You are on page 1of 140

UNIVERSITY OF EXETER

Scoping study for the


installation of floating
tidal stream in the
Pentland Firth
William English, Isy Hammond, Jonty Haynes, Barnaby King, Sharath
Kumar, Rob Spice

CSMM404
st
21 March 2018

0
Abstract
The following document provides a comprehensive assessment of the feasibility of a tidal stream
device in the Pentland Firth region, suggesting a development that contributes to current UK targets.
The document concludes by stating that unless a CFD strike price or similar financial system be
implemented in the near future, it would be unwise to advance this project in the current economic
climate.

Analysis of current commercially viable tidal range devices resulted in the selection of
Scotrenewables’ SR2000, 2MW floating turbine. Standard methodologies, industry GIS tools and
existing data contribute to the selection of the following site location: 58.698, -2.966, which was
estimated to provide the desired tidal current range for limited wave climate. The available power of
the chosen site was calculated to be between 0.587 MW and 0.88MW per device which, it is noted,
is likely to underestimate the true value. Further environmental calculations were undertaken
allowing for the estimation of an extreme wave of 16m. Informed from the current and wave
calculations, a study guided by industry standards was concluded by suggesting a total mooring line
length of 370m comprised of both chain and synthetic rope, and a foundation of piles at depth 20m
would result in a safety factor of 3 for the device in an extreme wave climate. Electrical connections
and transmission were assessed via the analysis of current and projected states of local networks. It
was found that the offshore infrastructure would not exceed 20 million pounds, whilst the onshore
infrastructure could range between 7 million and 30 million pounds. A detailed investigation into
O&M strategy based on similar projects and utilising industry standard Mermaid, found that
quarterly inspections are required to limit risk and that any major replacements made between
January and March are likely to double the cost of operations.

The above studies allowed for a detailed assessment of the economics of the development via a
sensitivity and Monte Carlo analysis. A 14% chance of profit but potential loss of 30 million pounds
was estimated for the installation of a single device. An LCOE of £280 for the installation of three
devices was calculated with stipulation that this figure may reduce but is severely limited by the
material cost of the device. Thus the authors conclude that the project is unlikely to be economically
feasible without aid from a financial system.

1
Contents
Table of Tables ........................................................................................................................................ v
Table of Figures ...................................................................................................................................... vi
Table of Equations ............................................................................................................................... viii
1 Introduction .................................................................................................................................... 1
1.1 Aim and Objectives ................................................................................................................. 1
2 Site Characterisation and Selection ................................................................................................ 2
2.1 Characterisation of site ........................................................................................................... 2
2.1.1 Tidal resource .................................................................................................................. 2
2.1.2 Summary ......................................................................................................................... 7
2.1.3 Local Wave Climate ......................................................................................................... 8
2.1.4 Site selection ................................................................................................................. 18
3 Site specific assessment ................................................................................................................ 23
3.1 Resource assessment at site ................................................................................................. 23
3.1.1 Methodology ................................................................................................................. 23
3.1.2 Results ........................................................................................................................... 24
3.1.3 Discussion ...................................................................................................................... 27
3.1.4 Summary ....................................................................................................................... 29
3.2 Device choice ........................................................................................................................ 29
3.3 Mooring design ..................................................................................................................... 30
3.3.1 Design Considerations................................................................................................... 30
3.3.2 Mooring Design Methodology ...................................................................................... 31
3.3.3 OrcaFlex Numerical Simulation Overview .................................................................... 36
3.3.4 Implications for Upscaling ............................................................................................. 41
3.3.5 Risks............................................................................................................................... 42
3.3.6 Mooring Summary ........................................................................................................ 42
3.3.7 Foundation Design ........................................................................................................ 43
3.4 Electrical Connections ........................................................................................................... 46
3.4.1 Spatial Analysis .............................................................................................................. 46
3.4.2 Planned Expansion ........................................................................................................ 48
3.5 Offshore Infrastructure ......................................................................................................... 49
3.5.1 Cables ............................................................................................................................ 49

i
3.5.2 Risks............................................................................................................................... 52
3.6 Onshore................................................................................................................................. 53
3.6.1 Landing Site ................................................................................................................... 54
3.6.2 Cable Transition Pit ....................................................................................................... 54
3.6.3 Developer Plant Compound .......................................................................................... 54
3.6.4 Onshore Power Transmission Line ................................................................................ 55
3.6.5 Risk ................................................................................................................................ 56
3.7 System Diagrams ................................................................................................................... 57
3.7.1 System Diagram 2MW .................................................................................................. 57
3.7.2 System Diagram 6MW .................................................................................................. 58
3.8 Electrical Connections Implications on Cost ......................................................................... 59
3.9 Electrical Connections Summary .......................................................................................... 60
4 Installation .................................................................................................................................... 61
4.1 Design.................................................................................................................................... 61
4.2 Preparation ........................................................................................................................... 61
4.3 Method ................................................................................................................................. 61
4.4 Location ................................................................................................................................. 62
4.5 Planning................................................................................................................................. 62
4.6 Mermaid................................................................................................................................ 63
4.6.1 Challenges to be addressed .......................................................................................... 63
4.6.2 Power cable ................................................................................................................... 64
4.6.3 Pin pile installation ........................................................................................................ 64
4.6.4 Mooring operation ........................................................................................................ 64
4.6.5 Vessels ........................................................................................................................... 64
4.6.6 Installation of Turbine Task ........................................................................................... 65
4.6.7 Mermaid Results ........................................................................................................... 66
4.7 Summary ............................................................................................................................... 71
5 Operation and Maintenance (O&M)............................................................................................. 72
5.1 Maintenance Strategies ........................................................................................................ 72
5.2 SR2000 Operation and Maintenance Strategy ..................................................................... 74
5.2.1 The SR2000 Device ........................................................................................................ 74
5.2.2 SR200 Subassemblies .................................................................................................... 75
5.2.3 Reliability Block Diagram (RBD) .................................................................................... 77
5.3 Reliability Data ...................................................................................................................... 78

ii
5.4 Reliability Data Analysis ........................................................................................................ 79
5.4.1 Correlating Reliability Data ........................................................................................... 79
5.4.2 Percentage Likelihood of Failure per Category ............................................................. 80
5.4.3 Probability of failure over lifetime ................................................................................ 81
5.4.4 Mean time to repair (MTTR) ......................................................................................... 83
5.4.5 Average Cost per Failure ............................................................................................... 84
5.4.6 Average Number of Technicians Required per Repair .................................................. 86
5.5 Proposed O&M Strategy ....................................................................................................... 87
5.6 Mermaid Analysis .................................................................................................................. 88
5.6.1 O&M port location ........................................................................................................ 88
5.6.2 Major Replacements: Towing back to shore ................................................................ 88
5.6.3 Minor onsite maintenance and repairs......................................................................... 89
5.6.4 Major Onsite repairs ..................................................................................................... 89
5.6.5 Metocean data. ............................................................................................................. 89
5.6.6 Other Simulation Assumptions ..................................................................................... 89
5.6.7 Simulation Inputs .......................................................................................................... 89
5.6.8 Results ........................................................................................................................... 90
5.7 Scaling Considerations .......................................................................................................... 95
5.8 Risks ...................................................................................................................................... 95
5.9 Summary ............................................................................................................................... 96
6 Economic appraisal ....................................................................................................................... 97
6.1 Capital Expenditure ............................................................................................................... 97
6.1.1 Device Cost .................................................................................................................... 97
6.1.2 Power ............................................................................................................................ 97
6.1.3 Support.......................................................................................................................... 98
6.1.4 Installation .................................................................................................................... 98
6.1.5 Project Management .................................................................................................... 98
6.1.6 Total capital expenditure .............................................................................................. 98
6.2 Operational Expenditure ..................................................................................................... 100
6.3 Revenue .............................................................................................................................. 100
6.4 Grants and subsidies ........................................................................................................... 101
6.5 Discounted Cash Flow Model.............................................................................................. 102
6.5.1 Funding........................................................................................................................ 102
6.5.2 Project life ................................................................................................................... 102

iii
6.5.3 Tax rate ....................................................................................................................... 102
6.5.4 Inflation ....................................................................................................................... 102
6.5.5 Grants and subsidies ................................................................................................... 102
6.5.6 Discount rate ............................................................................................................... 102
6.5.7 Results ......................................................................................................................... 102
6.6 Sensitivity Analyses ............................................................................................................. 105
6.6.1 Objective Analysis ....................................................................................................... 105
6.6.2 Subjective Analysis ...................................................................................................... 106
6.6.3 Discussion .................................................................................................................... 109
6.7 Scenario Analysis ................................................................................................................. 109
6.7.1 Standard scenario analysis .......................................................................................... 110
6.7.2 Fixed price scenario analysis ....................................................................................... 110
6.7.3 Discussion .................................................................................................................... 110
6.8 Monte Carlo Analysis .......................................................................................................... 110
6.8.1 Methodology ............................................................................................................... 110
6.9 Results and discussion ........................................................................................................ 111
6.10 Levelised Cost of Energy ..................................................................................................... 112
6.11 Payback period .................................................................................................................... 112
6.12 Scaling ................................................................................................................................. 112
6.13 Conclusions ......................................................................................................................... 114
6.14 Recommendations for future work .................................................................................... 114
7 Project Conclusions ..................................................................................................................... 115
8 References .................................................................................................................................. 117
9 Appendix A .................................................................................................................................. 122
10 Appendix B .............................................................................................................................. 127

iv
Table 3-17: Offshore Electrical Connections
Table of Tables Summary and Preliminary Costing. .................. 59
Table 2-1: Characteristic of a “good” site ...........3 Table 3-18: Onshore Electrical Connections
Table 2-2: mean maximum and standard Summary and Preliminary Costing. .................. 59
deviation data for the locations shown in Figure Table 4-1 Planning Factors informed from
2-8 .......................................................................9 Section 2. .......................................................... 63
Table 2-3: Extreme wave calculation results ....12 Table 4-2 Summary of minimum and maximum
Table 2-4: input parameters for Jonswap days required for tasks (winter). ...................... 66
spectrum ...........................................................15 Table 4-3 Summary of minimum and maximum
Table 2-5: 100-year extreme sea state .............17 days required for tasks.(summer) .................... 69
Table 3-1 Summary of limitations associated Table 5-1: SR2000 Subsystem categories and
with this resource assessment ..........................24 number ............................................................. 79
Table 3-2: Summary of outcomes of resource Table 5-2: Subsystems accessible for major
assessment of FoW data ...................................26 repair onsite or offsite. ..................................... 82
Table 3-3: Comparison of available and Table 5-3: Average MTTR and failure rates per
extractable resource predicted from FoW data category ............................................................ 87
with Neil et al power density values for the area Table 5-4: Vessel operating specifications ....... 88
around the Pentland Skerries, location closest to Table 6-1: Base case discounted cash flow
the chosen site. For the determination of power model for a single device deployment. .......... 103
and energy the swept area of SR2000 is used at Table 6-2: Base case discounted cash flow
402 m2. ..............................................................28 model for a three-device deployment............ 104
Table 3-4: SolidWorks Variables for Model Table 6-3: Scenario analysis for both installation
Geometry. .........................................................35 sizes, defining the absolute minimum and
Table 3-5: Scotrenewables SR2000 - Device maximum project net present values, given the
Characteristics (Scotrenewables, 2018)............35 parameter ranges defined during the preceding
Table 3-6: Environmental conditions for subjective sensitivity analysis. ........................ 110
OrcaFlex. ...........................................................36 Table 6-4: Scenario analysis for both installation
Table 3-7: Input Variables for OrcaFlex mooring sizes, defining the minimum and maximum
design simulation. .............................................37 project net present values in the scenario of a
Table 3-8: Preliminary Mooring Design Summary fixed £150/MWh CfD, given the parameter
..........................................................................40 ranges defined during the preceding subjective
Table 3-9: Mooring Design Associated Risks.....42 sensitivity analysis. ......................................... 110
Table 3-10 Suitability of different foundation Table 9-1: Failure rate sub-assembly breakdown
types ..................................................................44 adapted from Delorm (2014).......................... 122
Table 3-11: Landing sites A and B for Offshore Table 9-2: Failure rates split into categories:
Infrastructure. ...................................................48 major replacement, major repair and minor
Table 3-12: Offshore cable parameters for repair. ............................................................. 123
landing site A and B...........................................50 Table 9-3: Failure rate over lifetime (20 years)
Table 3-13: Available Subsea Connections. ......51 ........................................................................ 124
Table 3-14: Risk Assessment for Offshore Table 9-4: MTTR.............................................. 124
Electrical Infrastructures ...................................52 Table 9-5: MTTR over 20-year lifetime ........... 125
Table 3-15: Onshore MVAC and HVDC cable Table 9-6: MTTR over 20-year lifetime ........... 125
parameters for reinforced and un-reinforced Table 9-7: Average material cost over lifetime
local networks. ..................................................55 ........................................................................ 126
Table 3-16: Onshore System Risks ....................56 Table 10-1 Winter cost data (installation) ...... 127

v
Table 10-2 Summer cost data (installation) ....131 Figure 2-15: Rayleigh Distribution of Extreme
wave heights for the 15-year extreme sea state.
.......................................................................... 13
Figure 2-16: Rayleigh Distribution of Extreme
Table of Figures wave heights for the 50 year extreme sea state
Figure 2-1: Efficiency curve of generic tidal .......................................................................... 14
energy converter (Cornett, 2006) .......................2 Figure 2-17: Rayleigh Distribution of Extreme
Figure 2-2: Site of energy converter rows in wave heights for the 100 year extreme sea state
Adcock et als’ study (Adcock et al., 2013)...........3 .......................................................................... 14
Figure 2-3: Variation in extracted power at Figure 2-18: Johnswap Spectrum for all sites
location investigated in Draper et al (Draper et locations and extreme wave period ................. 16
al., 2014). ............................................................4 Figure 2-20: Map of local substations (Scottish
Figure 2-4: Normalised mean current speed and Southern energy networks, 2018). ............ 19
estimation from Easton, Woolf, and Bowyer Figure 2-19: Mean significant wave high in all
(Easton, Woolf and Bowyer, 2012) .....................5 peak tidal modes. ............................................. 19
Figure 2-5: Extract from UK Renewable Atlas Figure 2-21: Map showing the areas which is not
showing the Pentland Firth and Fall of Warness 6 suitable for installation of SR200, and areas
Figure 2-6: Output model for the Pentland Firth which are suitable pending analysis of the soft
and Orkney measuring mean power ..................6 constraints ........................................................ 20
Figure 2-7: Mean energy dissipation due to Figure 2-22: map of the hard constraints ......... 20
friction, located at the Pentland Firth (Draper et Figure 2-23: mean power density of the tidal
al., 2014) .............................................................7 resource shown alongside the unavailable area
Figure 2-8: site map of the data collection points from Figure 2-21 ............................................... 21
............................................................................9 Figure 2-24: The frequency of shipping through
Figure 2-9: M2 tidal flow velocity, left. the Pentland Firth, shown alongside the
Directional analysis of the 7 year swan unavailable area from Figure 2-21 ................... 21
simulation, right. .................................................9 Figure 2-25: site selection map, showing all
Figure 2-10: Statistical extreme wave analysis for relevant constraints .......................................... 22
Billia Croo by (Lawrence, Kofoed-Hansen and Figure 2-26: Site map showing the constraints
Chevalier, 2013- ‘peaks over threshold method’ alongside the current energy density ............... 22
..........................................................................11 Figure 3-1: Results of harmonic analysis, top line
Figure 2-11: Validation of Rayleigh Distribution from left to right shows velocity components,
for with M0 shape factor for Orkney Island Area. resultant velocity and tidal ellipse for raw data,
..........................................................................11 bottom row show the same but for predicted
Figure 2-12: Statistical comparison of the data for one year. ............................................. 25
measured and modelled wave data used by Figure 3-2: Velocity rose for raw data (left) and
(Lawrence, Kofoed-Hansen and Chevalier, 2013) predicted data (right) for the FoW ................... 25
for the extreme wave analysis. .........................11 Figure 3-3: Annual power availability from
Figure 2-13: maximum wave height and predicted FoW data series, top showing power
significant wave height for 20 year Mike 21 density variation throughout the year, middle
model ................................................................11 showing this variation as a cumulative
Figure 2-14: Rayleigh Distribution of Extreme probability curve and bottom showing the
wave heights for the 6-year extreme sea state. power density as a histogram........................... 26
..........................................................................13

vi
Figure 3-4: Shear profile in velocity and power Figure 3-24: Multiple Devices (6MW) Electrical
density for predicted FoW data series, and System Overview .............................................. 58
comparison with Pentland Firth ADCP data .....27 Figure 4-1 Location of the port, site and
Figure 3-5 Wave force regimes (Charkrabarti, substation (orange represents site, Green the
1987 cited by, (DNV C205, 2010)). D = port and blue the substation), while this does
characteristic dimension, H = Wave height, Y = not represent the chosen site, the timings would
Wave length. .....................................................33 not change significantly due to the proximity. . 62
Figure 3-6: SolidWorks model providing Figure 4-2 Multipurpose Vessel (damen.com,
geometry for OrcaFlex. .....................................34 (2018)................................................................ 64
Figure 3-7: Final mooring design, shown at 10s Figure 4-3 Various stages of SR2000 installation
point of the 100year wave. ...............................37 (YouTube, 2018). .............................................. 65
Figure 3-8: Example mooring line diameter Figure 4-4 Tasks 1-3 Mermaid installation time
optimisation. .....................................................38 (winter). ............................................................ 67
Figure 3-9: Normalised Tension at point 1 on the Figure 4-5 Duration Exceedance for winter
line A over 50s of the 100 year design wave. ...38 installation (winter). ......................................... 67
Figure 3-10: Normalised Tension in Line A for the Figure 4-6 Cost exceedance chart for winter
100 year wave sea state....................................39 installation show a fair probability of increase in
Figure 3-11: Render of Final Preliminary Mooring the cost (Appendix B data). .............................. 68
System. ..............................................................40 Figure 4-7 Progress burn up chart for winter
Figure 3-12: Current SR2000 Mooring Design installation. ....................................................... 68
(Hamilton, 2018). ..............................................40 Figure 4-8 Cost duration chart for summer
Figure 3-13: Device Spacing for 64MW Array. ..41 installation(July)................................................ 69
Figure 3-14: Bathymetry of the Pentland Firth Figure 4-9 Duration exceedance chart for
Region ...............................................................43 summer installation(July) . ............................... 70
Figure 3-15: sea bed in the east Pentland Firth Figure 4-10 Cost exceedance chart for summer
(Marine.gov.scot, 2018) ....................................43 installation(July)................................................ 70
Figure 3-16: (Starling, Scott and Parkinson, 2009) Figure 4-11 Progress burn up chat for summer
..........................................................................44 installation(July)................................................ 71
Figure 3-17: ultimate capacity od Pin Piles as Figure 5-1: illustration of balance between cost
anchors ..............................................................45 and lost revenue (The Crown Estate, 2013) ..... 72
Figure 3-18: Current Transmission Network for Figure 5-2: Reliability Block Diagram for SR2000
North West Scotland (National Grid, 2015). .....46 (Delorm, 2014).................................................. 77
Figure 3-19: Existing Local Distribution Network Figure 5-3: Failure rates ‘Bathtub Curve’ (Rinaldi,
for North East Scotland, (Parsons Brinckerhoff, 2017). ................................................................ 78
2012). ................................................................47 Figure 5-4: Failure rate correlation between
Figure 3-20: Spatial Analysis of Site Location and Delorm (2014) and Carrol et al. (2015)............. 80
Available Network Connections. .......................48 Figure 5-5: Percentage likelihood of failure per
Figure 3-21: Methodology for Cable Route category ............................................................ 81
Selection as suggested by (DNV, 2016). This Figure 5-6: Probability of failure over 20-year
preliminary stage is highlighted by the red box. lifetime.............................................................. 82
..........................................................................49 Figure 5-7: Failure rate and MTTR for minor
Figure 3-22: Typical Requirement for Onshore repairs ............................................................... 83
Infrastructure. ...................................................53 Figure 5-8: Failure rate and MTTR for major
Figure 3-23: Single Device (2MW) Electrical repairs ............................................................... 84
System Overview ..............................................57 Figure 5-9: Average material cost of repairs .... 85

vii
Figure 5-10: Average material cost of repair over Figure 6-9: Monte Carlo analysis, resulting from
device lifetime ...................................................86 3000 simulations, illustrating a range of
Figure 5-11: Duration variation – major probable outcomes for a three-device scenario.
replacement/ major repair offsite ....................91 The 86th percentile is indicated as the percentile
Figure 5-12: Cost variation from Mermaid – in which the NPV becomes positive. .............. 111
major replacement/ major repair offsite ..........91 Figure 6-10: Levelised costs of energy at scales
Figure 5-13: Duration variation from Mermaid – between 2 and 20 MW (1 to 10 devices),
major repair onsite ...........................................92 indicating a steady decrease in LCOE at
Figure 5-14: Cost variation from Mermaid – increasing scale. .............................................. 113
major repair onsite ...........................................93
Figure 5-15: Duration variation from Mermaid –
minor repair ......................................................94
Figure 5-16: Cost variation from Mermaid – Table of Equations
minor repair ......................................................94 Equation 2-1 ..................................................... 10
Figure 6-1: Pie charts displaying the proportion Equation 2-2 ..................................................... 10
of CAPEX for each capital expenditure, Equation 2-3: calculation of Omega_p ............. 15
illustrating the difference between the two Equation 2-4 calculation of Omega .................. 15
possible scales of installation............................99 Equation: 2-5 estimation of lambda ................. 15
Figure 6-2: Objective sensitivity analysis for a Equation 2-6: Jonswap spectrum formula ........ 16
three-device installation, with steeper gradients Equation 3-1 Power density equation used in
indicating a greater impact on NPV. ...............105 resource assessment ........................................ 23
Figure 6-3: Objective sensitivity analysis for a Equation 3-2 Velocity profile equation based on
single device installation, with steeper gradients the power law, extracted from Lewis et al (Lewis
indicating a greater impact on NPV. ...............105 et al., 2017) ....................................................... 23
Figure 6-4: Graph illustrating the UK corporation Equation 3-3: Reynolds number ....................... 31
tax rate from 1981 to 2018. The red area Equation 3-4: Keulegan-Carpenter number ..... 32
indicates the chosen potential range of Equation 3-5: Stokes Parameter ....................... 32
corporate tax rate for the subsequent modelling. Equation 3-6: Wake Amplification Factor......... 32
Modified from tradingeconomics.com ...........107 Equation: 3-7 .................................................... 32
Figure 6-5: Graph illustrating the UK corporation Equation 3-8: Dimension Correction for Marine
tax rate from 1981 to 2018. The red area Growth .............................................................. 32
indicates the chosen potential range of Equation 3-9: Drag Coefficient ......................... 33
corporate tax rate for the subsequent modelling. Equation 3-10 Thrust Force .............................. 36
Modified from tradingeconomics.com ...........108
Figure 6-6: Subjective sensitivity analysis for a
single device installation, with steeper gradients
indicating a greater impact on NPV. ...............108
Figure 6-7: Subjective sensitivity analysis for a
three-device installation, with steeper gradients
indicating a greater impact on NPV. ...............109
Figure 6-8: Monte Carlo analysis, resulting from
3000 simulations, illustrating a range of
probable outcomes for a single device scenario.
........................................................................111

viii
1 Introduction
The UK marine energy resource is often quoted as having the potential to provide 20% of the
country’s energy demand (BEIS, 2013). Under current demand levels this would equate to
approximately 60 TWh generated (BEIS, 2016). The sector has been identified by areas of the UK
Government (BEIS, 2013), European Union, and local authorities as a sector of significant potential in
developing local supply chain and economy (Marine Hub Cornwall, no date; Ocean Energy Forum,
2016).

