 Stakeholder consultation – outputs

Maps summarising the desirability of outputs from the stakeholder feedback are shown in Figure 17.

Figure 17: Maps of the desirability of aquaculture within in the FLAG region. (a) Seaweed aquaculture; (b) Finfish aquaculture; (c) Lobster aquaculture; (d) Bivalve aquaculture. Points have been georeferenced from the locations indicated by participants in the 3 consultation meetings.

A report on the workshops is provided in Appendix 4.

7.2 Mapping refinement and revision - post consultation

Following the consultation exercise, where possible the findings from the workshops were either included into the application of hard non-environmental constraints (see Section 7.3 below) or considered within the further discussions and recommendations.

7.3 Application of other constraints to determine potential aquaculture areas

In addition to constraints placed on potential aquaculture areas due to environmental and physical conditions, the impact of non-environmental constraints was addressed. Our approach was based on the crown estate model (The Crown Estate, 2019) and adopted to specific types of aquaculture system. So, we have considered areas of busy traffic, ship mooring areas as hard constraints and placed suitable buffers around them. The factors considered are shown in Table 4 alongside the available GIS data layers associated with these features. For those features that are single point or line features a buffer zone

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 48 of 185 was created around the feature and the resulting polygon used as the constraining data layer.

Other designations in the FLAG region may place constraints on the development of aquaculture but present no physical barrier. These soft exclusions include regulatory regions such as Marine Protected Areas and the MoD firing ranges (Lulworth and ), the extents of which can be found in Section 5.1.4. Additionally, the location and intensity of fishing grounds and fishing effort were not considered in the exclusion layers.

Table 4: Potential non-environmental constraints on aquaculture development.

Constraining Factor Comments A 2 km buffer was generated around the main transport routes as defined by UKHO transport and Commercial routes line data*. Recommended routes were not transport routes included as constraints as they are not defined routes and could be repositioned. Anchorage and berthing sites were extracted from UKHO transportation and routes point data*, and a 2 Commercial km buffer zone placed around them due to the berthing and potential size of commercial vessels using these anchorage points. Anchorage and berthing sites were extracted from the UKHO transportation and routes polygon layer*. A 1.0 km buffer was generated around the point Recreational location of the mooring, warping or berthing site to Transportation Mooring and allow vessel movements beyond the 500m directed Berthing areas in the UN Convention on the Law of the Sea (United Nations, 1982) Axe racing club racing marks were obtained as grid references denoting vertices of the bounding area, from which a racing area polygon was derived. WPNSA racing marks were obtained as grid Recreation sailing references of the centre points, along with individual areas diameter values, for each racing zone. The resulting overlapping circles were merged into one polygon layer clipped to the coastline but retaining the subdivisions for each of the 4 racing zones. The constraint of other marine traffic was assessed AIS vessel traffic using MMO data recorded from AIS Point locations and polygons of wreck sites in the Historical sites Wreck sites FLAG region were obtained from UKHO and Historic England* A 0.5 km buffer was generated around the route of Cables and the pipeline or cable as directed in the UN Obstructions pipelines Convention on the Law of the Sea (United Nations, 1982)

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 49 of 185

A 0.5 km buffer was generated around the point Buoys location of the buoy as directed in the UN Convention on the Law of the Sea (United Nations, 1982) Obstructions – Areas occupied by obstructions were extracted from polygons the UKHO ‘obstructions’ polygon layer*. A 0.5 km buffer was generated around points Obstructions - extracted from the UKHO ‘obstructions’ point layer*, points as directed by the UN Convention on the Law of the Sea (United Nations, 1982) Using the output from the wastewater discharges Modelled spread (section 5.2.3) a polygon layer was generated Water quality of bacteria from containing all the points within the model that had a sewage outfalls temporal maximum of 2 colony forming units or above The Portland Port boundary area was excluded due to an application for a Several Order (see Section. Portland Port 8.1.7.) to manage aquaculture in the area. Future Boundary aquaculture is not excluded but is controlled by the port authority. Other A 5 km exclusion zone was placed around the Current current mussel farm to prevent competition for aquaculture resourcea A polygon layer representing dredged areas, derived Dredged Areas from the Defra Marine DEM bathymetry layer* * Obtained by Cefas via the Defra SPIRE Agreement a Proposed new aquaculture sites such as that proposed by ScallopTech Limited have not been included in this exculsion layer but should be considered in any future application.

Once all the constraining data layers had been generated, they were combined, and clipped to the FLAG region to produce a single GIS polygon layer demarcating area within which aquaculture is deemed not feasible. These “exclusion” layers were specific to the species group and the culture method employed as set out in Table 5. These exclusion layers were converted to raster layers (10 x 10 m resolution) and used to reclassify all areas of the aquaculture potential raster maps generated in Appendix 6 that overlap the exclusion area and set the aquaculture potential value to zero.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 50 of 185

Figure 18: Map showing the location and extent of the constraining factors used to exclude regions in the FLAG area.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 51 of 185

Table 5: List of the constraining factors used to produce the exclusion layer for each species group-culture system.

Finfish Bivalves Crustacean Macroalgae

Bottom Bottom Bottom (motile) Bottom - Bottom - secured secured Open Pen Suspended Constraining factor Suspended (sedentary) Suspended

Routes        Berthing and          Transport anchorage Sailing    

AIS vessel     density Historical Shipwrecks       

Cables/pipelines        Obstructions Other       

Water quality    

Portland Port          Others Current          aquaculture Dredged areas         

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 52 of 185

8 Results 8.1 Potential areas for aquaculture

Marine near shore and offshore areas

An aquaculture potential model was developed to determine possible marine near shore and offshore areas that could support aquaculture. An outline of the process is detailed in Section 6 and the results for each species and aquaculture system combination is shown in Appendix 6. The outcomes of these maps were combined for each species group and each aquaculture system, and finally combined for each species group to indicate the maximum potential for aquaculture of that species group in any given area of the FLAG region (see Section 6.3 for details). The following sections present these results.

8.1.1.1 Potential areas for macroalgae aquaculture Potential areas for macroalgae aquaculture are summarised in Table 6 and mapped in Figure 19 The results from the modelling indicate that there are only very small areas suitable for the bottom-secured culture of macroalgae in the FLAG region. Furthermore, conditions appear to be unsuitable for winged kelp due to maximum temperature of the seawater exceeding that which the species can physiologically withstand. By contrast, the model suggests that there is an area of up to 639 km2 that has suitable conditions for the growth of sugar kelp, dulse and oarweed by suspended culture and does not interfere with current hard constraint activities in the FLAG area. Furthermore, there is 362 km2 that has suitable conditions for the growth of sea lettuce by floating (suspended) containers.

Table 6: List of Aquaculture potential for FLAG region for Macroalgae. FLAG area showing Major environmental factor(s) Culture potential (km2) constraining potentiala Species Method No (percentage of area Exclusions exclusions unsuitable) • Maximum Temperature (100) Bottom- Winged kelp 0 0 • Bathymetry (84) secured (Alaria • Exposure (53) esculenta) • Maximum Temperature (100) Suspended 0 0 • Substrate (23) Oarweed Bottom- • Bathymetry (84) 0.7 0 (Laminaria secured • Exposure (53) digitata) Suspended 791 614 • Substrate (23) Dulse Bottom- • Bathymetry (84) 0.7 0 (Palmaria secured • Exposure (53) palmata) Suspended 768 604 • Substrate (23) Bottom- • Bathymetry (84) 0.7 0 secured • Exposure (53)

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 53 of 185

FLAG area showing Major environmental factor(s) Culture potential (km2) constraining potentiala Species Method No (percentage of area Exclusions exclusions unsuitable) Sugar kelp (Saccharina Suspended 841 639 • Substrate (23) latissima) Sea lettuce Floating • Substrate (23) 528 362 (Ulva lactuca) containers • Exposure (53) a A major factor was defined as any showing more that 20% of the area not suitable for aquaculture

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 54 of 185

Figure 19: Areas of potential for the aquaculture of macroalgae in the FLAG region by bottom-secured, suspended and floating-container systems.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 55 of 185

8.1.1.2 Potential areas for finfish aquaculture The result of modelling the potential of aquaculture of the 6 finfish species is shown in Table 7 and Figure 20. These results show that there is a potential area of 333 km2 that could support the aquaculture of the salmonids (Atlantic salmon, rainbow trout and sea trout). By contrast there appears to be no suitable areas for the culture of Atlantic cod, sea bass and gilthead bream. Atlantic cod is constrained by the maximum temperature of the sea being too high for efficient culture, whereas sea bass and sea bream are constrained by currents speeds that are too high. There is a total of 14 km2 that could support sea bass or sea bream, but this area is excluded by the current aquaculture and transport activities.

Table 7: List of Aquaculture potential for FLAG region for Finfish. FLAG area showing Major environmental Culture potential (km2) factor(s) constraining Species Method No potentiala (percentage of Exclusions exclusions area unsuitable) Atlantic salmon Open 446 333 • Current speed (51) (Salmo salar) pen Rainbow trout Open (Oncorhynchus 446 333 • Current speed (51) pen mykiss) Sea trout (Salmo Open 446 333 • Current speed (51) trutta trutta) pen • Maximum Temperature Atlantic cod Open 0 0 (100) (Gadus horhua) pen • Current speed (51) Sea bass Open (Dicentrarchus 14 0 • Current speed (91) pen labrax) Gilthead bream Open 14 0 • Current speed (91) (Sparus aurata) pen a – A major factor was defined as any showing more that 20% of the area not suitable for aquaculture

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 56 of 185

Figure 20: Areas of potential for the aquaculture of finfish in the FLAG region by open pen systems.

8.1.1.3 Potential areas for crustacean aquaculture

The result of modelling crustacean aquaculture in the FLAG region is shown in Table 8 and Figure 21. For ranching of lobster there is a total area of 437 km2 that could support aquaculture, whereas for suspended container culture the total area is slightly less at 388 km2. The areas showing most potential lie to the west of the FLAG region but there is also a significant area offshore to the east of the region.

Table 8: List of Aquaculture potential for FLAG region for Crustacean. FLAG area showing Major environmental factor(s) Culture potential (km2) constraining potentiala Species Method No (percentage of area Exclusions exclusions unsuitable) European Bottom 513 437 • Current speed (47) lobster (Homarus Suspended 512 388 • Current speed (47) gammarus) a A major factor was defined as any showing more that 20% of the area not suitable for aquaculture

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 57 of 185

Figure 21: Areas of potential for the aquaculture of crustaceans in the FLAG region by bottom (ranching) and suspended systems.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 58 of 185

8.1.1.4 Potential areas for bivalve aquaculture

Potential areas for bivalve aquaculture are shown in Table 9 and in Figure 22. Suspended aquaculture of oyster and mussel shows the greatest potential with a total area of 345 km2, 799 km2 and 755 km2 for Pacific oyster, native oyster and blue mussel respectively across the FLAG region. Furthermore, bottom culture of these 3 species also shows promise with total areas of 437 km2, 437 km2 and 801 km2. In contrast, bottom-secured culture is highly constrained by the bathymetry and exposure levels across the region. Aquaculture of king scallop shows potential across an area of 327 km2 and 280 km2 for bottom culture and suspended culture, respectively. There is only a small region identified for Manila clam aquaculture which is severely reduced when competing activities are excluded.

Table 9: List of Aquaculture potential for FLAG region for Bivalves. FLAG area showing Major environmental factor(s) Culture potential (km2) constraining potentiala Species Method No (percentage of area Exclusions exclusions unsuitable) Bottom 592 437 • Current speed (45) Pacific oyster • Bathymetry (98) Bottom- (Crassostrea 0.7 0 • Exposure (97) secured gigas) • Current speed (45) Suspended 987 740 None Bottom 592 437 • Current speed (45) • Bathymetry (98) Native oyster Bottom- 0.7 0 • Exposure (97) (Ostrea edulis) secured • Current speed (45) Suspended 1144 799 None Bottom 1033 801 None Blue mussel Bottom- • Bathymetry (98) (Mytilus 0.7 0 secured • Exposure (97) edulis) Suspended 1035 755 None Manila clam • Bathymetry (98) (Ruditapes Bottom 2.5 0.2 • Exposure (53) philippinarum) • Substrate (27) King scallop • Current speed (45) Bottom 407 327 (Pecten • Substrate (23) maximus) Suspended 390 280 • Current speed (46) a – A major factor was defined as any showing more that 20% of the area not suitable for aquaculture

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 59 of 185

Figure 22: Areas of potential for the aquaculture of bivalves in the FLAG region by bottom, bottom-secured and suspended systems.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 60 of 185

The above mapping having considered environmental and physical requirements and having included exclusions around hard constraints may now allow for creation of zones for aquaculture. The process of incorporating zoning into marine spatial planning may lead to the designation of allocated zones for aquaculture (AZA), which are specific areas where aquaculture development is prioritized over other uses (Sanchez-Jerez et al., 2016). AZA are a promising technique for managing water quality, particularly in areas where other cumulative uses impact water quality (industrial discharges, microbial loading in wastewater). However, the creation of an AZA does not in and of itself mandate limits on farms or stocking densities, which are key for sustainably managing water quality. This can be overcome through regulation establishing allowable zones of effect (AZE) for fish cages in coastal systems, or a maximum total area (Bone et al., 2018).

Portland Harbour

Portland Harbour and its surrounding areas offer some off the best shelter from storms and waves provided by , the and the harbour breakwaters. This is attractive to aquaculture operations due to the ability to not only secure aquaculture systems against breakaway in storms but for day-to-day operation of the site. This was clearly reflected in the stakeholder consultation exercise where this area was highlighted as being desirable for both bivalve and macro-algae cultivation.

The harbour has previously been used for culture of mussels, oysters and scallops although within different plots than those in use today. Mussels were first classified & harvested from the harbour in 1995 [A. Younger (Cefas) pers. comm] and scallops were first classified in the Portland Harbour Several Order in 1998 (Kershaw et al., 2002).

More recently the areas available within the harbour for aquaculture were relocated and split into four available lease areas. These have recently been used for further growth trials of scallops and is currently being used for a pilot scale research project into the culture of macro-algae. These sites within the harbour are small for commercial viability but are ideal for pilot scale projects. Currently, licensed locations for aquaculture in the East Devon and FLAG area (Lyme Bay - 360 ha; Portland Harbour - 8 ha; the Fleet - 13 ha) comprise only 0.28% of the entire FLAG area.

Portland Harbour Authority Ltd (PHA), is a UK registered company which owns the commercial port operating out of Portland Harbour. It is a subsidiary of Langham Industries (Goulden, 2018). An order in 1997 established Portland Port as the statutory harbour authority for Portland Harbour with powers of control, operation, management and regulation in relation to the harbour and the

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 61 of 185 harbour premises. It provides for the development and for the safe and efficient operation of a commercial port, for harbour conservancy and maintenance; for the management and encouragement of recreation and commerce; and for the conservation of the natural beauty of the harbour and its flora and fauna and applies 24 hours a day, seven days a week in perpetuity.

With regards to aquaculture, the harbour authority’s policy is one of “being supportive of improving the cultivation/ management of the fishery within its jurisdiction but in doing so it must be mindful of the activities of the harbour’s existing users and the environment.”

The harbour authority will treat any application for an aquaculture business on the merits of the proposal, and in doing so it recognises that at present its users manage to go about their business without any negative impact on one another (in the main), and therefore any potential new aquaculture business, should be able to co-exist in a manner that is acceptable to the users of the harbour and the environment. The Harbour Consultative Committee, and other users and interested parties will be consulted, as necessary.

Decision makers and aquaculture businesses considering aquaculture activity outside the harbour authority’s jurisdiction need to be mindful of the joint pilotage area for , access and egress to and from the port from the English Channel and the need for vessels to anchor within Weymouth Bay. (S. Wilson, Portland Port pers comm. 07.10.2019)

Disclaimer

Portland Harbour Authority Ltd have recently commissioned a high-level review of aquaculture opportunities within the Portland Port area (MEP, 2018). Furthermore, they have made an application for a Several Order to manage aquaculture in the Harbour. For these reasons and the limitations of the biological data for nearshore and shallow areas within the harbour the potential areas for aquaculture within the jurisdiction of the harbour authority have not been mapped using the Cefas model.

Portland Mussel Seed Fishery

The area to the east of Portland Bill was subject to regular mussel dredge fishing activity between 1991 and 2014. Between 2001 and 2010 a total of 19,426.8 tonnes of mussels were fished from the Portland mussel beds and landed in the local port of Weymouth. In 2011, due to the proposal for Studland to Portland SAC, the fishery became subject to an annual Habitats Regulations Assessment (HRA) and subsequently adopted a series of restrictive measures for fishing within the SAC. These measures were designed to maintain the integrity of the

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 62 of 185 site. These measures included the introduction of two restricted fishing areas, measuring a total area of 4.638 km2, and a total allowable annual catch of 2,000 tonnes of mussels. At this point the total area of the mussel bed was estimated to be 48.67 km2 and based on a 25% coverage extrapolated to 1 km, the overall biomass of the mussel bed was estimated to be 190,000- 438,000 tonnes. Annual seabed surveys were undertaken by the Authority, examining the condition of the restricted fishing areas, together with two control sites. Over this period there was no significant difference identified in the density and population structure of the mussel beds in response to fishing activity between 2011 and 2013. Anecdotally, prior to 2011, the mussel bed appeared to show no deterioration despite more intense fishing pressure. Between 2011 and 2014 the average annual catch was 830 tonnes and fishing activity predominantly took place between March and July each year. Despite showing relative long-term stability over several years, over the winter of 2014-15 the mussel bed almost completely disappeared. The most likely cause for this is believed to be the impact of the severe 2014-15 winter storms. Since this date, the mussel bed has shown signs of recovery and anecdotal reports from local divers and static gear fishers now suggest that mussel density is approaching pre-2014 levels. (Source: Southern IFCA, 2018).

In 2018 an application to remove seed mussels from east of Portland Bill to be laid in Poole Harbour was received by Southern IFCA. The site is adjacent to the Studland to Portland Special Area of Conservation (SAC). In accordance with the Habitats Regulations, Southern IFCA considered the potential effects of the proposed activity and completed a Test of Significant Likely Effect (TSLE) for the proposed activity. The applicant conducted their own survey of the seabed and looked at the impact to other users of the site. Natural England was consulted. The original seed fishery site fell partially within the SAC but had not recovered sufficiently from the storm attrition of 2014 – 2015. Southern IFCA approved removal of seed mussel from a limited area outside the SAC for one year to 31st December 2019.

Assuming the mussel seed bed continues to recover, then following application and assessment a proportion of this may be available for removal as a resource to support existing and potential new mussel aquaculture ventures within the Southern IFCA district.

Ringstead Bay

Ringstead Bay offers reasonably sheltered area for potential aquaculture ventures and has a previous designation for a wreck to reef project. A recent project located just to the east of the wreck to reef site aimed to set up a trial longline site for ear hanging of scallops. This site will fall within the boundary of

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 63 of 185 an existing MPZ and SAC and is currently going through the process of finalising its position and associated permissions.

The Fleet

The area of the Fleet has an existing Pacific oyster farm, and its owner has an option for another area of further development subject to environmental consenting. The area has been classified since bivalve mollusc hygiene classifications were first introduced in 1992. However, there is probably limited potential for further aquaculture due to the environmental designations and the area being a private fishery.

Active management of bivalve mollusc production areas Active management of shellfisheries in relation to deteriorations in water quality, has been proposed by some stakeholders as a future strategy for some harvesting areas. This is envisaged as involving the closure of shellfisheries for a period in response to a forecast or actual contamination event. Any such strategy would be most effective by considering the causes of contamination and not just reacting to changes in shellfish flesh results. Actions aimed at responding to predicable deteriorations in water quality (e.g. changes in river flow, rainfall, in catchment use etc.) are likely to be most useful in this respect. Closures would need to be long enough to ensure the in-situ removal of pathogens such as viruses. However, management actions in relation to a series of intermittent events could effectively result in almost continuous closures, thus negating the strategy. In most cases and wherever possible, securing of long-term improvements in underlying water quality is preferable (Kershaw et al., 2002).

Fishery Orders The Shellfish Act (1967) made provision for ‘the establishment or improvement, and for the maintenance and regulation, of a fishery for shellfish.’ Under this Act, members of the public or agencies, including local authority bodies, may apply for ‘several’ or ‘regulating’ orders. These allow the management of private and natural fisheries. Whilst challenging for a member of the public to obtain an order this may be a potential option in conjunction with local authorities or agencies. This model has been used to enable fishermen to diversify and have an area for aquaculture.

Cleaner fish The Dorset, Devon and Cornwall stretches of coast have in recent years been targeted for the capture of Wrasse to act as cleaner fish within the Salmon industry. This has proved to be a lucrative option for local fishermen to supplement their traditional catches but has caused concerns around Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 64 of 185 sustainability. Southern IFCA has issued guidance to the fishing industry on sustainable fishing practices for wrasse. The guidance includes a series of species-specific minimum and maximum conservation reference sizes, no take and pot-free zones, closed seasons, and other forms of effort limitation. The industry has demonstrated compliance with these measures and continues to submit daily catch data to the Authority. The Southern IFCA intends to develop further controls and regulation for the removal of fish for use as ‘cleaner fish’ in aquaculture as necessary by 2021.

There has also been a strong driver to produce these fish through aquaculture. One successful local farm now operates within Portland Harbour producing Lumpfish for use as cleaner fish. This has produced several local jobs.

Oysters, Mussels and bioremediation Oyster beds often make a valuable contribution to biodiversity, most obviously by increasing habitat complexity and providing a hard surface for other organisms to settle, in some cases significantly increasing species diversity (Dame, 1996). Being suspension feeders of phytoplankton, bacteria, particulate detritus and dissolved organic matter their production of faeces and psuedofaeces enriches the underlying sediment and provides a food source. Oysters also decrease water column turbidity by filtering large volumes of water and stabilising the sediment which they cover. Destruction of oyster beds, as well as other bivalves, can lead to increased nutrient levels and stronger algal blooms. Experience from New Zealand indicates that fish and crustacea stocks usually aggregate around anchored aquaculture sites meaning that sport/recreational fishing is encouraged at certain sites where boats can moor to the longline buoys. Line caught fish catches at these sites are substantial meaning in New Zealand that new aquaculture developments are increasingly welcomed by local stakeholders (Graham Fielder, MD of FMS pers comm).

Land-based mariculture

Whilst the feasibility of commercial grow out of marine species within land-based systems has been demonstrated, the economics are often challenging (especially for RAS systems). However, the RAS sector has been moving closer towards financially viable and reliable system design for the last few years and is now viewed optimistically and favourably for the future. For either pump ashore or RAS mariculture systems there are certain requirements.

Generic vessel and land-side facilities likely to be required by an aquaculture business have been identified as follows:

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 65 of 185

a. Vessel support and services. b. Vessel mooring facilities c. Vessel in-water servicing facilities. d. Vessel lifting and repair services. e. Access to slipway/loading unloading jetty. f. Access to providers of marine supplies/provisions. g. Fuelling and watering services. h. Access to appropriate operating licences and marine permits. i. Access to pollution/contamination control equipment

Land based marine culture systems. a. Space - may require between 0.1-2.0 ha suitable for erecting buildings and storage areas. b. Seawater intake (suction pump intake not to be more than more than 5 m above MHWM) usage likely to be between 100-3000 m3/day c. Seawater farm effluent return to sea – avoiding recycling through seawater intake. d. Freshwater supply – potable, usage likely to be between 10-100 m3/day e. Single phase electricity supply f. Three phase electricity supply g. Connections to grey water/sewage network or grey water/sewage storage and removal facilities h. Surface water drainage i. Site access control/security j. Vehicular access k. Vehicle parking l. Telecoms and broadband connections m. 24-hour access with scope for staff remaining onsite 24/7 on occasion n. Disposal of non-hazardous biological waste (carcasses/biofouling waste) o. Disposal of dry waste p. Hire of lifting/loading equipment

Land based depuration, packing, processing and distribution. q. Space - may require between 0.1-1.0 ha r. Seawater supply – pumped, usage likely to be between ~10-100 m3/day s. Freshwater supply – potable, usage likely to be between ~10-300 m3/day t. Single phase electricity supply u. Three phase electricity supply v. Connections to grey water/sewage network or grey water/sewage storage and removal facilities. w. Process water storage and disposal/removal x. Surface water drainage y. Site access control/security z. HGV vehicular access aa. Vehicle parking bb. Telecoms and Broadband connections Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 66 of 185

cc. 24-hour access with scope for staff remaining onsite 24/7 on occasion dd. Disposal of non-hazardous biological waste (processed carcasses/shells/exoskeletons/skins) ee. Disposal of dry waste. ff. Hire of lifting/loading equipment gg. Access to secure storage facilities hh. Access to temperature-controlled storage facilities

During the activities to identify potential sites for land-based mariculture in the FLAG area the following were contacted:

Sandie Wilson Environmental & Planning Manager, Portland Port Joanna Rufus Inward investment team, Dorset Local Enterprise Partnership Martin Sutcliffe Dorset & East FLAG Animateur, Dorset Coast Forum. Martin Syvret Aquaculture Development Officer, Dorset Coast Forum Natalie Poulter Dorset Coast Forum Co-ordinator, Dorset Coast Forum.

The most significant location for potential land-based mariculture in the FLAG area exists within the confines of Portland Port with its associated infrastructure and the possibility of using existing buildings or vacant plots. The Port, which already hosts tenants with mariculture interests, can offer more land-based aquaculture opportunities subject to obtaining the necessary consents. Portland Port has the potential to provide sheltered sites with good quality seawater at suitable temperatures for land-based mariculture and has an active management team seeking to develop this sector.

Currently there is a marine support services sector developing in and around Portland Port (e.g. Quest Marine and Carlin Boat Charters) who are adapting their skills associated with offshore wind farms to help aquaculture businesses with cranage, diving expertise and establishment of shellfish farming infrastructure. Though this does not necessarily translate directly to the development of land-based mariculture, the presence of these emerging companies can act as an incentive to attract more specialised aquaculture businesses which will be able to provide the skills and expertise required in the development of land-based facilities within the FLAG area.

In addition to Portland Port the FLAG area may well offer opportunity along the coast for small scale hatcheries within existing areas of infrastructure or buildings at existing harbours, ports or adjoining industrial areas. However, for the creation of larger scale grow out farms then the access to this kind of space will be challenging and probably limited to one or two locations.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 67 of 185

Aquaculture Park concept

In July 2018, the FLAG Aquaculture Development Officer (ADO), met with local stakeholders to discuss the potential for a project that could provide an aquaculture zone with all the necessary permissions, licenses and services in- place to culture shellfish or seaweeds within a controlled water body. At the time of reporting (December 2019) discussions are on-going with a local stakeholder about the Aquaculture Park concept and how this could fit in with local marine stakeholders. Constructive dialogue has helped to develop the concept further and a new project brief will be available shortly. In wider terms, the FLAG Aquaculture Development Officer has discussed this concept with the Sea Fish Industry Authority who may be able to provide funding to develop a local case study. This approach also matches recommendations under the Seafood 2040 strategy in terms of developing Priority Aquaculture/Coastal Development Zones around England’s coast.

Assuming that Portland Harbour would be central an aquaculture park (with its local connections to stakeholders such as Cefas and DCF) it should be noted that:

• Bivalve mollusc cultivation and sanitary survey profiling of the harbour in relation aquaculture and wild production is has previously been undertaken by Cefas (Cefas 2009, Cefas, 2011; Cefas, 2012; Cefas, 2013). • More recently Portland port commissioned a high-level review of aquaculture opportunities within the Harbour and around its’ breakwaters. This review considers the potential aquaculture opportunities only and states that future aquaculture opportunities will require much broader consideration taking account of the harbour’s existing users and the environment (Goulden, 2018). • Mussels were first classified in and harvested from the harbour in 1995 [A. Younger (Cefas) pers. comm].