While national support has reduced with the removal of ring fenced CFD last year (Regen, 2016)
recently developments on the local level in areas such as Cornwall and Scotland, with the
development of industry bodies such as Wave Energy Scotland and Marine Hub (Marine Hub
Cornwall, no date; Wave Energy Scotland, no date) shows continued support. Globally the marine
sector has been identified as having a global potential £570bn between 2010 and 2050 (Ocean
Energy Forum, 2016). To this extent the European Union has set ambitious targets for installed
marine energy of 850 MW by 2021(Ocean Energy Forum, 2016), of this 32 MW are expected to be
taken up my tidal-stream over the next few years from projects currently under development
(Ocean Energy Forum, 2016). Out of the 850 MW suggested by the European union, 100MW of tidal
stream is expected to be installed by the mid 2020’s (Ocean Energy Forum, 2016), with further
deployment of ten 20-30 MW arrays by 2030 (Ocean Energy Forum, 2016). While these targets are
ambitious development towards these goals has taken place through the development of
MeyGen(Ocean Energy Forum, 2016). The project proposed in this study builds upon the starting
point of Meygen as a demonstration/Pre-commercial part of Action plan 2 of the Ocean Energy
Forum’s roadmap, with potential deployment during the 2018-2025 action plan period (Ocean
Energy Forum, 2016).

1.1 Aim and Objectives


This aim of this project is to assess the high level feasibility for a commercial floating tidal stream
device in the Pentland Firth. Scope for the expansion of the sites to accommodate an array is also
included. The technical challenges of site selection, installation, device mooring, grid connectivity,
and operations and management all feed into an assessment of economic viability. As a team of
renewable energy consultants, scenarios encased within the installation of a single device, up to an
array of devices are considered and analysed via the use of industry standard software and
published data. In order to provide a comprehensive assessment of the feasibility, several tasks were
allocated and assigned to members of the consultancy for detailed analysis. The tasks, outlined
below, were then split into the sub-groups that form the contents of this report.

1 Investigation into the environmental information of the site to inform engineering design.
2 Engineering and design of systems including, mooring and electrical transmission, informed
from task 1.
3 Installation, planning and the assessment of risks.
4 Operations and maintenance strategies.
5 Economic viability, informed from tasks 1 to 4.

1
2 Site Characterisation and Selection
2.1 Characterisation of site
2.1.1 Tidal resource

2.1.1.1 Methodology
The identification of high yield resource locations builds upon the work of literature, using previous
resource assessments of the Pentland to determine locations of “good” resource. The assessment
considered two main constraints at this stage; consideration of tidal current speeds and energy
dissipation at a location. For the former, the generic tidal current power curve presented by Cornet
(Figure 2-1) is used in replacement of a power curve for the SR2000 device. This helps identify a
viable current velocity range for a device, with current speeds outside of this range are considered
not-viable and omitted as potential sites. To this extent, a viable tidal range can be considered as 0.5
ms-1 to 3.5 ms-1 with values above this considered too energetic for a device and below as too
lethargic for power generation.

Figure 2-1: Efficiency curve of generic tidal energy converter (Cornett, 2006)

Energy dissipation occurs due to the interactions of the tidal current with the seabed, with the
resulting friction the main mechanism for energy dissipation (Easton, Woolf and Bowyer, 2012).
Greater seabed-current interactions are associated with sharper velocity shear profiles at locations
as well as complex flow regimes undesirable for power extraction. To this extent where possible
these locations are avoided. Table 2-1 summarises what the study considered a “good” resource
location in this study.

2
Table 2-1: Characteristic of a “good” site

Constraint Desired Characteristic


Tidal current speed Mean current speed within the range
of 0.5 ms-1 to 3.5 ms-1
Energy dissipation Low dissipation

The site selection process aggregates the results of past studies to suggest potential sites for power
extraction. However, the approach is limited to a high order assessment of the resource available.
This approach fails to consider the influence of tidal ellipses in the site selection as well as the impact
of maximum tidal current speed, being overly reliant on mean values instead. While this impact is
not carried forward into later energy yield calculations for the chosen site, they apply for the high
order site selection process.

2.1.1.2 Literature review

2.1.1.2.1 Tidal currents


Adcock et al considered the potential energy resource of the Pentland, modelling the available
power as a result of M2 and S2 constituent tidal components. Adcock considers energy conversion
devices at multiple locations with multiple row arrays, these locations are shown in Figure 2-2: Site
of energy converter rows in Adcock et als’ study (Adcock et al., 2013) with up to 5 rows of devices.
Single row case see the deployment of devices at site B1,C1,D1; with two row and three-row arrays
seeing deployment at B1,B2 C1,C2,D1,D2 and B1,B2,B3,C1,C2,C3,D1,D2,D3 respectively. Any further
rows are deployed at location A1, for four rows, and A2 for five rows (Adcock et al., 2013).

Figure 2-2: Site of energy converter rows in Adcock et als’ study (Adcock et al., 2013)

Adcock suggests that over a range of blockage ratios of 0.1 to 0.4, power available for extraction
ranges from 0.17 GW to 2.16 GW (Adcock et al., 2013). Further to this, Adcock et al conclude

3
suggesting a potential extractable power resource of 1.9 GW across the stretch of water (Adcock et
al., 2013). However, Adcock et al also suggest that generation devices would have to account for a
large variation in power over the spring-neap cycle and changes in flow rate up to 30% (Adcock et
al., 2013).

Adcock highlights the variation present in previous resource estimations, highlighting the work
Garret and Cummins in showing the flaws of previous methodologies (Adcock et al., 2013). To this
extent, Adcock builds upon the work of Vennel in using simple tidal constituents (Adcock et al.,
2013). These sentiments are repeated by Draper (contemporary of Adcock) who cite the work done
by Garret and Cummins, Salter, and MacKay in disproving previous estimations (Draper et al., 2014).
Due to this, this study considered multiple resource assessment from literature to draw a conclusion
in site selection, attempting to avoid reliance on a single source and the inaccuracies present.

Draper et al develops upon the work of Adcock investigating each sub-channel identified in Adcock.
Draper et al concludes that section A holds the greatest potential, at 3.7 GW available power (Figure
2-3). Draper et al continues to suggest that these values are strongly dependent on development in
neighbouring channels. They suggest that sites in series would experience a decline in available
power as more devices are deployed, while devices developed in parallel would experience slight
increases in each site’s individual power availability. This is shown in Figure 2-3 where site E
experiences appreciable decline in power availability once westward devices are added, meanwhile
sites such as C see an increase in available power as sites are introduced in parallel.

Figure 2-3: Variation in extracted power at location investigated in Draper et al (Draper et al., 2014).

4
Easton, Woolf, and Bowyer presents the results of a Navier-Stokes numerical model of the Pentland
Firth, from which it is possible to locate high potential yields locations from their normalised mean
current speed predictions (Figure 2-4). From these predictions it can be said that areas of high
current speed, 11 times that of the mean current velocity, can be found off the southern and
northern tip of Stroma and Swona; but also South East for the tip of South Ronaldsay. From Easton,
Woolf and Bowyer it can also be said that the high current flow between the island of Swona and
Stroma persist eastward and that the Inner Sound between Stroma and the mainland also sees a
slightly high current speed than its surroundings. From this areas of high current speeds can be
identified, with sites located in the western area of the Pentland typically experiencing higher tidal
currents. However, what cannot be identified is if these values exceed the upper limit and lower
limits suggested by Cornet.

Figure 2-4: Normalised mean current speed estimation from Easton, Woolf, and Bowyer (Easton, Woolf and Bowyer, 2012)

Neil et al presents the results of a reassessment of tidal resources around Scotland, using a model
developed by Goward-Brown (Figure 2-6). This confirms the suggestions from Easton, Woolf and
Bowyer of energetic flow north and south of Stroma and the persistence of this high velocities north-
west of Stroma. In contrast to Easton, Woolf and Bowyer however, there is a suggestion of energetic
flows east of Stroma and Swona providing a power density of c. 12 kWm-2 potentially suitable for
power extraction(Neill et al., 2017).

Also presented in Figure 2-5 is power density at the Fall of Warness (FoW), of around 8-10 kWm-2.
Due to the availability of long-term data for the FoW it is assumed that the current regime within the
Pentland is similar to that for the FoW. As such the FoW data is assumed representable for a chosen
within the Pentland. While a crude assumption its validation is attempted by comparing the max
current velocity at the FoW and peek current speed presented in the UK Renewable Atlas (Figure D)
(FoW: 3.6 ms-1 UK Atlas 2.5-3ms-1). Areas with similarities in current velocity are assumed to
representable by the FoW data, with an assumed power density of 8-10 kWm-2. Due to the
similarities between the current speeds and the current speed suggested by Cornet, areas which
meet the above criteria are considered suitable for power extraction.

5
Figure 2-5: Extract from UK Renewable Atlas showing the Pentland Firth and Fall of Warness

To this extent areas West and East of the high energy flow between Stroma and Swona would be
suitable for development, with areas beyond the Pentland Skerries (Westward) and Hoy (Eastward)
unsuitable due to the reduced current speed. However for sites located between the Pentland
Skerries and Stronsay there is a risk that future development in the Pentland may reduce the energy
yield from devices (Draper et al., 2014).

Figure 2-6: Output model for the Pentland Firth and Orkney measuring mean power

6
2.1.1.2.2 Energy Dissipation
From Easton’s, Woolf’s, and Bowyer’s study an estimation of the energy dissipation via friction can
be determined (Figure 2-7). This shows large amounts of energy dissipation off the immediate tips of
Stroma and Swona, in the region of 100 Wm-2. There is also energy dissipation between the islands,
in the region of 60 Wm-2 suggesting developments may aim to avoid these areas.

Figure 2-7: Mean energy dissipation due to friction, located at the Pentland Firth (Draper et al., 2014)

2.1.2 Summary
The suggested locations appropriate for the deployment of a tidal stream device (as presented by
Connett) is situated around the Pentland Skerries, or south of Hoy. In addition to this there is
potential for development within the inner sound. Locations between the island of Stroma and
Swona should be avoided due to the high energy flows that exceed generic energy converter cut out
speed and the high energy dissipation that occurs at the tips of the islands. Additionally, locations
eastward of the Pentland Skerries or west of Hoy are inappropriate due to their lethargic flows.

7
2.1.3 Local Wave Climate
The Pentland Firth has the best tidal stream resource in the UK and is near the European Marine
Energy Centre (EMEC), therefor the oceanographic data is reported extensively in literature, and
such wave data has been used to support this study.

(Lewis et al., 2014) simulated the wave climate at 3 sites in the Pentland Firth with SWAN wave
model simulation utilizing 7 years of data from the northwest European shelf seas. According to this
report, the extreme wave climate through the Pentland Firth is more extreme towards the west and
reduces in both T(s) and Hm0 towards the east, while the mean wave climate of the lowest magnitude
at site 2, Table 2-2, and Figure 2-8. The maintenance window, which is defined as the percentage of
time the sea state has a Hs below 2 (Lewis et al., 2014), is also shown in Table 2-2. It is clear that the
east of the Pentland Firth (site 3) provides a higher percentage of maintenance availability windows;
although only 1 percent separates sites 2 and 3 Figure 2-8. The limited availability windows in site
one would be a concern to developer, and significant benefits would need to be present in other
areas to make this site suitable for a tidal installation.

Figure 2-9 shows the directional spread of the same swan simulation, alongside a 30-day simulation
of the M2 component of the tidal current. Over 90% of the sea states occur within +-40% of the tidal
direction, however a further study in the same journal using data from a Wave rider buoy taken
between 16th January and 17th July 2012, showed that only 10% of actual wave events occurred
oblique to the current direction. The effects of these oblique waves on the tidal current resource
must be further analysed through field measurements to accurately predict the tidal resource over
the life span of the project and to determine the loading requirements for the mooring structure

Besides the implications for Annual Energy Production, the main implications of the wave climate on
a tidal stream turbine are; the survivability during an extreme sea state, the fatigue loading which
the turbine and mooring components will experience, and the installation and maintenance
availability. To calculate the extreme wave, preliminary statistical analysis from (Lawrence, Kofoed-
Hansen and Chevalier, 2013) of the extreme sea state at Billia Croo is used. The methodology of this
work constructed a dynamic wave model using Mike 21, this was validated against hindcast data
Figure 2-12, which showed a scatter index of 0.15. This model was then used to generate a statistical
extreme sea state prediction for a hundred-year return period using the ‘peaks over threshold
method’ as set out by (Mathiesen et al., 1994) Figure 2-10.

The maximum wave height observed over a 20 year period at the northern section of the Billia Croo
EMEC test site is ≈21m according to the 20 year Mike 21 study by (Lawrence, Kofoed-Hansen and
Chevalier, 2013) Figure 2-13. However, field data observed by EMECs north wave rider buoy,
recorded the highest wave recording at Bilia Croo north at 17 m (Emec.org.uk, 2018) over a period of
17 years. All the above information has been used to calculate the extrema wave at sites one, two,
and three, Figure 2-8.

8
Figure 2-8: site map of the data collection points

Table 2-2: mean maximum and standard deviation data for the locations shown in Figure 2-8

Mean sea state Maximum Sea state Maintenance


Window (%)
site Hs T(s) Hs T(s) m0 (calculated from1)

1 1.38 2.7 9.59 8.7 5.748006 79

2 1.13 2.4 7.08 7.4 3.1329 87

3 1.15 2.5 6.53 6.4 2.6569 88

Figure 2-9: M2 tidal flow velocity, left. Directional analysis of the 7 year swan simulation, right.

9
The Billia Croo test site data has been used to estimate the 15, 50 and 100, year maximum sea state
for the test sites 1-3. Firstly the 6 year extreme Hs at Billia Croo has been read From Figure 2-10
(Lawrence, Kofoed-Hansen and Chevalier, 2013) taking the centre line value. This value was
compared to the extreme Hs values as simulated by (Lewis et al., 2014) for the three sites of interest,
to calculate a linear correction factor which relates the extreme sea state at Billia Croo to each of the
other sites. The resultant extremes for each site can be found in Table 2-2.

If a time series of wave height data was available, a three parameter Weibull distribution could have
been fitted to this data, and the resultant parameters carried forward to the modelled extreme
states to calculate the maximum singular wave height. However due to the lack of such data, a
Raleigh Distribution with a shape factor of m0 has been used to develop a wave height distribution
of the maximum sea states, this can provide a good estimation for most sea states (Gōda, 2010).
Furthermore, this was validated in the Orkney island area by comparing the measured distribution
from Figure 2-12 with a Rayleigh distribution of the same sea state Figure 2-11, this comparison
showed a good correlation. m0 can be approximated by Equation 2-1, (Cahill and Lewis, 2013) while
the Rayleigh distribution is shown in Equation 2-2

Equation 2-1

𝐻𝑚0 2
𝑓2 − 1) 𝑚0 = [ ]
4
Equation 2-2

−𝐻2
𝐻 [
8𝑚0
]
𝑝(𝐻) = 𝐸
4𝑚0

The resulting Raleigh distributions are shown in Figure 2-14, Figure 2-15, Figure 2-16, and Figure
2-17. The probability density which represents the largest wave at each site was estimated by
correlation with real data. Two data types were considered, both for the Billia Croo test site; wave
Buoy data, and Mike 21 modelled data. After approximately 15 years of recordings the maximum
wave height was taken from both data sets. An iterative approach was then taken to determine
which probability density function, provided the corresponding wave height Figure 2-15. This
method was preferred over reading from the graph, as the distribution becomes close to horizontal
in these regions and accuracy was of concern. The probability densities were 5.3×10-4 and 4.7×10-4
for the modelled and wave buoy data respectively. It was assumed these values would be constant
across all the assessed sites, and therefore these probability densities defined the extreme wave at
each of the sites Table 2-3.

10
0.35

0.3

0.25

Probability Density
0.2

0.15

0.1

0.05 Wave Height (m)

0
0 2 4 6 8 10

Figure 2-11: Validation of Rayleigh Distribution for with M0


shape factor for Orkney Island Area.

Figure 2-12: Statistical comparison of the measured and


modelled wave data used by (Lawrence, Kofoed-Hansen and Figure 2-10: Statistical extreme wave analysis for Billia
Chevalier, 2013) for the extreme wave analysis. Croo by (Lawrence, Kofoed-Hansen and Chevalier, 2013-
‘peaks over threshold method’

Figure 2-13: maximum wave height and significant wave height for 20 year Mike 21 model

11
2.1.3.1 Results
Table 2-3: Extreme wave calculation results

Extreme wave height (wave buoy data)

site 15 50 100

1 15.69792 16.43664 16.84294

2 11.97917 12.54289 12.85294

3 11.125 11.64853 11.93647

Extreme wave height (Modelled)

site 15 50 100

1 18.37617 19.20686 20.18778

2 14.02295 14.65686 15.40541

3 13.02306 13.61176 14.30693

Extreme sea state

site 15 50 100

1 9.989583 10.45968 10.71824

2 7.375 7.722059 7.912941

3 6.791667 7.111275 7.287059

12
Rayleigh Distrobution of wave hights for 6 year extreme sea state
0.2

0.18

0.16
Site 1
0.14
Probability Density

Site 2
0.12

Site 3
0.1

0.08 Billia Croo

0.06 Design Wave Probability (Wave


Bouy Data)
0.04 Design Wave Probability (modeled
data)
0.02

0
0 2 4 6 8 10 12 14 16 18 20
Wave Hight (m)

Figure 2-14: Rayleigh Distribution of Extreme wave heights for the 6-year extreme sea state.

0.2
Rayleigh Distrobution of wave hights for 15 year extreme sea state
0.18

0.16 Site 1

0.14 Site 2

0.12 Site 3
Probability Density

0.1 Billia Croo

0.08 Design Wave Probability (Wave


Bouy Data)
0.06 Design Wave Probability
(modeled data)
0.04

0.02

0
0 2 4 6 8 10 12 14 16 18 20
Wave Hight (m)

Figure 2-15: Rayleigh Distribution of Extreme wave heights for the 15-year extreme sea state.

13
Rayleigh Distrobution of wave hights for 50 year extreme sea state
0.18

0.16

0.14 Site 1

0.12
Probability Density

Site 2

0.1
Site 3
0.08
Billia Croo
0.06
Design Wave Probability (Wave
0.04 Bouy Data)
Design Wave Probability (Modeled
0.02 Data)

0
0 5 10 15 20 25
Wave Hight (m)

Figure 2-16: Rayleigh Distribution of Extreme wave heights for the 50 year extreme sea state

Rayleigh Distrobution of wave hights for 50 year extreme sea state


0.18

0.16

0.14 Site 1

0.12
Site 2
Probability Density

0.1
Site 3

0.08
Billia Croo

0.06
Design Wave Probability (Wave
Bouy Data)
0.04
Design Wave Probability (Modeled
0.02 Data)

0
0 5 10 Wave hight (m) 15 20 25

Figure 2-17: Rayleigh Distribution of Extreme wave heights for the 100 year extreme sea state

14
2.1.3.2 Time period (Tz) calculation

A Jonswap spectrum was developed using the Sea state parameters from Table 2-3. Because of the
type of data available several assumptions must be made (Research.dnv.com, 2018) these are
summarized below:

 The chosen values for peak enhancement factor γ, and β were suggested by
Research.dnv.com, (2018) for extreme wave conditions.

 ωp and ω are calculated using Equations 2-3 and 2-4 respectively, it has been assumed that
the peak frequency (Tp) is the mean frequency T(p). However, this would not be the case in
Bimodal seas, which can occur during extreme states (Boon, Green and Suh, 1996). Therefor
a further investigation into the spectral distributions observed at the site during storm
conditions is required to validate this assumption.

Equation 2-3: calculation of Omega_p

2𝜋
ω𝑝 =
𝑇𝑝

Equation 2-4 calculation of Omega

2𝜋
ω=
𝑇
 Recommended values for σ have been assumed as reported across all the literature
consulted including Research.dnv.com, (2018).

 Slope factor (α) has been fitted to peak enhancement factor γ using Equation 2-5. This is
suggested by (Research.dnv.com, 2018).
Equation: 2-5 estimation of lambda

0.0027
γ = 7(1 − )
α
Table 2-4: input parameters for Jonswap spectrum

Tp T(s)
β 5/4
γ 2.8
0.7 when ω < ωp
σ
0.9 when ω > ωp

Using the above notation the Jonswap Equation can be written as follows

15
Equation 2-6: Jonswap spectrum formula

α × 𝑔2 ω4𝑝 𝑎
𝑆(𝜔) = [−β 4 ] γ
ω5 ω

𝑤ℎ𝑒𝑟𝑒 𝑎 𝑖𝑠 𝑔𝑖𝑣𝑒𝑛 𝑏𝑦

(ω − ω𝑝 )2
𝑎 = 𝑒𝑥𝑝 [ ]
2ω𝑝 5 σ2

Figure 2-18 shows the Jonswap Spectrum distribution according to the methodology highlighted
above. The frequency of the extreme wave is given by the lowest frequency on the spectrum, this
ensures the wave is real wave, and provides the worst possible scenario to perform calculations.
The resultant extreme waves are found in Table 2-4; these are used to perform hydrodynamic
calculations later in the report regarding the moorings.

Figure 2-18: Johnswap Spectrum for all sites locations and extreme wave period

16
2.1.3.3 Summary

 The extreme sea state magnitude reduces toward the east of the Pentland Firth.
 The ‘100 year return period’ extreme wave heights reduce as much as 6 meters across the
Pentland Firth. The calculated extreme waves are shown in Table 2-4.
 A Jonswap spectrum was generating making relevant assumption for stormy seas. The
highest time period was then chosen to represent the extreme wave
 Sites 1 and 2 have a higher maintenance availability window due to the wave climate
compared to site 3.