8.2 FLAG aquaculture areas, footprints and capacity

Current aquaculture within the FLAG area

Pacific Oysters are currently produced within the Fleet within a small spatial footprint. Current production figures representing only a small % of English & Wales production.

On the edges of the Dorset and East Devon Flag situated between three and six miles offshore in Lyme Bay is a mussel farm (Offshore Shellfish Ltd) that is expected to become the largest of its type in European waters covering a total Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 68 of 185 area of 15.4 square km and producing up to 10,000 tonnes per year of native blue mussels (Plymouth University, 2019).

Within Portland Harbour is a RAS system producing significant and increasing numbers of lumpfish for the Scottish Salmon sector.

Other production facilities for macroalgae and scallop within and around Portland are still at pilot or experimental stages.

Footprints for example aquaculture installations

Example footprints for commercial sized aquaculture installations for each of three different species groups representing: finfish (in pens); bivalve shellfish (mussels on suspended long lines) and macroalgae (on suspended rafts) are shown in Figure 23 below. These provide an indicative spatial context of a farm size relative to the overall FLAG sea area and the potential total optimal area available for each aquaculture species within the FLAG sea area.

Location of example aquaculture facilities within overall FLAG district for illustrative purposes (relative footprint size) only i.e. this is not intended to infer suitability of facilities location in other respects (shelter/proximity to port/other facilities, etc). Figure 23: Indicative footprints for potential commercial scale aquaculture (finfish, bivalves & macroalgae).

Statistics on aquaculture and its potential in the FLAG area.

Table 10 below, provides summary statistics on MCZs and aquaculture potential in the FLAG sea area.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 69 of 185

Table 10: Spatial area statistics for the sea area in the E. Devon & W. Dorset FLAG region. % of Area descriptor ha km2 nm2 total† FLAG total sea area to 6 nm*† 135,700 1,357.0 395.64 100.00

Marine Conservation Zones (MCZs) in FLAG sea area. • Studland Bay 400 4.0 1.17 0.29 • Chesil Beach & Stennis Ledges 3,770 37.7 10.99 2.78 • Purbeck coast 27,900 279.0 81.34 20.56 • South of Portland 1,750 17.5 5.10 1.29 • Axe Estuary 33 0.3 0.096 0.02 TOTAL MCZs in FLAG sea area 33,900 339.0 98.84 24.98

TOTAL area covered by one or more conservation designations (MCZ, SAC, SPA, SSSI, Lyme Bay SI) in 81,390 813.9 237.3 60.0 FLAG sea area.

Existing aquaculture in FLAG sea area. • Fleet oyster farm 13 0.13 0.04 0.01 • Lyme Bay mussel lease site 3 (Offshore Shellfish) 360 3.6 1.05 0.27 TOTAL Existing aquaculture in FLAG sea area 373 3.73 1.09 0.28

High (optimal) aquaculture potential in FLAG sea areaa • Bottom-secured seaweed 0 0 0 0 • Suspended seaweed 77,800 778 226.83 57.33 • Floating container 0 0 0 0 • Open pen finfish 33,300 333 97.09 24.54 • Bottom (ranching) of lobster 43,700 437 127.40 32.20 • Suspended culture of lobster 38,800 388 113.12 28.59 • Bottom culture of bivalves 83,500 835 243.74 61.53 • Bottom-secured culture of bivalves 0 0 0.00 0.00 • Suspended culture of bivalves 79,900 799 232.95 58.88 • Combined Aquaculture Biological Potential: Seaweed 63,900 639 186.30 47.09 • Combined Aquaculture Biological Potential: Finfish 33,300 333 97.09 24.54 • Combined Aquaculture Biological Potential: Lobster 43,700 437 127.40 32.20 • Combined Aquaculture Biological Potential: Bivalves 83,300 883 257.44 65.07 TOTAL area of high aquaculture potential (for one or 92,700 927 270.27 68.31 more of seventeen species).

TOTAL area of high aquaculture potential (for one or more of seventeen species) not within a conservation 54,400 544 158.60 40.08 designation.

* FLAG sea area = Beer to Swanage (Poole Entrance) out to 6 nm. a Aquaculture potential score greater than 0.5 (over 50% of the environmental variables are at an optimum level)

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 70 of 185

9 Discussion 9.1 Areas identified with potential

The areas highlighted on the different output maps for specific individual species and their culture systems give both potential aquaculture operators and regulators a better indication of where aquaculture can potentially develop in the future with least conflict.

For bottom culture of bivalve shellfish, the potential lies with mussel, oyster and to a lesser extent scallop. Results from the GIS data indicates little potential for bottom culture of manila clam although smaller localised areas close to shore cannot be ruled out. The areas identified for mussels and oysters cover extensive areas to both the West and East of Portland whilst scallop potential is predominantly to the West and further offshore in Lyme Bay with another smaller potential patch on the eastern side of Portland. However, operating a bottom culture system with potentially damaging dredges would need to be located away from priority habitats and features which at this point have not been clearly defined. The model shows minimal opportunity for bottom secured culture of shellfish.

For suspended bivalve culture the results indicate significant areas with potential for mussels, pacific and native oysters spread to both the East and West of Portland. The potential for scallop aquaculture via ear hanging is currently being explored within the FLAG area. Large areas for potential scallop culture are flagged as suitable to the West of Portland with smaller areas to the East. The ecosystem services and environmental benefits of these types of systems are now being more widely recognised, hence location of these culture types may have a wider fit within areas with environmental designations.

For bottom secured culture of macroalgae there are little, or no areas identified within shallower areas for kelp, oarweed and dulse primarily for depth and light reasons. However, for suspended culture significant areas are shown to be available across the FLAG area for these three species and additionally sea lettuce. The results indicate the area to be unsuitable for the culture of bladderlock (alaria) primarily due to the maximum sea temperature being too high for the species to thrive.

For open net pen culture of finfish, the model shows potential for salmon, rainbow trout and native sea-trout production primarily in the Western side of the FLAG area. The Eastern side shows limited potential due to unsuitable current speeds in the offshore regions. Across the region the maximum sea temperature is above the optimum and this should be considered in the light of climate change. For warmer water species such as seabass and gilthead Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 71 of 185 bream which reside in our waters for much of the warmer months the model indicates the current ongrowing conditions as less than optimal. However, with climate change the situation could change. For Cod, the temperatures are considered sub optimal.

Potential areas were identified for lobster culture either on the bottom (ranching) and via suspended containers across the FLAG areas. These areas are quite extensive to the Western side of the FLAG area but there is also scope to the Eastern side, especially for ranching.

These outputs of the project differ from previous work because we have included and integrated hard constraints with the biological constraints for individual species and used a greater range of larger scale base datasets and information. It has expanded on the species numbers previously considered within UK marine plans and has identified potential not only by species but also by system.

The study also involved three stakeholder consultations at the edges and centre of the FLAG area to allow for thoughts and input from stakeholders. The results and comments provided during the three workshops are provided in Appendix 4. Where possible we have reflected and incorporated the stakeholder comments into the mapping but also into the recommendations for further work. For example, ensuring that the maps of potential areas concur with the limitations due to areas of high tidal energy or that have high levels of recreational angling and commercial vessel movements. Future revisions would need to incorporate areas of high fishing activity should the data become available. However, whilst the range of stakeholder engagement was good the numbers that attended the workshop was low. The attendance from fishermen was disappointing due primarily to the good weather and time of year and possibly from stakeholder fatigue.

Whilst the project has been able to identify biological potential and exclude hard constraints (MMO, 2014) it has highlighted the need for further work to define some of the softer constraints such as what are priority habitats within MPA’s and what type of aquaculture can fit with which type of habitat. Whilst some work has begun on clarifying the compatibilities it was concluded it was beyond the scope of the project to incorporate as it stands. Further work in this area would be beneficial in enabling the correct sort of sustainable aquaculture in the right places but ultimately may still come down to a case- by-case assessment. Another consideration was that of the military areas either used as firing ranges or training areas. It was considered that as these areas are only used periodically this should not be modelled as an exclusion when permission may be possible on the many non-operational days.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 72 of 185

One concern often voiced is about the scale of enterprises and their impacts. Some of the larger potential aquaculture areas identified may raise concern from stakeholders. However, the demonstration of the footprints of typical farms may help to alleviate these concerns and demonstrate that only small footprints are required to produce significant quantities of seafood. These areas are a mere fraction of the areas within the FLAG area or of those being designated as MPA’s. It would also be useful in further studies to demonstrate the minimal visual impact that different types of offshore aquaculture would most likely have at various distances from the shoreline.

Whilst we have demonstrated the likely spatial footprint of a typical suspended longline or finfish farm the project has not had the resource to model various types of carrying capacity (Ecological, Physical and Social) should a number of these farms begin to develop. This would need to be assessed by regulators either as part of the process of designating a licensed aquaculture zone or for smaller individual applications. In addition, any attempt at this stage would be premature in trying to identify farm management areas or any form of disease management zones. However, this would be important for regulators should an aquaculture sector begin to develop, or any problems emerge.

A recent analysis of the previous MMO 1040 aquaculture potential model and recommendations for future improvements (Franco, 2017) made many recommendations for future improvements. Whilst this FLAG mapping project does not try to emulate the MMO’s requirement for a comprehensive planning tool it does address many of the recommendations made in the above analysis. For example:

• it includes a wider range of environmental variables • it maps areas of aquaculture potential within MPA’s assuming that developments would need to be assessed on a case-by-case basis • it maps species specific areas rather than generic aquaculture areas • it covers commonly cultured native species and some emerging species • emerging seaweed interest was included following stakeholder consultation for similar projects focussing of cultivation potential • emerging systems such as IMTA are considered • it covers management of competition for space • it covers areas of exclusion for aquaculture due to hard constraints • includes buffer zones around some hard exclusions but not around MPA’s • considers and includes areas of existing aquaculture.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 73 of 185

However, it was not with the project scope to cover recommendations (Franco, 2017) around: • the likely effects of climate change and its scenarios • the application of carrying capacity assessments • inclusion of disease management zones • areas of exclusive aquaculture zones • production potential for the area • potential likelihood of biofouling • decision making matrices or planning advice.

In addition to previous recommendations the project also took significant steps forward by using hydrodynamic modelling for offshore discharge points thus enabling identification of the areas with the highest risk of contamination of bivalve shellfish. Also, figures are presented for biotoxin testing across the FLAG which helps to provide understanding for future operators.

The model and mapping now combine both the traditional approach where areas of least conflict are identified, as well as preliminary elements of an ecosystem approach and integrated coastal management. Within the report there are many helpful indicators and suggestions for development of the sector within an ecosystem approach.

Data limitations must be taken into consideration. For example, some environmental parameters are unavailable for very near shore locations, and data sourced as priority fish spawning areas were unsuitable or incompatible for use in the project. One key limitation was the lack of data on priority areas used by the under 10 m fishing fleet via iVMS data. Users of the information should be aware that whilst every effort was made to obtain best available data for mapping and modelling, the results come with caveats as to data limitations and imperfections which, in many cases, have not been validated by field surveys. End users are, therefore, encouraged to perform more rigorous investigation, supported by detailed site surveys, when considering all new licence applications.

Whilst the benefits of the project are to highlight potential opportunities for local businesses, any prospective investors will still currently need to go through the statutory permissions that can be found on the Seafish webpage. https://www.seafish.org/article/aquaculture-regulatory-toolbox-for-england. However, after this project there is now potential for the next steps to facilitate the development of aquaculture exclusive zones, where carrying capacities have been estimated, production potential has been assessed and disease management zones implemented.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 74 of 185

9.2 Management & co-location opportunities

With an ever-growing population and increasing demands on the marine environment it makes sense to look for opportunities for management via co- location and sharing of space. Previously the Dorset C-Scope project examined a matrix for co-location and the area has subsequently been examined again (MMO, 2014). Both projects recognise the complexities of co- location and whilst some areas such as exclusions for safety reasons around wind farms are apparent and clear, opportunities for aquaculture co-location are often not so straight forward. This project has used buffers for exclusions around hard constraints by developing a matrix of differing types of aquaculture systems. However, for co-location with softer constraints, different rules or approaches need to be developed and applied for aquaculture.

Co-location with MPA’s

As previously discussed, co-location for aquaculture within MPA’s perhaps offers a significant opportunity but requires clarification of which type of aquaculture can co-locate with which habitat type.

Co-location within military training areas or firing ranges

Opportunity for co-location of aquaculture within military areas will ultimately depend on the types of operation being carried out by the military and its frequency. If a firing range is only active for two days a week it may be possible to operate in a financially viable way within the five days of no activity given suitable permissions and understandings. Permissions should be sought from the military within these areas.

Co-location within Lyme Bay Statutory instrument

The opportunity for development of aquaculture within statutory instruments such as at Lyme Bay was mentioned as part of the stakeholder workshop and warrants further investigation.

Co-location within no take zones.

Another potential opportunity that would warrant further consultation and consideration would be to look at options within no take zones. It was mentioned within the consultation that it would be helpful to compare mapped areas to the potting study as this could potentially identify areas of ‘no take’ which may be ideal.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 75 of 185

Integration of fishing and aquaculture

Recent workshops with fishermen across several different fishing areas identified a level of interest to explore and develop integration with aquaculture opportunities (Jeffery et al, in publication). Examples were found where fishermen have been able to be both fishermen and aquaculturists given a suitable space for their aquaculture.

9.3 Strategic development opportunities

The Marine and Coastal Access Act, 2009 ensures that the exploitation for developments such as aquaculture is carried out in a sustainable way. Subsequently, the South Marine Plan states that “Proposals for sustainable aquaculture in identified areas of potential sustainable aquaculture production will be supported”.

In addition to identifying areas as having aquaculture potential in terms of biological and environmental suitability a High Potential Opportunity for (inward investment in) Aquaculture has been awarded to Dorset from the Department for International Trade. Consequently, Dorset will be promoted as a place where aquaculture may expand. The Southern IFCA has a role in managing that expansion and intend to develop a framework strategy for aquaculture in their district.

In addition, fisheries management plans are being developed for MPA’s as part of FLAG community planning and cover following areas;

o Studland to Portland SAC; o Lyme Bay and Torbay SAC; o Chesil Beach and Stennis Ledges MCZ; o Chesil and the Fleet SAC.

Defra are working closely with the Aquaculture Leadership Group to develop strategies that help develop the seafood sector within England.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 76 of 185

10 Conclusions & Recommendations

10.1 Conclusions

a) This study has highlighted significant potential areas for different species and systems within the FLAG area in which aquaculture could contribute to increasing the share of ‘low-carbon’ protein from the ocean as a substitute for emission-intensive land-based animal proteins and also as a climate change mitigation strategy.

b) This approach and the model could now be replicated across other stretches of the coastline whilst being improved and updated as additional data becomes available and refinements are made.

c) The areas identified for stakeholders, planners and local authorities has mapped species-specific areas via biological and physical potential against constraints for differing systems of aquaculture. Whilst this approach has combined the least conflict approach with an ecosystem approach, any new aquaculture businesses will still be obliged to go through the complexities and challenges of obtaining the necessary permissions.

d) Access to and use of hydrodynamic modelling, whilst not originally incorporated into the study design proved invaluable for assessing the impact of long sea outfalls on microbiological water quality, in this tidally dynamic and hydrographically complex area. Inclusion is a positive step in identifying suitable areas for bivalve shellfish.

e) The engagement of local stakeholders in workshops provided useful additional information and pointers for further research. The attendance of fishers was limited despite a geographic spread of events and timings set to minimise impacts on the working day. Understandably, some reluctance to reveal exact fishing marks was also expressed. The advent of inshore vessel monitoring for the inshore fleet in the next few years should provide valuable data with which to refine spatial assessments and planning considerations.

f) The analysis presented on potential areas for aquaculture in this report is based on best available information within necessarily limited project resources and data availability. It is not the purpose of this work to recommend one area above another area or to try and resolve competing uses for space and marine resources within the 0-6 nm inshore zone, and we would emphasise that the MMO is the competent authority for marine spatial planning and the Crown Estate for granting seabed leases.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 77 of 185

10.2 Recommendations

a) Further examination around the rules for aquaculture within MPA’s should be undertaken to understand potential interactions of different aquaculture species and system combinations with different habitats with special attention to priority habitats. This has been completed for different fishing gear in European marine sites https://www.gov.uk/government/publications/fisheries-in-european- marine-sites-matrix and some work has been done for specific aquaculture sectors https://www.stir.ac.uk/research/hub/publication/618812. This would evaluate potential ecological impacts of differing aquaculture species/system combinations and their likely footprints or areas of effect and would thus aid site selection and more detailed (finer scale) assessments. This would address concerns of stakeholders with an ecological conservation interest. A guidance document should be produced specific to MPA’s in England.

b) Consideration should be given by planners and authorities towards the creation of either aquaculture exclusive zones or ‘Allowable Zones of Aquaculture’ (AZA’s) within the FLAG area to reduce the challenge for new entrants in accessing space and obtaining permissions. This would enable work to model the carrying capacity of the selected zones/areas/species/systems and provide for further development of an ecosystem approach something that is often beyond the abilities of new operators. This study now provides an indication those areas within which aquaculture prioritised zones could potentially be designated with permissions already sought by an appropriate body.

c) The model should be refined further when further data sets become available. For example, the incorporation of iVMS data for the inshore fleet when it becomes available.

d) The use of hydrodynamic modelling could be considered in assessments of other coastal locations especially where there is limited published current data and where contaminant transport is a constraint on development of bivalve shellfish. In addition, hydrodynamic modelling can also identify those areas that are unsuitable for aquaculture systems due to current speed.

e) Land-based planning should be considered to support aquaculture development (via an in-depth study of facilities and support capability) within sections of a coastline that are identified as important for seafood production.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 78 of 185

11 Final Outputs from Project 11.1 Project report

This report is available via the Cefas Data Hub: http://data.cefas.co.uk/ and via the Dorset & East Devon FLAG Virtual Aquaculture Hub: https://www.dors etcoast.com/projects/flag/.

11.2 GIS Shapefiles

Downloadable shapefiles of potential areas for aquaculture pertaining to seventeen species for use in Geographical Information Systems (GIS) have been made available via the Cefas Data Hub: http://data.cefas.co.uk/and via the Dorset & East Devon FLAG Virtual Aquaculture Hub: https://www.dorsetcoast.com/projects/flag/.

11.3 Interactive mapping tool

A ‘layered pdf’ interactive mapping tool is included in the web-based project report (Appendix 8). This comprises layers for seventeen species on Cefas Data Hub https://www.Cefas.co.uk/Cefas-data-hub/ and linked to the FLAG virtual hub https://www.dorsetaquaculture.co.uk/which will be supported for 5 years (2020-2025).

11.4 Stakeholder feedback on potential aquaculture areas maps

A 'Comments Box' is provided on the Dorset & East Devon FLAG's Virtual Aquaculture Hub web site for stakeholders to provide additional feedback which can be considered in any future update of the information provided in the maps of potential aquaculture areas or read in conjunction with these at the time of future updates to the MMO's South Marine Plan.

______

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 79 of 185

References

ABPmer (2015). A Spatial Assessment of the Potential for Aquaculture in Welsh Waters. Report 2384, May 2015, APBmer Marine Environmental Research td, 87 pp. URL: https://gov.wales/sites/default/files/publications/2018-05/assessment-of-the- potential-for-aquaculture.pdf. ABPmer (2017). WEBvision – Renewables V1.0. Interactive data and information portal. ABPmer Marine Environmental Research Ltd. URL:https://vision.abpmer.net/renewables/. ABPmer (2019). UK Atlas of Marine Renewable Energy Gateway. URL: http://www.renewables-atlas.info/. Austin, A. et al. (2017). Seafood 2040 – A Stategic framework for England. Government Appointed Expert Working Group publication, 1-60. Available at gbr188176.pdf (fao.org). Austun (2014). FishMap Môn – bespoke application development case study. Astun Technology Case Study 01/14. URL: https://astuntechnology.com/media/website/ downloads/case-studies/FishMapMonCaseStudy.pdf. Astrium (2011). Creation of a high-resolution Digital Elevation Model (DEM) of the British Isles continental shelf: Final Report. Prepared for Defra, Contract Reference: 13820. 26pp. Bailey, Liz; Owen, K. (2014). Seaweed Harvesting Natural England’s Advice. Natural England, Peterborough. July 2014. URL: https://www.westgateonsea.gov.uk/share d/attachments.asp?f=ffe10000%2D2ad4%2D4f3e%2D90f1%2D2e1e8e3e8103%2 Epdf&o=Algae%2Dharvesting%2DNE%2DSeaweed%2DHarvesting%2DAdvice% 2D2014%2Epdf . Beal, B.F. and Protopopescu, G.C. (2012). Ocean-based nurseries for cultured lobster (Homarus americanus Milne Edwards) postlarvae: Field experiments off the coast of eastern Maine to examine effects of flow and container size on growth and survival. Journal of Shellfish Research, 31, 177 - 193. Benson, B.B. and Krause, D. (1984). The concentration and isotopic fractionation of oxygen dissolved in freshwater and seawater in equilibrium with the atmosphere. Limnology Oceanography, 29(3), pp. 620–632. Berx, B, Hughes, S. (2009). Climatology of Surface and Near-bed Temperature and S alinity on the North-West European Continental Shelf or 1971-2000 dataset. DOI: 10.7489/1900-1. URL: https://data.marine.gov.scot/dataset/climatology-surface- and-near-bed-temperature-and-salinity-north-west-european-continental. Bignell, J.P., Daniels, C., Hendrikz, L. and Auchterlonie, N. (2016). Lobster Grower: Developing the technology to fast track the aquaculture potential for the European Lobster: WP2 Literature Review: Factors affecting growth and survival of Homarus sp., Cefas and National Lobster Hatchery Report. Cefas Contract Report C6261. Issue Date March 2016, 53 pp. Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 80 of 185

Björnsson, B., Steinarsson, A. and Oddgeirsson, M. (2001). Optimal temperature for growth and feed conversion of immature cod (Gadus morhua L.). ICES Journal of Marine Science, 58(1), pp. 29–38. Bœuf, G. and Payan, P. (2001). How should salinity influence fish growth? Comparative Biochemistry and Physiology - C: Toxicology and Pharmacology, 130, pp. 411-423. Bone, J., Clavelle, T., Ferreira, J.J., Grant, J., Ladner, I., Immink, A., Stoner, J., Taylor, N.G.H. (2018). Best Practices for Aquaculture Management - Guidance for Implementing the Ecosystem Approach in Indonesia and Beyond. URL: https://ww w.google.co.uk/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=2ahUKEwjp3 7qI5I_mAhUUQ0EAHeyBBR4QFjAAegQIARAC&url=https%3A%2F%2Fwww.sust ainablefish.org%2FMedia%2FFiles%2FAquaculture%2F2018-Best-Practices-for- Aquaculture-Management&usg=AOvVaw0d_mPEIV5Z_T-T5sgduD2E. Boogert, F.J., Cubillo, A.M., Nunes, J.P., Ferreira, J.G., Corner, R.A. (2018). Online Environmental Feasibility Application (Deliverable 2.5) and Milestone MS15 Release of on-line Environmental Feasibility Application. URL: http://www.aquaspace-h2020.eu/wp-content/uploads/2018/09/Aquaspace- D2-5-WATER-v2Aug2018.pdf Bourget, E. (1983). Seasonal variations of cold tolerance in intertidal molluscs and their relation to environmental conditions in the St. Lawrence Estuary. Canadian Journal of Zoology, 61, 1193-1201. Bourlès, Y., Alunno-Bruscia, M., Pouvreau, S., Tollu, D., Leguay, D., Arnaud, C., Goulletquer, P. and Kooijman, S.A.L.M. (2009) Modelling growth and reproduction of the Pacific oyster Crassostrea gigas: Advances in the oyster-DEB model through application to a coastal pond. Journal of Sea Research, 69(2–3), pp. 62–71. Buchholz, C. M., Krause, G., and Buck, B. H. (2012). Seaweed Biology. Seaweed Biology, 219: 471–493. Capuzzo, E. and McKie, T. (2016). Seaweed in the UK and abroad – status, products, limitations, gaps and CEFAS role. CEFAS Contract Report FC0021. URL: www.gov.uk/government/publications/the-seaweed-industry-in-the-uk-and-abroad. Capuzzo, E., Collingridge, K., Mitchell, P. and Aldridge, J. (2018). Feasibility study: areas suitable for seaweed growth and farming around the English Coast. 1-79. Capuzzo, E., Mengo, E., Kennerley, A. (2019). Seaweed farming feasibility within the Dorset and East Devon Fisheries Local Action Group (FLAG) area. Report 1 - Potential species, farming methods, end uses and benefits. Cefas, 2019. Cardia, F. and Lovatelli, A. (2015). Aquaculture operations in floating HDPE cages. FAO Fisheries and Aquaculture Technical paper, pp. 15–77. URL: www.fao.org/3/a- i4508e.pdf. Cefas (2009). Sanitary survey of Portland Harbour. Cefas report on behalf of the Food Standards Agency, to demonstrate compliance with the requirements for classification of bivalve mollusc production areas in England and Wales under of EC

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 81 of 185

Regulation No. 854/2004.URL: https://www.portlandrnsa.org/uploads/2/3/7/6/2376 7476/final-portland-harbour-sanitary-survey-report-2009.pdf Cefas (2011). Sanitary survey of Portland Harbour. Cefas report on behalf of the Food Standards Agency, to demonstrate compliance with the requirements for classification of bivalve mollusc production areas in England and Wales under of EC Regulation No. 854/2004. URL: https://www.cefas.co.uk/media/5hnpne2b/final- portland-harbour-sanitary-survey-report-2011.pdf. Cefas (2012). Sanitary survey of Poole Harbour 2009 (amended 2012). Cefas report on behalf of the Food Standards Agency, to demonstrate compliance with the requirements for classification of bivalve mollusc production areas in England and Wales under of EC Regulation No. 854/2004. URL: https://www.Cefas.co.uk/Cefas- data-hub/food-safety/sanitary-surveys/england-and-wales/sanitary-survey-reports- 2007-present/. Cefas (2013). Sanitary survey of The Fleet. Cefas report on behalf of the Food Standards Agency, to demonstrate compliance with the requirements for classification of bivalve mollusc production areas in England and Wales under EC regulation No. 854/2004. URL: https://www.cefas.co.uk/media/yxibtupp/fleet- sanitary-survey-2013-final.pdf. Cefas (2013-18). Annual reports on the results of the Biotoxin and Phytoplankton Official Control Monitoring Programmes for England and Wales. Contract Reference: FS115006. URL: Harmful Algal Blooms (HABS) Surveillance Programmes and Monitoring - Cefas (Centre for Environment, Fisheries and Aquaculture Science). Cefas and FSA (2017). Selection of a bivalve shellfish cultivation site: Consideration of faecal pollution issues. Cefas Weymouth Laboratory, March 2017, 9 pp. URL: 202005-selection-of-a-bivalve-shellfish-cultivation-site-suggestions.pdf (cefas.co.uk). Cefas (2018). Annual report on the results of the Biotoxin and Phytoplankton Official Control Monitoring Programmes for England and Wales – 2018. Contract Reference: FS115006. URL: Harmful Algal Blooms (HABS) Surveillance Programmes and Monitoring - Cefas (Centre for Environment, Fisheries and Aquaculture Science). Chopin, T. (2006). Integrated multi-trophic aquaculture. What it is and why you should care …and don’t confuse it with polyculture. Northern Aquaculture, 12, 4. Chopin, T., Buschmann, A.H., Halling, C., Troell, M., Kautsky, N., Neori, A., Kraemer, G.P., Zertuche-Gonzalez, J.A., Yarish, C. and Neefus, C. (2001). Integrating seaweeds into marine aquaculture systems: a key towards sustainability. Journal of Phycology 37: 975-986. Chopin, T., Troell, M., Reid, G. K., Knowler, D., Robinson, S. M. C., Neori, A., et al. (2010). Integrated multi-trophic aquaculture (IMTA)—A responsible practice providing diversified seafood products while rendering biomitigating services through its extractive components. In N. Franz and C. C. Schmidt (Eds.), Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 82 of 185

Proceedings of the Organisation for economic co-operation and development (OECD) workshop “Advancing the Aquaculture Agenda: Policies to Ensure a Sustainable Aquaculture Sector”, Paris, France, 15–16 April 2010: 195– 217 (p. 426). COM (2013). Strategic Guidelines for the sustainable development of EU aquaculture. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions. Brussels, Belgium. COM/2013/229 final. URL: EUR-Lex - 52013DC0229 - EN - EUR-Lex (europa.eu). Costello, C., Cao, L. and Gelcich, S. (2019). The Future of Food from the Sea. Washington D.C.: World Resources Institute. URL: https://oceanpanel.org/sites/def ault/files/2019-11/19_HLP_BP1%20Paper.pdf. Dame, R. (1996). Ecology of Marine Bivalves: An Ecosystem Approach. CRC Press, New York. pp.272. Dare P.J. and Davies G., 1975. Experimental suspended culture of mussels (Mytilus edulis L.) in Wales using spat transplanted from a distant settlement ground, Aquaculture, 6(3), 257-274. Dell, A.I., Pawar, S. and Savage, V.M. (2011). Systematic variation in the temperature dependence of physiological and ecological traits. Proceedings of the National Academy of Sciences, 108(26), pp. 10591–10596. Donlon, C.J., Martin, M., Stark, J., Roberts-Jones, J., Fiedler, E. and Wimmer, W. (2012). The Operational Sea Surface Temperature and Sea Ice Analysis (OSTIA) system. Remote Sensing of Environment, 116, pp. 140-158. EATiP (2017). A Review of the Strategic Research and Innovation Agenda –Our Vision for the future of European Aquaculture. The European Aquaculture Technology and Innovation Platform, 2017. URL: http://eatip.eu/wp-content/uploads/2018/02/EATIP-SRIA-2017.pdf. EAFRO (2017). A Vision on the future of European Aquaculture - where will European aquaculture be in 2030, 2050 and beyond, outlining the main drivers and game ch angers. URL: https://www.nkfv.se/images/pdf/EFARO_vision_future_EU_aquacult ure_2016.pdf. 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), pp. 1793–1817. EU Working Group (2018). Microbiological Monitoring of Bivalve Harvesting Areas. Guide to Good Practise Technical Application. EU Working Group on the Microbiological Monitoring of Bivalve Mollusc Harvesting Areas. Issue 7: December 2018 (under revision). URL: International Guidance - Cefas (Centre for Environment, Fisheries and Aquaculture Science).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 83 of 185

Falconer, L., Telfer, T., Pham, K.L., Ross, L., Kingdom, U. (2017). GIS Technologies for Sustainable Aquaculture. Comprehensive Geographic Information Systems. Elsevier Inc. URL: https://doi.org/10.1016/B978-0-12-409548-9.10459-2. FAO (2006). State of world aquaculture 2006. Food and Agriculture Organization of the United Nations, FAO Fisheries Technical Paper 500, Rome. FAO (2009). Integrated mariculture: a global review. FAO Fisheries and Aquaculture Technical Paper. No. 529. Food and Agriculture Organization of the United Nations, Rome, 183p. FAO (2010). Aquaculture development. 4. Ecosystem approach to aquaculture. FAO Technical Guidelines for Responsible Fisheries. No. 5, Suppl. 4. Rome, FAO. 53 pp. (also available at www.fao.org/docrep/013/i1750e/i1750e00.htm). FAO (2011). Code of Conduct for Responsible Fisheries. Rome, FAO, 91 pp. URL: www.fao.org/docrep/013/i1900e/i1900e00.htm. FAO and World Bank (2015). Aquaculture zoning, site selection and area management under the ecosystem approach to aquaculture.URL: http://www.fao.org/documents/ card/en/c/4c777b3a-6afc-4475-bfc2-a51646471b0d/.

FAO (2017a). Fisheries and Aquaculture topics. Cultured Aquatic species fact sheets- Ostrea edulis. In: FAO Fisheries and Aquaculture Department [online]. Rome. Updated 2017. URL: http://www.fao.org/fishery/culturedspecies/Ostrea_edulis/en. FAO (2017b). Fisheries and Aquaculture topics. Cultured Aquatic species fact sheets- Mytilus edulis. In: FAO Fisheries and Aquaculture Department [online]. Rome. Updated 2017. URL: http://www.fao.org/fishery/culturedspecies/Mytilus_edulis/en. Ferreira, J.G., Hawkins, A.J.S., Bricker, S.B. (2007). Management of productivity, environmental effects and profitability of shellfish aquaculture — the Farm Aquaculture Resource Management (FARM) model. Aquaculture 264, 160–174. URL: https://www.sciencedirect.com/science/article/abs/pii/S0044848606009094. Fitridge, I., Dempster, T., Guenther, J. and de Nys, R. (2012). The impact and control of biofouling in marine aquaculture: A review Biofouling. The Journal of Bioadhesion and Biofilm Research, 28(7), pp. 649–669. Folk, R.L. (2013). The Distinction between Grain Size and Mineral Composition in Sedi mentary - Rock. J. Geol. 62, 344-359. URL: https://www.jstor.org/stable/30065016 ?seq=1#metadata_info_tab_contents. Franco, S.C. (2017). Developing the MMO model for mapping aquaculture potential. Analysis of model MMO 1040 and recommendations for improvement. Valuing Nature Placement. Deliverable 1. Garcia, H. and Gordon, L. (1992). Oxygen solubility in seawater: Better fitting equations, Limnology Oceanography, 37(6), pp. 1307–1312. Garen P., Robert S., and Bougrier S. (2004). Comparison of growth of mussel, Mytilus edulis, on longline, pole and bottom culture sites in the Pertuis Breton,

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 84 of 185

France. Aquaculture, 232(1), 511-524. URL: https://www.sciencedirect.com/scienc e/article/abs/pii/S0044848603005350.

Gentry, R.R., Froehlich, H.E., Grimm, D., Kareiva, P., Parke, M., Rust, M., Gaines, S.D., Halpern, B.S. (2017). Mapping the global potential for marine aquaculture. Nature Ecology & Evolution. 1, 1317–1324. URL: https://doi.org/10.1038/s41559-017- 0257-9. Getchis T.E., (2014). Northeastern U.S., Aquaculture Management Guide, A manual for the identification and management of aquaculture production hazards, USDA, 288pp. Gimpel, A., Stelzenmüller, V., Töpsch, S., Galparsoro, I., Gubbins, M., Miller, D., Murillas, A., Murray, A.G., Pınarbaşı, K., Roca, G., Watret, R. (2018). A GIS-based tool for an integrated assessment of spatial planning trade-offs with aquaculture. Sci. Total Environ. 627, 1644–1655. URL: https://doi.org/10.1016/j.scitotenv.2018.01.133. Goulden, M., 2018. Portland Harbour. A High-Level Review of Aquaculture Opportunities.pdf. Lymington. URL: https://www.portland-port.co.uk/public/files/2C _%2020190530_Annex-C_PortlandAquaculturePotential%20(3123R01C)11_06_1 9.pdf.

Gosling E. (ed.) (1992). The Mussel Mytilus: ecology, physiology, genetics and culture. Developments in Aquaculture and Fisheries Science, 25. Elsevier Science Publishers, Amsterdam, Netherlands. 590 pp. URL: https://www.researchgate.net/ publication/281256336_The_Mussel_Mytilus_Ecology_Physiology_Genetics_and_ Culture. Hall, S.J., Delaporte, A., Phillips, M.J., Beveridge, M. and O’Keefe, M. Blue Frontiers: Managing the Environmental Costs of Aquaculture (The WorldFish Center, Penang, Malaysia, (2011). URL: https://www.researchgate.net/publication/227642849_Blue _Frontiers_Managing_the_Environmental_Costs_of_Aquaculture. Hanelt, D. and Figueroa, F.L. (2012). Physiological and photomorphogenic effects of light on marine macrophytes. In Wiencke, C. and Bischof, K. (eds) Seaweed Biology. Novel insight into ecophysiology, ecology and utilization., pp. 510. Harrison, P.J. and Hurd, C.L. (2001). Nutrient physiology of seaweeds: Application of concepts to aquaculture., Cahiers de Biologie Marine, 42(1-2), pp. 71–82. Hawkins, A.J.S., Pascoe, P., Parry, H., Black, K., Davis, C., Lan, D., McGonigle, C., Moore, H., Newell, C., O‘Boyle, N., O’Carroll, T., O’Loan, B., Pastres, R., Service, M., Smaal, A., Zhang, Z., Zhu, M. (2012). ShellSIM: a user-friendly software tool predicting growth and environmental interrelations in wide range of bivalve shellfish. p. 464, Book of Abstracts, Prague, 1-5 September 2012. World Aquaculture Society, Louisiana State University, Baton Rouge, USA, 1227 pp. Hawkins, A.J.S., Pascoe, P.L., Parry, H., Brinsley, M., Black, K.D., McGonigle, C., Moore, H., Newell, C.R., O’Boyle, N.T., O'Carroll, T., O’Loan, B. Service, M., Smaal,

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 85 of 185

A.C., Zhang, X.L., Zhu, M.Y. (2013a). ShellSIM: a generic model of growth and environmental effects validated across contrasting habitats in bivalve shellfish. Journal of Shellfish Research, 32 (2): 237-253. Hawkins, A.J.S., Pascoe, P.L., Parry, H., Brinsley, M., Black, K.D., McGonigle, C., Moore, H., Newell, C.R., O’Boyle, N., O'Carroll, T., O’Loan, B., Service, M., Smaal, A.C., Zhang, X.L., Zhu, M.L. (2013b). ShellSIM: predicting growth and environmental effects across contrasting habitats in bivalve shellfish. Abstract, World Aquaculture Society Nashville 20-26 February 2013. Heal. R and Capuzzo E. (2019). Identification of Areas of Aquaculture Potential in English Waters, MMO Project No: 1184, May 2019, 109 pp. Hild A. and Günther C-P. (1999). Ecosystem engineers: Mytilus edulis and Lanice conchilega. In: The Wadden Sea Ecosystem: stability properties and mechanisms (ed. S. Dittmann), pp. 43-49. Berlin: Springer-Verlag. Holt T.J., Rees E.I., Hawkins S.J. and Seed R. (1998). Biogenic reefs (Volume IX). An overview of dynamic and sensitivity characteristics for conservation management of marine SACs. Scottish Association for Marine Science (UK Marine SACs Project), 174 pp. HM Government (2011). UK Marine Policy Statement – HM Government Northern Ireland Executive, Scottish Government, Welsh Assembly Government. The Stationery Office, London. March 2011. pp51. URL: http://www.official- documents.gov.uk/. Huang, G., Goblick, G.N., Burkhardt III, W., Woods, J., Calci, K.R. (2009). Exploring the role of GIS in assessing shellfish growing areas in proximity to wastewater treatment plant (WWTP) discharges. URL:https://www.researchgate.net/publication/237324165. Jakobsen, H.H. and Markager, S. (2016). Carbon-to-chlorophyll ratio for phytoplankton in temperate coastal waters: Seasonal patterns and relationship to nutrients. Limnology and Oceanography, 61, pp. 1853–1868. Jeffery, K., Mangi, S.C., Conejo-Watt, H., Muench, A., and Hyder, K. Integrating the English inshore fishery fleet with aquaculture into a more holistic seafood production system. (In Publication). JNCC (2018). Marine Evidence-based Sensitivity Assessment. URL: http://jncc.Defra.g ov.uk/default.aspx?page=6517. Kershaw, S., Morgan, O., Vogt, F. (2002). Initial Review of Priority Bivalve Production Areas. Cefas, Weymouth. Korringa, P. (1952). Recent advances in oyster biology. Quarterly Review of Biology, 27: 266-308. URL: https://pubmed.ncbi.nlm.nih.gov/13014276/. Laing, I. (2002). Scallop Cultivation in the UK: a guide to site selection. Centre for Environment, Fisheries and Aquaculture Science (CEFAS), Lowestoft, Available online at: www.CEFAS.co.uk/publications/files/scallop_cultivation.pdf.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 86 of 185

Laing, I., Lees, D.N., Page, D.J. and Henshilwood, K., 2004. Research on shellfish cultivation. Science Series Technical Report. 122, pp70. URL: http://www.Cef as.co.uk/publications/techrep/tech122.pdf. Laing I. and Spencer B.E. (2006). Bivalve cultivation: criteria for selecting a site. Cefas Science Series Technical Report No. 136, 34 pp. URL: www.Cefas.co.uk/publicati ons/techrep/techrep136.pdf. Lessin, G., Artioli, Y. and Wakelin, S. (2018). NEMO v3.2 ERSEM Model Simulation Output from the NERC-DEFRA-DECC funded project “Regional Ecosystem and Biogeochemical Impacts of Ocean Acidification - a Modelling Study”. Plymouth Marine Laboratory and National Oceanography Centre. Lieberknecht, L.M.; Hooper, T.E.J.; Mullier, T.M.; Murphy, A.; Neilly, M.; Carr, H.; Haines, R.; Lewin, S.; and Hughes, E. (2011). Finding Sanctuary final report and recommendations. A report submitted by the Finding Sanctuary stakeholder project to Defra, the Joint Nature Conservation Committee, and Natural England. Lyngsgaard, M.M., Markager, S., Richardson, K., Møller, E.F. and Jakobsen H.H. (2017). How Well Does Chlorophyll Explain the Seasonal Variation in Phytoplankton Activity? Estuaries and Coasts. Estuaries and Coasts, 40, pp. 1263–1275. McKay D.W. and Fowler S.L., (1997). Review of the exploitation of the mussel Mytilus edulis, in Scotland. Scottish Natural Heritage Review no. 68. Marine Planning Consultants Ltd. (2014a). Lyme Bay Fisheries and Conservation Reserve: Integrated Fisheries Management Plan. Appendix F: GIS processing applied in Habitat Risk Assessment. A report produced for the Lyme Bay Fisheries and Conservation Reserve Working Group, UK, September 2014, pp. 10. URL: https://www.lymebayreserve.co.uk/download-centre/management-plan.php Marine Planning Consultants Ltd. (2014b). Lyme Bay Fisheries and Conservation Reserve: Integrated Fisheries Management Plan. Final report, prepared for Blue Marine on behalf of the Lyme Bay Fisheries and Conservation Reserve Working G roup, UK, September 2014, pp. 64. URL: https://www.lymebayreserve.co.uk/downl oad-centre/management-plan.php. MEP (2018). Portland Harbour – A High Level Review of Aquaculture Opportunities. Report for Portland Harbour Authority Limited. Macalister, Elliott and Partners Ltd., Lymington. MMO (2013a). Spatial trends in aquaculture potential in the South and East Inshore and Offshore Marine Plan Areas. A report produced for the Marine Management Organisation, 222 pp. MMO Project No: 1040. ISBN: 978-1-909452-18-3. URL: htt ps://www.researchgate.net/publication/267209340_Spatial_trends_in_aquaculture _potential_in_the_South_and_East_inshore_and_offshore_marine_plan_areas. MMO (2013b). Spatial trends in aquaculture potential in the South and East Inshore and Offshore Marine Plan Areas – procedures and guidance. A report produced for the Marine Management Organisation, 63 pp. MMO Project No: 1040. ISBN: 978-1-90 9452-20-6. URL: https://webarchive.nationalarchives.gov.uk/20140305103959/http

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 87 of 185

://www.marinemanagement.org.uk/evidence/documents/1040-procedures- guidance.pdf. MMO (2014). Scoping of a robust approach to the assessment of co-existence of activities in marine plan areas. A report produced for the Marine Management Organisation, pp 119. MMO Project No: 1049. ISBN: 978-1-909452-23-7. MMO (2016). South Marine Plan Areas Sustainability Appraisal Report Part 2, November 2016. Marine Management Organisation, Newcastle-upon-Tyne, pp 171. MontyMariner, n.d. Admiralty Derived Tide Stream Charts. URL: http://montymariner.c o.uk/tide-stream-charts/tide-stream-charts-2/. Munro, C. D., and Baldock, B. M. (2012). Lyme Bay Closed Area - Measuring Recovery of Benthic Species in cobble reef habitats – analysis of data collected by SCUBA divers September 2008, August 2009 and July 2010. A Marine Bio-images report, February 2012, Marine Bio-images, Exeter, Devon, UK, pp 38. URL: Lyme Bay - A case study: measuring recovery of benthic species, assessing potential spill-over effects and socio-economic changes (marine-bio-images.com) Natural England, 2018. Designated Site View. South Dorset MCZ, Supplementary advice. URL: https://designatedsites.naturalengland.org.uk/Marine/SupAdvice.asp x?SiteCode=UKMCZ0022&SiteName=south+dorset&SiteNameDisplay=South+Do rset+MCZ&countyCode=&responsiblePerson=&SeaArea=&IFCAArea=. Nehls G. and Thiel M. (1993). Large-scale distribution patterns of the mussel Mytilus edulis in the Wadden Sea of Schleswig-Holstein: Do storms structure the ecosystems? Netherlands Journal of Sea Research, 31, 181-187. URL: https://ww w.sciencedirect.com/science/article/pii/007775799390008G. Neori, A., Chopin, T., Troell, M., Buschmann, A.H., Kraemer, G., Halling, C., et al. (2004). Integrated aquaculture: Rationale, evolution and state of the art emphasizing seaweed biofiltration in modern aquaculture. Aquaculture, 231, 361–391. New Forest District Council, 2017. Update of Carter, D., Bray, M., and Hooke, J. (2004). SCOPAC Sediment Transport Study. URL: https://www.scopac.org.uk/sts/. Newell, C.R., Davis, C., Hawkins, A J.S., Richardson, J., Getchis, T., Morris, K., Cheney, D. (2012a). ShellGIS – a new GIS tool for oyster farm site selection, oyster growth simulation and production carrying capacity. Journal of Shellfish Research, 31(1): 327-327 (Abstract). Newell, C.R., Hawkins, A.J.S., Morris, K., Richardson, J., Davis, C., Getchis, T. (2012b). ShellGIS: a GIS software tool for predicting the growth and environmental impacts of oysters as a function of site selection. p. 462, Book of Abstracts, Prague, 1-5 September 2012. World Aquaculture Society, Louisiana State University, Baton Rouge, USA, 1227 pp. Newell, C. R., Hawkins, A.J.S., Morris, K., Richardson, J., Davis, C., Getchis, T. (2013). ShellGIS: a dynamic tool for shellfish farm site selection. World Aquaculture, 44 (3): 52-55.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 88 of 185

Pastres, R., Borga, F., Brigolin, D., Galspsoro, I., Tett, P. (2018). Guide to the AquaSpace Toolbox. AquaSpace: Ecosystem Approach to making Space for Aquaculture. EU Horizon 2020 project grant no. 633476, 13 pp. URL: AquaSpace- ToolBox-Guide (aquaspace-h2020.eu). Perry, F. and Jackson, A. (2017). Ostrea edulis (Native oyster). In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. URL: www.marlin.ac.uk/species/detail/1146. Pham, L.K. (2012). GIS-based modelling of agrochemical use, distribution and accumulation in the Lower Mekong Delta, Vietnam: A case study of the risk to aquaculture. PhD thesis, University of Stirling, Stirling, 236 pp. Plymouth University (2019). Response Of predators to Protection and Enhancement (ROPE). Project description. URL: https://sheehanresearchgroup.com/rope/ Ren, J.S. and Schiel, D.R. (2008). A dynamic energy budget model: parameterisation and application to the Pacific oyster Crassostrea gigas in New Zealand waters. Journal of Experimental Biology, 361, pp. 42–48. Rivera-Ingraham, G.A. and Lignot, J.H. (2017). Osmoregulation, bioenergetics and oxidative stress in coastal marine invertebrates: raising the questions for future research. Journal of Experimental Biology, 220, pp. 1749–1760. Rödström, E., Jonsson, P.R. (2000). Survival and feeding activity of oyster spat (Ostrea edulis l) as a function of temperature and salinity with implications for culture policies on the Swedish west coast. J. Shellfish Res. 19, 799-808. URL: https://www.resear chgate.net/publication/288448100_Survival_and_feeding_activity_of_oyster_spat_ ostrea_edulis_l_as_a_function_of_temperature_and_salinity_with_implications_for _culture_policies_on_the_swedish_west_coast. Ross, L.G. (2014). The use of Geographical Information Systems in aquaculture: a review. URL: https://www.researchgate.net/publication/237720946. RYA (2007). Identifying Recreational Cruising Routes, Sailing and Racing Areas within the SEA 8 Area. A Report for the Department of Trade and Industry by the Royal Y achting Association. URL: https://www.rya.org.uk/sitecollectiondocuments/legal/W eb%20Documents/Environment/SEA%208%20recreational%20boating%20compr essed.pdf. RYA (2016). UK Coastal Atlas of Recreational Boating GIS dataset. Royal Yachting As sociation (RYA) UK Coastal Atlas of Recreational Boating IS dataset (RYA, 2016). URL: UK Coastal Atlas of Recreational Boating | Planning & Environment | Knowledge & Advice | RYA - Royal Yachting Association. Sanchez-Jerez, P., Karakassis, I., Massa, F., Fezzardi, D., Aguilar-Manjarrez, J., Soto, D., Chapela, R., Avila, P., Rivero, J.C., Tomassetti, P., Marino, G., Borg, J.A., Franicevic, V., Yucel-Gier, G., Fleming, I., Biao, X., Nhhala, H., Hamza, H.A., Forcada, Aitor and Dempster, T. (2016). Aquaculture's struggle for space: The need for coastal spatial planning and the potential benefits of Allocated Zones for

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 89 of 185

Aquaculture (AZAs) to avoid conflict and promote sustainability. Aquaculture Environment Interactions. Volume 8. 41-54. 10.3354/aei00161. Schulte, P.M. (2015) The effects of temperature on aerobic metabolism: towards a mechanistic understanding of the responses of ectotherms to a changing environment. Journal of Experimental Biology, 218, pp. 1856–1866.

Scott, D.C.B. and Muir, J.F. (2000) Offshore cage systems: A practical overview. In: Muir J. and Basurco B.(eds.). Mediterranean offshore mariculture. Zaragoza: CIHEAM, 2000. pp. 79-89 (Options Méditerranéennes: Série B. Etudes et Recherches; n. 30).

Scottish Natural Heritage (2011). The siting and design of aquaculture in the landscape: visual and landscape considerations. Prepared by Alison Grant, Landscape Architect in collaboration with the ASSG and the SSPO URL: The siting and design of aquaculture in the landscape - visual and landscape considerations | NatureScot.

Seafish (2019a). Clam cultivation. Aquaculture Methods Leaflet. Seafish Industry Authority. Seafish (2019b). Lobster hatcheries. Aquaculture Methods Leaflet. URL: www.seafish. org/media/Publications/Lobster_Cultivation.pdf. Seafish (2019c). Native oyster cultivation. Aquaculture Methods Leaflet. Available online at: www.seafish.org/media/Publications/Native_Oyster_Cult_Leaflet.pdf. Seafish (2019d). Pacific oyster cultivation. Aquaculture Methods Leaflet. Available online at: www.seafish.org/media/Publications/PacificOysterCult_leaflet.pdf. Seafish (2019e). Rope-grown Mussel Cultivation Site selection. Aquaculture Methods Leaflet, pp. 52–53. Seafish (2019f). Scallop cultivation. Aquaculture Methods Leaflet. Seafish (2019g) Seabed cultivation of mussels. Aquaculture Methods Leaflet. Seed R. and Suchanek T.H. (1992). Population and community ecology of Mytilus. In The mussel Mytilus: ecology, physiology, genetics and culture, (ed. E.M. Gosling), pp. 87-169. Amsterdam: Elsevier Science Publ. [Developments in Aquaculture and Fisheries Science, no. 25.]. Seed R. (1969). The ecology of Mytilus edulis L. (Lamellibranchiata) on exposed rocky shores 2. Growth and mortality. Oecologia, 3, 317-350. Shellfish Industry Development Strategy [SIDS] (2008). A Case for Considering MSC Certification for Shellfish Cultivation Operations. Sicilian, J.M., Hirt C.W, Harper, R.P. (1987). FLOW-3D: Computational modelling power for scientists and engineers, Flow Science report, 1987 (FSI-87-00-1). Silva, T., Biermann, L., Rees, J. (2016). Suspended Sediment Climatologies around the UK, Report for the UK Department for Business, Energy and Industrial Strategy

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 90 of 185

offshore energy Strategic Environmental Assessment programme. December 2016, pp. 40. Silva, T., Ferreira, J.G., Bricker, S.B., DelValls, T.A., Martín-Díaz, M.L., Yáñez, E. (2011). Site selection for shellfish aquaculture by means of GIS and farm-scale models, with an emphasis on data-poor environments. Aquaculture (2011), doi: 10.1016/j.aquaculture.2011.05.033. Snyder, J., Boss, E., Weatherbee, R., Thomas, A.C, Brady, D. and Newell, C. (2017). Oyster Aquaculture Site Selection Using Landsat 8-Derived Sea Surface Temperature, Turbidity, and Chlorophyll a. Frontiers in Marine Science, 4, pp. 190- 200. Spencer, B.E. (2002). Mussel cultivation. Molluscan Shellfish Farming, pp.147-165. Strand, Ø., Louro, A. and Duncan, P. (2016). European Aquaculture. In Shumway, S.E. and Parsons, G.J. (eds.) Scallops: Biology, Ecology, Aquaculture, and Fisheries. 3rd Edition. Elsevier Science. Chapter 20. Tacon, A.G.J. and Metian, M. Fish matters: importance of aquatic foods in human nutrition and global food supply. Rev. Fish. Sci. 21, 22–38 (2016). The Crown Estate (2019). Resource and Constraints assessment for offshore Wind: Methodology Report 38255-TCE-REP-024, pp 174. URL:https://www.thecrownest ate.co.uk/media/3331/tce-r4-resource-and-constraints-assessment-methodology- report.pdf. Tillin, H.M., Hull, S.C., Tyler-Walters, H. (2010). Development of a Sensitivity Matrix (pressures-MCZ/MPA features). Report to the Department of Environment, Food and Rural Affairs from ABPMer, Southampton and the Marine Life Information Network (MarLIN) Plymouth: Marine Biological Association of the UK. No. MB0102 Task 3A, Report No. 22. URL: https://www.marlin.ac.uk/assets/pdf/MB0102_Task 3-PublishedReport.pdf. Tyler-Walters H. (2008). Mytilus edulis Common mussel. In Tyler-Walters H. and Hiscock K. (eds) Marine Life Information Network: Biology and Sensitivity Key Information Reviews, [on-line]. Plymouth: Marine Biological Association of the Unit ed Kingdom. URL: http://www.marlin.ac.uk/species/detail/1421. United Nations (2017). Department of Economic and Social Affairs, News 21 June 2017. URL: https://www.un.org/development/desa/en/news/population/world- population-prospects-2017.html. United Nations (1982). Convention on the Law of the Sea. URL: https://www.un.org/de pts/los/convention_agreements/texts/unclos/unclos_e.pdf . Warren, I.R., Bach, H.K. (1992). MIKE 21: a modelling system for estuaries, coastal waters and seas. Environmental Software, 7 (4), 229-240. https://www.sciencedire ct.com/science/article/pii/026698389290006P. Wieland, T. and Paul, J.D. (1983). Potential for scallop culture in the UK. S.F.I.A. Tech. Rep. No. 224. 24 pp.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 91 of 185

Williams R.J. (1970). Freezing tolerance in Mytilus edulis. Comparative Biochemistry and Physiology, 35, 145-161. URL: https://www.sciencedirect.com/science/article/ pii/0010406X70909187. Wood, C. (2007). Seasearch surveys in Lyme Bay, June 2007, A Report to Natural England. 26 pp. URL: https://www.google.co.uk/url?sa=t&rct=j&q=&esrc=s&source =web&cd=1&ved=2ahUKEwjc_r7L4o_mAhVGTcAKHSQsCZcQFjAAegQIBBAC&u rl=http%3A%2F%2Fpublications.naturalengland.org.uk%2Ffile%2F188304&usg=A OvVaw3VljpECK1mIa8aq14lCwBa.