Table 2-5: 100-year extreme sea state

Tz (s) Hs low (m) Hs high (m)

Site 1 13.33333 16.4263 20.18778

Site 2 11.62791 12.535 15.40541

Site 3 10 11.6412 14.30693

2.1.3.4 Future work


A large sample of field data should be recorded using a wave buoy, this time series would then be
used to inform the following studies.

 In conjunction with Mermaid, the hours available for operations and maintenance would be
defined, and an installation strategy could be developed to minimize cost according to the
wave climate.
 a more accurate deign wave analyses would be conducted to inform the detailing
engineering design on the mooring system

2.1.3.4.1 Methodology for calculating the design wave with time series data.
Using a large sample of Sea elevation data do analysis with a similar process done by (Lewis et al.,
2014) to calculate the extreme sea states at the specific site more accurately and with given
uncertainty.

1) Using samples of data calculate the three parameter Weibull distribution which fits the local
wave climate
2) Repeat the extreme wave analysis above with the new Weibull frequency distribution and
extreme sea state analysis.
3) Use the data to generate a wave frequency spectral distribution for storm conditions, select
the highest observed time period

17
2.1.4 Site selection
A preliminary GIS study has been performed to analyse the local constraints and suggest a suitable
site. The following parameters were considered when assessing the area:

Grid connections and constraints: All of the substations in the area (Figure 2-19) are constrained;
however several have a small margin available. A further section in this report discusses the grid
connection and cabling in more detail. While grid connection is a crucial factor in the site selection of
any distributed generation, in this case, over the small area being analysed, there is no significant
benefit arising from any installation at any specific location regarding the grid connection, therefore
it is not considered in the GIS study.

Current marine licenses: there is a current marine licence in place for MeyGen who have turbines
already installed in this location, however this has not been considered in the GIS study, as it falls in
an area which is to shallow for the deployment of the SR2000. There are no other current marine
licences of concern in the area.

Underwater obstacles and shipwrecks: underwater obstacles make installation and maintenance
more complex and risky and therefore should be avoided; a buffer of 100m has been used in this
report.

Protected areas: protected areas have been avoided to mitigate, potential marine licence and
environmental licence issues.

Water depth: The SR 2000 device can deploy in water depth over 25m, for the purpose of the
scoping study, a 30m water depth constraint has been applied to allow for error in the low-
resolution bathymetry data.

Shipping lanes: shipping data from (Marinetraffic.com, 2018) has been used in the GIS study as a soft
constraint. Deploying in an area of high marine activity, will increase the cost and risk of
maintenance and installation, furthermore, it will add complexities to the application for the marine
licence.

Available energy: according to the resource assessment in section 2.2.2.1 desirable locations are
those with a mean power density of 8-10 kW/m2. The date for this section is from (Neill et al., 2017)
(Lewis et al., 2014)

Wave climate: according to Figure 2-20 at all tidal modes, the mean significant wave height reduces
towards the east of the Pentland Firth. The frequency of rough sea-states has implications for power
generation, mooring systems and installation windows.

Access: an assessment of the Harbours, ports and docks concluded that Aberdeen port (150 miles
away) could be used for installation and any maintenance requiring a deepwater vessel, while local
ports such as Scrabster Harbour and Copland Dock are both within 20 miles of the Pentland firth,
and could be used for more regular maintenance. There is no significant benefit to a specific location
regarding ports and access, therefore this is not included in the GIS study.

18
Figure 2-20: Map of local substations (Scottish and Southern energy networks, 2018).

Figure 2-19: Mean significant wave high in all peak tidal modes.

19
2.1.4.1 Hard Constraints Map

Firstly the hard constraints are used to eliminate areas which are wholly unsuitable for installation
Figure 2-22. The resultant available area is shown in Figure 2-23.

Figure 2-22: map of the hard constraints

Figure 2-21: Map showing the areas which is not suitable for installation of SR200, and areas which are suitable pending
analysis of the soft constraints

20
2.1.4.2 Soft constraints
The soft constraints were visualised using GIS software alongside the available area. This allowed for
analysis of potentially suitable locations.

Figure 2-23: mean power density of the tidal resource shown alongside the unavailable area from Figure 2-21

Figure 2-24: The frequency of shipping through the Pentland Firth, shown alongside the unavailable area from Figure
21
2-21
2.1.4.3 Site selection
The site was then selected considering both the hard and soft constraints; the resultant site is
located in energy-dense waters whilst also avoiding the major shipping lane, and all underwater
obstacles. Furthermore, it is towards the east where the wave climate is least severe (Figure 2-25
and 26). The main concern with this site is the high amount of marine traffic which passes between
the sites the shore on either side, which is a concern for the installation and Operations and
Maintenance phases of the project. Furthermore, a detailed Geotechnical survey will need to be
careered out to ensure the site is suitable, and beyond that to design a mooring system and cable
laying strategy.

Figure 2-25: site selection map, showing all relevant constraints

Figure 2-26: Site map showing the constraints alongside the current energy density

22
3 Site specific assessment
3.1 Resource assessment at site
3.1.1 Methodology
The process for the site-specific resource assessment follows the procedure presented by Cornett in
his “Guidance on assessing tidal current energy resources”. This suggested methodology is built
upon the harmonic analysis add-in tool for MatLab; ut_solve and ut-reconstr. This was suggested
through the tidal analysis workshop as part of the 3rd Year Marine energy module (Smith, 2016). This
involved the determination of the tidal constituents from a known data, from which tidal flows could
be predicted for a set period at the location where data was collected. Power density is then
determined using the kinetic energy in flow Equation (Equation 3-1).

Equation 3-1 Power density equation used in resource assessment

𝑃
𝐴
:Power Density (Wm-2)
𝑃 1 𝜌: Water Density (1025 kgm-3)
= 𝜌𝐶 𝑈3 𝐶𝑝 ; Efficiency coefficients taken from (Cornett,
𝐴 2 𝑝
2006)
𝑈: Flow velocity

By using an established method, it is hoped that the uncertainties associated with an unproven
methodology are avoided, improving the study’s validity. However, where Cornett’s guidance is less
clear, in areas such as in the determination of current shear profile, the work of Lewis et al is used to
estimate the shear profile of (FoW) data. This approach uses the suggested equation (Equation 3-2)
and coefficients α=7 and β=0.4, (Lewis et al., 2017) determined for the site of Anglesey. While it is
acknowledged that the use of the coefficients is not sufficient for a resource assessment associated
with a full feasibility study (due to the potential differences in seabed at the two locations). For the
purpose of a high order scoping study it is assumed that the limitations introduced by the
differences in geographical locations are acceptable, if mitigation methods of these limitations are
considered. This mitigation occurs by use of a secondary, measured, shear profile; taken from ship
mounted acoustic Doppler current profiler (ADCP) data for the Pentland Firth. This allows for the
impacts of differing locations (a factor highlighted by Lewis et al as significant) to be mitigate by
allowing a comparison of the two data sets to draw conclusions.

Equation 3-2 Velocity profile equation based on the power law, extracted from Lewis et al (Lewis et al., 2017)

1 𝑈𝑧 : Velocity down the water column


𝑧 𝛼
𝑈𝑧 = ( ) 𝑢̅ 𝑧: Water depth
𝛽ℎ
𝑢̅: Reference velocity at 20m sea depth

As stated in the site selection section; data from the FoW was used to determine energy yields, due
to the lack of data relating to the Pentland Firth. As such this is acknowledged as a significant
limitation within the study, even with established similarities between the Pentland and FoW. While
not having data related to the chosen site is a poor substitution for a feasibility study, in the context
of a scoping study it is considered acceptable if the limitations are clearly stated.

23
An additional limitation that must be considered is the use of a harmonic tool used in tidal
prediction. While the use of tidal constituents offers the potential for long term analysis it does not
consider bespoke aspects at the site such as bathymetry. While the approach taken is highlighted by
Cornett as appropriate, it is acknowledged that this approach has limitations especially when
compared to more sophisticated numerical models.

Table 3-1 Summary of limitations associated with this resource assessment

Limitation Impact and Mitigation Significance


Assumption that Fall of Introduces limitations, as site investigated is High
Warness Data is representative not in the FoW. Mitigation takes place
of chosen site through acknowledgement of the
limitations, and use only in a scoping study
Coefficients used from Lewis et Introduced limitations on the accuracy of the High
al shear profile prediction, mitigated by use of
additional data sets
Non-consideration of location This occurs as a result of the approach Low
bathymetry chosen, however impact on to the study is
considered minimal due to the recognition
surrounding this approach

A summary of highlighted limitations are presented in Table 3-1, for use as a reference.

An attempt to quantify the limitations above is done through the use of literature. In regards to the
shear profile this is done through Lewis et al’s own analysis which suggests a total error ranging from
0.005% to 0.137%. Bryden et al meanwhile suggests significantly higher errors for predictions based
on kinetic energy (Bryden et al., 2007), an approach used in this study. They suggest that for current
velocities taken close to the surface, an expected power value could be from 110% or 90% of the
true value (Bryden et al., 2007). This variation from the true value increases for measurements taken
as a mean depth to 50% to 70% of the true value (Bryden et al., 2007). It is unknown at what depth
FoW is taken from, as such it is assumed to be taken at depth. As such Bryden et al’s errors of 50% to
70% have been considered and applied to the results of this analysis as a correction factor. As a
result of the uncertainties considered above, the results of the analysis will be compared to
literature to act as final verification.

3.1.2 Results
Figure 3-1 shows the results of the harmonic analysis prediction and raw data from the FoW. From
this there is a strong suggestion that the y components of velocity are appreciably larger, strongly
influencing the tidal ellipse. The ellipse itself suggests a flow mainly flows along the NW, SE axis, but
is not completely bi-directional with slight, ‘s’ shape and oval present in the predicted data set. This
suggests some deviation from a simple bidirectional flow. These sentiments are reflected in the
velocity roses for the raw data and predicted data series, with the strongest velocities of 3-3.5 ms-1
occurring along the SE direction with some variation towards the east and south at lower velocities
(Figure 3-2).

24
Figure 3-1: Results of harmonic analysis, top line from left to right shows velocity components, resultant velocity and tidal
ellipse for raw data, bottom row show the same but for predicted data for one year.

Figure 3-2: Velocity rose for raw data (left) and predicted data (right) for the FoW

From Figure 3-1 it can also be determined that max velocity is within the region of 3.46 ms-1 to 3.64
ms-1. This is suggestive of a site close to the upper limits of the generic energy converter suggested
by Cornett. However, with a mean current speed of c.1.6 ms-1 this counters suggestion, implying that
the velocity range from max velocity to min velocity is potentially within the limits suggested by
Cornett (Cornett, 2006).

25
Figure 3-3: Annual power availability from predicted FoW data series, top showing power density variation throughout the
year, middle showing this variation as a cumulative probability curve and bottom showing the power density as a
histogram.

Figure 3-3 shows the clear effect of applying the Cornett’s efficiency curve to the available resource,
seeing a reduction in peak power density from c.25 kWm-2 to c.9 kWm-2. This is also seen in the
cumulative probability, with available resources reporting a power density of less than c.10 kWm-2
100% of the time and extractible power seeing this reduced to c.5k kWm-2 100% of the time. This is
reflected in histogram by the trailing of probabilities earlier for extractable power, however, it
should be noted that the probability stays consistent at the lower velocities of less than 4 ms-1. This
trend is additionally shown in the annual average power and annual energy yield which reduce 61%
between available and extracted power (Table 3-2).

Table 3-2: Summary of outcomes of resource assessment of FoW data

Annual average power density Annual energy yield (kWhm-2)


(kWm-2)
Available Resource 3.71 32500
Extractable Resource 1.46 12800

Figure 3-4 suggests that in terms for shear profile there is appreciable difference between the FoW
and the Pentland Firth with a difference of c.22% between the two-data series at a normalised depth
of 0.8. Similarly, there is a discrepancy between the relationships of the shear profile suggesting
different power relationship between the two sites.

26
Figure 3-4: Shear profile in velocity and power density for predicted FoW data series, and comparison with Pentland Firth
ADCP data

3.1.3 Discussion
The results of the eclipse data and velocity rose suggest a predominantly bi-directional flow with
some eccentricities resulting in the formation of an s shape flow and a slightly wider ellipse.
Typically, tidal current devices are more suitable to bi-directional flow, avoiding the need for
complex yawing mechanisms. While Figure 3-1 does suggest some deviation from a true bi-
directional flow, it is not significant enough in this study to be of major concern for site suitability.
However, when Figure 2-6 is considered the formation of the islands around the FoW goes some
way to explain the strong NW-SE direction of the flow, as this is the general orientation of the
channel. The presence of the island south of Eday would also explain the suggested ‘s’ shape of the
flow as the currents flow around the island.

In regard to the Cornett’s generic energy converter, the mean and max values for velocity (Figure
3-1) would strongly suggest that the FoW fits well into the operational rage of the converter. Using
the assumption stated in the methodology, that the chosen site is analogous to the FoW , this would
suggest a good fit between the operational range of the device and the chosen site with the stated
caveats acknowledged.

Figure 3-3 clearly shows the importance of sizing a site to a specific device, with a large portion of
the available power not available for power extraction. This drop in extractable power in comparison
to available resource is less surprising when the theoretical maximum extractable power is
considered at c.60% of the available resource (The Betz limit). Therefore, considering the immaturity
of the technology the drop from available to extracted power is expected. In general, there are
losses with the device that could be reduced with improved efficiency, even considering Betz limit.
However; while there is a limit to the power extracted, the device must be capable of surviving the
power present in the resource with a current velocity in the region of 3.64 ms-1.

As stated in the methodology, the results of the analysis are compared to literature. This is done in
Table 3-3, with consideration to the findings of Bryden in regards to the under prediction of 30% and

27
50% (Table 3-3). This analysis does confirm that the available resource predicted using the FoW data
is somewhat comparable; with all values, at each error correction and base value, within the range
of power densities suggested by Neil et al. However, it is also clear that the predictions are
consistently on the lower end of the suggested range. To this extent the predictions can be
considered extreme conservative estimates, with a maximum annual average power density of 2.19
kWm-2 and annual energy yield of 19,200 kWhm-2 per a device.

This analysis however is acknowledged as having significant limitations still present and should not
be relied upon for further study. The results are not strongly confirmed, with the UK renewable atlas
suggesting peek current speeds up to 4 ms-1 around the Pentland Skerries. This is only a slight
increase from the max value predicted of 3.64 ms-1 (Figure 3-1), however, this difference of 9% in
velocity would potentially increasing power density predicted to 6 kWm-2; a value still considered on
the low side of the range suggested by Neil et al. The remaining differences in predicted max power
density and those suggested by Neil et al could be a result of the underestimation of non-kinetic
energy sources. These would include the slight gravitational potential present in tidal flow.
Alternatively, there is potential for inaccuracies in Bryden et al’s estimation of resource prediction.
This could then result in an under compensation of the errors present. While not within the scope of
this study to investigate further, the results do suggest the non-suitability of the method used for
future resource assessments as part of a feasibility study.

Table 3-3: Comparison of available and extractable resource predicted from FoW data with Neil et al power density values
for the area around the Pentland Skerries, location closest to the chosen site. For the determination of power and energy
2
the swept area of SR2000 is used at 402 m .

Annual average Annual energy Annual average Annual Energy


power density yield (kWhm-2) power per a yield per a
(kWm-2) device (MW) device (GWh)
Available Resources 3.71 32,500 1.49 13.1
Available Resources 5.57 48,750 2.24 19.6
(+50% error) (Bryden
et al., 2007)
Available Resources 4.82 42,250 1.94 17.0
(+30% error) (Bryden
et al., 2007)
Extractable 1.46 12,800 0.587 5.15
Resources
Extractable 2.19 19,200 0.880 7.72
Resources (+50%
error) (Bryden et al.,
2007)
Extractable 1.90 16,640 0.764 6.69
Resources (+30%
error) (Bryden et al.,
2007)
Pentland Skerries, 10 - - -
Low (Neill et al.,
2017)
Pentland Skerries, 3 - - -
High (Neill et al.,
2017)

28
In regard to the shear profile suggested by Figure 3-4, while it is difficult to confirm without further
data, it is likely that sheer profile at the chosen site follows a relationship similar to the ADCP data.
However, it is expected to have velocities between the two trend lines presented. Differences in the
power laws describing the trends are to be expected, due to the changes in location and the implied
changes to seabed roughness. This is confirmed by literature, with Easton, Woolf, and Bowyer
suggesting drag coefficients for the Pentland seabed ranging from 0.0015 to 0.0025. In comparison
the drag coefficient used in Lewis et al’s study (and within this analysis) was 0.005; this may explain
the variation seen in the trends for the measured ADCP data and the predicted data.

The impact on the power production of a device deployed at the location, as a result of a shear
profile, is expected to be more in line with the results of the ADCP data. This would see a device
experiencing a faster decline in velocities as the structure proceeds down the water column. For a
device operating near the surface this may bean a greater range of loads induced in the blade as it
passes through the water column resulting in potential for fatigue loading on the blade structure.

3.1.4 Summary
While there is a high potential that predictive result from the FoW data are under representative of
the tidal flow at the site. The results suggest an annual average extractable power within a region of
0.587 MW and 0.88 MW per a device, with an annual extractable energy yield of 5.15 GWh and 7.72
GWh per a device. Based on FoW data, tidal flow is likely to be bi-directional; following a rough
NWSE axis. Shear profile is expected to follow a steep trajectory, similar to the recorded ADCP data
in Figure 3-4, but as a higher velocity in keeping with the predictive FoW data. Based on the
Cornett’s generic device, while there are suggestion of under prediction by the FoW data, peek
current speed is not expected to typically exceed 4 ms-1 (ABPmer, 2018). This value is outside of the
efficiency curve of the generic device; however, mean current speeds are well below the limit
suggesting that the majority of the flow regime matches well to the generic device. However even at
the higher estimation of speeds this is still within the rated cut-out current speed for the SR2000.

It is suggested that future work focuses on the collection of measured data at the site to confirm of
disprove assumptions made in this study. The results of this analysis should not be used as a
replacement of measured data in a feasibility study due to the large uncertainty surrounding the
results. The results can be considered as extreme conservative estimations of resource at the chosen
site location. In addition to this investigation on the impact of future arrays at westward location in
the Pentland should be considered as bar the suggestion of Drapper et al towards reduced energy
availability with the installation of devices in series.

3.2 Device choice


As the study requires the use of a commercially-ready floating tidal device, the Scotrenewables
SR2000 device (Scotrenewables, 2018) has been utilised for the assessments throughout this report.

29
3.3 Mooring design
This early scoping study will not provide a definitive design of the mooring and foundations of the
proposed device. However, conclusions drawn by the Carbon Trust suggest that requirements for
mooring and installation are left too late in the design process (Starling and Scott, 2009). For this
reason a preliminary concept design of the mooring is suggested in this section allowing for a more
accurate judgment of the cost of the mooring system.

3.3.1 Design Considerations


Whilst a range of variables govern the success of the project as a hole the success of the foundations
and mooring system may be attributed to specific criteria that govern the risks involved with
mooring:

- Lifetime - Operations and Maintenance


- Loading - Cost
- Instillation

3.3.1.1 Lifetime
Typically it is shown that leases for tidal energy sites run for 20 year periods (Starling and Scott,
2009). Therefore, assuming no intent to replace the foundation and mooring system, a lifetime of
over 20 years must be designed for.

3.3.1.2 Loading
The loads acting on the device and hence the mooring and foundation is dependent on a number of
factors:

- Tidal current speed - Turbulence at the site


- Current direction relative to device - Induced vibrations
- Wave loading - Impacts and other events
- Wind loading

The above factors are covered comprehensively by the DNV-OS-E301 and DNV-RP-C205 standards
(DNV, 2010; DNV C205, 2010). The following sections show adherence to these standards covering
suggested loading conditions.

3.3.1.2.1 Wave loading


The design of mooring systems for offshore marine devices is covered in DNV-OS-E301 (DNV, 2010),
wherein it is stated that the JONSWAP wave spectrum may be used to define the environmental
states present at the device location. This was implemented for the design site in Section 2.2.3.
Moreover, in accordance to E301, the 100 year design wave was used in this investigation for the
ultimate limit state (ULS). E301 states that whilst the fatigue limit state (FLS) is important to
consider, it is more crucial for steel and has therefore been excluded from the investigation which,
as stated, is used to provide an estimate for the cost and material required to moor the device (DNV,
2010).

30
3.3.1.2.2 Wind loading
The SR2000 device exists primarily below water level when in operational mode (Scotrenewables,
2018). Whilst it is noted that the wind will not greatly affect the mooring design, this study includes
the typical value suggested by E301 for North Sea (Greater Ekofisk area), see for the wind speed
value (DNV, 2010).

3.3.1.2.3 Current direction relative to unit


Standard suggest that four load directions should be regarded for symmetrically moored devices
(DNV C205, 2010). However, this has been omitted from the study as it was shown in Section 3.1.2
that the wave and current direction at the design location is largely unidirectional.

3.3.1.3 Marine growth


Marine growth has been accounted for via the implementation of standards outlined in (DNV, 2010;
DNV C205, 2010). Limited by the lack of known data for marine growth at the design location, the
standard increase in outer diameter of 50mm has been used in this investigation.

3.3.1.4 Safety factors


DNV-OS-E301 states several safety factors for the design of the mooring system (DNV, 2010). The 2.5
value assigned to the ULS load case, with a correction factor of 1.1 to account for the lack of vortex
induced motion has been chosen as the value for this investigations, this can be found in Table 3-7.

3.3.2 Mooring Design Methodology

3.3.2.1 Drag Coefficient for Circular Cylinders


The calculation of drag coefficients for the device was completed in accordance to the methods
suggested by (DNV-C205, 2010). The process requires the previous calculation of several variables,
Reynolds number, Keulegan-Carpenter number and the Stokes Parameter.

The Reynolds number flow characteristics are well defined for slender circular members. Equation
3-3 defines the calculation of Reynolds number (Re) used in this investigation; the diameter supplied
by Scotrenewables has been used as the dimension characteristic.

𝑈𝐷
𝑅𝑒 = Equation 3-3: Reynolds number
𝑣

𝑈 = 𝑀𝑒𝑎𝑛 𝑎𝑝𝑝𝑟𝑎𝑜𝑐ℎ 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦 𝑜𝑓 𝑓𝑙𝑢𝑖𝑑


𝐷 = 𝐶ℎ𝑎𝑟𝑎𝑐𝑡𝑒𝑟𝑖𝑠𝑡𝑖𝑐 𝑑𝑖𝑚𝑒𝑛𝑠𝑖𝑜𝑛 𝑜𝑓 𝑏𝑜𝑑𝑦
𝑣 = 𝐾𝑖𝑛𝑒𝑚𝑎𝑡𝑖𝑐 𝑣𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦 𝑜𝑓 𝑓𝑙𝑢𝑖𝑑

31
The Keulegan-Carpenter number (KC) is described by Equation 3-4. The time period used for the
calculation was estimated via the wave spectra produced in Section 2.2.3.2.