Acknowledgements

This research was funded by grant from the EMFF Dorset and Devon Fisheries Local Action Group (FLAG), with matched funding from the Department of Environment, Food and Rural Affairs (Defra). Additional project enhancement work including hydrodynamic water quality modelling work was funded by Cefas Science Futures and the Fishmongers’ Company’s Fisheries Charitable Trust.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 92 of 185

Appendix 1: Literature Review

Review of previous aquaculture suitability studies and tools

A1.1 Data collection

ABPmer (2015), Appendices A and B, provides a comprehensive list of data layers and sources collated as part of the spatial assessment, including identification of unused data, with reasons for their exclusion where relevant.

A series of surveys conducted in the Lyme Bay area led to locally detailed data collection within the Dorset FLAG area over the period 2007-2010. These surveys were arranged (i) to measure recovery of benthic species in cobble reef habitats of the Lyme Bay Closed Area (Munro and Baldock, 2012); and (ii) on behalf of Natural England as an element of gathering and collation of information by recreational SCUBA divers, on seabed habitats and wildlife, under the voluntary Seasearch (www.seasearch.org.uk) initiative (Wood, 2007). The latter survey arose partly from interest in the effects of grounding of the MSC Napoli off Beach, East Devon, in January 2007, but also repeated 2004 and 2006 surveys of the pink sea fan population at same site, to assess reported impact of scallop fishing on diversity of sessile fauna on low lying reefs.

The Lyme Bay closed area monitoring analysis (Munro and Baldock, 2012) used data collected by SCUBA divers at 10 fixed sites across Lyme Bay in 2008, 2009 and 2010. The report includes detail on monitoring and survey design. Each site surveyed represented a different type of closure, but all on similar habitat, with visually prominent taxa recorded by divers in situ on fixed transects. Changes in benthic assemblages, including species richness and abundance over time, were investigated using multivariate statistical methods, which indicated distinct differences between different closure methods. While the case study aimed to measure recovery of benthic species, it also tried to assess potential spill-over effects and resulting socioeconomic impacts of the closure.

The Lyme Bay Natural England/Seasearch (Wood, 2007) survey employed methods to record qualitative information on seabed composition and features of individual habitats, including species associated with each habitat, using the SACFOR (Superabundant-abundant-common-frequent-occasional-rare) scale. Dive times were recorded, dive positions were captured by GPS and depths and temperatures were logged using dive computers. During dives, extensive use was made of digital photography and post-survey, depths were adjusted to chart datum, JNCC biotopes were identified for each habitat and all data were entered into the Marine Recorder database (http://jncc.Defra.gov.uk/page-1599). Sites covered during the survey were: Beer Home Ground (50° 38.27’N 003° 02.79’W); West Tennants Reef (50° 38.80’N 002° 57.78’W); Dogleg Reef and Sunset Ledge, West

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 93 of 185

Bay (50° 40.76’N 002° 50.14’W; 50° 41.08’N 002° 48.03’W); Budleigh Outer Reef (50° 37.204’N 003° 18.798’W); Woodbury’s Ground (50° 34.760’N 003° 18.496’W); ‘In between’ (50° 33.922’N 003° 19.685’W); Tardis Reef (50° 33.265’N 003° 16.475’W). Findings from these sites are detailed fully in Wood (2007), but there was evidence of a general decline in condition in Lyme Bay since 2001/2002, due to benthic fishing and disease. Main litter was from angling waste and abandoned lobster pots, while main damage was due to impact from bottom trawling for scallops. Disease in sea fans was believed to be bacteriological, possibly linked to increased water temperature.

A1.2 Data synthesis and mapping

The current work examines several studies where existing data have been used to synthesise new information, including incorporation within a GIS for analysis and/or mapping to derive spatially explicit information. A small selection of such studies is summarised here.

There are several interactive mapping/viewing tools, including:

• C-SCOPE Coastal Atlas: output from the EU Interreg IV ‘Two Seas’ programme (http://www.cscope.eu/en/) (Combining Sea and COastal Planning in Europe), in partnership with Dorset Coast Forum – currently supports static mapping only, centred on Belgium with limited spatial coverage – http://www.compendiumkustenzee.be/geoviewer/#!/ o the tool uses Open Street Map as its base map; o and includes interactive graphs and charts – e.g. http://www.compendiumkustenzee.be/en/interactive-graphs-charts

• Dorset Explorer: https://explorer.geowessex.com/ o interactive online mapping tool for the region o essentially a ‘viewing’ tool for directions, locations, features of interest o allows map layers to be switched on/off, manual editing using drawing tools, etc.

More topic-oriented mapping exercises included, in 2010, development by Cefas of a set of national data layers providing information on the spatial extent of fishing activities, including intensity, for inshore waters around England and Wales. An attempt was also made to provide a layer on the economic value of these fishing grounds. Inshore fishing vessels are often <15 m long, therefore their movements are not captured in the same way as larger vessels, which can be monitored via Vessel Monitoring Systems (VMS). However, the collated inshore data included the distribution of MMO sightings from aerial surveillance and Royal Navy Fishery Patrol Vessels. The layers produced were accompanied by further layers indicating confidence in the data. Outputs were compared with those of the Finding Sanctuary initiative (Lieberknecht et al., 2011), which worked with stakeholders to design

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 94 of 185

Marine Conservation Zones (MCZs) in the seas around southwest England as part of a wider network of Marine Protected Areas (MPAs). The rationale for this comparison was to ascertain differences between the two approaches, and to develop an approach which may allow information from a questionnaire-based approach to be integrated into a future update of the National Inshore Fisheries Data layer.

Huang et al. (2009) used GPS and data-logger technology to collect sample data and visualise process results within a GIS framework, to assess shellfish growing areas in proximity to wastewater treatment plant discharges. Their use of time- sequence maps aided decision-making on appropriate size of prohibited zone, conditional zone, restricted zone and approved zone.

Other GIS assessments include Falconer et al. (2017), who make general remarks on issues surrounding data collection, database construction, spatial analysis, model development and decision support. They highlight the facility of Python scripting within the GIS for automated modelling processes, and the ability to use GIS with different degrees of complexity, across a range of stakeholders in the aquaculture sector, from visualisation and capacity-building to decision-making. This type of system is further facilitated by the growing use of open-source software such as QGIS, removing the requirement for expensive licences to run GI tools. Falconer et al. (2017) emphasise how employing GIS in this way allows stakeholder feedback to be incorporated into model structure, results and outcomes, and opens the door for web-based applications for data- and information-sharing.

Of direct relevance to the E. Devon and Dorset FLAG area is a report to the Lyme Bay Fisheries and Conservation Reserve Working Group, Appendix F of which provides a detailed methodology of GIS processing applied in the Lyme Bay Habitat Risk Assessment (Marine Planning Consultants Ltd, 2014a). This includes details of layers used/produced (habitat map, benthic species, bird foraging areas/distances, fishing activity) and risk matrix-based mapping procedures. Features of conservation interest were combined with fishing activity to produce maps of (a) baseline risk that a feature (species or habitat) is to a fishing activity; and (b) risk based on feature-fishing gear interaction, according to spatial overlap.

A1.3 Natural and anthropogenic environmental considerations and resource pressures

In terms of environmental considerations, analyses have shown that suboptimal conditions can occur because of both natural events and human activity. A study by Silva et al. (2016) analysed the remote sensing observations over an 18-year period to assess after-effects of construction of two large wind farms on non-algal suspended particulate matter (SPM). Although SPM appeared to increase during the period after construction, it was not possible to quantify the effect at the spatial

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 95 of 185

and temporal scales assessed. It was also concluded that neither aeolian deposition of Saharan dust nor dust from actual (Icelandic), or potential, volcanic eruptions were likely to affect the turbidity of UK waters. In their assessment of GIS technologies for sustainable aquaculture, Falconer et al. (2017) examined potential consequences of farming certain species in particular locations. Potential negative consequences included: habitat destruction; loss of biodiversity; damage to facilities; loss of stocks; introduction of non-native species; emergence and spread of disease within the sector. As well as potential environmental impacts and interactions with the wider ecosystem, they advised that attention should be paid to the needs of other resource users and activities within the area in question. This is supported by findings of the integrated management report (Marine Planning Consultants Ltd, 2014b) in consideration of the Lyme Bay Fisheries and Conservation Reserve, which is named as a marine biodiversity hotspot with nationally-acclaimed subtidal reefs: the reserve’s biodiversity makes it important for commercial fisheries, this area having supported some of the best scallop fishing grounds in the UK prior to introduction of management measures in 2014, under candidate Special Area of Conservation (cSAC) designation.

A1.4 Spatial modelling and defining potential areas for aquaculture

Various tools and techniques have been developed and utilised to aid the assessment of areas suitable for aquaculture. ABPmer (2015) suggest that a good starting point is to evaluate areas where aquaculture is already present which are, by default, already considered to be suitable. The main constraint in these locations, then, would be the capacity to expand within the present boundaries. Other restrictions would come from: Marine Spatial Planning (MSP) constraints (e.g. nature conservation designated sites, areas of other marine industry activity, infrastructure and exclusion zones, recreational activity etc.); and investment- dependent constraints (e.g. proximity to landing ports and depuration facilities).

In a still relevant (in terms of the basic concepts) review, Ross (2014) noted the slow uptake of GIS for aquaculture modelling purposes. Examples presented (many of which were from the 1990s) included regional investigations for areas of Africa and Latin America, using relatively simple resource availability and environmental models. There were also several national or state-level investigations that incorporated environment, infrastructure, resource availability and socio-economics - elements particularly useful for guiding national plans, addressing food security issues and investigating conflict and trade-off between different economic activities. Site selection was assessed from meso-scale to local, and facility locations were determined after broad preliminary assessments. Ross argued that, in conjunction with remote sensing and direct data collection, a GIS framework could form a foundation for ongoing monitoring of a site. The Ross review examined the range of GIS tools available (overlay of thematic data; highly flexible data management and archiving; inclusion of more complex computational Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 96 of 185 models; a range of outputs including static maps, table, 3D visualisations). A wide variety of data source types were available to the GIS, as well as a range of analytical tools and methods, and the facility to link with external modules. Illustrative applications in aquaculture included: suitability models (environmental, species-specific, system-specific, facility location); dynamic modelling (waste dispersal, fish growth models); time-series modelling (e.g. mangrove degradation in Brazil).

Huang et al. (2009) used dye studies in a spatial framework to determine exclusion zones around wastewater treatment plants (WWTP). Within a GIS, ‘zones of dilution’ were derived from GPS data, aiding site selection for monitoring stations, thereby saving labour costs and increasing data efficiency. They found that the GIS was also useful for finding outlier data in their analysis.

In an integrative methodology for site selection for shellfish aquaculture, Silva et al. (2011) combined GIS and dynamic farm-scale carrying capacity modelling in three stages: (i) analysis of regulatory and social spatial restrictions to generate a constraints map; (ii) Multi-Criteria Evaluation considering the spatial criteria (sediment, water and ecological quality data) and constituent factors (physical, growth and survival, product quality and environmental sensitivity) to generate a final map based on a spatial suitability index3; (iii) detailed analysis of production, socio-economic outputs and environmental effects of suitable areas using the FARM model (Ferreira et al., 2007). To emphasise the benefits of applying their methodology in environments burdened with a combination of social difficulties, data scarcity and aquaculture expansion pressures, Silva et al. (2011) tested the application for Pacific oyster suspended longline culture in south central Chile, a developing country where aquaculture expansion has led to significant environmental (e.g. organic enrichment of sediments; eutrophication) and management problems. For example, previous uncontrolled expansion of salmon aquaculture in this area led to: significant loss of benthic biodiversity; localised changes in physico-chemical characteristics of sediments; chemical contamination (e.g. from pharmaceuticals); increase in frequency and duration of dinoflagellate blooms; potential impacts of farmed fish escapes on native species; a 2 to 5-fold increase in omnivorous diving and carrion-feeding marine birds in salmon farm areas. At the time of their publication, Silva et al. (2011) suggested that current GIS approaches didn’t include dynamic models (such as DEPOMOD see, https://www.s ams.ac.uk/science/projects/depomod/; FARM (Ferreira et al., 2007); MOM - http://w ww.ancylus.net/Filbas/MOM/Manual_MOM_v3_2.pdf) to estimate carrying capacity or determine temporal variability of environmental effects. They argued that the integrative methodology presented could assist the decision-making process and

3 Although there was the option to apply a weighted mean to each of the spatial variables, ‘expert’ weightings can be inconsistent or subjective, therefore no weightings were applied in the Silva study. Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 97 of 185 reduce socio-economic and environmental problems associated with aquaculture expansion in developing countries. (N.B. Variables, data range values, etc, for Pacific Oyster production, are provided in Silva et al. (2011), Table 1.)

The Silva et al. (2011) analysis judged the following to be important factors governing different categories of site suitability, including observed seasonal differences in water quality parameters and spatial variability in faecal coliform levels and organic matter content:

• Physical suitability water circulation, currents, bathymetry, sediment grain size • Species growth and survival temperature, food concentration, salinity • Water quality criteria temperature, salinity, total particulate matter, dissolved oxygen, chlorophyll-a, particulate organic matter • Product quality faecal coliforms, metals • Environmental sensitivity benthic macrofauna diversity considered to be an indicator

From a more commercial and/or management perspective, the site selection software, ShellGIS (http://www.shellgis.com/Default.aspx), was developed as a mapping tool for shellfish growers, water body managers and regulators. The tool is a custom application of STEMgis (http://www.discoverysoftware.co.uk/STEMgis.htm), which handles additional dimensions of time and depth. Within this tool are embedded state-of-the-art models: ShellSIM (Hawkins et al., 2012; 2013a, b); MIKE 21 (Warren and Bach, 1992); and FLOW 3D (Sicilian et al., 1987) that account for interactive effects of culture type (i.e. suspended or bottom), shellfish density and hydrodynamics (and consequences for food supply, seawater temperature, salinity, seawater oxygen saturation and aerial exposure). ShellGIS is a transferable tool which, for a choice of 14 commonly cultured shellfish species (http://shellsim.com/Species.aspx), enables dynamic predictions of shellfish production, environmental effects and profitability, according to culture practice and site selection within a given GIS domain (Newell et al., 2012a, b; 2013). It is designed to assess site suitability once potential locations have been identified and is available to shellfish growers for a nominal fee, subject to licence and collaborative work agreement, appropriate training and preparation for a given geographic area.

The MMO (2013a) developed a model to identify aquaculture trends in the context of the growing importance of aquaculture in supporting food security, in line with the objectives of the UK Administrations (UK Marine Policy Statement – https://www.gov.uk/government/publications/uk-marine-policy-statement). The MMO model, which included a high-level cost benefit analysis of potential economic returns for each aquaculture type, was the first assessment of its type on the spatial potential for aquaculture in East and South MPAs. The objective was to produce a baseline model, that could be further developed as more data become available, to

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 98 of 185 enable industry to identify broad suitability areas that could then be the subject of more in-depth analysis and site-specific investigations, as required. Considering six main types of aquaculture (bottom culture of bivalve shellfish; trestle/bag culture of shellfish; rope grown bivalve shellfish; macro algae production; hatchery-reared lobster restocking; marine finfish cage culture) the tool highlighted the spatial distribution of natural resource limitations, as well as features and activities that interact with aquaculture, and proximity to local services. For each aquaculture type, a model was designed to be adaptable for different types of equipment. The models did not focus on assigning scores of importance but provided flexibility to aid the decision-making process (e.g. buffers were assigned around areas not normally defined by distinct boundaries, such as sensitive habitats or protected areas). The models also incorporated fewer static data to reflect, e.g. the density of mobile species and intensity of fishing and shipping activity. A temporal dimension was added to identify areas currently suitable and those with potential for future development. The analysis identified significant potential for future expansion of the industry in the East and South Plan areas, although the East Plan area was limited in terms of bottom culture and lobster restocking due to lack of rocky habitat. Mussel culture was found to give the best potential economic returns. Macro algae production was still at an early stage and, therefore, the economics of such were poorly defined. It was acknowledged that developments further offshore were currently limited due to technological constraints, operating costs and risks associated with operating in open, unprotected waters.

The main structure of the MMO (2013a) model was as follows:

• GIS layers were developed for each constraining feature (e.g. substrate, depth, energy) and overlaid to identify initial suitability areas; • Proximity to essential infrastructure was added to account for economic and investment constraints (suitable landing ports; locations of hatcheries, purification units, processing plants; location of potential business base; proximity to invasive non-native species); • Cost-benefit models were applied to inform economic and commercial considerations; • Quality assurance was performed (data accuracy/collection methods; supporting evidence on interpretation of final maps; evaluation of contributing data; independent ground-truthing of modelled or interpolated data; expert opinion; support for model users); • Compiled metadata were compliant to MEDIN standards.

Final outputs of the MMO (2013a) model identified and mapped areas of current, near future and future, potential development per species/aquaculture method, including infrastructure and planning constraint areas, investment factors and relative investment required. Detailed modelling steps, including data processing requirements, are given in MMO (2013b).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 99 of 185

With a focus on sustainable aquaculture, Falconer et al. (2017) reviewed the application of GIS in aquaculture planning and management in its entirety, from data collection and database construction, through spatial analysis, model development and decision support, to capacity building and stakeholder engagement. They also highlighted the relatively recent innovation of web-based applications, opening a range of new possibilities for GIS. Selected applications illustrated:

• site selection and carrying capacity assessment o weighting of factors (baseline, terrestrial, environmental, infrastructural and societal constraints) o ranking of site suitability using different reclassifications (Boolean, user- defined, hard, Fuzzy) • spatial analysis with temporal and time series data • alternative and worst-case scenario modelling • competition for space and natural resources • environmental impact assessment • socio-economic and livelihood analysis • climate change (assessment of potential climate impacts).

Falconer et al. (2017) provided an example of multicomponent GIS via a case study of the Mekong Delta. The case study noted the potential for aquaculture systems to be impacted by other activities in the surrounding area, such as agrochemicals used in crop production in neighbouring terrestrial areas. This study formed one element of the thesis of Pham (2012), who discussed how GIS can be used to investigate the wider environmental pressures and potential cumulative issues in such assessments.

A guidance document produced by Cefas and the Food Standards Agency recommends site-specific investigations prior to making significant investment in potential production sites for bivalve shellfish, and that the local Environment Agency (EA) office should be contacted for information on location of combined sewer overflows (CSOs), storm overflows and related water quality issues. Furthermore, the document recommends that sites are situated as far as possible from sewage outfalls, since the effects of larger discharges can extend several km away. It was also highlighted that septic tanks may have significant localised effect if poorly maintained. Other risks identified included: proximity of farming activity to watercourses (slurry stores, muck spreading – especially just before, or during, heavy rain); access to streams by cattle and other livestock; boating activity, including marinas – the Good Practice Guide (EU Working Group, 2018) suggests a >300 m buffer zone around marinas should be excluded from establishment of a shellfishery. Although difficult to monitor and/or control, wildlife could also be a potential source of faecal contamination.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 100 of 185

A1.5 Management requirements

The Lyme Bay Integrated Fisheries Management Plan analysis (Marine Planning Consultants Ltd, 2014b) comprised: a desk review; data collection and consultation; habitat and species risk assessment; fisheries sustainability assessment; final synthesis, review of management practice and options for integrated management. A comprehensive SWOT analysis (Strengths, Weaknesses, Opportunities and Threats) brought together the findings of the assessment findings and identified attention needed in the following areas: • an understanding of how fisheries affect mobile species of conservation importance both inside and outside the Reserve, and the need for implementation of monitoring and strategy to reduce risks; • an analysis of how recreational fishing contributes to the total fishing effort for bass; • an acknowledgement that some stocks are not within sustainable limits and that all fisheries need clear management plans; • the need for guidance with respect to commercial diving to minimise risk to sensitive seabed habitats and species.

A risk matrix was compiled for nine different types of fishing activity versus the full range of habitats, benthic organisms, fish, mammals, turtles and birds in the area (Marine Planning Consultants Ltd, 2014b, Table 1). Additionally, the report presented fisheries sustainability assessment results for five key fisheries. The range of management options considered were:

• Jurisdiction (within 0-12 nm offshore); • Designated Areas (Lyme Bay and Torbay Special Area of Conservation); • Current (IFCA) Byelaws (Devon and Severn IFCA; Southern IFCA District); • National Management (MMO licence conditions); • EU Fisheries Management guided by CFP (EC 850/98 – Technical measures for the conservation of fishery resources); • Other regulatory measures, including: Wildlife and Countryside Act 1981, Natural Environment and Rural Communities Act 2006, EU Habitats Regulations 2010; • Voluntary Management (conservation reserve; voluntary codes of conduct; commercial fishing – use of VMS, effort limitation, escape hatches, ban on use of undersized fish and shellfish as bait, moving nets to prevent significant bycatch of crustaceans; recreational sea angling – return of undersized fish, catch and release, prevention of ‘anchor drag’, fishing for own consumption only).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 101 of 185

A1.5 Co-location opportunities, and constraints to development ABPmer (2015) proposed additional opportunities for aquaculture expansion, including co-location with offshore wind farms and/or tidal lagoons. However, these would have to be considered in the context of: • Natural resource constraints (e.g. water depth, substrate, temperature); • Investment-related constraints, which would affect the viability of an aquaculture business, including: o distance of cultivation site from landing port o distance to depuration facility o location of invasive non-native species o confirmed designation areas of disease (e.g. Bonamia ostreae); • Licensing and consent of aquaculture with respect to: o deposition of equipment on seabed o navigational risk o bottom cultivation of shellfish out to 6 nm o planning permission o Fish Health Inspectorate (FHI) authorisation o consent for discharges from a fish farm, or marine licence for discharge from a boat (finfish farming only) o abstraction licence (for finfish/plant aquaculture) o mussel seed licence; • Interaction with other marine sector activities and associated infrastructure; • Disease (transfer to/from other sites, human health concerns with respect to mollusc consumption); • Available technology.

Appendix B of the ABPmer (2015) report details spatial model parameters and rules, and includes tables listing constraint data layers, with reasons for exclusion from the model, where relevant.

A1.6 Model and data limitations

Several limitations were identified in the MMO (2013a) model, including: • absence of water quality parameters; • no consideration of visual impacts; • a marine recreation layer was under development, but not available in time for inclusion in the model; • the need for more detail on size and spacing of aquaculture sites; • no assessment of carrying capacity and production levels; • limited availability of surveyed seabed substrate data.

It was also noted that the sensitivity of mobile species to aquaculture developments is likely to be temporally variable, but that gaps in mobile species observation data made this hard to evaluate. There were also gaps in knowledge on the interaction between aquaculture and fishing activity, and loss of spatial detail occurred due to

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 102 of 185 different aquaculture types being combined into the same groups, even though they had different ranges of physical/environmental requirements. The MMO report highlighted that model outputs can only be as good as the lowest-quality inputs, and that output quality was largely affected by data limitations, availability etc. This added to the challenge of building a repeatable, reusable model. Particular attention was drawn to the following classes of data, which were deemed unsuitable for robust modelling: • data layers with too coarse a resolution (e.g. the 30 x 30 nautical mile gridded layer for fish spawning grounds); • point data used to represent features that vary in size and extent; • datasets for which further manipulation would be required to convert to suitable model input; • data whose descriptive nature was too coarse (e.g. “net fishing” could represent one of several types of fishing activity).

The Lyme Bay report (Marine Planning Consultants Ltd., 2014b) summarised several shortfalls to be addressed in terms of integrated management for fisheries and conservation, including: • the need for better understanding, monitoring and management of fishery interactions with Endangered, Threatened and Protected (ETP) species; • need for fishery-specific management practice; • lack of protection of reef outside the Lyme Bay Reserve; • need to understand the impact of netting fisheries; • monitoring of recreational bass fishery removals is required; • lack of guidance on how to avoid diving impacts on conservation features.

The report contains a section on Adopting Best Practice Management, including case studies. Data availability for analysis was found to be problematic for several reasons: • data were scarce, difficult to access, commercially sensitive, or of an unsuitable resolution or format; • there were insufficient spatial activity data – maps were too broadscale and detail was patchy; • permission was required from individual fishers, making collection time- consuming, and individuals were not comfortable sharing commercially confidential data.

It was suggested that initiatives such as the Fully Documented Fisheries (FDF) Project and spatial data from inshore VMS could offer improvement to data availability. Also, FDF may allow for monitoring of ETP species through its App. Table 5 of Marine Planning Consultants Ltd. (2014b) provides a summary of possible fisheries assessment options for addressing shortfalls (including those relating to data collection) to support management of the Lyme Bay Reserve.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 103 of 185

A1.7 Recommendations

General recommendations for model improvement relate to consideration of environmental tolerances and preferences, sensitivity to future aquaculture developments and taking account of likely technological advances (MMO, 2013a). Ross (2014) advises that field verification is an essential element of a GIS analysis, seeing data sourcing and quality verification as fundamental to success of the project. ABPmer (2015) recommend that a site selection model should incorporate point source (sewage) discharges, which will impact on water quality, as well as including sensitive benthic habitats e.g. from the FishMap Môn project (Austun, 2014) and carrying capacity. As well as the inclusion of essential fish habitats, and improvement and expansion of model input data, Falconer et al. (2017) also see carrying capacity as central to site assessment, subdividing this theme into: • physical carrying capacity, to quantify total area potentially available for aquaculture, leading to more specific site selection and further carrying capacity assessment at a later stage; • production carrying capacity (maximum capacity that can be supported in an area) and ecological carrying capacity (amount of production that can be maintained without leading to unsustainable changes) – these two are closely linked; • social carrying capacity (amount of aquaculture that can be developed without having adverse social impacts).

Cefas and FSA (2017) highlight the importance of faecal contamination potential in site selection for bivalve shellfish and raise the following points: • Large areas of population will bring with them a greater faecal contamination risk. In general, therefore, it is best to locate aquaculture sites as far as possible from residential areas. • Human faecal inputs (e.g. sewer pipes, septic tanks, boats) are sources of norovirus, whereas human and agricultural inputs (e.g. livestock, wildlife) can contribute to E. coli loadings. • Small, localised sources of contamination can be just as significant as large distant sources, so all should be investigated, considered and, where possible, their effects mitigated. This could involve, for instance, implementing an appropriate buffer distance between source and site, or by liaising with the local Environment Agency officer to determine whether remedial action can be taken in relation to the polluting source. • Shellfish can concentrate contamination in the surrounding water by up to 100-fold, so even a comparatively small amount of pollution can have a significant effect on shellfish E. coli results. This could mean the difference between site classifications B and C.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 104 of 185

• The classification of areas that only just conform to the upper limits for Class B may be more likely to change (i.e. to potentially be downgraded to Class C) as monitoring continues. • As the extent of contamination reduces towards Class A compliance, the risk of illness associated with harvested shellfish will also decline, although zero risk may be unobtainable for shellfish if eaten raw.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 105 of 185

Appendix 2: Overview of principal aquaculture production systems

A2.1 Bottom culture

o Usually, this method is primarily for bivalve shellfish (clams, oysters Scallops and mussels). Seedstock from hatcheries or relayed from other locations. o Need a suitable substrate (avoiding fine mud and silt and bare rock). o Can occur in water up to 20 m deep - harvested by dredging or pumping.