𝑈𝑇
𝐾𝐶 = Equation 3-4: Keulegan-Carpenter number
𝐷

𝑈 = 𝑀𝑒𝑎𝑛 𝑎𝑝𝑝𝑟𝑎𝑜𝑐ℎ 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦 𝑜𝑓 𝑓𝑙𝑢𝑖𝑑


𝐷 = 𝐶ℎ𝑎𝑟𝑎𝑐𝑡𝑒𝑟𝑖𝑠𝑡𝑖𝑐 𝑑𝑖𝑚𝑒𝑛𝑠𝑖𝑜𝑛 𝑜𝑓 𝑏𝑜𝑑𝑦
𝑇 = 𝑃𝑒𝑟𝑖𝑜𝑑 𝑜𝑓 𝑓𝑙𝑜𝑤

The Stokes Parameter (𝛽) is then calculated via Equation 3-5.

𝑅𝑒
𝛽= Equation 3-5: Stokes Parameter
𝐾𝐶

A value for surface roughness (K) is required by (DNV C205, 2010) and was chosen based on the
standard values suggested. Table 3-7 contains the value used and accounts for marine growth on the
device.

A correction to the drag coefficient as a function of KC is suggested by (DNV C205, 2010). A wake
amplification factor (ϕ) based on Equation 3-6 is used.

ϕ = 𝐶𝜋 + 0.10(𝐾𝐶 − 12) Equation 3-6: Wake Amplification


Factor

Where C π is defined via Equation: 3-7 and may be found in Table 3-7.

12
𝐶𝜋 = 1.50 − 0.024 − Equation: 3-7
𝐶𝐷𝑆 − 10
𝐶𝐷𝑆 = 𝐷𝑟𝑎𝑔 𝑐𝑜𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡 𝑓𝑜𝑟 𝑠𝑡𝑒𝑎𝑑𝑦 𝑓𝑙𝑜𝑤

Finally, a correction to the dimensional characteristic of the body was applied in order to account for
the effect of marine growth on the device. This is described by Equation 3-8 where growth thickness
(t) is the suggested value for water depths greater than 40m. The corrected characteristic may be
found in Table 3-7.

Equation 3-8: Dimension


𝐷 = 𝐷𝐶 + 2𝑡 Correction for Marine Growth

𝐷𝐶 = 𝑂𝑟𝑖𝑔𝑖𝑛𝑎𝑙 𝑜𝑢𝑡𝑒𝑟 𝑑𝑖𝑎𝑚𝑡𝑒𝑟


𝑡 = 𝑇ℎ𝑖𝑐𝑘𝑛𝑒𝑠𝑠 𝑜𝑓 𝑚𝑎𝑟𝑖𝑛𝑒 𝑔𝑟𝑜𝑤𝑡ℎ

32
The final drag coefficient of the body is then found via Equation 3-9. It should be noted that the
calculations specified above were run for two of cases:

- Normal drag coefficient.


- Axial drag coefficient.

𝐶𝐷 = 𝐶𝐷𝑆 ∗ ϕ Equation 3-9: Drag Coefficient

𝐷𝐶 = 𝑂𝑟𝑖𝑔𝑖𝑛𝑎𝑙 𝑜𝑢𝑡𝑒𝑟 𝑑𝑖𝑎𝑚𝑡𝑒𝑟


𝑡 = 𝑇ℎ𝑖𝑐𝑘𝑛𝑒𝑠𝑠 𝑜𝑓 𝑚𝑎𝑟𝑖𝑛𝑒 𝑔𝑟𝑜𝑤𝑡ℎ

3.3.2.2 Solver Regime


Figure 3-5 describes the limits of solver regimes suitable for the dimension characteristics of the
body and wave state (DNV C205, 2010). The characteristic dimension of the body (D) was chosen as
the diameter of the main body of the device. The wavelength (λ) was chosen as the lower and upper
limits of the spectrum defined in Section 2.2.3.2. Table 3-7 contains these values and shows that the
simulation falls within the negligible diffraction regime, thus Morrisons equations apply.

Figure 3-5 Wave force regimes (Charkrabarti, 1987 cited by, (DNV C205, 2010)). D = characteristic dimension, H = Wave
height, Y = Wave length.

33
3.3.2.3 MODEL GEOMETRY
OrcaFlex requires the center of mass (COG) and the moment of inertia values for the simulation. In
order to calculate these variables, SolidWorks CAD package was used. The model was constructed
based on the dimensions and mass provided by (Scotrenewables, 2018). Figure 3-6 shows an
overview of the CAD model.

Figure 3-6: SolidWorks model providing geometry for OrcaFlex.

Figure 3-6 displays the model used in order to calculate the variables required for OrcaFlex. It should
be noted that the entire body was shelled to a thickness of 0.02 m and the material set to Annealed
Stainless Steel. Table 3-4 contains the details of the SolidWorks model including variables required
by OrcaFlex.

34
Table 3-4: SolidWorks Variables for Model Geometry. Table 3-5: Scotrenewables SR2000 - Device
Characteristics (Scotrenewables, 2018).

SolidWorks Variables SR2000 Characteristics


Variable Value unit Variable Value Unit
Mass 4000000 Kg Hull Length 64 m
Material Annealed Stainless Hull Max Diameter 3.8 m
N/A
Steel
Ixx 1545355561 Kg/m2 Transport draught 6 m
Ixy 49996 Kg/m2 Operational draught 25+ m
Ixz 21760 Kg/m2 Rotor diameter 16 m
COG (x,y,z from Swept Area 2*201 m2
(25,0,3) m
end A)
Cut-in speed 1 m/s
Cut-out speed 4.5 m/s
Max rotor speed 16 rpm

3.3.2.4 Environmental and Device Influences on Mooring Design


The Pentland Firth is known to have a seabed of hard rocks (Starling and Scott, 2009) and is
therefore capable of causing degradation to mooring lines through abrasion. For this reason the,
lower section of the mooring will be modelled as a chain as suggested by (Karimirad et al., 2013).

Moreover, the cycling of rope mooring lines from wet to dry conditions is known to reduce the life of
lines and should be considered in more detailed design iterations (Weller et al., 2015). A further
assumption that has been drawn for this preliminary study is that due to the energy present at the
design location and the weight of the device, traditional catenary mooring systems that rely heavily
on the weight of mooring lines on the seabed and reduce the load at the anchor point not suitable
for the conditions.

The design presented in this preliminary study is a multiple taut line system, as anchor points are
capable of loading. Moreover, the costs associated with multiple taut lines are known to be lower
than those for multiple catenary lines (Karimirad et al., 2013).

35
3.3.3 OrcaFlex Numerical Simulation Overview
Selecting a mooring system has large implications on the cost of the chosen design. The main criteria
for a successful mooring system are that it can withstand the 100 year wave and fatigue loading with
a high factor of safety. These criteria will then inform the volume of material required to hold the
device/devices in place. In order to calculate the required volume of material the chosen device,
Scotrenewables’ SR2000 was modelled in SolidWorks based on the dimensions supplied by
Scotrenewables. This was combined with the extreme wave data calculated in Section 2.2.3 and
supplied to OrcaFlex. Some variables required original calculation, in this case the DNV205
document was followed. Table 3-7 provides an overview of the variables supplied to OrcaFlex.

3.3.3.1 OrcaFlex Model


The model has been shown to exist within the negligible diffraction regime, hens it has been
modelled within OrcaFlex as a 6 degrees of freedom (6DOF) buoy split into 10 equal sections of 6.4m
length to form the total 64m device. Data contained within Table 3-5 allowed for the design of the
draught of the model. The general variables for the model where left as standard however, for the
final simulation a stage 1 simulation time of 3hours was used in accordance to (DNV C205, 2010).

The Environmental variables of the model are described in Table 3-6 and have been defined in
previous sections of this report. As the wire frame used in the model does not include the physical
dimensions of the turbines, a further thrust force is applied to the buoy. For the purpose of this
investigation, actuator disk theory has been used to calculate the thrust. Equation 3-10 describes
this phenomenon.

Equation 3-10 Thrust


𝐹𝑇 = 𝜌𝐴 ∗ (1 − 𝑎) ∗ 𝑈 − (𝑈(1 − 2𝑎) Force

𝐹𝑇 = 𝑇ℎ𝑟𝑢𝑠𝑡 𝑓𝑜𝑟𝑐𝑒
𝜌 = 𝑊𝑎𝑡𝑒𝑟 𝑑𝑒𝑛𝑠𝑖𝑡𝑦
𝐴 = 𝑅𝑜𝑡𝑎𝑟 𝐴𝑟𝑒𝑎
𝑈 = 𝑊𝑎𝑡𝑒𝑟 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦
𝑎 = 𝐼𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑓𝑎𝑐𝑡𝑜𝑟

Table 3-7 contains the results of the thrust force calculation and the variables used. It should be
noted that further studies should not relay on the actuator disk theory as it requires a large number
of caveats however, for the purpose of this preliminary investigation it has been deemed suitable.
Moreover, overestimating the thrust applied to the device will increase the design strength of the
mooring system.

Table 3-6: Environmental conditions for OrcaFlex.

OrcaFlex Environmental Variables


Variable Value Unit
Seabed 75 m
Hs 7.3 m
Tz 6.4 s

36
Table 3-7: Input Variables for OrcaFlex mooring design simulation.

OrcaFlex -Mooring Requirement Data


Variables Symbol Value Unit
Axial Drag Coeff 1.2 N/A
Normal Drag Coeff 1.0 N/A
Device Mass 400000 Kg
1 hour mean wind speed 34 m/s
πD/λ 0.3 – 0.9
Factor of Safety FOS 2.75 N/A
Surface roughness K 5e-3 M
Marine growth thickness t 50 Mm
Clean water outer diameter DC 3.8 M
Marine growth outer diameter D 3.9 M
Roughness dimension ratio K/D (Δ) 1.32e-6 N/A
Thrust force FT 1303 KN
Water density 𝜌 1029 kg/m2
Rotor Area A 402 m2
Water velocity U 4.5 m/s
Induction factor a 0.3 N/A

3.3.3.2 Simulation Limitations


The investigation omitted the inclusion of vortex induced motion (VIM), this is known to increase the
tension in the mooring lines (DNV, 2010) and should be accounted for in the factor of safety value in
this investigation.

3.3.3.3 ORCAFLEX RESULTS


The mooring system design has been modelled to allow tension at the anchor. Therefore, it does not
rely on the weight of lines to hold the device in place. Primarily this is due to the unidirectional
nature of the environmental conditions. A three line system is explored in this preliminary
investigation. The results shown in this section are a line diameter optimisation and the ULS case.
Mooring system analysed in this study can be seen in Figure 3-7.

Line C
Line B Line A

Figure 3-7: Final mooring design, shown at 10s point of the 100year wave.

37
3.3.3.4 Line diameter optimisation
An optimisation process for the diameter of the mooring lines was used. Figure 3-8shows the results
of the line diameter optimisation. The simulation was run with the full mooring design in place for
50s of the 100year wave sea-state. It can be seen that the mooring line reaches the desired factor of
safety at a diameter of 0.2m. This diameter was then chosen as a starting point for the full
preliminary design of the mooring system.

Line A - Diameter Optimisation Maximu


5 m
Normalis
4
Normalised Load & FOS

ed Load
FOS
3

1 Design
FOS
0
0 0.1 0.2 0.3 0.4 0.5
Mooring Line Diameter (m)

Figure 3-8: Example mooring line diameter optimisation.

3.3.3.5 Ultimate Limit State (ULS)


Each of the three mooring lines was simulated with the other lines removed. The simulations ran for
50s of the 100 year wave event. An 8s build time was used in order for the initial dynamics to
resolve. As suggested by Figure 3-9 in isolation, under the 100 year wave event line A and B only
reach 30% of their breaking tension, thus it can be assumed that it would be extremely unlikely that
line C must support the device in isolation.

The 30% load corresponds to a factor of safety (FOS) of 3.3, well above that suggested in Section
3.3.2 as a correction factor for the standard presented in (DNV, 2010). It should be noted that whilst
line C does very little to support the device in the extreme wave event, it must be included in the
design to maintain its orientation.

Normalised Tension
0.3
Normailsed Tension

0.25
0.2
0.15
0.1
0.05
0
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Figure 3-9: Normalised Tension at point 1 on the line A over 50s of the 100 year design wave.

38
3.3.3.6 300hour Simulation
One final simulation was run for the chosen mooring design. The 100year wave sea-state was run for
300hours in accordance with (DNV, 2010). It can be seen in Figure 3-10 that for the described design
case the tension in line A never exceeds a normalised tension of 0.4. This gives reasonable
verification for the suggested mooring design and will allow for estimates of cost. It should be noted
that further, more detailed analysis of the loads experienced in the mooring system should be
undertaken before the project is progressed.

Normalised Tension
0.4
0.35
0.3
Normalised Tension

0.25
0.2
0.15
Normalised
0.1 Tension
0.05
0
0 2000 4000 6000 8000 10000
Time (s)

Figure 3-10: Normalised Tension in Line A for the 100 year wave sea state.

39
3.3.3.7 Mooring Design Validation
The final characteristics of the suggested mooring design are summarised by Table 3-8 and Figure
3-11. Interestingly, it can be seen by comparing Figure 3-11 with Figure 3-12, an image of the
SR2000’s current mooring design, that there are similarities between the two. This suggests that
whilst the mooring system proposed in this study may not be detailed, it is reasonable.

Table 3-8: Preliminary Mooring Design Summary

Variable Length (m) Diameter (m)


Chain Nylon Rope
Line A 20 100 0.25
Line B 20 100 0.25
Line C 20 80 0.25

Figure 3-11: Render of Final Preliminary Mooring System.

Figure 3-12: Current SR2000 Mooring Design (Hamilton, 2018).

40
3.3.3.8 Further work
If the design suggestions are to carried forward to feasibility stage, more detailed investigations
according to (DNV, 2010) should be considered, including the three extra directional load cases
omitted from this investigation. Omitting vortex induced motion from the investigations reduced
the validity of the results and should be improved upon in further investigations.

3.3.4 Implications for Upscaling


Given the suggested increase in capacity via the instalment of several SR2000 devices, it is important
to discuss the implications of two scenarios, multi-mooring systems for a number of devices, and the
spacing required between each device. The latter is a risk mitigation technique.

3.3.4.1 Multi-Device Mooring


This scenario would require an increase in mooring lines and foundation points. However it is noted
that in the detailed design stage of the project, mooring two or more devices via the same line
should be considered as this would reduce the costs of the overall system.

3.3.4.2 Device Spacing


The spacing of each device is defined via the geometry of the device and the lengths of the mooring
lines. Figure 3-13 represents the distance achieved in the worst case scenario. When the device has
lost two mooring lines and has drifted to reach its full extent. It can be seen that in this case the total
length of the system becomes 160m. A buffer zone of 40m is then applied to this radius resulting in a
minimum spacing between devices of 200m.

Based on the nominal power of the SR200 of 2MW per device (Scotrenewables, 2018) this results in
a packing density of 64MW/km2. This can be seen to be a reasonable value when compared to
(Hamilton, 2012) who state a packing density of 60MW/k m2 .

Figure 3-13: Device Spacing for 64MW Array.

41
3.3.5 Risks
Risks associated with mooring design are predominantly due to the lack of data available. General
mitigation for this is to produce more a detailed model and undertaken scale model testing, which is
a planning requirement.

The key risks associated with the mooring system are outlined in Table 3-9. These are informed
similar projects and standards used to design mooring systems (DNV, 2010; Taaffe and Manager,
2016).

Table 3-9: Mooring Design Associated Risks

Category Risk Description Severity Mitigation


Mooring Excessive Seabed may cause Could result in the Adequate operations and
Line corrosion. corrosion in the failure of one or more management systems and
mooring lines. mooring lines. following of standards in
design stage.
Lines causing Position and Could result in the Detailed survey of the
harm to sea movement of termination of site surrounding areas and
life. mooring lines could operation. rigorous operation and
harm sea life. management scheme.
Continued monitoring of
the site after instalment
will help inform.
Connection A mooring line may May result in the loss Adherence to known
to device or become unattached of power generation standards and installation
seabed is from the device or or damage to the procedures.
lost. the seabed. device. Adequate operations and
maintenance systems.

3.3.6 Mooring Summary


A preliminary mooring design has been informed via, data extrapolated to define the site device
specific variables and the output from a CAD model of the SR200. Several summative points
regarding the mooring design may be drawn.

- Further assessments must be undertaken to confidently design the mooring system,


location specific data is crucial to the model.
- The bathymetry of the site and conclusions drawn from similar studies suggest that
mooring should be a taught system.
- DNV design procedures were followed and the system analysed for the 100year wave
event allowing for a design with a factor of safety of over 3.
- Given the wave and tidal data provided, the mooring lines are made up of both chain
and synthetic rope.
- An estimated total line length of 370m and diameter of 0.25m is suggested allowing for
the estimation of cost.
- A preliminary design for the upscaling of site allowing for multiple devices was included
and was found to predict a packing density of 64MW/km2, similar to that predicted by the
device developers.

42
3.3.7 Foundation Design
The results and conclusions of the preliminary mooring design shown in Section 3.3 allow for a more
contained selection of foundation system. Moreover, the environmental conations discussed in
Section 2 further refine this selection process. The key variables driving the foundation selection are
listed below:

- Sea bed composition: in Pentland Firth the ‘bedrock geology is dominated by Middle Devonian
sedimentary deposits consisting of Old Red Sandstone and conglomerates (Figure 3-14)’
(Starling, Scott and Parkinson, 2009).
- Vertical loads present in mooring lines: the extreme load which the mooring lines have to
withstand according to the Orcaflex study conducted is 2800 KN (Martin-Short et al., 2015)
- Sea Bed Bathymetry: Figure 3-14 shows that the sea bed in the Pentland Firth is uneven; at
places there are steep slopes and deep spots, this is specifically prevalent in the chosen site
location.

Figure 3-15: sea bed in the east Pentland Firth Figure 3-14: Bathymetry of the Pentland Firth Region
(Marine.gov.scot, 2018)

Table 3-10 offers an overview of the available foundation types and their suitability for the above
conditions, based on the work presented by (Starling and Scott, 2009) for Carbon Trust.

- Green signifies, suitable.


- Orange signifies, Possible.
- Red signifies, unsuitable.

43
Table 3-10 Suitability of different foundation types

Foundation Hard Seabed Uneven Seabed Sloping sea bed


Type Composition Bathymetry

Driven
Monopile

Drilled
Monopile

Driven Pin-
Pile
Drilled Pin-
Pile
Suction Pin-
Pile

Gravity Base

The SR2000 Deployment at EMEC’s Fall of Warness test site used modular gravity anchors, due to
compatibility with their own vessels and the incompatibility of sea bed composition for driven piles
(All-energy.co.uk, 2018). However, due to the bathymetry shown in Figure 3-14 it is likely that
several moorings would be located on sloping sea bed, which renders gravity mooring incompatible.
In addition to this the Old Red Sandstone present at the site location does not permit suction or
driven piles to be installed (Figure 3-16). Therefore, drilled pin piles remain the only option, however
(Starling, Scott and Parkinson, 2009) state this would be difficult for the specific Seabed composition.

Figure 3-16: (Starling, Scott and Parkinson, 2009)

44
Figure 3-17: ultimate capacity od Pin Piles as anchors

Figure 3-17 shows that a pile length of 20 meters would provide sufficient tensile load bearing
capacity assuming a FOS of 2 grouting pressure of 0.7 MNm-2(Bruce and Nicholson, 1998), buckling
of the pile is unlikely to be an issue considering the hard rock in which the piles are to be driven.

For the purpose of clarity, this report has considered drilled pin pile anchors of 20 meters in length
for each of the anchoring points. ‘Drilling of the sockets can be conducted from the surface through
a conductor or using subsea drilling rigs’ (Starling, Scott and Parkinson, 2009)

45
3.4 Electrical Connections
The success of the project and its continuation past this preliminary study is highly dependent on the
availability of grid capacity and the ability to successfully connect the device/devices to the grid. For
small scale project such as those suggested in this study, existing grid networks should be used as
minor works will render them suitable. Making use of existing 11KV and 33KV networks reduces the
construction required but limits the location and capacity. The project is designed such that
progressive increases may be made in the number of devices installed at the site. It is therefore
crucial to consider this when planning connectivity.

The electrical connections designed in latter stages of the project will adhere closely to all suggested
DNV standards, as well as any introduced BSI standards as deemed likely by the MeyGen Lessons
learned project report (MeyGen, 2017). Relevant standards for this section of the design are listed
below.

- DNV-OS-D201 Electrical Installation (DNV, 2003).


- DNV-RP-0360 Subsea Power Cables in Shallow Water (DNV, 2010).

3.4.1 Spatial Analysis


Connecting the device/devices to the network in a manner which minimises the need for
reinforcement and new infrastructure will reduce risk and cost in the project. Therefore an overview
of the current connectivity of the design area is included. Figure 3-18 provides an overview of the
connectivity for the north west of Scotland. Crucially, the can be seen that the north coast of
Scotland is served from a single 275KV line at Dounreay and a smaller 132KV line at Thurso.

Figure 3-18: Current Transmission Network for North West Scotland (National Grid, 2015).

46
Further information gathered from The Crown Estate shows the extent of the smaller 11KV networks
and importantly, current lease areas around the design location. This information is presented in
Error! Reference source not found.3-6 (Parsons Brinckerhoff, 2012).

Figure 3-19: Existing Local Distribution Network for North East Scotland, (Parsons Brinckerhoff, 2012).

The inclusion of the data provided by (Parsons Brinckerhoff, 2012) and (National Grid, 2015) allows
for the creation of a spatial assessment of the geographical locations suitable for connection for the
project. This is shown in Error! Reference source not found.3-7 and allows for a preliminary design
of the electrical connection system. It should be noted that areas of the Pentland Firth are known to
have deeper pockets and more versatile bathymetry (Martin-Short et al., 2015).

These areas should be avoided for cable routing. It is suggested that a further seabed study is
undertaken before the final cable route if chosen. Two locations have been selected as suitable for
landing, points A and B. Table 3-11 summarises the key variables of connecting at the two points.
Both lie within a 10km distance from the offshore site and can be seen to lie close to an 11KV circuit.
Point A is the closest location however, in order to land cables to that location they would have to
pass through lease areas for exiting projects; this can be mitigated in several ways:

1 Access granted to our cables for a charge that must be weighed up against re-routing the cable.
2 A co-sharing strategy may be achieved allowing our project and the current lease-holders to
share space and infrastructure.
3 Access is refused and the cables must be routed to point B.

Most leaseholders have planning consent for capacity well above the current installed capacity of
the sites shown in Error! Reference source not found. .Therefore, the possibility of co-sharing the
nearshore and onshore infrastructure of point A should be explored.

47
Figure 3-20: Spatial Analysis of Site Location and Available Network Connections.

Table 3-11: Landing sites A and B for Offshore Infrastructure.

Variable Distance from Site (Km) Closest 11KV Connection


Point A 7 John O’Groats
Point B 9 Huna

3.4.2 Planned Expansion


With the level of activity in the Pentland Firth region, as evident from the current lease locations in
Figure 3-20, there are plans in place for development and reinforcement of grid connections and
transmission networks (Parsons Brinckerhoff, 2012). These should be considered as the design
process is taken forward as currently all local substations are constrained. For the assessment of this
project’s electrical infrastructure two scenarios for cable landing at point A will be considered:

- Local network at point A is insufficient and new infrastructure is required.


- Local network at point A is reinforced alongside project allowing for minimal new infrastructure.

48
3.5 Offshore Infrastructure
3.5.1 Cables
The DNV standard for subsea cables in shallow water should be adhered to, focusing on the required
areas and lengths of lines and cables, as suggested by Figure 3-21. This aspect was covered in
Section 3.4.1, Spatial Analysis. Conclusions from the 2017 publication from MeyGen also suggest
that a new BSI technical standard may come into effect in the near future and advise that this be
adhered to closely as it will provide further detail (MeyGen, 2017).

Figure 3-21: Methodology for Cable Route Selection as suggested by (DNV, 2016). This preliminary stage is highlighted by
the red box.

The design of the cable and cable layout should be enhanced from this early preliminary stage of the
project through analysis. Care should be taken to produce a solution/solutions that are not
detrimental to the project at any point in its lifecycle (DNV, 2016). Past this preliminary stage an
iterative design process that considers the following should be undertaken:

- Site conditions (Operations & Maintenance).


- Technical design choices (advancements in technology).
- Supply of material.
- Ease of construction (access, multi-operations).
- Changes in scope (upscaling capacity).
- Commercial factors (flexibility of prices for material/services).
- Opportunities for financing (Construction and operational flexibility).

If the system is designed to consider the above variables for all stages including, roll-out, transport,
installation, operations and decommissioning risks should be minimised. Beyond the
recommendations made by the DNV standard, the Lessons Learned report produces by MeyGen
provides further suggestions outlined below:

49
- Increased inspections on cables should be planned for as no testing methodology was found to
accurately predict their behavior.
- Design of cable laying should be undertaken with more detail at an earlier stage in order to
reduce costs.
- Specially sized cable drums may have a higher upfront cost but can save money as they reduce
instillation costs (Important for this project due to the intensity of the environmental conditions
in the Pentland Firth).
- Use of cable armoring instead of increasing cable density was desirable to reduce cost.
- Cable junction box design should be undertaken at an earlier stage to reduce weight and cost.
- Remote Operated Vehicles are desirable for overseeing cable installation.

Table 3-12 provides an overview of the cable parameters required to reach landing sites A and B
along with high level quantities.

Table 3-12: Offshore cable parameters for landing site A and B.

Landing A - Quantity Landing B – Quantity


Parameter
Medium Voltage AC (MVAC)
100 100
export system overview
Maximum number of single
1 1
core cables
MVAC route length (km) 7 9
Cable weight (kg/m) 40 40

3.5.1.1 Cable Design


The design of the cable system should be able to cope with the addition of devices to the site. As the
design location remains less than 10km from the landing site, it is suggested that an offshore
substation is not considered at this stage. However, two options for cable design may be discussed.

1 Each device has its own medium voltage single core cable that runs to the landing site.
2 Each device has its own single core cable that joins before land at a subsea connection point
where a larger high voltage cable connects the site to the onshore grid network.

The design choice is highly dependent on the planned expansion discussed in Section 3.4.2. If plans
to increase the capacity near the landing site from 11KV to 33KV are implemented option 1 is highly
recommended as this will greatly decrease the losses that occur in low voltage lines and allow for
ease of expansion of the site. An overview of the required infrastructure for scenario 1 and 2 will be
given in Section 3.7 It is further recommended that the offshore cable be sufficient to allow for
additional devices’ power take-off, as this will reduce costs if the project is expanded (MeyGen,
2017).

50
3.5.1.2 Cable Laying
The installation and layout of the cables should be after extensive surveying of the areas concerned.
Section 2.2.4 has already outlined risks such as unexploded ordnance’s (UXO’s), shipwrecks and
similar boundaries. Specific strategies or mitigating layouts should be outlined in further studies in
order to reduce the risks imposed by these boundaries. Strategies for dealing with UXO’s may be as
follows:

- If possible, avoid contact (minimum of 100km).


- Remove obstacles.

In both of these scenarios and the instillation of the cables, sufficient strategies should be
implemented as part of O&M. this may include:

- Human Divers.
- Remote Operated Vehicles (ROV).
- Specialised surface equipment.

3.5.1.3 Subsea Connection


Any connectors designed for the project should be in accordance with DNVGL-ST-F301 standards
(DNV, 2016). Within these are contained further codes that should be adhered to if risks are to be
minimised. Several connection types are available and are summarised in Table 3-13. Though a
variety of connectors are shown, detailed discussion should be held with sector experts in order to
optimise the design.

Table 3-13: Available Subsea Connections.

Name Company Description Reference


SpecTRON 10 Siemens 11KV Subsea connector (Colby, 2012)
Subsea junction MacArtney 11KV Subsea junction box (MacArtney’,
box no date)

The selection of subsea connector may be further informed by the merits of each connector
attachment method (Pye, 2014). It is suggested that a pressure-balanced compression gland is used
for deeper waters and long lifetime applications. These devices should be further explored in further
studies.

51
3.5.2 Risks
Several risks have been identified connected to the design and lay of subsea cables. Whilst this
preliminary stage of the project cannot assume to capture the total range of risks that will occur,
Table 3-14 provides an overview of scenarios that must be mitigated.

Table 3-14: Risk Assessment for Offshore Electrical Infrastructures

Category Risk Description Severity Mitigation


Subsea Excessive Considerable marine Break of connection External plating
Connectors corrosion. growth and between device and or coating.
corrosion. onshore network.
Mishandling Improper Breakage of equipment Detailed
in Instalment. instalment of leading to loss of surveying of
connectors. transmission. installation
process.
Cables Excessive Considerable marine Break of connection External plating
corrosion. growth and between device and or coating.
corrosion. onshore network.
Proximity to Cable drift/lay too Could result in the Detailed
UXO + close to undesirable complete destruction of surveying.
objects subsea transmission Effective cable
laying techniques.

52
3.6 Onshore
The required onshore infrastructure for projects of this size is well documented. Figure 3-22 provides
an overview of the entire electrical topology (Parsons Brinckerhoff, 2012). Similarities can be found
to the suggested onshore electrical system used for the Dogger Bank windfarm

Figure 3-22: Typical Requirement for Onshore Infrastructure.

Further to the infrastructure suggested in Figure 3-22 are several additional systems that may need
installing depending on the current compatibility at the local network connection station; this
includes:

- Service corridors such as telecommunication.


- Local electrical connection for works.

This preliminary scoping study will not present a detailed design for the onshore infrastructure.
However, a basic overview will be suggested based on similar projects along with the risks
associated. Furthermore, it has been suggested that reinforcement of local networks may lead to
higher voltage lines closer to the landing site of the project, with this in mind two onshore power
transmission line scenarios are outlined.

53
3.6.1 Landing Site
Offshore export cables will come onshore in a manor depending on the environmental and
geological restraints of the location. Locations A and B suggested in Section 3.4.1 have been chosen
as suitable landing sites for the offshore connections of the project. Moreover, it has been discussed
that two methods of cable layout are applicable to the project allowing for a range of power take-off
capacities. In both of these cases the landing site infrastructure is similar and may be covered in one
design.

Section 2.2.4 offered an overview of the bathymetry of the design locations showing that a sloping
gradient is present. This is further agreed upon by (Scott et al., 2009) who suggest that the cliff
structures surrounding the areas in the Pentland Firth can cause issue for cable landing. The
suggested solution to this is given by (Taaffe and Manager, 2016) in a report of the MeyGen project.
The recommended infrastructure for cable landing in this project is directional drilling.

3.6.1.1.1 Directional Drilling


This Solution has been sown to have a high accuracy when used by MeyGen (MeyGen, 2017). The
success of this solution in the MeyGen project makes it a likely candidate for use in this project.

3.6.2 Cable Transition Pit


Offshore export cables will be connected to the onshore transmission cables via a specifically
designed joint pit. This must be located on dry land and buried to specific depth depending on the
geology.

Typical dimensions for the transition are 1.5m depth, 4m width and 12m length (ForeWIND, 2014).
However the dimensions stated are taken from a large scale wind site and should be scaled
appropriately for the systems suggested for this project.

Once installed, usual backfilling processes should be used and the top surface reinstalled to its
original condition. This should allow for the surface area to return to its original usage, minimising
the environmental impact of the transition pit.

3.6.3 Developer Plant Compound


The compound is another system that is the responsibility of the developer and should therefore be
the focus of detailed design past this preliminary stage. Usual components of the compound can
include the following (Parsons Brinckerhoff, 2012):

- Substation.
- Power control/conditioning.
- Onshore generating station.
- Parking.
- Staff welfare facilities.
- Fencing.

A general size of the compound is provided by (ForeWIND, 2014) who suggest a minimum of 0.25ha.
Importantly it should be noted that the compound and transition pit are located on the same
envelope.

54
3.6.4 Onshore Power Transmission Line
Whilst it is accepted that designs may change past this preliminary scoping study, it has been
decided that the onshore cables for this project will be buried and not carried overhead. This
decision has been made as an attempt to reduce permanent environmental and ecological impacts.

The exclusion of overhead cables is a decisive move within this project. As further investigations are
undertaken and further information gathered regarding the expansion of local networks its
suitability should be explored.

Each cable is to be buried in trenches similar to those specified by (Planning and Forms, 2014). Each
cable will require burial in a trench at a depth of approximately 1.2m and backfilled with material
with adequate thermal properties to achieve efficient cable performance. Materials commonly used
for this process are ether native soil or a stabilised material such as Cement Bound Sand (CBS).

Table 3-15 provides an overview of the onshore cable parameters with and without local network
reinforcement.

Table 3-15: Onshore MVAC and HVDC cable parameters for reinforced and un-reinforced local networks.

Without local Reinforcement With local Reinforcement


Parameter
Medium Voltage AC (MVAC) Single three-phase circuit Single three-phase circuit
export system overview comprising of a single core. comprising of a single core.
HVDC system overview Single HVDC three-phase
Single HVDC three-phase circuit.
circuit.
Maximum number of single
1 1
core cables MVAC
Maximum number of triple
3 3
core cables HVDC
MVAC route length (km) 2.0 2.0
HVDC route length (km) 30 0
MVAC Cable weight (kg/m) 20 20
HVDC Cable weight (kg/m) 40 40

3.6.4.1 Onshore Cables Without Reinforced Local Network


In this case the required infrastructure is much larger. Medium voltage AC cables will run from the
landing site compound to a local substation. Here further infrastructure will be required to increase
the voltage and transform from AC to DC allowing for connection to the 132KV Thurso station.

3.6.4.2 Onshore Cables With Reinforced Local Network


If suggested improvements are made to the local network nearby the landing site the added
infrastructure and HCDC lines will not be required allowing for a MVAC line to run the short distance
to the substation and some power electronics to connect to the network.

55
3.6.5 Risk
The risks associated with the onshore transmission lines are shown in Table 3-14. Further risk
assessments should be included past this preliminary design stage.

Table 3-16: Onshore System Risks

Category Risk Description Severity Mitigation


Onshore Cable damage Damage to cable Loss of power Risk
Transmission during whilst being transmission and assessments
Lines trenching. trenched from potential health and and procedure
landing to final safety risk. governing
network connection. trenching
process.
Cable damage Damage to cables Loss of power Warning tape
in future caused by transmission and and protective
excavation. redevelopment of potential health and tiles to be
local area. safety risk. installed in
original
trenching
process.

56
3.7 System Diagrams
As previously discussed, a detailed view of the electrical system will not be given in this study.
However, given the offshore and onshore systems previously discussed suggestions for the overall
systems can be estimated. This will allow for a high level identification of the costs involved with
both the installation of the system and its maintenance.

3.7.1 System Diagram 2MW

Figure 3-23: Single Device (2MW) Electrical System Overview

57
3.7.2 System Diagram 6MW

Figure 3-24: Multiple Devices (6MW) Electrical System Overview

58
3.8 Electrical Connections Implications on Cost
A summary of the discussed offshore and onshore electrical systems is provided in Table 3-17 and
Table 3-18 respectively. A preliminary cost estimate has also been provided in order to ascertain
whether the suggested engineering solutions for the project are economically feasible.

Data contained in an electrical transmission costing study produced by (Parsons Brinkerhoff, 2012)
was used as a guide for the trenched HVDC lines connecting the developer’s compound to the
network in the eventuality of the local network having not undergone any reinforcement by the time
of project development.

In the case of the subsea cables a cost breakdown for 75km of infrastructure is presented. These
costs were then scaled by adjusting to the 7km and 9km systems in this project. An additional 10%
cost for each variable was included as a crude method of accounting for the loss of cost depreciation
due the scale of the presented 75km system.

The medium voltage AC cables for onshore application were estimated in a similar manor to the
subsea cables. A 3km route presented by (Parsons Brinkerhoff, 2012) was extrapolated down to
match the 2km requirements for this project. In this case, due to the relative closeness in lengths of
systems, no correction for scaled costs was applied.

The costs for the subsea cable and infrastructure, and the onshore medium voltage AC has been
extrapolated from examples also provided by (Parsons Brinkerhoff, 2012).

Table 3-17: Offshore Electrical Connections Summary and Preliminary Costing.

Landing A Landing B
Parameter
3-phase
Cable Type 3-phase MVAC
MVAC
Cable Length (km) 7 9
Cable studies and assessments (m£) 0.02 0.02
Cable landing costs and materials (m£) 2.1 1.9
Fixed Build Costs
Cable mobilisation costs (m£) 0.4 0.4
Converter project launch (m£) 4 3.5
Cable contractor (m£) 0.8 0.7
Variable Build Costs Cables studies and assessments (m£) 11 9.4
Cable material and instillation (m£) 1.4 1.2
Total (m£) 19.7 17
Table 3-18: Onshore Electrical Connections Summary and Preliminary Costing.

Without local With local


Reinforcement Reinforcement
Parameter
Cable Type Trenched Trenched Trenched MVAC
MVAC Cable HVDC Cable
Cable Length (km) 2 30 2
Cable Cost (million £/km) N/A 24.1 N/A
Fixed Build Transmission compound (m£) 4.9 N/A 4.9
Costs Cable termination and testing (m£) 1.1 N/A 1.1
Tunnelling (m£) 14 N/A 14
Total (m£) 31.5 6.4

59
3.9 Electrical Connections Summary
Whilst there remains an amount of uncertainty around the electrical connections for the project,
several summary points may be drawn from this preliminary study.

- A general length of offshore cable has been specified based on the ability to land in two separate
locations, both under 10km.
- Further site specific assessments must be made to ensure offshore cable position due to the
bathymetry of the Pentland Firth.
- Offshore Cables must be sufficiently sized to allow for the addition of several devices.
- The success of directional drilling has been proven for similar projects and should be a strong
consideration for landing the cables.
- Based on the preliminary values for the two landing sites estimated offshore electrical
infrastructure coats are between 17 and 20 million pounds.
- Infrastructure of the onshore electrical requirements has been shown to dependant on future
reinforcements at local substations as all current stations are limited.
- A scenario based on the eventuality that local reinforcements do not take place produces an
estimated cost for onshore electrical connections of over 30 million pounds.
- If the project is able to build alongside the reinforcement of local infrastructure, the onshore
electrical connection cost comes in at 6.5 million pounds.

60
4 Installation
Tidal turbine installation presents a challenge for offshore marine construction because of the need
to work in water that is deep, fast flowing and susceptible to adverse conditions. Currents impose
significant drag loads on the vessels that are used for installation and may induce vibration in the
entire structure.(American Bureau of Shipping, 2013)

Success of the project will rely on planning of installation time, installation vessels and installation
weather windows.

Installation for simplicity is broken into the following different segments:

4.1 Design
 The SR2000 significantly reduces the installation cost compared to ground mounted
systems, by using smaller vessels which can be locally accessible.
 Since the floating device incorporates retractable legs that have rotors mounted on it, it will
be easy to tow the turbine to the nearest port and this reduces the draft of the turbine
(Barnaby et al., 2018).
 The turbine functions in the highest energy profile region as tidal range is highest expected
in the surface water.

4.2 Preparation
 Although the foundation requirements are lessened for a floating turbine. Anchoring and
mooring of the tidal turbine vessel involves weather challenges. A 3 pointed taut mooring
system is proposed for the plan and this will involve drilling of pin piles of 20 m into sea bed
to support the 509-ton tidal turbine.
 Seabed conditions have a major impact on the installation of the tidal turbine. The mooring
and anchoring of installation device depends on the sea bed surface.
 Pentland Firth comprises of a very uneven seabed conditions, boulders and uneven slopes
(Maygen, 2012) and drilling of 4 pin pile at a depth of 20m under severe conditions will
propose a problem to drilling operation. The drilling operation takes a significant amount
labour, time, vessel.
 Laying of cables from site to the substation is another time-consuming task of the
installation and will involve the aid of ROV’s and deep divers and experienced crew.

4.3 Method
Number of devices to be installed will determine the time and crew required for installation.

For our scoping we simulate the planning process for 2 different climates, at different times of the
year, summer and winter.

Also, simulation is carried out for same time of the year for the installation of a single tidal turbine
and an array comprising of 3 such tidal turbines of 2MW each.

61
4.4 Location
Depth of the water, port location and substation location are important aspect of installation
planning (S, K and G.C, 2008).

The vessels would make many travels from substation to the site for cable installation. Therefore, a
suitable substation location is assumed for the scoping. The nearest port accessible for docking and
operation is Scrabster port.

Figure 4-1 Location of the port, site and substation (orange represents site, Green the port and blue the substation), while
this does not represent the chosen site, the timings would not change significantly due to the proximity.

4.5 Planning
 This involves what time of the year the installation is set to start and the tasks to be carried
out for an efficient installation.
 The kind of vessel utilized for various tasks of installation, the crew required, apparatus for
installation and safety measures are all part of the planning.
 The Stimulation of installation planning is made and analysed using a planning and
optimization tool developed by Mojo maritime called Mermaid.

The installation plan will consider numerous factors for worst case scenario such as tidal speed, flow
conditions, Water depth etc. The following criteria are accounted for:

62
Table 4-1 Planning Factors informed from Section 2.

Conditions Comment
Tidal speed 3.5-4.5m/s
Flow Conditions Turbulent
Variation across the site Not significantly seen
Depth (astronomical) 30-40m
Tidal range Highest tidal range as it is floating tidal turbine
Survivability/transportation extremes 12-14m significant wave height recorded over 50
years.
Wave and wind window for installation Constrained to seasons.
Type of sea bed Sea bed needs preparation (Hard rock).

4.6 Mermaid
Mermaid is used for the following purposes:

 Develop a project baseline based on initially selected vessels, ports, methods and processes.
 Simulate the operation against hind cast weather data.
 Evaluate task durations, identify the weather risk drivers.
 Optimise the project – changeout vessels, ports, methods or processes.
 Re-simulate and compare with the baseline.
 Develop an optimised project with the lowest possible weather risk profile.
 Make procurement and project decisions based on solid analysis – reduce the risk of project
cost over runs.

4.6.1 Challenges to be addressed


 Due to strong currents and the use of smaller vessels, dynamic positions of the vessel may
be real challenge (Scotrenewables.com, 2018)
 Installing cables and handling of cables might be challenging task as it will mean calculating
the radius of curvature for the specific cable.
 Pentland firth site is a significantly more difficult and challenging compared to neighbouring
sites. The exposure of the site is a significant factor. YouTube. (2018), (Barnaby et al., 2018).

63
4.6.2 Power cable
Consideration must also be given to the power cable. In fixed foundations this is relatively simple as
the cable can be connected to the base and there tends to be no relative movement, complications
will only arise if the cable needs to be disturbed during any maintenance or repair of the device. The
main issue is scour wear and protection of the cable (Carbontrust.com, 2009).

4.6.3 Pin pile installation


Pin piles are commonly used for more permanent moorings, but they tend to cost more to install.
Piles can be either drilled, driven or use suction to be put in place. Drilling is used in the harder
bottom conditions, with piling and suction techniques being used in the softer sediments. The
advantage of suction techniques over piling is that with piling the life of the pile can be reduced if
too much energy is expended in the piling operation. Once a pile has been placed a mooring
structure is then inserted and can be cemented in place or ball and roller type fittings can be used
(Carbontrust.com, 2009).

4.6.4 Mooring operation


The major obstacle in the design of a mid-water device is the optimisation of the buoyancy of the
device to counteract the weight of the turbine and provide adequate vertical component in the
mooring line tension. This may restrict the weight and ultimately the output of the device. Additional
advantages may be the speed with which the devices could be deployed as the moorings could be
rapidly deployed using suitable anchor handling vessels. While these vessels are expensive, if the
installation could be carried out in a short period, this method may become cost effective
(Carbontrust.com, 2009).

4.6.5 Vessels
There are various vessels in the market that serve a lot of purposes, but for our model we choose to
minimize the cost of installation by using Vessels that could be used for multiple tasks. Vessel
selection is made based on environmental conditions, crew members, accessibility to port,
availability of the vessel, cost per day of the vessel etc.

The 3 main vessels apart from smaller boats and ribs will include the following:

 Renewable service vessel Damen RSV 3315, Voe vanguard: The vessel is capable of anchor
handling, towing and survey Figure 4-2.

Figure 4-2 Multipurpose Vessel (damen.com, (2018).

64
 Various oil and gas industries cater drilling services and we propose to use drillship for drilling of
pin piles for anchoring and mooring of the tidal turbine. The proposed Drillship is capable of
drilling pin piles and is available for the time we intend to install. The size and length determine
the type of Drillship for the task.