A2.2 Bottom secured culture

o Includes trestles, trays and secured mesh bags installed to posts/poles driven into the seabed so they are accessible at mean low water spring tides. o Need regular attention so must have good access. o Need to be in relatively protected sites safe from excessive wave/wind action.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 106 of 185

A2.3 Suspended longline – rope/textile/raft

o A suspended rope frame is created using surface buoys, submersed buoys, weighted or screw anchors and strong rope (usually 35-50 mm diameter). o Ropes, rafts or textiles (for seaweed culture) are suspended from main rope framework to provide attachment for the cultured species. o Can be established in low to medium energy environments in up to 50 m of water.

A2.4 Suspended longline – container culture

o A suspended rope frame is created using surface buoys, submersed buoys, weighted or screw anchors and strong rope (usually 35-50 mm diameter). o Net or composite containers are suspended from main rope framework on droppers to contain and protect the cultured species. o Can be established in low to medium energy environments in up to 50 m of water.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 107 of 185

A2.5 Floating/submersible cage culture.

o Open or net covered pens (usually round with diameter between 20-90 m) are secured to the seabed by the aid of weighted or screw anchors and strong rope or chains. Submersible cages can be temporarily submerged to avoid heavy seas/bad weather. o Cultured species will need regular feeding, grading and harvesting – usually from working boats moored alongside the pens. o They can be more visible than longline culture systems and can be established in low to medium energy environments in up to 50 m of water.

A2.6 Submerged cage culture

o Completely enclosed positively buoyant net pens (usually spherical, hexagonal, or trapezoidal) secured to the seabed by the aid of weighted or screw anchors and strong chains. Submerged cages can be raised to the surface for grading, maintenance or harvesting. o Cultured species will need regular feeding by vessel based remote feeding systems. o Can be established in medium to high energy environments in water up to 100 m deep and do not have as high visibility as floating pens. Not suitable for all finfish species. Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 108 of 185

Appendix 3: Species and culture system threshold values

The following tables indicate the threshold levels that were used to assess the suitability of areas within the FLAG region for aquaculture of 17 different species. The species were:

A. Macroalgae 1. Saccharina latissimi (sugar kelp) 2. Laminaria digitata (oarweed) 3. Alaria esculenta (winged kelp) 4. Palmaria palmata (dulse) 5. Ulva lactuca (sea lettuce) B. Finfish 6. Salmo salar (Atlantic Salmon) 7. Oncorhynchus mykiss (Rainbow Trout) 8. Salmo trutta (Brown Trout) 9. Gadus morhua (Atlantic Cod) 10. Dicentrachus labrax (Sea bass) 11. Sparus aurata (Gilthead sea bream) C. Crustaceans 12. Homarus Gammarus (Lobster) D. Bivalves 13. Mytilus edulis (Blue Mussel) 14. Crassostrea gigas (Pacific oyster) 15. Ostrea edulis (Flat oyster) 16. Venerupis philippinarum (Manila clam) 17. Pecten maximus (King Scallop)

For each species, the common method for aquaculture was assessed and environmental thresholds obtained from the scientific literature and trade press where possible. These were translated into a series of classification levels:

Optimal – species survives and grows best under these conditions Sub-optimal – the species can survive under these conditions, but growth is reduced Unsuitable – the species cannot survive or grow when exposed to these conditions in the medium to long term (not acute exposure)

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 109 of 185

A3.1 Threshold tables

Table A 1: Threshold levels used for Saccharina latissima aquaculture. Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Minimum sea surface Optimal >5 >5 temperature Sub-optimal 2-5 2-5 (°C) Unsuitable 0-2 0-2 Maximum sea surface Optimal <16 <16 temperature Sub-optimal 16-18 16-18 (°C) Unsuitable >18 >18 Optimal >24 >24 Minimum salinity Sub-optimal 15-24 15-24 Unsuitable 0-15 0-15 Optimal NA NA Maximum salinity Sub-optimal NA NA Unsuitable NA NA Optimal >2 >2 Light level Sub-optimal 1-2 1-2 Unsuitable 0-1 0-1 Optimal >10 >10 Total Oxidised Nitrogen Sub-optimal 4-10 4-10 Unsuitable 0-4 0-4 Optimal 0-2.1 0-2.1 Average significant wave height Sub-optimal 2.1-3.1 2.1-3.1 (m) Unsuitable >3.1 >3.1 Optimal 0.25-1.5 0.25-1.5 Current speed (m/s) Sub-optimal 1.5-2 & 0-0.25 1.5-2 & 0-0.25 Unsuitable >2 >2 Optimal 2-7 10-40 Depth (m) Sub-optimal 1-2 & 7-20 5-10 & >40 Unsuitable <1 & >20 <5 Optimal NA NA Dissolved oxygen concentration Sub-optimal NA NA Unsuitable NA NA Optimal NA NA Chlorophyll a concentration Sub-optimal NA NA Unsuitable NA NA Optimal RH RH/CS/XS/SM/S/MS Substrate Sub-optimal CS/XS/SB SB/FM Unsuitable SM/S/MS/FM NA Optimal Low Low Exposure level Sub-optimal Moderate Moderate/High Unsuitable High NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 110 of 185

Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Optimal 0-14.34 0-14.34 Suspended particulate Sub-optimal 14.34-29.27 14.34-29.27 inorganic matter Unsuitable >29.27 >29.27

Table A 2: Threshold levels used for Laminaria digitata aquaculture. Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Minimum sea surface Optimal >5 >5 temperature Sub-optimal 2-5 2-5 (°C) Unsuitable 0-2 0-2 Maximum sea surface Optimal 0-16 0-16 temperature Sub-optimal 16-18 16-18 (°C) Unsuitable >18 >18 Optimal >20 >20 Minimum salinity Sub-optimal 15-20 15-20 Unsuitable 0-15 0-15 Optimal NA NA Maximum salinity Sub-optimal NA NA Unsuitable NA NA Optimal >2 >2 Light level Sub-optimal 1-2 1-2 Unsuitable 0-1 0-1 Optimal >10 >10 Total Oxidised Nitrogen Sub-optimal 4-10 4-10 Unsuitable 0-4 0-4 Optimal 0-2.1 0-2.1 Average significant wave Sub-optimal 2.1-3.1 2.1-3.1 height (m) Unsuitable >3.1 >3.1 Optimal 0.25-1 0.25-1 Current speed (m/s) Sub-optimal 0-0.25 & 1-1.5 0-0.25 & 1-1.5 Unsuitable >1.5 >1.5 Optimal 2-7 10-40 Depth (m) Sub-optimal 1-2 & 7-20 5-10 & >40 Unsuitable <1 & >20 <5 Optimal NA NA Dissolved oxygen Sub-optimal NA NA concentration Unsuitable NA NA Optimal NA NA Chlorophyll a concentration Sub-optimal NA NA Unsuitable NA NA Substrate Optimal RH RH/CS/XS/SM/S/MS

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 111 of 185

Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Sub-optimal CS/XS/SB SB/FM Unsuitable SM/S/MS/FM NA Optimal Moderate Low Exposure level Sub-optimal Low Moderate and High Unsuitable High NA Optimal 0-14.34 0-14.34 Suspended particulate Sub-optimal 14.34-29.27 14.34-29.27 inorganic matter Unsuitable >29.27 >29.27

Table A 3: Threshold levels used for Alaria esculenta aquaculture. Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Minimum sea surface Optimal >4 >4 temperature Sub-optimal 2-4 2-4 (°C) Unsuitable 0-2 0-2 Maximum sea surface Optimal <16 <16 temperature Sub-optimal 16-17 16-17 (°C) Unsuitable >17 >17 Optimal >20 >20 Minimum salinity Sub-optimal 15-20 15-20 Unsuitable 0-15 0-15 Optimal NA NA Maximum salinity Sub-optimal NA NA Unsuitable NA NA Optimal >2 >2 Light level Sub-optimal 1-2 1-2 Unsuitable 0-1 0-1 Optimal >10 >10 Total Oxidised Nitrogen Sub-optimal 4-10 4-10 Unsuitable 0-4 0-4 Optimal 0-2.1 0-2.1 Average significant wave height Sub-optimal 2.1-3.1 2.1-3.1 (m) Unsuitable >3.1 >3.1 Optimal 0.5-1.5 0.5-1.5 Current speed (m/s) Sub-optimal 0-0.5 0-0.5 Unsuitable >1.5 >1.5 Optimal 2-7 10-40 Depth (m) Sub-optimal 1-2 & 7-20 5-10 & >40 Unsuitable <1 & >20 <5 Optimal NA NA Dissolved oxygen concentration Sub-optimal NA NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 112 of 185

Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Unsuitable NA NA Optimal NA NA Chlorophyll a concentration Sub-optimal NA NA Unsuitable NA NA Optimal RH/CS RH/C/S/XS/SM/MS Substrate Sub-optimal XS/SB FM/SB Unsuitable SM/S/MS/FM NA Optimal Low Low Exposure level Sub-optimal Moderate Moderate & High Unsuitable High NA Optimal 0-14.34 0-14.34 Suspended particulate Sub-optimal 14.34-29.27 14.34-29.27 inorganic matter Unsuitable >29.27 >29.27

Table A 4: Threshold levels used for Palmaria palmata aquaculture. Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Minimum sea surface Optimal >6 >6 temperature Sub-optimal 2-6 2-6 (°C) Unsuitable <2 <2 Maximum sea surface Optimal <15 <15 temperature Sub-optimal 15-18 15-18 (°C) Unsuitable >18 >18 Optimal >32 >32 Minimum salinity Sub-optimal 30-32 30-32 Unsuitable <30 <30 Optimal NA NA Maximum salinity Sub-optimal NA NA Unsuitable NA NA Optimal >1 >1 Light level Sub-optimal 0.5-1 0.5-1 Unsuitable <0.5 <0.5 Optimal >10 >10 Total Oxidised Nitrogen Sub-optimal 4-10 4-10 Unsuitable <4 <4 Optimal 0-2.1 0-2.1 Average significant wave Sub-optimal 2.1-3.1 2.1-3.1 height (m) Unsuitable >3.1 >3.1 Optimal 0.25-1 0.25-1 Current speed (m/s) Sub-optimal 0-0.25 & 1-1.5 0-0.25 & 1-1.5 Unsuitable >1.5 >1.5

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 113 of 185

Threshold values used to assess potential Parameter Threshold Bottom-secured Suspended Optimal 1-7 10-40 Depth (m) Sub-optimal 7-20 5-10 & >40 Unsuitable <1 & >20 <5 Optimal NA NA Dissolved oxygen Sub-optimal NA NA concentration Unsuitable NA NA Optimal NA NA Chlorophyll a concentration Sub-optimal NA NA Unsuitable NA NA Optimal RH RH/C/S/XS/SM/MS Substrate Sub-optimal CS/XS/SB FM/SB Unsuitable SM/S/MS/FM NA Optimal Low Low Exposure level Sub-optimal Moderate Moderate & High Unsuitable High NA Optimal 0-6.88 0-6.88 Suspended particulate Sub-optimal 6.88-14.34 6.88-14.34 inorganic matter Unsuitable >14.34 >14.34

Table A 5: Threshold levels used for Ulva lactuca aquaculture. Threshold values used to assess potential Parameter Threshold Suspended (Floating) Minimum sea surface Optimal 10-15 temperature Sub-optimal 0-10 (°C) Unsuitable <0 Maximum sea surface Optimal 10-15 temperature Sub-optimal 15-28 (°C) Unsuitable >28 Optimal 20-30 Minimum salinity Sub-optimal 10-20 & >30 Unsuitable 0-10 Optimal 20-30 Maximum salinity Sub-optimal 30-40 Unsuitable >40 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal >13 Total Oxidised Nitrogen Sub-optimal 1-13 Unsuitable 0-1 Optimal 0-2.1

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 114 of 185

Threshold values used to assess potential Parameter Threshold Suspended (Floating) Average significant wave height Sub-optimal 2.1-3.1 (m) Unsuitable >3.1 Optimal 0.25-1.5 Current speed (m/s) Sub-optimal 0-0.25 & 1.5-2 Unsuitable >2 Optimal 10-40 Depth (m) Sub-optimal 5-10 & >40 Unsuitable <5 Optimal NA Dissolved oxygen concentration Sub-optimal NA Unsuitable NA Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal RH/C/S/XS/SM/MS Substrate Sub-optimal FM/SB Unsuitable NA Optimal Low Exposure level Sub-optimal Moderate Unsuitable High Optimal NA Suspended particulate inorganic Sub-optimal NA matter Unsuitable NA

Table A 6: Threshold levels used for Salmo salar aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >6 temperature Sub-optimal 2-6 (°C) Unsuitable <2 Maximum sea surface Optimal 12-16 temperature Sub-optimal 2-12 & 16-22 (°C) Unsuitable >22 & <2 Optimal 22-36 Minimum salinity Sub-optimal 0-22 & 36-40 Unsuitable >40 Optimal 22-36 Maximum salinity Sub-optimal 0-22 & 36-40 Unsuitable >40 Optimal NA Light level Sub-optimal NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 115 of 185

Threshold value used to assess potential Parameter Threshold Open pen Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal 0-1.5 Average significant wave height Sub-optimal 1.5-6 (m) Unsuitable >6 Optimal 0.25-0.5 Current speed (m/s) Sub-optimal 0.02–0.25 & 0.5-0.75 Unsuitable <0.02 & >0.75 Optimal 25-50 Depth (m) Sub-optimal 10-25 & >50 Unsuitable <10 Optimal >5 Dissolved oxygen concentration Sub-optimal 3-5 Unsuitable 0-3 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal CS/SM/S/MS/FM Substrate Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Optimal 0-25 Suspended particulate inorganic Sub-optimal NA matter Unsuitable >25

Table A 7: Threshold levels used for Oncorhynchus mykiss aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >4 temperature Sub-optimal 0-4 (°C) Unsuitable <0 Maximum sea surface Optimal 12-17 temperature Sub-optimal <12 & 17-24 (°C) Unsuitable >24 Optimal 0-24 Minimum salinity Sub-optimal 24-36 Unsuitable >36

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 116 of 185

Threshold value used to assess potential Parameter Threshold Open pen Optimal 0-24 Maximum salinity Sub-optimal 24-36 Unsuitable >36 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal 0-1.5 Average significant wave height Sub-optimal 1.5-6 (m) Unsuitable >6 Optimal 0.25-0.5 Current speed (m/s) Sub-optimal 0.02-0.25 & 0.5-0.75 Unsuitable <0.02 & >0.75 Optimal 25-50 Depth (m) Sub-optimal 10-25 & >50 Unsuitable <10 Optimal >6 Dissolved oxygen concentration Sub-optimal 5-6 Unsuitable 0-5 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal CS/SM/S/MS/FM Substrate Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Optimal 0-25 Suspended particulate inorganic Sub-optimal NA matter Unsuitable >25

Table A 8: Threshold levels used for Salmo trutta aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >4 temperature Sub-optimal 0.7-4 (°C) Unsuitable <0.7 Optimal 12-17

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 117 of 185

Threshold value used to assess potential Parameter Threshold Open pen Maximum sea surface Sub-optimal <12 & 17-24 temperature (°C) Unsuitable >24 Optimal 33-36 Minimum salinity Sub-optimal 0-33 Unsuitable >36 Optimal 33-36 Maximum salinity Sub-optimal 0-33 Unsuitable >36 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal 0-1.5 Average significant wave height Sub-optimal 1.5-6 (m) Unsuitable >6 Optimal 0.25-0.5 Current speed (m/s) Sub-optimal 0.02-0.25 & 0.5-0.75 Unsuitable <0.02 & >0.75 Optimal 25-50 Depth (m) Sub-optimal 10-25 & >50 Unsuitable <10 Optimal >6 Dissolved oxygen concentration Sub-optimal 4-6 Unsuitable <4 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal CS/SM/S/MS/FM Substrate Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Optimal 0-25 Suspended particulate inorganic Sub-optimal NA matter Unsuitable >25

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 118 of 185

Table A 9: Threshold levels used for Gadus morhua aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >9 temperature Sub-optimal 0-9 (°C) Unsuitable <0 Maximum sea surface Optimal <15 temperature Sub-optimal 15-16 (°C) Unsuitable >16 Optimal >28 Minimum salinity Sub-optimal 8-28 Unsuitable 0-8 Optimal >28 Maximum salinity Sub-optimal 8-28 Unsuitable 0-8 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal 0-9 Average significant wave height Sub-optimal NA (m) Unsuitable >9 Optimal 0.25-0.5 Current speed (m/s) Sub-optimal 0.02-0.25 & 0.5-0.75 Unsuitable 0-0.02 & >0.75 Optimal 25-50 Depth (m) Sub-optimal 10-25 & >50 Unsuitable <10 Optimal >5 Dissolved oxygen concentration Sub-optimal 3-5 Unsuitable 0-3 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal CS/SM/S/MS/FM Substrate Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 119 of 185

Threshold value used to assess potential Parameter Threshold Open pen Optimal 0-25 Suspended particulate inorganic Sub-optimal NA matter Unsuitable >25

Table A 10: Threshold levels used for Dicentrarchus labrax aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >18 temperature Sub-optimal 2-18 (°C) Unsuitable <2 Maximum sea surface Optimal 24-26 temperature Sub-optimal <24 & 26-32 (°C) Unsuitable >32 Optimal >12 Minimum salinity Sub-optimal 4-12 Unsuitable <4 Optimal 28-36 Maximum salinity Sub-optimal 13-28 & 36-40 Unsuitable >40 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal <9 Average significant wave height Sub-optimal NA (m) Unsuitable >9 Optimal 0.1-0.2 Current speed (m/s) Sub-optimal 0.05-0.1 & 0.2-0.4 Unsuitable <0.05 & >0.4 Optimal 20-50 Depth (m) Sub-optimal 10-20 & 50-100 Unsuitable <10 & >100 Optimal >6 Dissolved oxygen concentration Sub-optimal 4-6 Unsuitable 0-4 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Substrate Optimal CS/SM/S/MS/FM

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 120 of 185

Threshold value used to assess potential Parameter Threshold Open pen Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Optimal 0-10 Suspended particulate inorganic Sub-optimal 10-100 matter Unsuitable >100

Table A 11: Threshold levels used for Sparus aurata aquaculture. Threshold value used to assess potential Parameter Threshold Open pen Minimum sea surface Optimal >17 temperature Sub-optimal 6-17 (°C) Unsuitable <6 Maximum sea surface Optimal 25-27 temperature Sub-optimal <25 & 27-33 (°C) Unsuitable >33 Optimal >15 Minimum salinity Sub-optimal 5-15 Unsuitable <5 Optimal 28-36 Maximum salinity Sub-optimal <28 & 36-40 Unsuitable >40 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal <9 Average significant wave height Sub-optimal NA (m) Unsuitable >9 Optimal 0.1-0.2 Current speed (m/s) Sub-optimal 0.05-0.1 & 0.2-0.4 Unsuitable >0.4 Optimal 20-50 Depth (m) Sub-optimal 10-20 & 50-100 Unsuitable <10 & >100 Optimal >7 Dissolved oxygen concentration Sub-optimal 2.7-7

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 121 of 185

Threshold value used to assess potential Parameter Threshold Open pen Unsuitable <2.7 Optimal NA Chlorophyll a concentration Sub-optimal NA Unsuitable NA Optimal CS/SM/S/MS/FM Substrate Sub-optimal RH/XS/SB Unsuitable NA Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Optimal <25 Suspended particulate inorganic Sub-optimal 25-50 matter Unsuitable >50

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 122 of 185

Table A 12: Threshold levels used for Homarus gammarus aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Suspended Minimum sea surface Optimal >5 >5 temperature Sub-optimal 0-5 0-5 (°C) Unsuitable <0 <0 Maximum sea surface Optimal 10-22 10-22 temperature Sub-optimal <10 & 22-31 <10 & 22-31 (°C) Unsuitable >31 >31 Optimal >31 >31 Minimum salinity Sub-optimal 8-31 8-31 Unsuitable <8 <8 Optimal <32 <32 Maximum salinity Sub-optimal 32-46 32-46 Unsuitable >46 >46 Optimal NA NA Light level Sub-optimal NA NA Unsuitable NA NA Optimal NA NA Total Oxidised Nitrogen Sub-optimal NA NA Unsuitable NA NA Optimal 0-2.1 0-2.1 Average significant wave Sub-optimal 2.1-3.1 2.1-3.1 height (m) Unsuitable >3.1 >3.1 Optimal 0.2-0.9 0.2-0.9 Current speed (m/s) Sub-optimal NA NA Unsuitable <0.2 & >0.9 <0.2 & >0.9 Optimal 5-50 25-50 Depth (m) Sub-optimal 50-150 10-25 & >50 Unsuitable <5 & >150 <10 Optimal >6.2 >6.2 Dissolved oxygen Sub-optimal 4.5-6.2 4.5-6.2 concentration Unsuitable <4.5 <4.5 Optimal NA NA Chlorophyll a concentration Sub-optimal NA NA Unsuitable NA NA Optimal RH/XS/SB RH/XS/SB Substrate Sub-optimal CS/SM/S/MS/FM CS/SM/S/MS/FM Unsuitable NA NA Optimal Low & Moderate Low Exposure level Sub-optimal High Moderate & High Unsuitable NA NA Suspended particulate Optimal NA NA inorganic matter Sub-optimal NA NA Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 123 of 185

Threshold value used to assess potential Parameter Threshold Bottom Suspended Unsuitable NA NA

Table A 13: Threshold levels used for Mytilus edulis aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended Minimum sea surface Optimal >8 >8 >8 temperature Sub-optimal -4-8 -4-8 -4-8 (°C) Unsuitable <-4 <-4 <-4 Maximum sea surface Optimal 12-20 12-20 12-20 temperature Sub-optimal 8-12 & 20-25 8-12 & 20-25 8-12 & 20-25 (°C) Unsuitable >25 >25 >25 Optimal >18 >18 >18 Minimum salinity Sub-optimal 4-18 4-18 4-18 Unsuitable <4 <4 <4 Optimal 25-36 25-36 25-36 Maximum salinity Sub-optimal 18-25 & 36-40 18-25 & 36-40 18-25 & 36-40 Unsuitable >40 >40 >40 Optimal NA NA NA Light level Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA NA Total Oxidised Nitrogen Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA NA Average significant wave Sub-optimal NA NA NA height (m) Unsuitable NA NA NA Optimal 0.5-1.0 0.5-1.0 0.5-1.0 Current speed (m/s) Sub-optimal 0.2-0.5 & 1.0-1.5 0.2-0.5 & 1.0-1.5 0.2-0.5 & 1.0-1.5 Unsuitable <0.2 & >1.5 <0.2 & >1.5 <0.2 & >1.5 Optimal 10-30 0-5 25-50 Depth (m) Sub-optimal 5-10 & 30-50 NA 10-25 & >50 Unsuitable <5 & >50 >5 <10 Optimal >7 >7 >7 Dissolved oxygen Sub-optimal 1.5-7 1.5-7 1.5-7 concentration Unsuitable <1.5 <1.5 <1.5 Optimal >0.006 >0.006 >0.006 Chlorophyll a Sub-optimal 0.001-0.006 0.001-0.006 0.001-0.006 concentration Unsuitable <0.001 <0.001 <0.001 Optimal RH/CS/XS RH/CS/XS RH/CS/XS Substrate Sub-optimal SM/S/MS/FM/SB SM/S/MS/FM/SB SM/S/MS/FM/SB Unsuitable NA NA NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 124 of 185

Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended Optimal Low & Moderate Low Low & Moderate Exposure level Sub-optimal High NA High Unsuitable NA Moderate & High NA Optimal 5-100 5-100 5-100 Suspended particulate Sub-optimal 0-5 & 100-440 0-5 & 100-440 0-5 & 100-440 inorganic matter Unsuitable >440 >440 >440

Table A 14: Threshold levels used for Crassostrea gigas aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended Minimum sea surface Optimal >15 >15 >15 temperature Sub-optimal 4-15 4-15 4-15 (°C) Unsuitable <4 <4 <4 Maximum sea surface Optimal 15-25 15-25 15-25 temperature Sub-optimal 4-15 & 25-35 4-15 & 25-35 4-15 & 25-35 (°C) Unsuitable >35 >35 >35 Optimal >20 >20 >20 Minimum salinity Sub-optimal 13-20 13-20 13-20 Unsuitable 0-13 0-13 0-13 Optimal 25-36 25-36 25-36 Maximum salinity Sub-optimal 0-25 & 36-45 0-25 & 36-45 0-25 & 36-45 Unsuitable >45 >45 >45 Optimal NA NA NA Light level Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA NA Total Oxidised Nitrogen Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA 0-2.1 Average significant wave Sub-optimal NA NA 2.1-3.1 height (m) Unsuitable NA NA >3.1 Optimal 0.5-1.0 0.5-1.0 0.5-1.0 0.25-0.5 0.25-0.5 &1.0- Current speed (m/s) Sub-optimal 0.25-0.5 1.5 Unsuitable <0.25 & >1.0 <0.25 & >1.0 <0.25 & >1.5 Optimal 10-30 <5 25-50 Depth (m) Sub-optimal 5-10 & 30-50 NA 10-25 & >50 Unsuitable <5 & >50 >5 <10 Optimal >8.0 >8.0 >8.0 Dissolved oxygen Sub-optimal 4.5-8.0 4.5-8.0 4.5-8.0 concentration Unsuitable 0-4.5 0-4.5 0-4.5 Optimal >0.008 >0.008 >0.008 Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 125 of 185

Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended Chlorophyll a Sub-optimal 0.001-0.008 0.001-0.008 0.001-0.008 concentration Unsuitable <0.001 <0.001 <0.001 Optimal RH/CS/XS RH/CS/XS RH/CS/XS Substrate Sub-optimal SM/S/MS/FM/SB SM/S/MS/FM/SB SM/S/MS/FM/SB Unsuitable NA NA NA Optimal Low Low Low Exposure level Sub-optimal Moderate & High NA Moderate & High Unsuitable NA Moderate & High NA Optimal 0-90 0-90 0-90 Suspended particulate Sub-optimal 90-200 90-200 90-200 inorganic matter Unsuitable >200 >200 >200

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 126 of 185

Table A 15: Threshold levels used for Ostrea edulis aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended

Minimum sea surface Optimal >10 >10 >10 temperature Sub-optimal -1-10 -1-10 -1-10 (°C) Unsuitable <-1 <-1 <-1 Maximum sea surface Optimal 12-20 12-20 12-20 temperature Sub-optimal 10-12 & 20-25 10-12 & 20-25 10-12 & 20-25 (°C) Unsuitable >25 >25 >25 Optimal >25 >25 >25 Minimum salinity Sub-optimal 15-25 15-25 15-25 Unsuitable 0-15 0-15 0-15 Optimal 25-36 25-36 25-36 Maximum salinity Sub-optimal 20-25 & >36-40 20-25 & >36-40 20-25 & >36-40 Unsuitable >40 >40 >40 Optimal NA NA NA Light level Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA NA Total Oxidised Nitrogen Sub-optimal NA NA NA Unsuitable NA NA NA Optimal NA NA NA Average significant wave Sub-optimal NA NA NA height (m) Unsuitable NA NA NA Optimal 0.5-1.0 0.5-1.0 0.5-1.0 Current speed (m/s) Sub-optimal 0.25-0.5 0.25-0.5 0.25-0.5 Unsuitable <0.25 & >1.0 <0.25 & >1.0 <0.25 & >1.0 Optimal 10-30 <5 25-50 Depth (m) Sub-optimal 5-10 & 30-50 NA 10-25 & >50 Unsuitable <5 & >50 >5 <10 Optimal >8.0 >8.0 >8.0 Dissolved oxygen Sub-optimal 4.5-8.0 4.5-8.0 4.5-8.0 concentration Unsuitable <4.5 <4.5 <4.5 Optimal >0.008 >0.008 >0.008 Chlorophyll a Sub-optimal 0.001-0.008 0.001-0.008 0.001-0.008 concentration Unsuitable <0.001 <0.001 <0.001 Optimal RH/CS/XS RH/CS/XS RH/CS/XS Substrate Sub-optimal SM/S/MS/FM/SB SM/S/MS/FM/SB SM/S/MS/FM/SB Unsuitable NA NA NA Optimal Low Low Low Exposure level Sub-optimal Moderate & High NA Moderate & High Unsuitable NA Moderate & High NA Optimal <90 <90 <90

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 127 of 185

Threshold value used to assess potential Parameter Threshold Bottom Bottom-secured Suspended Suspended particulate Sub-optimal 90-200 90-200 90-200 inorganic matter Unsuitable >200 >200 >200

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 128 of 185

Table A 16: Threshold levels used for Ruditapes philippinarum aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Minimum sea surface Optimal >5 temperature Sub-optimal NA (°C) Unsuitable <5 Maximum sea surface Optimal 18-23 temperature Sub-optimal 5-18 & 23-30 (°C) Unsuitable >30 Optimal >25 Minimum salinity Sub-optimal 12-25 Unsuitable <12 Optimal 25-36 Maximum salinity Sub-optimal 12-25 & 36-40 Unsuitable >40 Optimal NA Light level Sub-optimal NA Unsuitable NA Optimal NA Total Oxidised Nitrogen Sub-optimal NA Unsuitable NA Optimal NA Average significant wave height Sub-optimal NA (m) Unsuitable NA Optimal 0.5-1.0 Current speed (m/s) Sub-optimal 0.25-0.5 Unsuitable <0.25 & >1.0 Optimal 0-5 Depth (m) Sub-optimal NA Unsuitable >5 Optimal >8.0 Dissolved oxygen concentration Sub-optimal 1.1-8.0 Unsuitable <1.1 Optimal >0.002 Chlorophyll a concentration Sub-optimal 0.001-0.002 Unsuitable <0.001 Optimal XS/SM/S/MS/FM Substrate Sub-optimal CS/SB Unsuitable RH Optimal Low & Moderate Exposure level Sub-optimal High Unsuitable NA Suspended particulate inorganic Optimal NA matter Sub-optimal NA Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 129 of 185

Threshold value used to assess potential Parameter Threshold Bottom Unsuitable NA

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 130 of 185

Table A 17: Threshold levels used for Pecten maximus aquaculture. Threshold value used to assess potential Parameter Threshold Bottom Suspended Minimum sea surface Optimal >10 >10 temperature Sub-optimal 6-10 6-10 (°C) Unsuitable <6 <6 Maximum sea surface Optimal 10-17 10-17 temperature Sub-optimal 17-25 17-25 (°C) Unsuitable >25 >25 Optimal 30-35 30-35 Minimum salinity Sub-optimal 29-30 29-30 Unsuitable <29 <29 Optimal 30-36 30-36 Maximum salinity Sub-optimal 36-40 36-40 Unsuitable >40 >40 Optimal NA NA Light level Sub-optimal NA NA Unsuitable NA NA Optimal NA NA Total Oxidised Nitrogen Sub-optimal NA NA Unsuitable NA NA Optimal NA 0-2.1 Average significant wave Sub-optimal NA 2.1-3.1 height (m) Unsuitable NA >3.1 Optimal 0.2-0.9 0.2-0.9 Current speed (m/s) Sub-optimal NA NA Unsuitable <0.2 & >0.9 <0.2 & >0.9 Optimal 10-30 25-50 Depth (m) Sub-optimal 5-10 & 30-50 10-25 & >50 Unsuitable <5 & >50 <10 Optimal >8.0 >8.0 Dissolved oxygen Sub-optimal 2.0-8.0 2.0-8.0 concentration Unsuitable <2.0 <2.0 Optimal >0.02 >0.02 Chlorophyll a concentration Sub-optimal 0.001-0.02 0.001-0.02 Unsuitable <0.001 <0.001 Optimal CS/XS/SM/S/MS RH/CS/XS Substrate Sub-optimal FM/SB SM/S/MS/FM/SB Unsuitable RH NA Optimal Low Low Exposure level Sub-optimal Moderate & High Moderate & High Unsuitable NA NA Suspended particulate Optimal <150 <150 inorganic matter Sub-optimal 150-700 150-700 Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 131 of 185

Threshold value used to assess potential Parameter Threshold Bottom Suspended Unsuitable >700 >700

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 132 of 185

A3.2 Scientific literature used to determine threshold levels.

A database which links references to specific threshold levels is held by Cefas to inform project work but does not form part of the outputs from this project.

1. Ale, M.T., Mikkelsen, J.D. and Meyer, A.S. (2011). Differential growth response of Ulva lactuca to ammonium and nitrate assimilation. Journal of Applied Phycology, 23, pp. 345-351. 2. Allen, S. M. and Burnett, L. E. (2008). The effects of intertidal air exposure on the respiratory physiology and the killing activity of haemocytes in the pacific oyster, Crassostrea gigas (Thunberg). Journal of Experimental Marine Biology and Ecology, 357, pp. 165–171. 3. Árnason, T. (2007). The effects of temperature on growth of the Atlantic cod (Gadus morhua). Thesis submitted for Master of Science in Natural Resource Science, University of Akureyri. 4. Árnason, T., Magnadóttir, B, Björnsson, B., Steinarsson, A. and Björnsson, B.T. (2013). Effects of salinity and temperature on growth, plasma ions, cortisol and immune parameters of juvenile Atlantic cod (Gadus morhua). Aquaculture, 380- 383, pp. 70-79 5. Artigaud, S., Lacroix, C., Pichereau, V and Flye-Sainte-Marie, J. (2014). Respiratory response to combined heat and hypoxia in the marine bivalves Pecten maximus and Mytilus spp. Comparative Biochemistry and Physiology - Part A: Molecular and Integrative Physiology, 175(1), pp. 135–140. 6. Baldwin, B. S. and Newell, R. I. E. (1995). Relative importance of different size food particles in the natural diet of oyster larvae (Crassostrea virginica). Marine Ecology Progress Series, 120, pp. 135–145. 7. Baluyut, E. A. (1989). Aquaculture Systems and Practices: A Selected Review. FAO, Rome. 8. Bayne, B. (1998) The physiology of suspension feeding by bivalve molluscs: an introduction to the Plymouth “TROPHEE” workshop. Journal of Experimental Marine Biology and Ecology, 1–2, pp. 1–19. 9. Beard, T. W. and McGregor, D. (2004). Storage and care of live lobsters. Cefas Laboratory Leaflet (Revised), pp. 1–27 10. Bekkby, T. and Moy, F.E. (2011). Developing spatial models of sugar kelp (Saccharina latissima) potential distribution under natural conditions are areas of its disappearance in Skagerrak. Estuarine, Coastal and Shelf Science, 95, pp. 477-483. 11. Beninger, P. and Lucas, A. (1984). Seasonal Variations in Condition, Reproductive Activity, and Gross Biochemical Composition of Two Species of Adult Clam Reared in a Common Habitat: Tapes decussatus and Tapes philippinarum. Journal of Experimental Marine Biology and Ecology, 79, pp. 19– 37. 12. Bergström, P. and Lindegarth, M. (2016). Environmental influence on mussel (Mytilus edulis) growth - A quantile regression approach. Estuarine, Coastal and Shelf Science, 171, pp. 123–132. 13. Bibby, A. (2015). Open ocean sea scallop culture in Canada, Aquaculture North America. Available online at www.aquaculturenorthamerica.com/research/open- ocean-sea-scallop-culture-in-chaleur-bay-canada-1230.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 133 of 185

14. Bignell, J. P., Daniels, C., Hendrikz, L. and Auchterlonie, N. (2016). Lobster Grower: Developing the technology to fast track the aquaculture potential for the European Lobster: WP2 Literature Review: Factors affecting growth and survival of Homarus sp., Cefas and National Lobster Hatchery Report. 15. Birkett, D. A., Maggs, C.A., Dring, M.J. and Boaden, P.J.S. (1998). Infralittoral reef biotopes with kelp species (volume VII). An overview of dynamic and sensitivity characteristics for conservation management of marine SACs. Scottish Association of Marine Science (UK Marine SACs Project). Available online at: http://www.ukmarinesac.org.uk/pdfs/reefkelp.pdf. 16. Björnsson, B. and Steinarsson, A. (2002). The food-unlimited growth rate of Atlantic cod (Gadus morhua). Canadian Journal of Fisheries and Aquatic Sciences, 59, pp. 494–502. 17. Björnsson, B., Steinarsson, A. and Oddgeirsson, M. (2001). Optimal temperature for growth and feed conversion of immature cod (Gadus morhua L.). ICES Journal of Marine Science, 58(1), pp. 29–38. 18. Björnsson, B., Steinarsson, A. and Árnason, T. (2007). Growth model for Atlantic cod (Gadus morhua): Effects of temperature and body weight on growth rate. Aquaculture, 271, pp. 216-226 19. Bœuf, G. and Payan, P. (2001). How should salinity influence fish growth? Comparative Biochemistry and Physiology - C: Toxicology and Pharmacology, 130, pp. 411-423. 20. Bolton, J. J. and Lüning, K. (1982) Optimal growth and maximal survival temperatures of Atlantic Laminaria species (Phaeophyta) in culture, Marine Biology, 6(1) pp. 89–94. 21. Boogert, F.J., Cubillo, A.N., Nunes, J.P., Ferreira, J.G. and Corner, R.A. (2018). Ecosystem Approach to making Space for Aquaculture. EU Horizon 2020 project grant number 633476: Deliverable 2.5. Available online at: www.aquaspace- h2020.eu/wp-content/uploads/2017/10/Online-Environmental-Feasibility- Application.pdf. 22. Bourlès, Y., Alunno-Bruscia, M., Pouvreau, S., Tollu, D., Leguay, D., Arnaud, C., Goulletquer, P. and Kooijman, S.A.L.M. (2009). Modelling growth and reproduction of the Pacific oyster Crassostrea gigas: Advances in the oyster-DEB model through application to a coastal pond. Journal of Sea Research, 69(2–3), pp. 62–71. 23. Brenner, M., Broeg, K., Wilhelm, C., Buchholz, C. and Koehler, A. (2012). Effect of air exposure on lysosomal tissues of Mytilus edulis L. from natural intertidal wild beds and submerged culture ropes. Comparative Biochemistry and Physiology – Part A, 161(3), pp. 327–336. 24. Broch O.J. and Slagstad, D. (2012). Modelling seasonal growth and composition of the kelp Saccharina latissima., Journal of Applied Phycology, 24(4), pp. 759- 776 25. Brown, J. R. (1986). The influence of environmental factors upon the growth and survival of the Pacific oyster, Crassostrea gigas Thunberg. Thesis submitted to Simon Fraser University. 26. Buck, B. H. (2007). Experimental trials on the feasibility of offshore seed production of the mussel Mytilus edulis in the German Bight: installation, technical requirements and environmental conditions. Helgoland Marine Research, 61, pp. 87–101.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 134 of 185

27. Buck B. H. and Langan R. (2017). Aquaculture Perspective of Multi-Use Sites in the Open Ocean. Springer Open. 28. Buck, B. H. and Buchholz, C. M. (2005). Response of offshore cultivated Laminaria saccharina to hydrodynamic forcing in the North Sea. Aquaculture, 250, pp. 674–691. 29. Buck, B. H. and Grote, B. (2018). Seaweed in high-energy environments. Protocol to move Saccharina cultivation offshore. In Charrier, B., Wichard, T. and Reddy, C.R.K. (Eds) Protocols for Macroalgae Research. CRC Press, pp. 485. 30. Bunker, F. S. D., Brodie, J.A., Maggs, C.A. and Bunker, A.R. (2017). Seaweeds of Britain and Ireland. Second Edition, Wild Nature Press 31. Capuzzo, E., Stephens, D., Aldridge, J. and Forster, R.M. (2014). Feasibility study – Potential locations for macro-algae farming off the East Anglian coast. The Crown Estate 37. ISBN: 978-1-906410-61-2. 32. Capuzzo, E., Collingridge, K., Mitchell, P. and Aldridge, J. (2018). Feasibility study: areas suitable for seaweed growth and farming around the English Coast. Cefas Report (unpublished) 33. Capuzzo, E. and McKie, T. (2016). Seaweed in the UK and abroad – status, products, limitations, gaps and Cefas role. Cefas Contract Report FC0021. Available online at: www.gov.uk/government/publications/the-seaweed-industry- in-the-uk-and-abroad. 34. Cardia, F. and Lovatelli, A. (2015). Aquaculture operations in floating HDPE cages. FAO Fisheries and Aquaculture Technical paper, pp. 15–77. Available online at: www.fao.org/3/a-i4508e.pdf. 35. Carter, M. (2003). Ruditapes philippinarum (Manila clam). In Tyler-Walters, H. and Hiscock, K. (Eds) Marine Life Information Network: Biology and Sensitivity Key Information Reviews. Plymouth: Marine Biological Association of the United Kingdom. Available on at: www.marlin.ac.uk/species/detail/2203. 36. Castro, V., Grisdale-Helland, B., Helland, S.J., Kristensen, T., Jørgensen, S.M., Helgerud, J., Claireaux, G., Farrell, A.P., Krasnov, A. and Takle, H. (2011). Aerobic training stimulates growth and promotes disease resistance in Atlantic salmon (Salmo salar). Comparative Biochemistry and Physiology, Part A, 160(2), pp. 278–290. 37. Chávez-Villalba, J., Pommier, J., Andriamisez, J., Pouvreau, S., Barret, J., Cochard, J.-C. and Le Pennec, M. (2002). Broodstock conditioning of the oyster Crassostrea gigas: Origin and temperature effect. Aquaculture, 214, pp. 115– 130. 38. Cheney, D., MacDonanald, B. and Elston, R. (2000). Summer Mortality of Pacific Oysters, Crassostrea gigas (Thunberg): Initial Findings on Multiple Environmental Stressors in Puget Sound, Washington, 1988. Journal of Shellfish Research, 19(1), pp. 353–359. 39. Christophersen, G. (2005). Effects of environmental conditions on culturing scallop spat (Pecten maximus). Thesis submitted to the Department of Biology, University of Bergen. Available online at: Bergen Open Research Archive: Effects of environmental conditions on culturing scallop spat (Pecten maximus) (uib.no) 40. Christophersen, G. and Strand, Ø. (2003). Effect of reduced salinity on the great scallop (Pecten maximus) spat at two rearing temperatures. Aquaculture, 215, pp. 79–92. 41. Chronis, G., Lykousis, V., Georgopoulos, D., Zervakis, V., Stavrakakis, S. and Poulos, S. (2000) Suspended particulate matter and nepheloid layers over the

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 135 of 185

southern margin of the Cretan Sea (N. E. Mediterranean): seasonal distribution and dynamics. Progress in Oceanography, 46, pp. 163–185. 42. Cranford, P. J., Ward, J. E. and Shumway, S. E. (2011). Bivalve filter feeding: variability and limits of the aquaculture biofilter. In Shellfish Aquaculture and the Environment. 1st Edition. John Wiley and Sons. 43. Dare, P. (1980). Mussel Cultivation in England and Wales. Laboratory Leaflet No. 50. Directorate of Fisheries Research. 44. Dégremont, L., Be´diera, E., Soletchnik, P., Ropert, M., Huvet, A., Moale, J., Samain, J.F. and Boudry, P. (2005). Relative importance of family, site, and field placement timing on survival, growth, and yield of hatchery-produced Pacific oyster spat (Crassostrea gigas), Aquaculture, 249, pp. 213–229. 45. Dmitry, A. (2013). Effect of water quality on rainbow trout performance Water oxygen level in commercial trout farm. Thesis submitted to the Department of Environmental Engineering, Mikkeli University of Applied Sciences. Available online at: http://www.theseus.fi/bitstream/handle/10024/57500/Avkhimovich%20Dmitry.pd f 46. Dolmer, P. and Svane, I. (1994). Attachment and orientation of Mytilus edulis L. in flowing water. Ophelia, 40(1), pp. 63-74. 47. Dutil, J.-D., Munro, J., Audet, C. and Besner, M. et al. (1992). Seasonal variation in the physiological response of Atlantic cod (Gadus morhua) to low salinity. Canadian Journal of Fisheries and Aquatic Sciences. 49, pp. 1149–1156. 48. EFSA (2008). Scientific Opinion of the Panel on Animal Health and Welfare on a request from the European Commission on animal welfare aspects of husbandry systems for farmed European seabass and Gilthead seabream. The EFSA Journal 844, pp. 1-21 49. 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), pp. 1793– 1817. 50. Evans, S. and Langdon, C. (2006). Effects of genotype × environment interactions on the selection of broadly adapted Pacific oysters (Crassostrea gigas). Aquaculture, 261, pp. 522–534. 51. Falconer, L., Hunter, D.-C., Scott, P.C., Telfer, T.C. and Ross, L.G. (2013). Using physical environmental parameters and cage engineering design within GIS- based site suitability models for marine aquaculture. Aquaculture Environmental Interactions, 4, pp. 223–237. 52. FAO (2018c). Species Fact Sheets: Gadus morhua. In: FAO Fisheries and Aquaculture Department. Rome. Available online at: http://www.fao.org/fishery/species/2218/en, accessed 28 November 2018. 53. FAO (2018d). Cultured Aquatic Species Information: Ruditapes philippinarum, In: FAO Fisheries and Aquaculture Department [online]. Rome. Available online at: www.fao.org/fishery/culturedspecies/Ruditapes_philippinarum/en, accessed: 28 November 2018. 54. FAO (2018e). Cultured Aquatic Species Information: Oncorhynchus mykiss, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Oncorhynchus_mykiss/en, accessed 28 November 2018.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 136 of 185

55. FAO (2018f). Cultured Aquatic Species Information: Salmo trutta, In: FAO Fisheries and Aquaculture Department [online]. Rome. Available online at: www.fao.org/fishery/culturedspecies/Salmo_trutta/en, accessed 28 November 2018. 56. FAO (2018g). Cultured Aquatic Species Information: Gadus morhua, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Gadus_morhua/en, accessed 28 November 2018. 57. FAO (2018h). Species Fact Sheets: Homarus gammarus. In: FAO Fisheries and Aquaculture Department. Rome. Available online at: http://www.fao.org/fishery/species/2648/en, accessed 28 November 2018. 58. FAO (2018i). Cultured Aquatic Species Information: Crassostrea gigas, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Crassostrea_gigas/en, accessed 28 November 2018. 59. FAO (2018j). Cultured Aquatic Species Information: Ostrea edulis, In: FAO Fisheries and Aquaculture Department [online]. Rome. Available online at: www.fao.org/fishery/culturedspecies/Ostrea_edulis/en, accessed 28 November 2018. 60. FAO (2018k). Cultured Aquatic Species Information: Mytilus edulis, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Mytilus_edulis/en, accessed 28 November 2018. 61. FAO (2018l). Species Fact Sheets: Mytilus edulis. In: FAO Fisheries and Aquaculture Department. Rome. Available online at: http://www.fao.org/fishery/species/2688/en, accessed 28 November 2018. 62. FAO (2018m). Cultured Aquatic Species Information: Dicentrarchus labrax, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Dicentrarchus_labrax/en, accessed 28 June 2018. 63. FAO (2018n). Cultured Aquatic Species Information: Sparus aurata, In: FAO Fisheries and Aquaculture Department. Rome. Available online at: www.fao.org/fishery/culturedspecies/Sparus_aurata/en, accessed 28 June 2018. 64. Figueroa, F.L., Israel, A., Neori, A., Martinez, B., Malta, J.-E., Ang, P., Inken, S., Marquardt, R. and Korbee, N. (2009). Effects of nutrient supply on photosynthesis and pigmentation in Ulva lactuca (Chlorophyta): responses to short-term stress. Aquatic Biology, 7, pp. 173-183. 65. Filgueira, R., Grant, J. and Petersen, J. K. (2018). Identifying the optimal depth for mussel suspended culture in shallow and turbid environments. Journal of Sea Research. 132, pp. 15–23. 66. Fischer, H. (1986). Influence of temperature, salinity, and oxygen on the cadmium balance of mussels Mytilus edulis. Marine Ecology Progress Series, 32, pp. 265– 278. 67. Flye-Sainte-Marie, J., Jean, J., Paillard, C., Ford, S., Powell, E., Hofmann, E. and Klinck, J. (2007). Ecophysiological dynamic model of individual growth of Ruditapes philippinarum. Aquaculture, 266, pp. 130–143. 68. Fortes, M.D. and Lüning, K. (1980). Growth rates of North Sea macroalgae in relation to temperature, irradiance and photoperiod. Helgoländer Meeresunters, 34, pp. 15-29.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 137 of 185

69. Franco, S.C. (2017). Mapping areas of aquaculture potential for marine spatial planning in the UK: Analysis of the MMO 1040 aquaculture model and recommendations for future improvements. [Unpublished report for the MMO.] 70. Fredersdorf, J., Müller, R., Becker, S., Weincke, C. and Bischof, K. (2009). Interactive effects of radiation, temperature and salinity on different life history stages of the Arctic kelp Alaria esculenta (Phaeophyceae). Oecologia, 160(3), pp. 483–492. 71. Freitas, C., Moland, E., Moland, E., Ciannellis, L. and Knutsen, H. (2015). Behavioral responses of Atlantic cod to sea temperature changes. Ecology and Evolution, 5(10), pp. 2070–2083. 72. Freites, L., Cote, J. and Lodeiros, C. (1999). Effects of wave action on the growth and survival of the scallops Euvola ziczac and Lyropecten nodosus in suspended culture. Journal of Experimental Biology, 239, pp. 47–59. 73. Frimodt, C. (1995). Multilingual illustrated guide of the world’s commercial coldwater fish. Fishing News Books, pp. 215. 74. Gernez, P., Barill, L., Lerouxel, A., Mazeran, C. and Lucas, A. (2014). Remote sensing of suspended particulate matter in turbid oyster-farming ecosystems. Journal of Geophysical Research: Oceans, 119, pp. 7277–7294. 75. Gordon, M. S. (1958). Osmotic and Ionic Regulation in Scottish Brown Trout and Sea Trout (Salmo trutta L.). Journal of Experimental Biology, 36(2), pp. 253–260. 76. Gosling, G. (2003). Bivalve Molluscs Biology, Ecology and Culture. Oxford, UK: Fishing News Books. 77. Goulletquer, P. (2018). Datasheet report for Crassostrea gigas (Pacific oyster), In Invasive Species Compendium. CAB International, Wallingford, UK. Available online at: www.cabi.org/ISC/datasheet/87296, accessed 28 November 2018. 78. Halswell, P., Daniels, C. L. and Johanning, L. (2018). Framework for evaluating external and internal parameters associated with Sea Based Container Culture (SBCC): Towards understanding rearing success in European lobsters (Homarus gammarus). Aquacultural Engineering, 83, pp. 109–119. 79. Handeland, S.O., Berge, A., Björnsson, B.Th. and Stefansson, S.O. (1998). Effects of temperature and salinity on osmoregulation and growth of Atlantic salmon (Salmo salar L.) smolts in seawater, Aquaculture, 168, pp. 289–302. 80. Handeland, S.O., Björnsson, B.Th., Arnesen, A.M. and Stefansson, S.O (2003). Seawater adaptation and growth of post-smolt Atlantic salmon (Salmo salar) of wild and farmed strains. Aquaculture, 220, pp. 367-384. Available online at: https://www.sciencedirect.co m/science/article/pii/S0044848698003561, accessed 28 November 2018. 81. Handeland, S.O., Imsland, A.K. and Stefansson, S.O. (2008). The effect of temperature and fish size on growth, feed intake, food conversion efficiency and stomach evacuation rate of Atlantic salmon post-smolts. Aquaculture, 283, pp. 36-42. 82. Hanelt, D. and Figueroa, F. L. (2012). Physiological and photomorphogenic effects of light on marine macrophytes. In Wiencke, C. and Bischof, K. (eds) Seaweed Biology. Novel insight into ecophysiology, ecology and utilization., pp. 510. 83. Hardy, R. (2018). Datasheet report for Oncorhynchus mykiss (rainbow trout), In Invasive Species Compendium. CAB International, Wallingford, UK. Available online at: www.cabi.org/ISC/datasheet/71813, accessed 28 November 2018.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 138 of 185

84. Harris, M. (2016). A study of the naturalisation and dispersal of a non-native bivalve, the manila clam, Ruditapes philippinarum (Adams and Reeve 1850) in estuaries along the south coast of England. Thesis submitted to the University of Portsmouth. 85. Harrison, P.J. and Hurd, C.L. (2001). Nutrient physiology of seaweeds: Application of concepts to aquaculture., Cahiers de Biologie Marine, 42(1-2), pp. 71–82. 86. Haure, J., Penisson, C., Bougrier, S. and Baud, J.P. (1998). Influence of temperature on clearance and oxygen consumption rates of the flat oyster Ostrea edulis: determination of allometric coefficients. Aquaculture, 169, pp. 211–224. 87. Heal. R and Capuzzo E. (2019). Identification of Areas of Aquaculture Potential in English Waters, MMO Project No: 1184, May 2019, 109 pp. 88. Hill, J. (2008). Laminaria digitata: Oarweed, In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1386, accessed on 30 November 2018. 89. Hughes, J. (2008). Magallana gigas Pacific oyster. In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1676, accessed 30 November 2018 90. Humphreys, J. and Yang, N. (2018). Datasheet report for Ruditapes philippinarum (Japanese carpet shell). In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/2203, accessed 28 November 2018. 91. Hunt, H. L. and Scheibling, R. E. (2001). Predicting wave dislodgment of mussels - variation in attachment strength with body size, habitat, and season. Marine Ecology Progress Series, 213, pp. 157–164. 92. Hutchinson, S. and Hawkins, L. (1992). Quantification of the Physiological Responses of the European Flat Oyster Ostrea edulis L. to Temperature and Salinity. Journal of Molluscan Studies, 58(2), pp. 215–226. 93. Jahn, A., Theede, H. and Oeschger, R. (1992). The significance of intracellular amino acid regulation for freezing tolerance of the mussel Mytilus edulis L. 12th Baltic Marine Biologists Symposium, Helsingör, Denmark. Conference Proceedings, pp. 81-85. 94. Johansson, D., Ruohonen, K., Kiessling, A., Oppedal, F., Stiansen, J.-E., Kelly, M. and Juell, J.-E. (2006). Effect of environmental factors on swimming depth preferences of Atlantic salmon (Salmo salar L.) and temporal and spatial variations in oxygen levels in sea cages at a fjord site. Aquaculture, 254, pp. 594– 605. 95. Jonsson, B., Forseth, T., Jensen, J. and Næsje, T.F. (2001.) Thermal performance of juvenile Atlantic Salmon, Salmo salar L. Functional Ecology, 15, pp. 701–711. 96. Jorgensen, C. (1996) Bivalve filter feeding revisited. Marine Ecology Progress Series, 142, pp. 287–302. 97. Kang, H.Y., Lee, Y.-J., Choi, K.-S., Park, H.J., Yun, S.-G. and Kang, C.-K. (2016). Combined effects of temperature and seston concentration on the physiological energetics of the Manila clam Ruditapes philippinarum. PLoS ONE, 11(3), pp. 1– 17.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 139 of 185