- The third vessel we intend to use is the cable laying vessel that facilitates the installation of
export power cable from the offshore site to the nearest substation (the substation is assumed
in this case since establishing a substation requires License). Cable laying vessels have certain
specifications for winches and they are the only vessel that could be used for laying of cables for
long distances. Cable laying is a time consuming and an expensive affair.

4.6.6 Installation of Turbine Task

1 Installation of export cable from proposed area to the Power Conversion facility onshore.
2 Drilling of 4 pin pile into the sea bed
3 Installation of floating turbine itself along with the jumper cables.
4 Jacking up and Towing of tidal turbine from port to offshore location
5 Anchoring and mooring of the tidal turbine using the multi cat vessel.

The step by step installation operations are illustrated by taking example of previous tidal turbine
installations.

Figure 4-3 Various stages of SR2000 installation (YouTube, 2018).

65
4.6.7 Mermaid Results
Results obtained for 3 different simulations conducted for installation during summer months (May
22nd and July 7th) and winter (Jan 12th) are compared.

4.6.7.1 Winter scenario data


 The starting day is chosen as 14th Jan and drilling operation is planned for 3 days with 2 days
for mobilizing and demobilizing at the port.
 The second task is the cable handling operation handled by both multipurpose vessel and
cable laying vessel and it runs for a period of 7 days.
 The final process of installation is the installation of turbine which involves mooring of the
tidal turbine at the 4 pin piles drilled points. This is carried out with the multipurpose vessel
Voe vanguard. The vessel consists of 2 cranes which facilitates mooring of the tidal turbine.

Table 4-2 Summary of minimum and maximum days required for tasks (winter).

P75 -
Unweathered Min Max Mean P10 P25 P75 P90 Max - Min
Task P25
(Days) (Days) Days (Days) (Days) (Days) (Days) (Days) (Days)
(Days)
Overall 33.08 33.95 97.48 71.54 34.83 45.77 92.18 97.04 46.42 63.53
mobilizing
1.00 1.00 6.66 3.32 1.00 1.00 4.97 6.55 3.97 5.66
port
drilling
3.42 10.55 72.92 43.58 11.98 25.59 67.62 72.59 42.03 62.36
operation
demobilizing
1.42 1.42 3.41 1.95 1.42 1.42 2.54 3.36 1.12 1.99
port
mobilizing 1.00 1.00 7.53 3.02 1.00 1.00 4.82 7.33 3.82 6.53
cable laying
7.42 8.79 81.02 52.46 10.55 27.32 72.16 80.52 44.83 72.23
opeation
demobilizing 1.42 1.42 2.15 1.49 1.42 1.42 1.42 2.07 0.00 0.73
mobilizing 0.04 0.04 1.68 0.46 0.04 0.04 1.00 1.64 0.96 1.64
Device
0.61 0.61 5.41 1.94 0.61 0.61 2.72 5.32 2.10 4.79
Installation
demobilize 0.39 0.39 7.69 1.27 0.39 0.39 0.84 7.04 0.45 7.30

66
Figure 4-4 Tasks 1-3 Mermaid installation time (winter).

Figure 4-5 Duration Exceedance for winter installation (winter).

67
Figure 4-6 Cost exceedance chart for winter installation show a fair probability of increase in the cost (Appendix B data).

Figure 4-7 Progress burn up chart for winter installation.

68
Table 4-3 Summary of minimum and maximum days required for tasks.(summer)

P75 - Max -
Unweathered Min Max Mean P10 P25
Name P75 (Days) P90 (Days) P25 Min
(Days) (Days) (Days) (Days) (Days) (Days)
(Days) (Days)
Overall 24.08 24.08 34.49 27.81 24.08 24.08 32.28 34.39 8.19 10.41

mobilizing port 1.00 1.00 1.82 1.13 1.00 1.00 1.12 1.79 0.12 0.82
drilling operation 3.00 3.00 10.66 6.24 3.08 3.83 7.92 10.65 4.09 7.66
demobilizing port 1.00 1.00 3.35 1.44 1.00 1.00 1.63 3.30 0.63 2.35

mobilizing 1.00 1.00 3.53 1.25 1.00 1.00 1.00 3.28 0.00 2.53
cable laying
7.42 7.42 16.41 10.76 7.42 7.42 16.41 16.41 8.99 8.99
opeation
demobilizing 1.42 1.42 4.04 1.84 1.42 1.42 1.83 3.94 0.41 2.63
mobilizing 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.00 0.00
Device
0.61 0.61 2.45 0.80 0.61 0.61 0.61 2.26 0.00 1.83
Instalation
demobilize 0.39 0.39 7.14 3.09 0.39 0.39 7.14 7.14 6.75 6.75

Figure 4-8 Cost duration chart for summer installation(July).

69
Figure 4-9 Duration exceedance chart for summer installation(July) .

Figure 4-10 Cost exceedance chart for summer installation(July)

70
Figure 4-11 Progress burn up chat for summer installation(July)

4.7 Summary
Based on the above simulation it is seen that summer installation offers benefits due to better
weather conditions. Even though winter installation is cheaper than summer installation, there are
tremendous risks involved with it. Delays in tasks could have knock on effects which in turn may
increase the installation costs.

Keeping the installation cost low will mean lower levelised cost of energy. This can be achieved by
the smart use of the available Vessels and proper study of weather phenomena (met ocean data)

A similar model consisting of 3 tidal devices will be cost effective based on the data obtained.
Further research on developing own vessel might be one way of looking forward to continuous
maintenance and operation. This will also act as permanent asset for all the future.

71
5 Operation and Maintenance (O&M)

A key factor impacting the competitiveness of Tidal Stream Devices (TSDs) and other renewable
technologies is the cost of maintenance. In the past, corrective maintenance strategies have been
implemented for O&M due to lack of reliability data and lack of experience within the industry.
Corrective maintenance strategies are known to be expensive and an inefficient method of
increasing device availability. Therefore, predictive maintenance strategies are now being developed
that anticipates major failures and their effect on power production. However, there remains a
trade-off between minimising the cost of O&M and maximising device availability. This trade-off is
illustrated in Figure 5-1, which highlights the theoretical operating point where cost is minimised,
and a corresponding level of availability is reached. However, in practise it is extremely challenging
to operate at this point due to the complexity of the interactions between components and sub-
systems within energy generation devices, compounded by factors such as weather conditions, lead
times and vessel limitations (Rinaldi et al., 2016).

Figure 5-1: illustration of balance between cost and lost revenue (The Crown Estate, 2013)

5.1 Maintenance Strategies


Corrective maintenance: In this strategy, maintenance is only performed reactively when a failure
occurs. This method gives a low-cost approach to O&M, however is likely to incur high amounts of
device downtime due to the unplanned nature of the repairs. This in turn would reduce the revenue
generation capability of the tidal array, impacting significantly on economic feasibility.

Preventative maintenance: A cyclic maintenance strategy based on each component manufacturer’s


recommendations. Maintenance is based on recommendations from component manufacturers,
based on metrics such as time intervals between service and prescribed procedures of maintenance.
Time intervals are usually in monthly/yearly periods or stated as a number of duty cycles for rotating
equipment. This strategy will ensure the device remains compliant with HSE guidance at a
component level, but it does not consider the components interactive with other systems or the
environment the component is operating in. This approach may also lead to high OPEX (Operational

72
Expenditure) costs, as an accumulation of each component’s maintenance recommendation would
be considered individually rather than a strategy developed at device level, therefore there is likely
to be inefficiency, through overlapping in maintenance tasks etc.

There are various types of maintenance tasks depending on the component. These can include and
are not limited to:

 Non-obtrusive monitoring: vibration, temperature, fluid flow conditions and noise.


 Oil and grease sampling.
 Wear measurements.
 Visual inspections: corrosion, biofouling, cracking, leakage etc.
 Torque and fastenings checks.

Condition based maintenance: Supervisory Control and Data Acquisition (SCADA) systems can be
used to provide detailed information on the performance of a device and provide non-obtrusive
monitoring, using vibration and temperature monitoring for example. However, these systems are
expensive to install and maintain, particularly in a harshly corrosive marine environment. Data from
SCADA systems can be used to perform pre-emptive maintenance on component failures that have
been detected at an early stage, before major failure occurs causing more damage and downtime.

Predictive maintenance: A reliability based approach to O&M that predicts the failure rate of
components based on data from previous tidal arrays and from analogous systems that are adjusted
to the marine environment, for example offshore wind turbines. This approach is becoming the most
favoured approach to O&M as it utilises previous experiences to produce a predictive maintenance
strategy without additional costs.

Probabilistic evaluation techniques have been developed specifically for use in the development of
O&M strategies for marine renewable energy systems. They are based on Markov chains and Monte
Carlo analysis (Rinaldi et al, 2017; Hofmann, 2011; Alexander, 2003) to simulate possible failure and
component life outcomes based on known reliability data, as well as a number of other inputs
including (Rinaldi et al., 2016).

 Metocean data: wind, wave and tidal data at a site.


 Port data: location, access limitations, logistics.
 Vessel information: number, speed, maximum operating conditions, response time, number
of crew.
 Number and power rating of the devices.
 Procurement and repair times.
 Spare parts stock.
 Redundancies and criticalities of device components: to enable evaluation of consequences
of a failure on the device’s operation and power production.
 Dependencies of components and subsystems; common causes and/or cascading failures

The outcomes of these simulations then aid the decision-making process to optimise an O&M
strategy for the device.

73
This type of predictive maintenance is what is focused on in this report. TSDs (Tidal Stream Devices)
are relatively new type of marine energy converter. As a result, there is very little operational and in-
service reliability data, making it difficult to critically evaluate the technology with much accuracy.
Due to the immaturity of the industry there has been little design convergence, so there are still a
wide range of TSD solutions being developed in the market, such as floating tethered, submerged
tethered, and seabed mounted turbines. Developing reliability models and tailored predictive
maintenance strategies could reduce the long-term risks and costs of the technology; making TSD
projects more economically viable and more attractive to investors (Delorm, 2014).

The follow section details the factors and challenges that need to be considered to develop a full
O&M strategy for a Scotrenewables SR2000 tidal array situated in the Pentland Firth. The software
package Mermaid is then used to simulate operations and provide data to aid the decision-making
process.

5.2 SR2000 Operation and Maintenance Strategy


The final operation and maintenance strategy should be compiled at a later phase of the project,
when more detailed information is available as a result of real world data acquisition and
component level device details should be utilised. However, the following considerations should lead
to the development of a cost-effective and safe maintenance strategy throughout the lifetime of the
project until decommissioning. Alongside the predictive maintenance strategy, the O&M strategy
should also include plans for emergency procedures, including an active safety management plan
and grid disconnection procedures, which are not included in this report.

5.2.1 The SR2000 Device


One of the main design drivers for the SR2000 device was reducing O&M costs. Its floating body and
retractable turbine legs make transportation simple as the device can be towed; negating the need
for expensive heavy lifting vessels such as jack-ups or dynamic positioning (DP) vessels. The majority
of the device’s electrical componentry and auxiliary systems are located in the device’s hull, allowing
easy access for maintenance, with the exception of the drive-train sub-assembly. All cable
penetrations, structural connections and leg actuation systems are readily inspectable and
serviceable above the waterline.

The majority of the SR2000’s system modules are installed in two removable containers in the hull.
This gives the potential to swap out faulty modules and increase availability, while the modules are
repaired onshore. This modular design also allows for easy maintenance access. Data from the wind
industry shows that 80% of maintenance interventions arise from failures of components weighing
less than 25kg. This means that the majority of failures can be fixed onsite without the need for
heavy lifting vessels (Scotrenewables, 2012).

An SR2000 has been operational at The European Marine Energy Centre (EMEC), located off Orkney,
Scotland, since October 2016: generating a cumulative total of 1.5GWh (EMEC, 2018). Over this
period Scotrenewables have demonstrated a number of successful O&M procedures. Multiple RHIB
(Rigid Hull Inflatable Boat) based interventions at vessel costs less than £1000 per intervention have
been demonstrated at the EMEC site. As well as this, two turbine recoveries and three turbine
installations were completed in 2017, with a total vessel cost of £25,000. This total is approximately

74
equivalent to 4 hours of charter cost for a heavy lifting vessel; illustrating the cost-effectiveness of
maintenance strategies available due to the device’s design (Scotrenewables, n.d.). Along with this,
hull accessibility of the SR2000 has been demonstrated in up to 2m significant wave height and was
therefore accessible over 90% of the year at EMEC with no weather delay (Scotrenewables, n.d.).

5.2.2 SR200 Subassemblies


There is still considerable variation in TSD systems and components and owing to commercial
sensitivity, no information is publicly available on the detailed system assemblies in the SR2000.
Therefore, a generic assembly is considered in this section, which includes the drive train, rotor
blades, gearbox, generator, power electronics, electrical systems, grid connection, ancillary system,
control and management and support structures. Each of these will be explored in more detail
below.

Drive Train

The drive train includes the turbine and generator sub-systems. The turbine sub-system includes the
rotor blades, seals, shaft, bearings, and mechanical brakes. The generator sub-system includes the
generator unit, inverter system and circuit breakers. The rotor blades of the SR2000 are located on a
large hinged arm that extends at 90 degrees to the floating platform when in operation.

The SR2000 device has fixed pitch blades with passive yaw (Scotrenewables, 2018) and it is assumed
that the device is direct drive, so there is no gearbox. The generator will be asynchronous or
synchronous and will require a cooling system. This may be an air circulating fan or water pump,
both of which requiring additional components and a separate control system.

Power Electronics and Electrical Systems

As the SR2000 will not be directly connected the grid, a converter is required to convert the
generator AC frequency to the gird frequency (50Hz in the UK). This converter consists of two
inverters, one generator side and one grid side, and a DC link capacitor. This allows the TSD to
operate at variable speeds, increasing power production capabilities.

A step-up transformer is likely to be required to reduce transmission losses to shore. The SR2000
operates at 6kV / 11kV (Scotrenewables, 2018). This transformer will also need a cooling system, a
circuit breaker and an isolator switch for maintenance access and safety.

A low voltage DC uninterrupted electricity supply system is often required as some TSDs need to
draw power from the grid to start the turbine. It is also needed to supply electricity for the braking,
communication and lubrication systems in the event of unexpected disconnection from the grid. This
back-up electricity supply is often provided by a low voltage battery system, charged from the grid.

75
Grid Connection

The electricity generated by the SR2000 device will be transmitted to shore via seabed cables. A
flexible umbilical cable will be required to connect the floating device to the seabed cable. The
cabling will also include a fibre-optic cable for communication (see Section 5.3.4 for details).

Control and Management

A control system is required to monitor components and lubrication and cooling systems etc. the
control system should be able to remotely stop the device for safety, inspection or maintenance
reasons. SCADA is used to provide non-obtrusive inspection and future predictive maintenance,
through vibration, temperature and noise sensors.

Support Structures

The SR2000 device is comprised of a floating hull, to which the turbines are attached, so a mooring
system is required to tether it to the seabed

76
5.2.3 Reliability Block Diagram (RBD)
In order to accurately assess subsystem interaction in the SR2000, a Reliability Block Diagram (RBD)
must be constructed. Figure 5-2 below shows the assumed RBD of the SR2000 device, taken from
Delorm (2014). The RBD shows how the subsystems interact and if they are in parallel or series.

Figure 5-2: Reliability Block Diagram for SR2000 (Delorm, 2014).

From Figure 5-2 it is clear that the two rotor and generator systems act in parallel, meaning that if
one fails then the device can theoretically continue operating but at 50% capacity. However, in
practice is it not known if the device can withstand the uneven loading caused by this, and therefore
all the subsystems are considered in series. This means that if a fault occurs in a critical subsystem
then the device will stop generating.

77
5.3 Reliability Data
Due to the relative novelty of the TSD market and subsequent lack of available reliability data
specific to TSDs, reliability data for each subsystem detailed above can be taken from other marine
industries including offshore wind and wave energy converters. As such, it is acknowledged that
using this data will give rise to greater uncertainty when applied to TSDs, as it is not specific to tidal
environments and may include data from manned installations, where maintenance checks may be
more regularly carried out. These increased uncertainties are of significant importance to investors
and developers; therefore, it is important to develop reliability models that can be updated and
verified with TSD industry data as it becomes available.

For the purposes of the feasibility study reliability data is taken form Delorm (2014). This data is
from a variety of surrogate sources with similar assemblies, with an environmental correction factor
applied. The method applied in Delorm (2014) calculated the average number of failures per year for
the SR2000 device, assuming no preventative maintenance is carried out. It assumes that the
subsystems are individual blocks, which can lead to an error in the predicted total failure rate of as
much as 20% (Faraci, 2006). However, due to the lack of data from the TSD industry, this approach
has produced a conservative estimate of the failure rate that can be validated as more accurate data
becomes available.

The method does not include failures due to degradation over time and “Infant Mortality” failures
due to faults during manufacturing. This is a well-documented trend best represented by the
‘Bathtub Curve’ (Figure 5-3).

Figure 5-3: Failure rates ‘Bathtub Curve’ (Rinaldi, 2017).

The failure rate 𝜆 (failures/year) and the reliability survival function R(t) are used to compare the
probability of failure of the SR2000 subsystems. The reliability survival function R(t) is a probability
that the subsystem or component will survive after a specified time and the failure rate 𝜆 can be
expressed as (Thies et al., 2012):

𝑓(𝑡) 𝑓(𝑡)
𝜆(𝑡) = =
𝑅(𝑡) 1 − 𝐹(𝑡)

78
Where F(t) is the cumulative distribution function and f(t) is the probability distribution function. If a
constant failure distribution is assumed, i.e. operating in the random failures section of the Bathtub
Curve, then an exponential relationship is used to describe the reliability survival function R(t) as:

𝑅(𝑡) = 𝑒 −𝜆𝑡

5.4 Reliability Data Analysis


The following section details the methodology and results of calculations of the SR2000 subsystems
failure rates.

Reliability data and the RBD information from Delorm (2014) were used, as described above, to
model the SR2000 device. See Appendix A for all reliability data used. From this data 12 subsystems
were categorised for the device, shown in Table 5-1 below.

Table 5-1: SR2000 Subsystem categories and number

Subsystem Subsystem
Number
Hub 1
Blades 2
Generator 3
Main circuit breaker 4
Main shaft, bearing, 5
seals
Brake system 6
DC/DC converter 7
Transformer 8
Electrical components 9
Controls 10
Structure and mooring 11
Ancillary system 12

5.4.1 Correlating Reliability Data


This reliability data was correlated with data from Carrol et al. (2015) who predicted failure rate,
repair time and unscheduled O&M costs for offshore wind turbines. Carrol et al. (2015) splits the
failure rate data into three categories: major replacement (material cost>€10,000), major repair
(€1,000<material cost<€10,000), and minor repair (material cost<€1,000). To compare to Delorm
(2014), these failure rate categories were summed. The results can be seen in Figure 4 (data in
Appendix A). The data in Carrol et al. (2015) was then combined into the subsystems listed in Table
5-1 to allow for it to be applied to the SR2000 device and the Delorm (2014) reliability data.

As Delorm (2014) shows, wind turbine subsystem assembly and TSD subsystem assembly are similar.
However, using Carrol et al.’s (2015) surrogate data from the offshore wind turbine industry

79
introduces a significant level of uncertainty due to the inherently different environment the
subsystems are installed into. However, Figure 5- 4 shows that there is a strong correlation between
Delorm’s (2014) environmentally corrected data and the data from Carrol et al. (2015), with an
average difference between data sets of 5%. Therefore, data on mean time to repair (MTTR), repair
costs, and percentage likelihood of failure from Carrol et al. (2015) is used in the following O&M
analysis; adjusting the data into the SR2000 subsystems shown in Table 5-1. When more accurate
data becomes available from floating TSDs, this data can be updated to better represent the SR2000
and its environment.

Failure rates of sub-assemblies


1.2
Failure Rate (Failures/year)

0.8

0.6

0.4

0.2

0
1 2 3 4 5 6 7 8 9 10 11 12
Subsystem Number

Delorm (2014) Carrol et al. (2015)

Figure 5-4: Failure rate correlation between Delorm (2014) and Carrol et al. (2015).

5.4.2 Percentage Likelihood of Failure per Category


Figure 5-5 shows the percentage likelihood of failure from each repair category: major replacement,
major repair and minor repair. Considering the entirety of the device, 80% of the number of failures,
are due to minor failures, needing only minor repairs; in line with Section 5.2.1 which states that
80% of maintenance interventions are minor and arise from component failures weighing less than
25kg. Figure 5 also shows that the likelihood of a major replacement needing to be undertaken is
very small compared to the other categories.

80
Percentage Likelihood of Failure per year in each Subsystem
100
percentage likelihood of failure

90
80
70
60
50
40
30
20
10
0
1 2 3 4 5 6 7 8 9 10 11 12
Subsytem Number

Major Replacement Major Repair Minor Repair

Figure 5-5: Percentage likelihood of failure per category

5.4.3 Probability of failure over lifetime


Applying this data to the failure rate data for SR2000 over its lifetime shows that the blades and the
controls are the most likely subsystems to require minor repairs, requiring a predicted 17 minor
repairs over the 20-year life time of the device. This can be seen in Figure 5-6 below (data in
Appendix A). The most likely assembly to fail within the control subsystem is the SCADA, with a
failure rate of 0.75 failures/year (Delorm, 2014). This is due to the complexity of the assembly as it
contains many small electrical components. A failure of this assembly may lead to large repair times
onsite as troubleshooting will have to be carried out to find the root cause of the fault. Having the
necessary spare electrical components in stock will reduce this repair time.

The generator is the most likely subsystem to require a major replacement, with a likelihood of
needing to be replaced once in its 20-year lifetime. Therefore, contingency may need to be built into
the financial model to account for the cost of this replacement. The major repair failure rate over
lifetime for all other subsystems are below 1 and therefore are not predicted to need replacing over
the device lifetime of 20 years.

Figure 5-6 also shows that failures that require minor repairs consist of 80% of the failure types. It is
assumed that these can be repaired onsite with a small maintenance vessel. 19% of failures are
predicted to need major repairs. The method of repair depends on the subsystem, as due to the
design of the device some subsystems are not fully accessible onsite as they are below the waterline.
Table 5-2 shows which subsystems can receive a major repair on site and which need to be repaired
onshore. For minor repairs below the waterline it is assumed that divers can carry out the repair.
Major repairs to the drive train and generator systems will therefore be much more costly to repair
than the subsystems that can be repaired onsite. It is assumed all major replacements will take place
onshore. Some minor repairs will not be immediately critical to the functionality of the device, for

81
example, corrosion of the nacelle or floating structure. In this case, repair of these will be suspended
until the device is brought back in to harbour for a major repair. This way downtime and costs will be
reduced.