98. Kapetsky, J. and Aguilar-Manjarrez, J. (2007). Geographic Information Systems, remote sensing and mapping for the development and management of marine aquaculture., FAO Fisheries Technical Paper, Number 458. Available online at: www.fao.org/3/a0906e/a0906e00.htm, accessed 30 November 2018. 99. Karna, D. W. (2003). A Review of Some of the Effects of Reduced Dissolved Oxygen on the Fish and Invertebrate Resources of Ward Cove, Alaska. Report for the Watershed Restoration Unit, Office of Water, U.S. Environmental Protection Agency. Available online at: https://nepis.epa.gov, accessed 30 November 2018. 100. Karsten, U., Dummermuth, A., Hoyer, K. and Wiencke, C. (2003). Interactive effects of ultraviolet radiation and salinity on the ecophysiology of two Arctic red algae from shallow waters. Polar Biology, 26(4), pp. 249–258. 101. Kasai, A., Horie, H. and Sakamoto, W. (2004). Selection of food sources by Ruditapes philippinarum and Mactra veneriformis (Bivalva: Mollusca) determined from stable isotope analysis. Fisheries Science, 70(1), pp. 11–20. 102. Kerrison, P. D., Stanley, M.S., Edwards, M.D., Black, K.D. and Hughes, A.D. (2015). The cultivation of European kelp for bioenergy: Site and species selection. Biomass Bioenergy, 80, pp. 229–242. 103. Kiceniuk, J.W. and Colbourne, E. (1997). Relating oxygen levels in the Newfoundland offshore waters to the physiology of Atlantic cod (Gadus morhua). Canadian Journal of Fisheries and Aquatic Sciences, 54, pp. 81-87. 104. Kim, W.S., Huh, H.T., Huh, S.-H. and Lee, T.W. (2001). Effects of salinity on endogenous rhythm of the Manila clam, Ruditapes philippinarum (Bivalvia: Veneridae). Marine Biology, 138(1), pp. 157–158. 105. King, J.W., Malham, S.K., Skov, M.W., Cotter, E., Latchford, J.W., Culloty, S.C. and Beaumont, A.R. (2006). Growth of Crassostrea gigas spat and juveniles under different environmental conditions at two sites in Wales. Aquatic Living Resources, 19(3), 289-297. 106. Kooijman, S.A.L.M. (2006) Pseudo-faeces production in bivalves. Journal of Sea Research, 56, pp. 103–106. 107. Korringa, P. (1952). Recent Advances in Oyster Biology. Quarterly Review of Biology, 27(3), pp. 266–308. 108. Laing, I. (2000.) Effect of temperature and ration on growth and condition of king scallop (Pecten maximus) spat. Aquaculture, 183, pp. 325–334. 109. Laing, I. (2002). Scallop Cultivation in the UK: a guide to site selection. Centre for Environment, Fisheries and Aquaculture Science (Cefas), Lowestoft, Available online at: www.Cefas.co.uk/publications/files/scallop_cultivation.pdf, accessed 30 November 2018 110. Laing, I. (2004). Filtration of king scallops (Pecten maximus). Aquaculture, 240, pp. 369–384. 111. Laing, I. and Spencer, B. E. (2006). Bivalve cultivation: criteria for selecting a site. Cefas Science Series (Technical report 136). Available online at: www.Cefas.co.uk/publications/techrep/techrep136.pdf, accessed 30 November 2018. 112. Laing, I., Utting, S.D. and Kilada, R.W.S. (1987). Interactive effect of diet and temperature on the growth of juvenile clams. Journal of Experimental Marine Biology and Ecology, 113, pp. 23–38. 113. Lambert, Y., Dutil, J.D. and Munro, J. (1994). Effects of intermediate and low- salinity conditions on growth rate and food conversion of Atlantic cod (Gadus

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 140 of 185

morhua). Canadian Journal of Fisheries and Aquatic Sciences, 51(7) pp. 1569– 1576. 114. Lanari, D, D’Agaro, E. and Ballestrazzi, R (2002). Growth parameters in European sea bass (Dicentrarchus labrax L.): effects of live weight and water temperature. Italian Journal of Animal Science, 1(3), pp. 181-185 115. Lancaster, D (2017). Developing scallop and geoduck aquaculture on British Columbia’s Central Coast: Recommendations from experts. Project report available at http://www.aerinjacob.ca/uploads/1/0/5/5/10559030/lancaster__2017__recomm endations_for_developing_shellfish_aquaculture_on_bc_central_coast.pdf. Accessed 30 November 2018. 116. Larsen, P.S. and Riisgård, H.U. (2016). Growth-Prediction Model for Blue Mussels (Mytilus edulis) on Future Optimally Thinned Farm-Ropes in Great Belt (Denmark). Journal of Marine Science and Engineering, 4(1), pp. 42–56. 117. Larsson, S. and Berglund, I. (2006). Thermal performance of juvenile Atlantic salmon (Salmo salar L.) of Baltic Sea origin. Journal of Thermal Biology, 31, pp. 243-246. 118. Lauzon-Guay, J.S., Barbeau, M.A., Watmough, J. and Hamilton, D.J. (2006). Model for growth and survival of mussels Mytilus edulis reared in Prince Edward Island, Canada. Marine Ecology Progress Series, 323, pp. 171–183. 119. Lewis, D.S. (2002). The effect of environmental temperature on growth and muscle development in the European lobster Homarus gammarus. Thesis submitted to the Institute of Biomedical and Life Sciences, University of Glasgow. Available online at: http://theses.gla.ac.uk/5507/, accessed 30 November 2018. 120. Longdill, P.C., Healy, T.R. and Black, K.P. (2008). An integrated GIS approach for sustainable aquaculture management area site selection. Ocean and Coastal Management, 51, pp. 612-624 121. Lucu, C. and Devescovi, M. (1999). Osmoregulation and branchial Na+, K+- ATPase in the lobster Homarus gammarus acclimated to dilute seawater. Journal of Experimental Marine Biology and Ecology, 234, pp. 291–304. 122. Lüning, K. (1984). Temperature tolerance and biogeography of seaweeds: the marine algal flora of Helgoland (North Sea) as an example. Helgoländer Meeresunters, 38(2), pp. 305–317. 123. Lüning, K. (1990). Seaweeds. Their environment, biogeography, and ecophysiology. In John Wiley and Sons, pp. 527. 124. Mainwaring, K., Tillin, H. and Tyler-Walters, H. (2014). Assessing the sensitivity of blue mussels (Mytilus edulis) to pressures associated with human activities. Report for the Joint Nature Conservation Committee. Number 506. Available online at: www.marlin.ac.uk/assets/pdf/JNCC_Report_506_web.pdf, accessed on 30 November 2018 125. Marshall, C.E. and Wilson, E. (2008). Pecten maximus Great scallop. In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1398, accessed 30 November 2018. 126. McCullough, D.A., Spalding, S., Sturdevant, D. and Hicks, M. (2001). Summary of Technical Literature Examining the Physiological Effects of Temperature on Salmonids. Issue Paper Number 5. United States Environment Protection Agency Report. Available online at: https://nepis.epa.gov, accessed 30 November 2018.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 141 of 185

127. McKay, L. and Gjerde, B. (1985). The effect of salinity on growth of rainbow trout. Aquaculture, 49, (3-4), pp. 325–331. 128. META (2019). Maritime and Environmental Thresholds for Aquaculture (META). Available online at https://longline.co.uk/meta/. Accessed 28 June 2019 129. van der Molen, J., Ruardij, P., Mooney, K., Kerrison, P., O-Connor, E., Gorman, E., Timmermans, K., Wright, S., Kelly, M., Hughes, A.D. and Capuzzo, E. (2018). Modelling potential production of macroalgae farms in UK and Dutch coastal waters. Biogeosciences, 15, pp. 1123–1147. 130. Morgan, K.C., Shacklock, P.F. and Simpson, F.J. (1980). Some aspects of the culture of Palmaria palmata in greenhouse tanks. Botanica Marina, 23(12), pp. 765–770. 131. Morgan, K.C. and Simpson, F.J. (1981). The cultivation of Palmaria palmata. Effect of light intensity and nitrate supply on growth and chemical composition. Botanica Marina, 24(5), pp. 273–277. 132. Msuya, F.E. and Neori, A. (2008). Effect of water aeration and nutrient load level on biomass yield, N uptake and protein content of the seaweed Ulva lactuca cultured in seawater tanks. Journal of Applied Phycology, 20, pp. 1021-1031. 133. Muss B.J. and Nielsen J.G. (1999). Sea fish. Scandinavian Fishing Yearbook, Hedehusene, Denmark, pp 340. 134. Nath, S.S., Bolte, J.P., Ross, L.G. and Manjarrex, J.A. (2000). Applications of geographical information systems (GIS) for spatial decision support in aquaculture. Aquacultural Engineering, 23, pp. 233-278. 135. Nel, A., Pitcher, G., Richoux, N.B. and Jackson, S. (2014). Culture environment and hatchery of origin influence growth, condition and feeding organ morphology in the Pacific oyster Crassostrea gigas in South Africa. African Journal of Marine Science, 36(4), pp. 481–491. 136. Nelson, T.A., Haberlin, K., Nelson, A.V., Ribarich, H., Hotchkiss, R., van Alstyne, K.L., Buckingham, L., Simunds, D.J. and Fredrickson, K. (2008). Ecological and physiological controls of species composition in green microalgal blooms. Ecology, 89, pp. 1287-1298. 137. Newcombe, C.P. and Jensen, J.O.T. (1996). Channel Suspended Sediment and Fisheries: A Synthesis for Quantitative Assessment of Risk and Impact. North American Journal of Fisheries Management, 16(4), pp. 693–727. 138. Oppedal, F., Dempster, T. and Stien, L.H. (2011). Environmental drivers of Atlantic salmon behaviour in sea-cages: A review. Aquaculture, 311, pp. 1–18. 139. OSPAR (2009). Background document for Ostrea edulis and Ostrea edulis beds. Biodiversity Series Report prepared for the OSPAR commission. Available online at: https://qsr2010.ospar.org/media/assessments/Species/P00428_ostrea_edulis_a nd_beds.pdf, accessed 30 November 2018 140. OSPAR (2010). Intertidal Mytilus edulis beds on mixed and sandy sediments., Quality Status Report 2010. Available online at: https://qsr2010.ospar.org/media/assessments/ Species/p0010_supplements/CH10_03_Intertidal_mytilus_edulis.pdf, accessed 30 November 2018. 141. Pauley, G., Van Der Raay, B. and Troutt, D. (1988) Species Profiles: Life Histories and Environmental Requirements of Coastal Fishes and Invertebrates: Crassostrea gigas. United States Fish and Wildlife Biological Report, 82(4), pp. 1–28.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 142 of 185

142. Perry, F. and Jackson, A. (2017). Ostrea edulis (Native oyster). In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1146, accessed 30 November 2018. 143. Pieterse, A. (2013). Growth, condition, survival and feeding rate of the Pacific oyster (Crassostrea gigas Thunberg) cultured in three distinct South African environments. Thesis submitted to the University of Stellenbosch. URL: http://scholar.sun.ac.za, accessed 30 November 2018. 144. Plaut, I. (2001). Critical swimming speed: its ecological relevance. Comparative Biochemistry and Physiology – Part A, 131, pp. 41–50. 145. Price, H. (1982.) An analysis of factors determining seasonal variation in the byssal attachment strength of Mytilus edulis. Journal of the Marine Biological Association of the United Kingdom, 62(1), pp. 147–155. 146. Quillet, E., Faure, A., Chevassus, B., Krieg, F., Harache, Y., Arzel J. and Metailler, R. (1992). The potential of brown trout (Salmo trutta L.) for mariculture in temperate waters. Icelandic Agricultural Sciences, 6, pp. 63–76. 147. Radiarta, N., Saitoh, S.-I. and Yasui, H. (2011). Aquaculture site selection for Japanese kelp (Laminaria japonica) in southern Hokkaido, Japan, using satellite remote sensing and GIS-based models. ICES Journal of Marine Science, 68(4), p. 773-780. 148. Ren, J.S. and Schiel, D.R. (2008). A dynamic energy budget model: parameterisation and application to the Pacific oyster Crassostrea gigas in New Zealand waters. Journal of Experimental Biology, 361, pp. 42–48. 149. Riisgård, H.U., Bøttiger, L. and Pleissner, D. (2012). Effect of Salinity on Growth of Mussels, Mytilus edulis, with Special Reference to Great Belt (Denmark). Open Journal of Marine Science, 2, pp. 167–176. 150. Riisgård, H.U., Egede, P.P. and Barreiro Saavedra, I. (2011). Feeding Behaviour of the Mussel, Mytilus edulis: New Observations, with a Minireview of Current Knowledge. Journal of Marine Biology, 2011, pp. 1–13. 151. Roberts, D. (2018a). Datasheet report for Mytilus edulis (common blue mussel). In CABI Invasive Species Compendium. Available online at: www.cabi.org/isc/datasheet/73755, accessed: 28 November 2018. 152. Roberts, D. (2018b). Datasheet report for Ostrea edulis (European oyster). In CABI Invasive Species Compendium. Available online at: www.cabi.org/ISC/datasheet/71177, accessed: 28 November 2018. 153. Roesijadi, G., Buenau, K.E., Coleman, A.M., Tagestad, J.D., Judd, C., Wigmosta, M.S., Van Cleve, B., Ward, J.A. and Thom, R.M. (2011). Macroalgae analysis. A national GIS-based analysis of macroalgae production potential. Summary report and project plan. Report prepared for the US Department of Energy (Contract DE-AC05-76RL01830). Available online at: www.pnnl.gov/main/publications/external/technical_reports/PNNL-21087.pdf, accessed 01 March 2019. 154. Rosland, R., Strand, Ø., Alunno-Bruscia, M., Bacher, C. and Strohmeier, T. (2009). Applying Dynamic Energy Budget (DEB) theory to simulate growth and bioenergetics of blue mussels under low seston conditions. Journal of Sea Research, 62, pp. 49–61. 155. Scott, D.C.B. and Muir, J.F. (2000). Offshore cage systems: A practical overview. In: Muir J. and Basurco B.(eds.). Mediterranean offshore mariculture. Zaragoza:

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 143 of 185

CIHEAM, 2000. pp. 79-89 (Options Méditerranéennes: Série B. Etudes et Recherches; n. 30) 156. Seafish (2015) Responsible Sourcing Guide: Farmed Sea Bass. Available online at: https://www.yumpu.com/en/document/view/24485412/seafish-responsible- sourcing-guide-sea-bass, accessed 01 December 2019 157. Seafish (2019a). Clam cultivation. Aquaculture Methods Leaflet. Seafish Industry Authority. Available online at: https://www.seafish.org/responsible- sourcing/aquaculture-farming-seafood/starting-an-aquaculture- business/selecting-a-species/, accessed 01 March 2019 158. Seafish (2019b). Lobster hatcheries. Aquaculture Methods Leaflet. Available online at: https://www.seafish.org/responsible-sourcing/aquaculture-farming- seafood/starting-an-aquaculture-business/selecting-a-species/, accessed 01 March 2019 159. Seafish (2019c). Native oyster cultivation. Aquaculture Methods Leaflet. Available online at: https://www.seafish.org/responsible-sourcing/aquaculture- farming-seafood/starting-an-aquaculture-business/selecting-a-species/, accessed 01 March 2019 160. Seafish (2019d) Pacific oyster cultivation. Aquaculture Methods Leaflet. Available online at: https://www.seafish.org/responsible-sourcing/aquaculture-farming- seafood/starting-an-aquaculture-business/selecting-a-species/, accessed 01 March 2019 161. Seafish (2019e). Rope-grown Mussel Cultivation Site selection. Aquaculture Methods Leaflet, pp. 52–53. Available online at: https://www.seafish.org/responsible-sourcing/aquaculture-farming- seafood/starting-an-aquaculture-business/selecting-a-species/, accessed 01 March 2019 162. Seafish (2019f). Scallop cultivation. Aquaculture Methods Leaflet. Available online at: www.seafish.org/media/401787/scallop_cultivation.pdf, accessed 01 March 2019 163. Seafish (2019g) Seabed cultivation of mussels. Aquaculture Methods Leaflet. Available online at: https://www.seafish.org/responsible-sourcing/aquaculture- farming-seafood/starting-an-aquaculture-business/selecting-a-species/ accessed 01 March 2019 164. Shpigel, M., Barber, B. J. and Mann, R. (1992). Effects of elevated temperature on growth, gametogenesis, physiology, and biochemical composition in diploid and triploid Pacific oysters, Crassostrea gigas Thunberg. Journal of Experimental Marine Biology and Ecology, 161(1), pp. 15–25. 165. Shpigel, M. and Spencer, B. (1996). Performance of diploid and triploid Manila clams (Tapes philippinarum, Adams and Reeve) at various levels of tidal exposure in the UK and in water from fishponds at Eilat, Israel. Aquaculture, 141, pp. 159–171. 166. Shumway, S.E., Cucci, T.L. and Newell, C. (1985). Particle selection, ingestion, and absorption in filter-feeding bivalves. Journal of Experimental Marine Biology and Ecology, 91, pp. 77-92. 167. Siriwardena, S. (2018a). Datasheet Report for Salmo salar (Atlantic salmon), In CABI Invasive Species Compendium. Available online at: www.cabi.org/ISC/datasheet/65307, accessed: 28 November 2018.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 144 of 185

168. Siriwardena, S. (2018b). Datasheet report for Salmo trutta (sea trout). In CABI Invasive Species Compendium. Available online at: www.cabi.org/ISC/datasheet/65308, accessed: 28 November 2018. 169. Snyder, J., Boss, E., Weatherbee, R., Thomas, A.C, Brady, D. and Newell, C. (2017). Oyster Aquaculture Site Selection Using Landsat 8-Derived Sea Surface Temperature, Turbidity, and Chlorophyll a. Frontiers in Marine Science, 4, pp. 190-200. 170. Sobral, P. and Widdows, J. (2000). Effects of increasing current velocity, turbidity and particle-size selection on the feeding activity and scope for growth of Ruditapes decussatus from Ria Formosa, southern Portugal. Journal of Experimental Marine Biology and Ecology, 245, pp. 111–125. 171. Solomon, D. and Lightfoot, G. (2008). The thermal biology of brown trout and Atlantic salmon. Environmental Agency, UK Report. Available online at: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/att achment_data/file/291741/scho0808bolv-e-e.pdf, accessed 30 November 2018. 172. Spencer, B. E. (1990). Cultivation of Pacific Oysters (Laboratory Leaflet Number 63). Directorate of Fisheries Research, Ministry of Agriculture, Fisheries and Food. 173. Spencer, B., Edwards, D. and Millican, P. (1991). Cultivation of Manila clams (Laboratory Leaflet 65). Directorate of Fisheries Research, Ministry of Agriculture, Fisheries and Food. 174. Steenbergen, J., Baars, J.M.D.D., van Stralen, M.R. and Craeymeersch, J.A. (2005). Winter survival of mussel beds in the intertidal part of the Wadden Sea. In: Monitoring and Assessment in the Wadden Sea. Proceedings from the 11th Scientific Wadden Sea Symposium, Denmark, Esbjerg 4.-8. April 2005 (Laursen, K. Ed.). NERI Technical Report No. 573, pp. 107-111. 175. Steeves, L.E, Filgueira, R., Guyondet, T., Chassé, J. and Comeau, L. (2018). Past, Present, and Future: Performance of Two Bivalve Species Under Changing Environmental Conditions. Frontiers in Marine Science, 5, pp. 184–198. 176. Stelzenmüller, V., Gimpel, A., Gopnik, M. and Gee, K. (2017). Aquaculture site- selection and marine spatial planning: the roles of GIS-based tools and models. In: Buck, B.H. and Langan, R.(eds.) Aquaculture perspective of multi-use sites in the open ocean, Chapter 6, pp. 131-148. 177. Stevens, A.M. and Gobler, C.J. (2018). Interactive effects of acidification, hypoxia, and thermal stress on growth, respiration, and survival of four North Atlantic bivalves. Marine Ecology Progress Series, 604, pp. 143–161. 178. Strand, Ø., Solberg, P.T., Andersen, K.K. and Magnesen, T. (1993). Salinity tolerance of juvenile scallops (Pecten maximus) at low temperature. Aquaculture, 115(1-2), pp. 169–179. 179. Strand, Ø., Louro, A. and Duncan, P. (2016). European Aquaculture. In Shumway, S.E. and Parsons, G.J. (eds.) Scallops: Biology, Ecology, Aquaculture, and Fisheries. 3rd Edition. Elsevier Science. Chapter 20. 180. Strohmeier, T. (2008). Feeding behavior and bioenergetic balance of the great scallop (Pecten maximus) and the blue mussel (Mytilus edulis) in a low seston environment and relevance to suspended shellfish aquaculture. Thesis submitted to the University of Bergen. 181. Sundene, O. (1962). The implications of transplant and culture experiments on the growth and distribution of Alaria esculenta. Nytt Magasin for Botanikk, 9, pp. 155–174.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 145 of 185

182. Szostek, C.L., Davies, A.J. and Hinz, H. (2013). Effects of elevated levels of suspended particulate matter and burial on juvenile king scallops Pecten maximus, Marine Ecology Progress Series, 474, pp. 155–165. 183. Toro, J., Paredes, P.I., Villagra, D.J. and Senn, C.M. (1999). Seasonal Variation in the Phytoplanktonic Community, Seston and Environmental Variables During a 2-Year Period and Oyster Growth at Two Mariculture Sites. Marine Ecology, 20(1), pp. 63–89. 184. Tudorache, C., de Boeck, G. and Claireaux, G. (2013). Forced and Preferred Swimming Speeds of Fish: A Methodological Approach. In Palstra, A.P. and Planas, J.V. (eds.) Swimming Physiology of Fish, pp. 81–108. 185. Tyler-Walters, H. (2008). Alaria esculenta Dabberlocks, In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1291, accessed: 28 November 2018. 186. Utting, S. and Spencer, B. (1992). Introductions of marine bivalve molluscs into the United Kingdom for commercial culture - case histories. ICES Marine Science Symposium, 194, pp. 84–91. 187. Valero, J. (2006.) Ostrea edulis Growth and mortality depending on hydrodynamic parameters and food availability. Thesis submitted to the Department of Marine Ecology, Goteborg University. 188. Vincenzi, S., Caramori, G., Rossi, R. and De Leo, G.A. (2006.) A GIS-based habitat suitability model for commercial yield estimation of Tapes philippinarum in a Mediterranean coastal lagoon (Sacca di Goro, Italy). Ecological Modelling, 193, pp. 90–104. 189. Wadsworth, P., Wilson, A. E. and Walton, W. C. (2019). A meta-analysis of growth rate in diploid and triploid oysters. Aquaculture, 499, pp. 9–16. 190. Walne, P. (1958). Growth of oysters. Journal of Marine Biology Association of the UK, 37, pp. 591–602. 191. Werner, A. and Dring, M. (2011). Aquaculture explained. Cultivating Palmaria palmata. No. 27. Project, PBA/SW/07/001. Available online at: www.bim.ie/media/bim/content/publications/Aquaculture,Explained,Issue,27,- ,Cultivating,Palmaria,palmata.pdf, accessed 01 March 2019. 192. White, N. and Marshall, C.E. (2007). Saccharina latissima Sugar kelp. In Invasive Species Compendium. Edited by Tyler-Walters, H. and Hiscock, K. Plymouth: Marine Biological Association of the United Kingdom. Available online at: www.marlin.ac.uk/species/detail/1375, accessed: 28 November 2018. 193. Wildish, D.J., Keizer, P.D., Wilson, A.J. and Martin, J.L. (1993). Seasonal Changes of Dissolved Oxygen and Plant Nutrients in Seawater near Salmonid Net Pens in the Macrotidal Bay of Fundy. Canadian Journal of Fisheries and Aquatic Sciences, 50(2), pp. 303–311. 194. Wildish, D.J., Kristmanson, D.D. and Saulnier, A.M. (1992). Interactive effect of velocity and seston concentration on giant scallop feeding inhibition. Journal of Experimental Marine Biology and Ecology, 155, pp. 161–168. 195. Wildish, D.J. and Saulnier, A.M. (1992). The effect of velocity and flow direction on the growth of juvenile and adult giant scallops. Journal of Experimental Marine Biology and Ecology. 155(1), pp. 133.143. 196. Wood, D., Capuzzo, E., Kirby, D., Mooney-McAuley, K. and Kerrison, P. (2017). UK macroalgae aquaculture: what are the key environmental and licensing considerations? Marine Policy, 83, pp. 29–39.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 146 of 185

197. Yang, C., Sierp, M.T., Abbott, C.A., Li, Y. and Qin, J.G. (2016). Responses to thermal and salinity stress in wild and farmed Pacific oysters Crassostrea gigas. Comparative Biochemistry and Physiology, Part A., 201, pp. 22–29. 198. Yin, X., Chen, P., Chen, H., Jin, W. and Yan, X. (2017). Physiological performance of the intertidal Manila clam (Ruditapes philippinarum) to long-term daily rhythms of air exposure. Scientific Reports, 7, article 41648. 199. Yoshikawa, T., Murata, O., Furuya, K. and Eguchi, M. (2007). Short-term covariation of dissolved oxygen and phytoplankton photosynthesis in a coastal fish aquaculture site. Estuarine Coastal and Shelf Science, 74(3), pp. 515-527. 200. Young, G.A. (1985) Byssus-thread formation by the mussel Mytilus edulis: effects of environmental factors. Marine Ecology Progress Series, 24, pp. 261–271. 201. Zippay, M.L. and Helmuth, B. (2012). Effects of temperature change on mussel, Mytilus. Integrative Zoology, 7(3), pp. 312–327.

Appendix 4: Stakeholder consultation workshops

Aquaculture Mapping workshops: Notes from Weymouth, Lyme and Swanage sessions

A4.1 Introduction

Dorset Coast Forum provided facilitation support for three workshops to inform the work relating to the Cefas Aquaculture Mapping project. The workshops were organised to coincide with the best opportunities to engage with fishers who are key stakeholders in the areas being mapped, and invitations were sent to a range of other key stakeholders across recreation, aquaculture, local authority, harbour and port management, charter operations and coastal management. Three locations were chosen to provide maximum opportunity for local engagement and minimise the need for participants to travel.

A4.2 Workshop aim

The purpose of the project is to identify and map areas best suited to specific types of sustainable aquaculture, within the boundaries of the Dorset and East Devon FLAG area.

These workshops provide input into this process, consulting with key stakeholders to identify potential management and mitigation measures that may facilitate sustainable aquaculture in the FLAG area without unduly restricting opportunities for fishing.

Following consultation with the industry, the project will produce an in-depth analysis which should benefit aquaculture interests in the FLAG area and act as an exemplar approach for others to adopt

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 147 of 185

A4.3 Workshop tasks

Participants were given a brief overview of modern aquaculture methods and which species and general conditions were best suited to each, in order to inform the discussions through the session.

Looking at the whole of the FLAG area (Dorset’s and East Devon’s coastal waters between Beer and Swanage), participants were tasked with identifying locations which would provide good opportunities (and/or minimum constraints or conflicts) for development of aquaculture across a range of species. This was introduced over a series of three maps with different layers of background data to drive the discussions.