Probability of failure over 20-year lifetime


20
18
Failure Rate over Lifetime

16
14
12
10
8
6
4
2
0
1 2 3 4 5 6 7 8 9 10 11 12
Subsystem Number

major replacement major repair minor repair

Figure 5-6: Probability of failure over 20-year lifetime

Table 5-2: Subsystems accessible for major repair onsite or offsite.

Subsystem Subsystem Major Repair


Number Accessible
Onsite
Hub 1 No
Blades 2 No
Generator 3 No
Main circuit breaker 4 No
Main shaft, bearing, 5 No
seals
Brake system 6 No
DC/DC converter 7 Yes
Transformer 8 Yes
Electrical components 9 Yes
Controls 10 Yes
Structure and mooring 11 Yes
Ancillary system 12 Yes

82
5.4.4 Mean time to repair (MTTR)
The mean time to repair (MTTR) a subsystem is estimated using data from Carrol et al. (2015).
Figures 5-7 and 5-8 compare the probability of failures needing minor and major repairs to the MTTR
for each subsystem. The MTTR is defined as the mean time technicians spend on the device carrying
out the repair and does not include travel time, procurement time, weather windows etc. It is also
worth noting this data is collected from wind turbines and is applied to TSDs for indication purposes
only. When TSD specific data becomes available, this analysis can be updated.

Figure 5-8 illustrate that although major repairs are less likely to occur, the MTTR is much higher so
their contribution to downtime will be increase. Figure 5-8 shows that although the hub and the
main shaft subsystems have low failure rates, the MTTR is large. The maintenance strategy should
reflect this; where regular maintenance of these subsystems is carried out, as if they do fail they will
increase downtime significantly. Failure of these two subsystems also requires the device to be
brought onshore for repair, as shown in Table 5-2; further increasing device downtime. Figure 5-8
shows the opposite is true for the generator and DC/DC converter; where although they have high
failure rates, the MTTR is lower, meaning their contribution to the device downtime will be reduced.
These results illustrate that it is not only the failure rate that will determine the O&M strategy, but
also the MTTR as this will have a direct impact of downtime and therefore device electricity
generation.

Including procurement time will increase the accuracy of the analysis. As mentioned previously, the
amount of spare parts kept in stock will greatly affect procurement time and could save valuable
time if a component fails during a peak operating time. However, keeping spare parts is costly so a
cost-benefit analysis should be undertaken. These results suggest that the subsystems most likely to
fail are the control systems, generator and blades. Therefore, it is recommended that spare parts of
these subsystems are kept in stock so that repair times can be minimised.

Failure rate over lifetime and MTTR for minor repairs


20 12
18
10
Failure rate over lifetime

16
14
8
MTTR (hours)

12
10 6
8
4
6
4
2
2
0 0
1 2 3 4 5 6 7 8 9 10 11 12
Component Number

minor repair failure rate over lifetime minor repair MTTR

Figure 5-7: Failure rate and MTTR for minor repairs

83
Failure rate over lifetime and MTTR for major repairs
5 45
4.5 40
Failure rate over lifetime

4 35
3.5
30
3
25

MTTR
2.5
20
2
15
1.5
1 10
0.5 5
0 0
1 2 3 4 5 6 7 8 9 10 11 12
Subsystem Number

major repair failure rate over lifetime major repair MTTR

Figure 5-8: Failure rate and MTTR for major repairs

5.4.5 Average Cost per Failure


This section presents the average material cost per failure of each subsystem depending on the type
of failure. This cost data is material cost only and does not include labour cost, vessel hire cost etc.
The data has been collected from Carrol et al. (2015), and therefore the same uncertainties apply as
with the MTTR data. These should be used as an indication of subsystem cost and updated with TSD
specific data when available.

Figure 5-9 shows the average material cost to repair each subsystem based on the failure type:
major replacement, major repair or minor repair, on a logarithmic scale. No data on major
replacement cost was available on the brake, structure and ancillary subsystems and therefore they
are not shown on the graph.

As shown in Figure 5-9, the largest material costs are the main shaft, bearing and seals, costing
€95,000 for a major replacement. A major replacement of the hub also costs €95,000. The third
largest cost for a major replacement is €60,000 for a blade replacement. However, Figure 5-10
shows that the probability of a major replacement being required of these subsystems over the
device’s lifetime is small.

As stated previously, the component most likely to require a major replacement is the generator
subsystem, with a failure rate of 1.3 over its lifetime. The cost of this replacement is estimated at
€60,000. This is reflected in Figure 5-10, which shows the estimated material cost of all the repairs
over the lifetime of the device. An exchange rate of 0.88 was used to convert Euros to GBP. This
figure reflects the likelihood of the generator needing a major replacement over its lifetime as it has
by far the largest cost of all the subsystems, at £70,700. Figure 5-10 was produced by multiplying the
cost of each subsystems failure by the failure rate of the subsystems over the device lifetime of 20
years (data in Appendix A).

84
Figure 5-10 also shows that the DC/DC converter subassembly has the highest material cost for
major repairs over device lifetime, totalling £19,400. This is likely to be due to the large number of
components in the subsystem; particularly electrical components exposed to the marine
environment. The maintenance strategy should take this into account and regular visual inspections
should be performed on the subsystem components. If there is any wear on a component it should
be switch out as early as possible in a minor repair procedure to reduce the likelihood of requiring a
major repair at a later date. SCADA data could also be used to assess the state of components and
indicate when a fault may occur.

From this data the total material cost of all the predicted failures across the device over its lifetime is
estimated at £145,000. With the implementation of the suggestions made above, this figure can be
reduced to reflect the improvements that can be made by performing predictive maintenance on
the device.

Average Material Cost of Repair (Euros) (logarthimic scale)


100000

10000
AVerage material cost (Euros)

1000

100

10

1
1 2 3 4 5 6 7 8 9 10 11 12
Subsystem Number

major replacement major repair minor repair

Figure 5-9: Average material cost of repairs

85
Average material cost of repair over device lifetime (£)
90000

80000

70000

60000
average material cost (£)

50000

40000

30000

20000

10000

0
1 2 3 4 5 6 7 8 9 10 11 12
Subsystem Number

major replacement major repair minor repair

Figure 5-10: Average material cost of repair over device lifetime

5.4.6 Average Number of Technicians Required per Repair


Another factor that affects the cost of the maintenance strategy is the number of technicians
required to complete each repair/replacement. Data from Carrol et al. (2015) shows that the
average number of technicians required for minor repairs is 2.2, with a very small range between 1.8
and 2.6. Major repairs have an average of 2.8 technicians required per repair, whereas major
replacements require on average 7.6.

These results show that the number of technicians required to carry out both minor and major
repairs are similar and therefore will not affect the overall repair costs as much as the material costs
described in the previous section. These results are for technicians carrying out the repairs only, and
do not include extra crew required to operate the vessels and provide support.

86
5.5 Proposed O&M Strategy
The results of the failure rate analysis suggest that 1.8 major replacements will be required over the
20-year device lifetime, and the most likely subsystem that will need replacement is the generator.
From this, it is assumed for the O&M strategy that 2 major replacements will take place over the
device lifetime.

The other two failure categories, major repair and minor repair, are likely to occur 1.0 and 4.0 times
respectively, across all the subsystems in each year of operation. The average of the MTTR over all
device subsystems is 6.6 hours for a minor repair. Therefore, it is recommended that maintenance
and inspection occurs every quarter. Any non-critical minor failures, i.e. failures that will not cause
the device to lose production, will be repaired in the summer, when the operation is less likely to be
restricted by weather (See Section 5.6.2). Therefore, reducing O&M costs and maximising device
availability.

The average MTTR for major repairs that needs to be towed back to shore (see Table 5-2) is 25.2
hours. A failure of this kind is estimated to occur 0.36 times per year. Whereas, a major repair
repaired on site has an average MTTR of 14 hours and is estimated to occur 0.61 times per year.
Table 5-3 below summarises this information.

Table 5-3: Average MTTR and failure rates per category

MTTR Failure rate failures/lifetime


average (failures/year) (20 years)
(hours)
Major replacement 100.3 0.09 1.8
major repair- tow to 25.2 0.36 7.2
shore
major repair -onsite 14 0.61 12.1
minor repair 6.6 4.02 80.4

Based on the results in Table 5-3 above, it can be assumed that over the lifetime of the device, it will
be towed back to the shore 9 times for major replacements and major repairs to subsystems that
cannot be accessed on site. It is assumed that quarterly maintenance and inspection will take place
for many minor repairs, with an average MTTR of 7 hours. Major repairs onsite are predicted to
occur less than once a year and therefore will be repaired when they arise. It is estimated that the
MTTR of these failures will be 14 hours on site.

These operations were simulated in Mermaid to assess the impact weather conditions throughout
the year will have on the duration and cost of each operation. The results can be seen in Section 5.6
below.

87
5.6 Mermaid Analysis
Although reliability analysis can help advise maintenance strategies, it does not take into account
site specific factors such as weather, vessel information and port location. To achieve this Mermaid
was used. The input parameters and results are described in this section.

5.6.1 O&M port location


Scrabster Harbour has been chosen as the O&M port. It is situated on the north coast of Scotland,
approximately 30km from the TSD site. It was chosen because it has already received significant
activity from both the offshore wind and TSD industries. It offers 24/7 access and there are plans to
expand the port that would make it a good place to base operations.

5.6.2 Major Replacements: Towing back to shore


Maintenance that requires the device to be towed back to shore consists of any major replacements
and any major repairs as specified by Table 5-2. Towing time is assumed to be 2.5 hours from site to
Scrabster Harbour.

The vessel used to tow the device to shore is a Dart Fisher vessel, with operating limits specified in
Table 5-4 (Rinaldi et al., 2017).

Table 5-4: Vessel operating specifications

Operating Specification Dart Fisher RIB MultiCat


2631
Day rate (including fuel) £6000 £1000 £3250
Stand-by rate £6000 £0 £3250
At site station keeping limits:
Tidal current 2m/s 4m/s 3m/s
Wave height 3m 2m 2m
Wind speed 15m/s 15m/s 15m/s
Duration:
Get on station 0.1h 0.1h 0.1h
Prepare to leave 0.1h 0.1h 0.1h
Get on station limits:
Tidal current 2m/s 4m/s 3m/s
Wave height 3m 2m 2m
Wind speed 15m/s 15m/s 15m/s
Prepare to leave limits:
Tidal current 2m/s 4m/s 3m/s
Wave height 3m 2m 1.8m
Wind speed 15m/s 15m/s 15m/s

88
5.6.3 Minor onsite maintenance and repairs
As stated previously, all minor repairs and inspections will take place onsite. For these a RIB vessel is
used, operational with a maximum access wave height of 2m. The cost of this vessel per intervention
is assumed to be £1,000 (Scotrenewables, 2012). The operating limits are assumed in Table 5-4.

5.6.4 Major Onsite repairs


Any major repairs that can be performed on site, for example the ancillary subsystem repairs (see
Table 5-2), are assumed to use the Damen Multicat 2631 as this is the vessel currently used by
Scotrenewables at EMEC. The vessel is assumed to cost £3,250 per day, taken from an average of
values in Scotrenewables (2012). However, the costs of all the vessels is likely to vary throughout the
year due to changes in demand.

This vessel’s operating limits (Paterson et al., 2018) are shown in Table 5-4.

5.6.5 Metocean data.


Due to lack of long-term data availability at the Pentland Firth, metocean data for a site off the coast
of Liverpool has been used to estimate the wave and wind data at site. This data was corrected so
that the average significant wave height was 1.15m; matching that of the East Pentland Firth. The
Mermaid simulations used the same tidal data as in the resource assessment section (Section 3.1).
The data will be updated as the project progresses to feasibility stage and metocean data from the
site will be collected.

5.6.6 Other Simulation Assumptions


Operations and repairs available 24/7, 365 days per year.

No suspendable tasks.

5.6.7 Simulation Inputs


Simulation 1: Major Replacement/ Major Repair Offsite

Vessel: Dart Fisher

Repeat: twice monthly

Mobilisation: 3h

Work: 4h turbine shutdown, disconnect from mooring and attach to vessel

Back to port: 0.5h

Onshore work: 25h

Mobilisation: 3h

89
Work: 4h connect to mooring, turbine switch-on.

Back to port: 0.5h

Simulation 2: Major Repair Onsite

Vessel: MultiCat 2631

Repeat: twice monthly

Mobilisation: 1h

Work: 14h

Back to port: 0.5h

Simulation 3: Minor Repair

Vessel: RIB

Repeat: twice monthly

Mobilisation: 1h

Work: 7h

Back to port: 0.5h

5.6.8 Results
The results are presented as box and whisker diagrams for each simulation to show the variation in
duration and cost due to weather.

5.6.8.1 Simulation 1: Major Replacement/ Major Repair Offsite


Figure 5- 11 and 5-12 show the duration and cost, respectively, of a major replacement or major
repair process where the device was towed back to shore to complete the repair. The task duration
varies from 2.14 days in the summer months, to 7.26 in the winter months; with a P90 of 4.49 days.
This duration has a direct effect on the cost of the operation, with the vessel cost for the operation
varying from £18,000 to £35,000 depending on the time of year. Figures 5-11 and 5-12 show that
the optimal time to carry out the operation is in quarter 2 (from April to June) as the duration and
cost is minimised.

90
Figure 5-11: Duration variation – major replacement/ major repair offsite

Figure 5-12: Cost variation from Mermaid – major replacement/ major repair offsite

91
5.6.8.2 Simulation 2: Major Repair Onsite
Figures 5-13 and 5-14 show the results of carrying out a major repair on site. Although this type of
failure is rare, predicted to occur only 12 times over the device lifetime, it is important to perform
this analysis to estimate cost throughout the year and build contingency into the O&M budget.

The major repair has a minimum duration of 0.88 days and a maximum of 10.27 days. However, it
can be seen from Figure 5-13 that the maximum of 10.27 days is only likely to occur in the first
quarter (January to March). At other times of year the range in repair time is similar; ranging from
0.88 to 2.2 days. This trend is also reflected in Figure 5-14, showing the cost of the operation.

From these results it can be concluded that the time of year has less of an impact of the duration
and cost of this operation compared to the major replacement (shown in Figures 5-11 and 5-12).
This is due to its simpler procedure and lower average MTTR.

Figure 5-13: Duration variation from Mermaid – major repair onsite

92
Figure 5-14: Cost variation from Mermaid – major repair onsite

5.6.8.3 Simulation 3: Minor Repair


Figures 5-15 and 5-16 show the results of the minor repair simulation. They illustrate large
differences in both duration and cost of the operation depending on the time of year. The results
suggest that any minor repair carried out in quarter 3 (July to September) will not be restricted by
weather. However, as the weather worsens in the winter months, the results show that the
operation may have to be delayed by a maximum of 6 days in quarter 1 (January to March). This
large difference can be attributed to the vessel used for the repairs, a small RIB, which has tight
operating limits compared to the larger vessels used in the other simulations.

Despite the large variation in operating windows throughout the year, the cost of the operation does
not fluctuate as greatly as in the other simulations; varying from £1000 to £3000. This is because of
the relatively cheap cost of the vessel. Although this vessel is restricted by tight operating limits, its
low cost and quick mobilisation time make it a good vessel to perform these minor repairs.

93
Figure 5-15: Duration variation from Mermaid – minor repair

Figure 5-16: Cost variation from Mermaid – minor repair

94
5.7 Scaling Considerations
As the TSD array increases in size it may be necessary to change the maintenance strategy to
perform more planned maintenance. This is because if a failure occurs in multiple devices at the
same time, there may not be the resources in harbour (crew, vessels, spares etc.) to repair the
failures at the same time.

As mentioned previously, data from the first phase of the TSD array, where only one device will be
deployed, will be used to further optimise the O&M strategy. This will include data on failure rates,
metocean data, MTTR and costs. This can be analysed during the next phases of the project, where
more devices are deployed at the site so the O&M strategy will continually evolve and become more
cost efficient.

Economies of scale exist in O&M strategies as multiple technician teams can be deployed during one
operation from harbour; thus saving fuel and vessel costs.

5.8 Risks
There is significant risk in the development of an O&M strategy without access to a complete
economic and technical model of the device in question, as any deviations created by assumption
will inherently lead to inaccuracies within the strategy, which may lead to unintended cost increases
when implemented.

As previously mentioned, the interaction between systems and components may result in their
operation outside of the manufacturer’s intended range, reducing the validity of any reliability
calculations. In order to mitigate against this risk, system interactions should be considered and
investigated experimentally if it is determined that they pose significant deviations from standard.

As the device is still being developed, detailed condition reporting should be undertaken during
testing in order to establish any unexpected failure mechanisms that may be specific to the
mechanisms of operation of the SR2000 device. These could then be factored in to the O&M
strategy, or design changes could be made to eliminate them.

Dependent on the financing structure of the SR2000 project, there may be risks associated with cost
estimations for maintenance. If the project is to be initially funded through borrowing, a fixed
amount for ongoing operations and maintenance may be agreed based on a developed O&M
strategy. If operating costs overrun, it may have a very significant impact on the financial outcome of
the project. Strategies to avoid escalations in operating cost can be utilised, such as fixed rate
contracts with contractors and component suppliers, avoiding unexpected price increases due to
fluctuations in material costs etc.

Insurance of the device should be carefully considered, as this may be able to cover some
unexpected failures or damage. Component manufacturer warranty should also be accounted for
when failure costs are being counted.

Another important risk to be considered in any O&M strategy is HSE. Manufacturer guidelines
should be followed, along with industry guidelines and standards, as to the safe practice of repairs.

95
The use of divers should be minimised as it has high risk to human life, particularly in floating arrays
as both the mooring lines and devices are subject to sudden movement.

Full Control of Substances Hazardous to Health (COSHH) assessments should be carried out on any
lubricants and oil used in the device. Manufacturer instructions should be followed to negate the
risk of any substances leaking into the marine environment.

All technicians should complete the necessary training for working in the marine environment, for
example the GWO training standards. Emergency procedures should be understood and practised to
maximise safety whilst on the device.

5.9 Summary
This section shows how reliability and metocean data can be used to inform O&M strategies. The
results show that the design of the SR2000 allows for most minor repairs, totalling 80% of all
unexpected failures, to be fixed on site, lowering the cost of the maintenance. These minor repairs
can be carried out using an inexpensive RIB vessel. Mermaid analysis showed that there is a large
variation in operation length depending on the time of year, with an average operation taking up to
6 days in the winter and only 1 day in the summer. However, the low cost of the repair and vessel
mean that this deviation does not have a huge effect on the cost of the repair; with vessel cost only
varying from £3000 to £1000 in these cases.

Major repairs in the drive train assemblies will need to be repaired on shore, as well as any major
replacements. Major replacements are predicted to occur only 1.8 times over device lifetime. The
SR2000 will require no heavy lifting vessels to perform these replacements as the device can be
towed back to shore, further reducing maintenance costs.

Mermaid analysis of a typical major repair onshore showed that the time of year of failure greatly
effects the duration and cost of the operation. The cost varied from £18,000 to £35,000 in summer
and winter, respectively. This large variation reflects the more complex operation, where larger
weather windows are required.

The reliability data suggests that the blades, generator and control systems are most likely to fail and
therefore the O&M strategy should reflect this. The analysis also showed how MTTR predictions can
be used assess the impact of subsystem failure on downtime. The results suggest that regular
maintenance of the hub and main shaft subsystems should be prioritised as although their failure
rates are low, the MTTR to repair is high, which will negatively impact the device availability.

From the results it is recommended that maintenance and inspection is carried out every quarter.
Any non-critical minor failures will be repaired in the summer, when the operation is less likely to be
restricted by weather. The results suggest that over the lifetime of the device, it will be towed back
to the shore 9 times for major replacements and major repairs to subsystems that cannot be
accessed on site. Therefore, contingency should be made available in the O&M budget to account
for these repairs.

96
6 Economic appraisal
An economic appraisal of the two proposed Pentland Firth tidal stream device deployments has
been undertaken. For both the 2 MW and 6 MW installations, estimations for capital expenditure
(CAPEX), operational expenditure (OPEX) and revenue have been made, informed through both the
preceding contents of this report, and industrial & academic sources. A simple discounted cash flow
model has been constructed to ascertain the estimated net present value (NPV) of the two potential
installations. Subsequently, sensitivity analyses have been performed, to assess the parameters
most influential on the NPV, while scenario and Monte Carlo analyses are also performed, as a
means of judging the potential risk and outcomes of the proposed project. Finally, a levelised cost of
energy (LCOE) is derived for each scenario, and an indication of the potential outcome of upscaling is
established.

6.1 Capital Expenditure


Capital expenditure for the proposed Pentland Firth tidal steam installations, both single and triple
device variants, was calculated using a combination of derived values, typical cost ranges and typical
cost proportions.

The capital expenditure is divided into five separate categories:

 Device: the cost of the full tidal dual-turbine power take-off unit.
 Power: consisting of the total costs of undersea power cables, overland power cable,
and any required substations or grid strengthening.
 Support: the total cost of mooring lines and mooring anchors.
 Installation: cost of installing the device & moorings, power transmission works etc.
 Other: miscellaneous project capital expenditure, such as insurance and management
costs.

6.1.1 Device Cost


The device used in this study is based upon the Scotrenewables SR2000. While device costs are not
publically available, the cost of tidal power take-off devices are typically in the range of 70-80% of
total capital costs for tidal stream devices, including mechanical, electrical, and structural aspects
(Carbon Trust, 2005, 2006). Given a Scotrenewables estimate of £2.6m/MW capital costs for a 10
MW installation (Hamilton, 2012), a broad scale estimate of £2m/MW device costs has been
assumed for the purposes of this analysis.

While one would expect a small amount of cost reduction per unit due to producing three devices
rather than a single unit, the majority of the device cost is simply the cost of the steel material
(Scotrenewables, 2018), and as such cost savings are deemed minimal.

6.1.2 Power
Power CAPEX is separated into two units. The undersea power cables, totalling 6.7 km of MVAC
cable for both 2 MW and 6 MW installations, give a total cost of £4.22m (Section 3.5).At the
intersection of the undersea cable with the shoreline, power is delivered via overland power cables
to the proposed medium voltage substation. At a distance of 2 km, this trenched MVAC cable has an
estimated total cost of £7.40m.

97
The total cost of generated power delivery hardware from the tidal device(s) to the proposed
substation is therefore approximated at £11.62m.

6.1.3 Support
The floating device is secured to the bedrock of the Pentland Firth via a nylon mooring rope and
vertical lift anchor system. The mooring ropes, totalling 340 m for a single device, and 1020 m for a
3-device array, give a total approximate cost of £114,000 and £343,000, respectively. A cost of £336
per meter for the nylon rope is derived from an indicative cost of £1,687 for a 49 m-long, 80 mm-
diameter taut nylon mooring line, scaled accordingly (Ridge, Banfield and Mackay, 2010). This is in
line with costs for high breaking load mooring line materials (Harnois et al., 2010).