Participants were also invited to highlight concerns or issues associated with different areas or types of aquaculture in specific locations.

Input was encouraged via colour-coded post-it notes* directly on the maps provided, but facilitators at Weymouth and Swanage also captured any additional comments separately; these are given below the images as a separate section (all comments at the Lyme Regis workshop were captured on the maps).

Post-it notes (and species stickers) were colour coded according to the sector of stakeholders: green- NGOs, yellow- fisher/ fisher representative, blue- Local Authority/IFCA/Government, Pink- others.

Map 1- Human activities and Aquaculture Space This map was provided as a base map with bathymetry and contours, vessel movements (MMO AIS densities), angling intensity and sailing areas.

Question: Where are the main areas of human activity, and where might be available for aquaculture space?

Swanage workshop – East Comments

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 148 of 185

1 Waste impact on MPA feature

2 MCZ feature will limit harvest methods

3 Fishing (pots). Busy for ferries in and out of Poole

Comment Swanage workshop – West No markers put on West map

Weymouth workshop – East Comments 1 Bass/ Gilthead/ bream- successful culture requires economic viability 2 Issues with alternative uses of area and navigational hazard 3 Heavily utilised fishing area 4 Cage systems- too early in development to offer practical use in exposed areas of coast 5 Possibly by negotiation with harbour authority and other users of the water space

Weymouth workshop – West Comments

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 149 of 185

1 Pots and nets (arrows to sites) 2 Fleet narrows. Seaweed. Needs to have no detrimental effects on protected habitats and species of high conservation value. 3 Suspended longline- water exchange required but some shelter needed. Navigation/fishing conflict in most areas.

Lyme Regis workshop – East Comments 1 Being considered for longline scallops 2 Possible but heavily used for sailing/windsurfing 3 Portland Harbour may be a good site for shaded and bottom culture in areas of existing or known zones. Potential for increase in bottom culture without too much displacement 4 There are only 4 existing zones along northern breakwater. There is significant pressure on other areas within the wider harbour.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 150 of 185

Map 2- Marine protected areas and modelled aquaculture suitability – biological and physical This map was overlain with habitat types and conservation areas with inset maps of modelled aquaculture suitability areas.

A scale of aquaculture suitability mapping ranging between 0 and 1 was employed where 0 = unsuitable, 0.5 = sub-optimal and 1 = optimal. Aquaculture suitability presented in Maps 2 and 3 utilised a scale range between 0.5-1.0 and were shaded accordingly.

Question: Add thoughts and concerns regarding the most suited (darker) areas for each aquaculture type and express views on any areas that should be avoided (and if possible, by system and species) and that could be prioritised.

Modelled aquaculture potential aggregated by species groups (darker shading represents increased potential)

Lyme Regis workshop – West Comments 1 Existing lobster fisheries exist here- maybe get buy-in from this sector. 2 might be sustainable for seaweed but need to consider conflict and impact on existing legitimate fisherman. 3 Known to have existing longline system here 4 need to consider MPAs and existing fishing activities 5 Possible expansion of existing oyster production- other species? Clams? 6 It would be helpful to compare mapping areas to the potting study. It would identify areas of ‘no take’ and to place near those areas would be ideal. Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 151 of 185

Swanage workshop- East Comments 1 Finfish not compatible with MPA features.

2 South of Handfast Point (apart from Swanage Bay which has fishing gear in all year round as it is a safe area for static fishing gear during bad weather) is not suitable due to strong tides.

Comment Swanage workshop – West No markers put on West map

Weymouth workshop- East Comments 1 Portland Harbour Authority aim to promote aquaculture on a crofting basis and not to disrupt existing grower using the water. 2 Portland Harbour Authority have applied for a Several Order that covers the port area. Is all suitable for seabed and hanging/ long line culture of seaweed and bivalve shellfish. 3 Nothing anywhere. 4 This might work. 5 This area is the most contentious for fishers for any type of scheme 6 This might work. 7 This area will impact on other operators from Grove Point to Lulworth if using surface hand lines. 8 This might work.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 152 of 185

Weymouth workshop- West Comments 1 This (location) might work. 2 Fishers will not support lobster farming- undermines own catch value. 3 Most aquaculture has potential to conflict with protected/ conservation species zones. Core matching of ‘least impacts’ should reveal best scope for aquaculture. 4 Congested with pots, nets and mobile gear in permitted areas. 5 Nutrients from Chesil outfall make the area around the

diffusers ideal for bottom shellfish culture.

Lyme Regis workshop- East Comments 1 Several Order being considered by Portland Port. 2 Shore bay site available balaclava bay. 3 There is potential at these (marked) points. 4 High potential area is probably in the main shipping lanes?

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 153 of 185

Lyme Regis workshop- West Comments 1 Many of these potential areas are in heavily fished areas and need more evidence and consultation. 2 Could more bivalve fisheries be considered in closed areas? 3 Need to consider existing small- scale inshore fisheries where communities exist around them. 4 M/A- what is NE’s policy on aquaculture in these sites? 5 SI area here (marked) could be aquaculture site but need to consider existing opportunities. 6 Always lobster potential in an area where lobster fishery exists. Buy-in from local community fisheries? 7 Statutory instrument extends beyond the MPA- not reef habitat but listed as biodiversity value. Could aquaculture occur here? What are the existing fisheries? What is the impact? 8 Commercial trawling and scalloping areas- important as large areas closed to mobile gear. 9 Near Beer and Axmouth there are many pots but not many trawlers. If equipment were to go outside the MPA I would suggest going closer to Beer and Axmouth. Less potting near Lyme and West Bay, but there are more trawlers. 10 Need to add Fleet Firing Zone.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 154 of 185

Map 3- Potential aquaculture areas with least conflicts This map showed potential aquaculture areas with least conflicts compared with hard constraints e.g. existing sites, wrecks (bottom aquaculture), transportation routes. However tidal current had not been considered at this stage of the analysis.

Participants were asked to add sticky dots (6 of each colour per person for both east and west areas) representing the following: Dark Purple = Definitely don’t approve; Blue = Maybe; Yellow = OK.

Swanage workshop – East

Comment Swanage workshop – West No markers put on West map

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 155 of 185

Map 3- Potential aquaculture areas with least conflicts (cont’d.)

Weymouth workshop- East

Comments 1 Not ideal for anyone other than the aquaculture operator. 2 Could possibly work. 3 Unlikely to work. 4 Different species of seaweed require different habitat. Whole coastline’s intertidal and shallow water has potential for harvesting. Deeper water has potential for long line culture. 5 Unlikely to work. 6 Could possibly work.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 156 of 185

Weymouth workshop- West

Comments - None

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 157 of 185

Lyme Regis workshop- West

Comments 1 Areas outside the SI and SAC not used for any aquaculture system will need to consider impact on other legitimate users (plus seaweed- both maps). 2 Seems to be a general potential are. 3 Lobster aquaculture could be a form of diversification for local lobster fishers, creating new markets rather than new businesses forming. 4 Purple areas for bivalves, lobsters and seaweed are heavily dredged. Trawl maybe- huge conflict. Need to be consulted, so not necessarily so. 5 NB- Ecosystem benefits of aquaculture increase biodiversity, protection areas and juvenile areas. 6 It would be helpful to have an overview of the fishing gear in the area, as well as routes the trawlers take to avoid static gear. Would be great to have fishermen give their views on this map. 7 Not opposed to finfish farms but I don’t have enough knowledge of what would be the impact? And priority? So not put stickers on.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 158 of 185

Lyme Regis workshop- East

Comments - None

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 159 of 185

Additional comments captured throughout workshops: Question/task Weymouth workshop Swanage workshop Project Q - Does project exclude shore-based introduction systems? A- not necessarily but is restricted to marine species. Study is really about marine space rather than where species are/can be grown. Q - How will end product of project be used to inform site selection? A - yes- to identify areas with least contention/best suitability. Could potentially lead to ‘presumption in favour’ BUT any development would still be subject to the relevant consents and licences. Q - Reassuring to hear that there is thought about where might be suitable, rather than just blanket ‘come to Dorset’ invitation. There is a clear need to make sure development is in harmony with the local fleet. A - Cefas acknowledges that there is intensive use of the Dorset Coast and any future aquaculture development needs to complement that. Aquaculture Suspended longline culture Q - is there data for what kind of impacts are methods- (contained) created by each type of system? familiarisation Q - How deep for ‘blue A - there is some data from Natural England, buoys’ (see image)? but there are lots of variables, so any A - species-dependant; can development would still need individual be all buoys on the surface. site/species specific impact assessment. Submerged cage Q - Can maintenance be done by SCUBA? A - Again, depends. Depth is limiting factor. ROVs can/could be used. Map 1 Tide is critical to placement The whole area (Dorset inshore waters) is Human activities of certain shellfish; mussels well-used. It’s difficult to pinpoint specific areas and Aquaculture need a 2knt tide minimum. where aquaculture wouldn’t displace other Space activity (recreation or fishing effort). Wherever you pick, somebody will likely be potting there.

Group was reluctant to place any development within the MPAs because of existing protections and fishing efforts. Map 2 Distance from nearest port Strong tides south of Peverel Point (Swanage); Marine protected vs economic viability high potential to lose gear. areas and Exclusions for fishing BUT modelled please include leisure in Interesting to note that some areas are shown aquaculture general narrative. in the modelling as higher potential than others suitability – (e.g. on the lobster/finfish maps). Could Cefas biological and explain what the influencing factors are in the physical model?

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 160 of 185

Question/task Weymouth workshop Swanage workshop Map 3 Generally, tides and nutrient flow will dictate Potential that systems work better in bays than channel aquaculture areas areas (evidence from potting) with least conflicts Could Cefas test please that the modelling constraints are ‘real’ (logic tested) so the study doesn’t rule in/out areas which should have constraints and don’t or vice versa in a real situation.

E.g. Swanage Bay- might be suggested as ‘suitable’ by model, but there’s so much varied and intense use of the area that it’s not realistic.

Need to find areas with less tide so it’s workable.

A4.4 Next steps

The Cefas team will now use the above information as part of the Aquaculture Mapping research currently being carried out; outputs on this research will be available in Autumn 2019 and will be presented at a workshop (January 2020) open to all invitees of the research workshops. The final report and data from the project will also be publicly available through the Cefas Data hub.

The data from the workshops will also feed into an interactive mapping tool on a new Virtual Aquaculture Hub which is being designed and set up by DCF’s Aquaculture Development Officer; information on this new resource will be available late Autumn 2019.

Both the projects described above are funded separately by the Dorset and East Devon Fisheries Local Action Group (FLAG).

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 161 of 185

Appendix 5: Sea outfall modelling

A5.1 Overview

The aim of this work is to predict the spreading and extent of nearshore/coastal effluents regarding the dispersion of the Escherichera coli (E. coli) bacteria.

A Telemac tidal model was constructed, calibrated and subject to validation. The validated model has then been used for a water quality study and has been forced with 6 nearshore effluents containing a number of E. coli bacteria. In a 4 second phase of the water quality modelling, a T90 decay rate is included to better replicate E. coli.

The results of the model are presented for critical tidal conditions, i.e. high water and low water for spring peak flood and peak ebb, and areas above limits are extracted.

A5.2 Mesh construction

As shown on Figure A-1, the mesh offshore boundary is curved as this improves the boundary tidal forcing. The inshore boundary follows the coast, while the very fine coastal features like the Fleet channel near Weymouth that cannot be reproduced within the current mesh coastal resolution, have been dropped. The bathymetry information stems from Astrium (Astrium, 2011), with bathymetry information ranging from 4 m deep, inshore to ~100 m deep, offshore (Figure A-1). The mesh refinement was based on the bathymetry, with fine resolution (100 m) in shallow water and coarser cell offshore (5 km).

A5.3 Steering file parameters

Model stability The regular mesh and smooth increase in size associated with numerical parameters enabled the model to run with a time step of 15 s. The results using a smaller time step of 5 s or 2 s are comparable with the 15 s timestep-results. Using such a large time-step enabled a swift calculation (~few minutes for simulating a complete tidal cycle).

Bed friction SCOPAC work (New Forest District Council, 2017) indicated sand and gravel sediment transport along the shore, while the offshore sediment transport mainly consisted of sand transport. It is hence assumed that sediment on seabed is non-cohesive sediment with a grain size in the range of a few

4 T90 is the time necessary for the time necessary for 90% of the E. coli bacteria to dye Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 162 of 185

millimetres. The roughness on the seabed is estimated using the formulation from Soulsby, (1997) with a grain size of 3 mm: Nikuradse roughness5 is 0.0075.

Minor tidal constituents The inclusion of the minor tidal constituents did not amend the predicted tides and hence has not been kept as an option in the steering file.

A5.4 Tidal calibration Observation The tidal observation sites from British Oceanographic Data Centre (BODC), indicated with a blue triangle and a purple diamond in Figure A-1 are located nearshore in the mesh, while the Admiralty data in small red squares provided information on tidal ranges close to the shore.

The purple diamond and the blue triangle indicate BODC velocity and elevation, respectively. The coastal small red dots locate Admiralty tidal range information at, from west to east, Turf Lock, Lyme Regis, Bridport (West Bay), Chesil Beach, and Swanage. Figure A-1: Telemac mesh coloured following water depth information.

5 Relative roughness based on the diameter of uniform sand grain particles Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 163 of 185

Figure A-2 Predicted water depth over tidal cycle at the BODC elevation point (blue triangle in Figure A-1) Calibration method The calibration method consisted of checking the model sensitivity to the friction parameter, as the friction can influence the model hydrodynamics. The model has been run over a tidal cycle, with results saved once the model is stabilised (no wiggling at the beginning of the calculation (model ‘spin-up’), see Figure A- 2. The tidal coefficients have been calibrated against Admiralty coastal tidal range observation (Figure A-1, red dots). The calibrated model predictions have then been compared against BODC velocity and elevation observations and predicted coastal recirculation was compared with website data.

Calibration results Sensitivity to friction coefficient Here, changing the friction hardly changed the hydrodynamics, and the friction used is calculated from the onsite grain size of ~3 mm.

Tidal coefficient The best fit was obtained with a forced Tidal Range Coefficient of 0.5. Using a coefficient for tidal range of 0.5 results in better tidal range prediction at the different coastal locations (Table A 18).

Table A 18: Table of tidal coefficient calibration Location Turf Lyme Bridport Chesil Mupe Swanage Lock Regis Beach Bay Admiralty (M2+S2) 1.65 1.67 1.64 1.58 0.90 0.58 Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 164 of 185

Location Turf Lyme Bridport Chesil Mupe Swanage Lock Regis Beach Bay No calibration 3 3 3 3 2.3 1.7 Calibration1 - Tidal range 1.5 1.5 1.5 1.5 1 0.7 0.5

Reproduction of local current velocity

While the velocity magnitudes are not always adequately reproduced in the model (Figure A-3), with some diurnal inequalities missing in model prediction, the prediction is perfectly in phase with the observations.

Figure A-3: Velocity from model prediction and observation in cm/s format X axis

Reproduction of local surface elevation

Over the ebbing phase of the tidal cycle (Figure A-4), the model prediction is perfectly in phase with the observation. Good match in elevation is observed for the low tide, while for the high tide some discrepancies exist, probably due to wind and wave effect.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 165 of 185

Figure A-4: Predicted and observed water elevation over the ebbing phase of the tidal cycle.

Validation results Reproduction of global current magnitude

The validation phase consisted in comparing the Telemac predicted velocities magnitude against model prediction from ABPmer (ABPmer, 2017) and against the ‘Monty Mariner‘ observations off Portland.

In Figure A-5 the predicted magnitude of the maximum current from Telemac model matches the predictions form ABPmer.

ABPmer predictions Current Telemac model

Figure A-5: Comparison ABPmer predictions with Telemac predictions.

Reproduction of coastal re-circulation

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 166 of 185

Eddy features around Weymouth Bill: Plots stemming from the (‘Monty Mariner’ n.d.) website http://montymariner.co.uk/tide-stream-charts/tide-stream-charts-2/ are compared with current predictions from the model. These tidal data are derived from Admiralty Tidal Stream Vector Data, with the peak speed and direction of the tidal stream indicated for spring and neap tide, respectively.

The velocity dimension is tenth of a knot, i.e. for the top figures in Figure A-6 (6 h before HW), at the extreme south of the Portland island the maximum global velocity is indicated in the Monty Mariner map as 32, or 3.2 knots (1.63 m/s), and the predicted velocity over this area in the model ranges between 1.4 and 2.81 m/s, with the highest predicted rate found close to the tip of the island as expected.

In the model, high water at Portland is predicted on the 23/01/2019 at 08:15:00 (Figure A5.6).

Time Admiralty tide vectors Validated model source: Monty Mariner, n.d.

Portland

HW-6 h

23/01/2019 02:15:00

Portland HW-5 h

23/01/2019 03:15:00

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 167 of 185

Time Admiralty tide vectors Validated model source: Monty Mariner, n.d.

Portland

HW-4 h

23/01/2019 04:15:00

Portland

HW 23/01/2019 08:15:00

Figure A-6: Comparison of current pattern in the vicinity of Weymouth Bay.

Over the area, both the current direction and the current velocities are matching the admiralty data.

A5.5 Validated model tidal peak velocity and elevation

From the validated model, a central point has been chosen (node 6470 in the mesh), see Figure A-7: The simulated water elevation and velocity time-series over the tidal cycle are extracted from that point and are presented in Figure A- 8 and Figure A-9, respectively.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 168 of 185

Figure A-7: Model mesh with the location where the tidal state is extracted indicated with a red square.

Figure A-8: Predicted water elevation over a tidal cycle, with low water indicated with the blue arrow, and high water with the green arrow.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 169 of 185

Figure A-9: Predicted velocity at the node - over tidal cycle, including the spring (green arrow) and neap (blue arrow) tide.

Table A-19 summarises the key tidal states assessed.

Table A-19: Time of High and Low water tidal state, and peak velocity over spring and neap tide. Tidal state Date in the simulation High water 2019/01/24 09:00:00 Low water 2019/01/16 01:30:00 Spring Peak velocity 2019/01/23 15:00:00 Neap peak velocity 2019/01/15 20:30:00

A5.6 Water quality results

Coastal/Nearshore discharges The validated model has been used to reproduce the discharge of E. coli indicator bacteria from a total of 6 wastewater sewage treatment works. Details of these are given in Table A-20.

Table A-20: Details of discharges included in the validated model. Wastewater Sewage DWF Lat Lon E. coli N Treatment Works (m3/s) (degrees) (degrees) (cfu /100 ml) Swanage 4,910 50.608 -1.941 398 49 Weymouth & Portland 32,141 50.568 -2.480 330,000 864 Bridport 8,050 50.699 -2.772 330,000 864 Charmouth 1,270 50.722 -2.897 330,000 864 Lyme Regis 3,022 50.720 -2.929 830 27 6,331 50.675 -3.230 237 66

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 170 of 185

Constant decay scenarios From earlier study we know that the decay rate over daytime is 8.56 hours and 30.40 hours for night for the E. coli T90. As a first estimate, an average value assuming 12 hours of both light and dark give us an average T90 of 19.48 hours.

The model prediction for the four different tidal states are presented in Figure A- 10, with plumes of E. coli concentration for the following four large discharges: Lyme Regis, Charmouth, Bridport, Weymouth & Portland. Even if the Lyme Regis discharge has very little impact it is presented with the later three discharges as the plume from Bridport discharge spreads off Lyme Regis (Figure A-11).

Predicted plume of E. coli concentration for the discharges at Sidmouth and Swanage are presented in Figure A-12 with a different scale, the discharges being less concentrated and having a smaller volume than the four central discharges.

The plume concentration at Sidmouth, Swanage and Charmouth discharges never cross the threshold of 5 cfu/100 ml over the simulated tidal cycle; for Bridport and Weymouth & Portland discharges the plume is more concentrated overall due to both larger discharge volumes and more bacterial concentration. The threshold of 5 cfu/100 ml is reached for Low water condition at Bridport; and spring peak velocity condition at Weymouth & Portland; under Neap peak velocity conditions the two later plume locations are over 5 cfu/100 ml.

Time Tidal E. coli prediction (cfu/100 ml) state 2019/01/24 High 09:00:00 water

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 171 of 185

2019/01/16 Low 01:30:00 water

2019/01/23 Spring 15:00:00 Peak velocity

2019/01/15 Neap 20:30:00 peak velocity

Figure A-10: Prediction of E. coli plume for several state of tide, assuming a constant decay rate of 19.48 hours. The discharges locations are indicated with black octagon.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 172 of 185

Time Tidal E. coli prediction (cfu/100 ml) state Sidmouth Swanage

2019/01/24 High 09:00:00 water

2019/01/16 Low 01:30:00 water

2019/01/23 Spring 15:00:00 Peak velocity

2019/01/15 Neap 20:30:00 peak velocity

Figure A-11: Prediction of E. coli spreading for four tidal states. The locations of the discharge are indicated with a black dot with the associated name above for Lyme Regis, Charmouth, Bridport and Weymouth & Portland.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 173 of 185

In the maximum extent of the 5 cfu/100 ml isoline over the simulated tidal cycle is presented in black. For each discharge, the plume extent is centred over the discharge location. The largest spread of the E. coli plume is found with the largest discharge at Weymouth & Portland and conversely for the smallest discharge at Charmouth, even if their concentration in bacteria is the same. The footprint of the plume over the spring and neap peak velocities are comparable, with the former slightly spreading more westward and the latter eastward. Low water conditions are always more critical than high water regarding the spreading of the plume, as it limits the diffusion of the bacteria.

The maximum extent over a tidal cycle is showed, as are the following tidal state: Spring Peak Flood, Neap peak food, Low water and High water. The background colour (BOTTOM) indicates the depth. Figure A-12: Extent of the E. coli 5 cfu/100 ml isoline off Charmouth, Bridport and Weymouth & Portland.

Hydrodynamic modelling conclusion

Near Bridport and Weymouth & Portland discharges the plume can induce localised water quality with a concentration in the bacteria E. coli larger than 5 cfu/100 ml. The presence of such concentrations is mainly reached over low water conditions. On the flood tide as the plume from Weymouth & Portland discharge reaches the tip of the Portland Island it spreads further east with the strong tidal flow.

Hydrodynamic modelling references

Mar. Environ. Res. Ltd. URL: https://vision.abpmer.net/renewables/

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 174 of 185

Astrium (2011). Creation of a high resolution Digital Elevation Model (DEM) of the British Isles continental shelf: Final Report. Prepared for Defra, Contract Reference: 13820. 26pp. MontyMariner (n.d.). Admiralty Derived Tide Stream Charts. URL: http://montym ariner.co.uk/tide-stream-charts/tide-stream-charts-2/.

New Forest District Council (2017). Update of Carter, D., Bray, M., and Hooke, J. (2004). SCOPAC Sediment Transport Study. URL: https://www.scopac.org.uk/sts/

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 175 of 185

Appendix 6: Aquaculture Potential Maps

A6.1 Overview

This section provides maps for the derived aquaculture potential across the FLAG region. These maps were derived by using environmental data, and the threshold levels outlined in Appendix 3, to determine the growth potential (suitability) of each species according to the specific method of aquaculture. Current constraints to aquaculture were assessed (see Sections 5 and 7) according to the impact the aquaculture system would have and areas with conflict between existing exploitation and modelled growth potential were resolved by excluding the area from the aquaculture potential data layer. The resulting layer is the aquaculture potential layer, and this section documents each layer for each species and possible aquaculture system.

Generation of species-specific suitability and aquaculture potential raster data was performed in the software package R (version 3.5.1) using the IDE RStudio (version 1.1.383). The packages raster (Hijmans, R.J., 2019; version 2.8-19), rgdal (Bivand, R., Keitt, J. and Rowlingson, B., 2019; version 1.4-3) and rgeos (Bivand, R. and Rundel, C., 2019; version 0.4-3) were used to mask, reclassify and display the raster data. Generation of raster layers to exclude areas from the aquaculture potential rasters were generated in ArcGIS 10.5 and all map outputs were generated in ArcMap 10.5.

Guidance on Interpretation of aquapotential maps

In the following maps the biological and environmental aquaculture potential of each species using appropriate culture systems is shown as a raster layer. This aquaculture potential ranges from 0 (unsuitable) to 1 (optimal) and the interpretation of these values is shown in the table below.

Aquapotential value 0.0 0.5 1.0 Colour on map Aquaculture Aquaculture Aquaculture Interpretation suitable but may unsuitable optimal not be optimal A score of zero indicates that one or more environmental variable is unsuitable for the growth or survival of that species at that location - This is shown as a white/grey area on the map

A score of 1 on the map indicates that all environmental variables considered in the model are at optimal values at that location - This is shown as a green area on the map

A score of 0.5 on the map indicates that 50% of the environmental variables are optimal and 50% are suboptimal at that location - This is shown as a yellow area on the map

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 176 of 185

A6.2 Aquaculture Potential: Bottom-secured seaweed

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 177 of 185

A6.3 Aquaculture Potential: Suspended seaweed

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 178 of 185

A6.4 Aquaculture Potential: Open pen finfish

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 179 of 185

A6.5 Aquaculture Potential: Bottom (ranching) and Suspended lobster

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 180 of 185

A6.6 Aquaculture Potential: Bottom culture of bivalves

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 181 of 185

A6.7 Aquaculture Potential: Bottom-secured culture of bivalves

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 182 of 185

A6.8 Aquaculture Potential: Suspended culture of bivalves

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 183 of 185

Appendix 7: Data layers

A catalogue of data layers used in the analysis, including derived suitability layers, is provided in the attached document:

AquacultureMappingProject_GIS_Catalogue.pdf

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 184 of 185

Appendix 8: Interactive mapping tool

A layered pdf comprising a map of the Dorset and East Devon FLAG area, showing all GIS layers used in the analysis and mapping process, including derived species/system specific suitability layers, is provided separately in the file:

Dorset&EDevon_FLAG_layered_aquaculture_suitability_areas_Final_r.pdf

This pdf document allows viewers to switch on/off individual map layers to examine the spatial overlap of different map features and areas. The document can be viewed, with layer functionality, using standard Adobe Acrobat Reader software, using the following procedure:

1. If, on opening the pdf, the navigation panel is not visible to the left of the document, go to the top menu bar and choose

View > Show/Hide > Navigation Panes > Layers

This will open the Layers pane, where an expandable folder named ‘Basemap’ should be visible. Expand this folder.

2. By default, Adobe Reader opens with all layers switched on (giving the map a very cluttered appearance). Use the ‘eye’ symbol beside each layer to toggle various layers on and off, according to overlay/visibility requirements.

3. Layers representing potential suitability for individual species reflect areas for which all suitability criteria score at least 50%.

Mapping of Areas of Potential Aquaculture within the Dorset and East Devon FLAG area

Page 185 of 185

World Class Science for the Marine and Freshwater Environment

We are the government’s marine and freshwater science experts. We help keep our seas, oceans and rivers healthy and productive and our seafood safe and sustainable by providing data and advice to the UK Government and our overseas partners. We are passionate about what we do because our work helps tackle the serious global problems of climate change, marine litter, over-fishing and pollution in support of the UK’s commitments to a better future (for example the UN Sustainable Development Goals and Defra’s 25-year Environment Plan). We work in partnership with our colleagues in Defra and across UK government, and with international governments, business, maritime and fishing industry, non-governmental organisations, research institutes, universities, civil society, and schools to collate and share knowledge. Together we can understand and value our seas to secure a sustainable blue future for us all and help create a greater place for living.

© Crown copyright 2021

______Pakefield Road, Lowestoft, Suffolk, NR33 0HT

The Nothe, Barrack Road, Weymouth DT4 8UB

www.cefas.co.uk | +44 (0) 1502 562244