For anchoring, the mooring system utilises four pin piles per device. At approximately £70,000 per
anchor, this is projected to cost in the order of £280,000 for a single device, and £840,000 for three
devices (Ridge, Banfield and Mackay, 2010).

6.1.4 Installation
Installation costs have been estimated using a number of sources, as a percentage of the final capital
cost. Figures from the Carbon Trust suggest an approximation of 13% of CAPEX for marine energy
devices (Carbon Trust, 2006), and 16% for tidal stream devices specifically (Carbon Trust, 2005).
However, the Low Carbon Innovation Co-ordination Group in 2012 suggested a significantly greater
proportional cost, with an installation cost at 35% COE (Lcicg, 2012). Astariz, Vazquez and Iglesias
(2015) assess marine energy installations, and while a breakdown is not given for tidal capital costs,
wave energy converters (WECs) are given a range of 16%-30% dedicated to installation costs. As the
SR2000 is a moored, floating device, comparison to WECs is reasonably applicable for this study.

Given that the Scotrenewables SR2000 device is designed for ‘low cost installation’, it would be
prudent to assume a reasonably low-end estimate for proportional installation costs – mitigated
somewhat by the difficult conditions of the Pentland Firth. For the purposes of this assessment, a
value of 20% CAPEX has been adjudged as a reasonable estimate for installation costs.

6.1.5 Project Management


Project management includes miscellaneous capital expenditure, including insurance, management,
legal fees, and further miscellany.

Meygen Phase 1c had an eventual proportional cost of 9% for project initiation, management and
insurance (Meygen Ltd, 2017). Carbon Trust estimates place total project management at 2% for an
example wave energy device, and 4% for an example tidal stream device (Carbon Trust, 2005, 2006).

Given the difficulty of estimating these costs at such an early stage of project planning, a mean
average of the above three sources has been used, giving a final project management estimate of 5%
of the total capital expenditure.

6.1.6 Total capital expenditure


The total CAPEX for a single device installation is estimated at 18.7 £m, 9.34 £m per MW capacity.

The total CAPEX for a three-device installation is estimated at 25.1 £m, 4.18 £m per MW capacity.

A breakdown of the capital costs can be seen in Figure 6-1.

98
Figure 6-1: Pie charts displaying the proportion of CAPEX for each capital expenditure, illustrating the difference between
the two possible scales of installation.

A 2010 report ‘Cost of and financial support for wave, tidal stream and tidal generation in the UK’
(Ernst & Young LLP, 2010) for the Department of Energy and Climate Change and the Scottish
Government listed various predicted CAPEX costs for tidal stream developments. For a
demonstration project (albeit based on a 10 MW array), capital expenditure is expected to total
£4.1m - £5.7m per MW capacity, while a commercial project is expected to be limited to £2.8m -
£3.9m per MW capacity.

In a 2016 report on marine energy by the World Energy Council (World Energy Council, 2016),
summary data capital expenditure indicates a typical CAPEX between $4.3m and $8.7m per MW for
a demonstration project (£3.1m - £6.2m per MW), and $3.3m to $5.6m per MW for a first
commercial-scale installation (£2.4m - £4.0m per MW).

Additionally, an estimate from Scotrenewables themselves suggests a range of between £25m and
£28m for the capital expenditure for a proposed 10 MW array (Hamilton, 2012)

As such, it is clear that our installations are at the higher end of the capital expenditure ranges. The
single device installation cost per MW is hugely beyond typical capital expenditure estimates; this is
certainly due to the power distribution cost, a result of significant distances of both overland and
undersea cabling, while also designed to transmit far more than the 2 MW capacity of the
installation. It may be more prudent to compare the capital costs of the three-turbine installation,
which are towards the lower end of typical demonstration project ranges, and at the top end of
typical commercial project ranges. This is somewhat expected, given the distance of cabling required
at such a remote location, coupled with higher than average mooring costs, and significant cost of
the device itself, due to the amount of steel required during fabrication.

The CAPEX/MW is roughly 50% greater than the equivalent value proposed by Scotrenewables for a
hypothetical 10 MW array of the SR2000 device. As there is not much information on the

99
methodology behind these derived values, it is difficult to ascertain where the differences in
estimation may lie. It is possible that the hypothetical scenario is far closer to a high voltage grid
connection, as power distribution has the greatest impact on the estimated capital expenditure for
this project.

6.2 Operational Expenditure


Annual operational expenditure is informed by a Cost of Energy analysis by Scotrenewables, based
upon a hypothetical first 10 MW array of SR2000 devices. The total annual operating cost per MW is
estimated at £198,200 in a typical case, £160,000 for an optimistic estimate, and £235,000 for a
pessimistic estimate (Hamilton, 2012). Given the relatively difficulty of maintenance within the
Pentland Firth (Section 5), and the reduced size of the proposed installations (2 MW and 6 MW), the
pessimistic estimate has been utilised for the purposes of this economic analysis.

These figures, produced using The Carbon Trust methodology, are in line with World Energy Council
estimates (World Energy Council, 2016) ,which suggest a range of 150k-530k $/MW/a (£107k-£377k)
for ‘second stage projects’, and 90k-400k $/MW/a (£64k-£285k) for commercial scale projects. Given
the small installation capacity, and difficulty of Pentland Firth conditions, it is expected that the
project’s annual OPEX estimate is aligned with the upper end of the commercial scale project
estimate.

6.3 Revenue
Revenue is derived using a gross generation of 5.39 GWh/device (Section 3.1),multiplied by an
estimated total loss factor of 0.85 to encompass energy loss through downtime, planned
maintenance, cable losses, turbine biofouling etc. This is based upon limited information available
for tidal turbine availability (Vazquez and Iglesias, 2015) and estimates for offshore wind turbines
(GL Garrad Hassan, 2013). This gives a total saleable output of 4.58 GWh per annum for a single
device, and 13.74 GWh per annum for three devices.

The total revenues of the installations are dependent on the procurement of a Contract for
Difference (CfD) agreement with the government. Current auctions put marine energy in
competition with offshore wind, with strike prices as low as £57.50/MWh for projects
commissioning in 2022/23 (Grubb and Newbery, 2018). This would clearly be unfeasible for tidal
stream energy, given the current high capital costs associated with the technology. The September
2017 CfD auctions placed an administrative strike price of £300/MWh for tidal power; however, no
bids were made from tidal power developers (Tidal Energy Today, 2017).

Atlantis Resources, owners of MeyGen, are currently aiming to secure a CfD from the UK
government of £150/MWh through bilateral negotiations. This would act as a form of subsidy for the
emerging UK tidal industry, aiming to contribute to decreasing the future costs of similar
installations (Tidal Energy Today, 2017).

This economic analysis is undertaken under the assumption that this project could secure a
£150/MWh CfD, similar to the proposed Meygen negotiation. As such, the total revenue for a single
device is estimated at £687,000/a, and £2,060,000/a for three devices.

100
6.4 Grants and subsidies
The Scottish Government has identified marine energy as a potentially significant industry, both in
Scotland, and on a global scale. There are a number of grant, loan and subsidy schemes aimed at
enticing wave and tidal energy projects in Scottish waters; this floating tidal array would be well
placed to apply for a number of these initiatives.

On February 28th, 2018, the Scottish National Investment Bank implementation plan was published,
aiming to support economic growth in Scotland. With the plan focussing on long-term investment
and emerging sectors, tidal stream renewable energy projects are well placed to benefit from the
bank’s investment (Tidal Energy Today, 2018).

Scottish Enterprise, a public body with an economic development remit, established three rounds of
funding earmarked for testing of wave and tidal energy prototypes in Scotland (Scottish
Government, 2014). The three rounds led to over £20m of support, with the third round focussing
on bringing advanced marine energy devices to a commercial level, ideally aiding in improvement
throughout the sector. While this support scheme appears to have concluded, there is clear and
obvious desire for marine renewables support in the Scottish government, and future investment
plans would not be unexpected.

Contemporaneously, Scottish Enterprise provides financial assistance via its Renewable Energy
Investment Fund for renewable projects with a “demonstrable funding gap” and “at a sufficient
stage of development” (Scottish Enterprise, 2014). Marine energy is explicitly stipulated as an area
of support for investment. The fund prioritised projects near to commercialisation, and innovative
projects with an aim to decreasing costs and installation risks and O&M risk. The fund was
significantly involved in the early stages of the Meygen tidal stream project, also located in the
Pentland Firth and at a similar scale (in the first phase), demonstrating that this project would likely
be a strong candidate for investment.

On a non-governmental level, the Carbon Trust has provided over £50m of investment in innovative
marine energy technology since 2003, with a number of programmes aimed at cost reduction, full-
scale technology proving, capital support for commercial-scale installations and aiding the
development of arrays for marine energy devices (Carbon Trust, 2018). All of the above programmes
involved supporting industrial projects, and the project detailed in this report would likely be
strongly involved in any further Carbon Trust schemes.

A vital determiner for this project’s economic viability is the revenue price obtained for any saleable
tidal power generated by the device/array. This would likely be heavily depending on securing a
beneficial Contract for Difference (CfD) deal with the UK government, to support to development of
tidal power towards potential competitive prices in the future. As of late 2017, Atlantis Resources
were aiming to negotiate a £150/MWh CfD deal for Phase 1c of the Meygen project, indicating a
similar deal could possibly be struck for our floating tidal development (Whiterow, 2017).

As evidenced by the above examples, there is significant support for advanced projects at both a
governmental and non-governmental level, with this project well placed to benefit from such
initiatives. While it would be prudent to assume as ‘worst case scenario’ of zero grants or investment
for determination of the economic prospects of this project, recent history suggests a high likelihood
of grant, subsidy or investment support.

101
6.5 Discounted Cash Flow Model
Discounted Cash Flow Models (DCFMs) have been constructed for both single device and triple
device installations, based upon the CAPEX, OPEX and revenue estimates detailed within sections
6.1, 6.2 and 6.3. This allows the Net Present Value (NPV) of both projects to be assessed, given an
indication of the value and overall economic feasibility of the proposed installations. In addition to
the estimated costs and revenue, a number of informed assumptions have been made:

6.5.1 Funding
The project is modelled as funded through standard business debt, with a loan of the total value of
the year zero capital expenditure. This leads to a projected loan of £18.7m for a single device set-up,
and £25.1m for three devices.

In reality, project funding is likely to be funded via a mixture of developer equity, debt, grants and
potentially other investment (Catapult Offshore Renewable Energy, 2014). A standard loan is used in
this model as a basic equivalent for the complex myriad of potential funding avenues.

A loan rate of 6% is assumed, paid back via straight-line amortisation over a twenty-year period,
based upon the design life of the device (Scotrenewables, 2018).

6.5.2 Project life


The Scotrenewables SR2000 device is designed to DNV-GL standards, specifying a twenty-year design
life. As such, this is the assumed lifespan of the project.

6.5.3 Tax rate


The model uses a corporation tax rate of 19%, based on the current tax rate (www.gov.uk, 2018).

6.5.4 Inflation
Inflation over the twenty-year project life is assumed at the Bank of England target inflation rate of
2% (Bank of England, 2018). This is applied to operational expenditure alone, as revenue is based
upon a hypothetical fixed CfD agreement, and capital expenditure occurs entirely within year zero.

6.5.5 Grants and subsidies


As detailed in the previous section, grants and subsidies are often available for marine renewable
developments; however, it is difficult to predict the value or likelihood of the impact of this aid on
the proposed project. With the exception of the CfD agreement detailed in the revenue estimate, a
‘worst case scenario’ of no grants or subsidies has been assumed for the purposes of this model.

6.5.6 Discount rate


Due to the risk of tidal stream investments, discount rates are typically high. Over recent years
discount rate assumptions in similar studies have ranged from 8% to 15% (Allan et al., 2011;
Stegman et al., 2017). Multiple sources suggest a standard discount rate of 10% (Allan et al., 2011;
Astariz, Vazquez and Iglesias, 2015), and this rate has been applied to this cash flow model.

6.5.7 Results
The DCFM for a single device installation gives an NPV estimate of £-13.1m.

The DCFM for three devices gives an NPV estimate of £-15.4m.

The full models follow this section.

102
Single Device Discounted Cash Flow Model

Table 6-1: Base case discounted cash flow model for a single device deployment.

End of Year 0 1 2 3 4 5 6
Revenue 700553 700553 700553 700553 700553 700553
OPEX 479400 488988 498767.76 508743.1152 518917.9775 529296.3371
Margin 221153 211565 201786 191810 181635 171257
Interest payments 1121145 1065088 1009031 952973 896916 840859
CAPEX 18685751 0 0 0 0 0 0
Capital allowance 934287.574 934288 934288 934288 934288 934288
Taxable income -1834279 -3622089 -5363622 -7059073 -8708641 -10312530
Loss c/f 1834279 3622089 5363622 7059073 8708641 10312530
Tax 0 0 0 0 0 0
Loan 18685751
Principle payment 934288 934288 934288 934288 934288 934288
Interest 1121145 1065088 1009031 952973 896916 840859
Total payment 2055433 1999375 1943318 1887261 1831204 1775146
Cashflow -1834279 -1787810 -1741533 -1695451 -1649568 -1603889
Cumulative cashflow -1834279 -3622089 -5363622 -7059073 -8708641 -10312530
Present value -1667527 -1477529 -1308439 -1158016 -1024252 -905354
Net present value -13097789

End of Year 7 8 9 10 11 12 13
Revenue 2101660 2101660 2101660 2101660 2101660 2101660 2101660
OPEX 1619647 1652040 1685081 1718782 1753158 1788221 1823985
Margin 482013 449620 416579 382878 348502 313439 277675
Interest payments 1052965 977753 902541 827329 752118 676906 601694
CAPEX 0 0 0 0 0 0 0
Capital allowance 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Taxable income -13705272 -15486934 -17226425 -18924406 -20581551 -22198547 -23776096
Loss c/f 13705272 15486934 17226425 18924406 20581551 22198547 23776096
Tax 0 0 0 0 0 0 0
Loan
Principle payment 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Interest 1052965 977753 902541 827329 752118 676906 601694
Total payment 2306494 2231282 2156071 2080859 2005647 1930435 1855223
Cashflow -1824481 -1781662 -1739491 -1697981 -1657145 -1616996 -1577549
Cumulative cashflow -13705272 -15486934 -17226425 -18924406 -20581551 -22198547 -23776096
Present value -936247 -831158 -737714 -654645 -580819 -515225 -456960

End of Year 14 15 16 17 18 19 20
Revenue 2101660 2101660 2101660 2101660 2101660 2101660 2101660
OPEX 1860465 1897674 1935628 1974340 2013827 2054104 2095186
Margin 241195 203986 166032 127320 87833 47556 6474
Interest payments 526482 451271 376059 300847 225635 150424 75212
CAPEX 0 0 0 0 0 0 0
Capital allowance 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Taxable income -25314912 -26815727 -28279283 -29706339 -31097671 -32454068 -33776335
Loss c/f 25314912 26815727 28279283 29706339 31097671 32454068 33776335
Tax 0 0 0 0 0 0 0
Loan
Principle payment 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Interest 526482 451271 376059 300847 225635 150424 75212
Total payment 1780012 1704800 1629588 1554376 1479165 1403953 1328741
Cashflow -1538817 -1500814 -1463556 -1427057 -1391332 -1356397 -1322267
Cumulative cashflow -25314912 -26815727 -28279283 -29706339 -31097671 -32454068 -33776335
Present value -405219 -359283 -318512 -282336 -250243 -221782 -196547

Net present value -13097789

103
Triple Device Discounted Cash Flow Model

Table 6-2: Base case discounted cash flow model for a three-device deployment

End of Year 0 1 2 3 4 5 6
Revenue 2101660 2101660 2101660 2101660 2101660 2101660
OPEX 1438200 1466964 1496303.28 1526229.346 1556753.933 1587889.011
Margin 663460 634696 605357 575431 544906 513771
Interest payments 1504235 1429024 1353812 1278600 1203388 1128176
CAPEX 25070588 0 0 0 0 0 0
Capital allowance 1253529.39 1253529 1253529 1253529 1253529 1253529
Taxable income -2094305 -4142162 -6144146 -8100845 -10012856 -11880791
Loss c/f 2094305 4142162 6144146 8100845 10012856 11880791
Tax 0 0 0 0 0 0
Loan 25070588
Principle payment 1253529 1253529 1253529 1253529 1253529 1253529
Interest 1504235 1429024 1353812 1278600 1203388 1128176
Total payment 2757765 2682553 2607341 2532129 2456918 2381706
Cashflow -2094305 -2047857 -2001984 -1956699 -1912012 -1867935
Cumulative cashflow -2094305 -4142162 -6144146 -8100845 -10012856 -11880791
Present value -1903913 -1692444 -1504121 -1336452 -1187209 -1054401
Net present value -15425228

End of Year 7 8 9 10 11 12 13
Revenue 2101660 2101660 2101660 2101660 2101660 2101660 2101660
OPEX 1619647 1652040 1685081 1718782 1753158 1788221 1823985
Margin 482013 449620 416579 382878 348502 313439 277675
Interest payments 1052965 977753 902541 827329 752118 676906 601694
CAPEX 0 0 0 0 0 0 0
Capital allowance 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Taxable income -13705272 -15486934 -17226425 -18924406 -20581551 -22198547 -23776096
Loss c/f 13705272 15486934 17226425 18924406 20581551 22198547 23776096
Tax 0 0 0 0 0 0 0
Loan
Principle payment 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Interest 1052965 977753 902541 827329 752118 676906 601694
Total payment 2306494 2231282 2156071 2080859 2005647 1930435 1855223
Cashflow -1824481 -1781662 -1739491 -1697981 -1657145 -1616996 -1577549
Cumulative cashflow -13705272 -15486934 -17226425 -18924406 -20581551 -22198547 -23776096
Present value -936247 -831158 -737714 -654645 -580819 -515225 -456960

End of Year 14 15 16 17 18 19 20
Revenue 2101660 2101660 2101660 2101660 2101660 2101660 2101660
OPEX 1860465 1897674 1935628 1974340 2013827 2054104 2095186
Margin 241195 203986 166032 127320 87833 47556 6474
Interest payments 526482 451271 376059 300847 225635 150424 75212
CAPEX 0 0 0 0 0 0 0
Capital allowance 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Taxable income -25314912 -26815727 -28279283 -29706339 -31097671 -32454068 -33776335
Loss c/f 25314912 26815727 28279283 29706339 31097671 32454068 33776335
Tax 0 0 0 0 0 0 0
Loan
Principle payment 1253529 1253529 1253529 1253529 1253529 1253529 1253529
Interest 526482 451271 376059 300847 225635 150424 75212
Total payment 1780012 1704800 1629588 1554376 1479165 1403953 1328741
Cashflow -1538817 -1500814 -1463556 -1427057 -1391332 -1356397 -1322267
Cumulative cashflow -25314912 -26815727 -28279283 -29706339 -31097671 -32454068 -33776335
Present value -405219 -359283 -318512 -282336 -250243 -221782 -196547

Net present value -15425228

104
6.6 Sensitivity Analyses
Both objective and subjective sensitivity analyses are performed. This aims to identify the variables
that have the greatest influence on the economic viability of the project, and thus indicate where
future efforts should be most focused.

6.6.1 Objective Analysis


An initial objective sensitivity analysis is performed, flexing the following parameters by an arbitrary
20% either side of the ‘best guess’ Figure:

- CAPEX
- OPEX
- Gross output
- Loss factor
- Price
- Loan rate
- Corporation tax
- Project life
- Inflation rate

By flexing each parameter in isolation, the NPV of the project is calculated for each parameter flex.

Figure 6-3: Objective sensitivity analysis for a single device installation, with steeper gradients indicating a
greater impact on NPV.

105 installation, with steeper gradients indicating a


Figure 6-2: Objective sensitivity analysis for a three-device
greater impact on NPV.
This produces the following analysis:

This initial analysis indicates that for a single device, the capital expenditure is the dominant factor in
determining the value of the project, with over twice the magnitude of impact of any other
parameter. The three parameters determining revenue (gross output, loss factor and price) have a
smaller but significant impact, while loan rate, project life and operational expenditure also have an
effect. Corporation tax has no impact on the base case, as no profit is ever made.

For an installation of three devices, the influence of CAPEX is reduced. While still the most significant
parameter in this analysis, the revenue factors (gross output, loss factor and price) have a similar
impact when flexed objectively. This is in line with the reduced CAPEX/MW of a triple device
installation, as other factors hold proportionally more weight. Operational expenditure becomes a
much more significant influencer, given the greater number of devices to be maintained, while
project lifespan and loan rate still have a significant impact. Once again, inflation and corporation tax
have minimal impact. This is somewhat fortuitous, seeing as they are arguably the parameters that
can be controlled the least.

From this initial analysis, it is clear that control of capital costs is an important consideration for
reducing the risk to the project. However, in installations of multiple arrays, the impacts of revenue
and O&M costs are amplified, and as such also bear significant risk when it comes to project
planning.

For a number of parameters, best-case and worst-case estimates can be derived. As such, a
subjective sensitivity analysis is also undertaken, to further identify significant parameters.

6.6.2 Subjective Analysis


While an objective analysis is useful for giving an indication of sensitive parameters, there are a
number of variables where reasonable best case, worst case and ‘best guess’ limits can be defensibly
derived.

6.6.2.1 Costs
Due to lack of primary data, capital and operational expenditure discussed in this section are broad-
scale estimates, with high associated risk and significant scope for further refinement in subsequent
stages of the project. As such, a range of ±20% of the best guess figure has been used as a
reasonable range of CAPEX and OPEX values over the lifespan of the project.

6.6.2.2 Total loss factor


A total loss factor has been estimated to give an idea of the proportion of the total possible gross
device output that is actually input into the National Grid system. This primarily encompasses
transmission losses in power cables, and turbine downtime for planned or unplanned maintenance.
A ‘best guess’ of 0.85, justified in Section 6.3, accounts for the estimated transmission losses, and a
small amount of unplanned maintenance, on top of scheduled planned maintenance. The SR2000
device is designed for easy access by maintenance crews, and as such it is expected that planned
maintenance will not be a dominant factor on availability.

In a worst-case scenario, an average annual total loss factor has been reasoned at 0.75; this would
suggest significant downtime due to unplanned maintenance requirement, exacerbated by the
restrictive access timeframes as a result of the Pentland Firth location. A best-case estimate places

106
the average total loss factor at 0.95; essentially, accounting for minimal transmission losses and
planned maintenance.

6.6.2.3 Price
This economic assessment assumes a ‘best guess’ CfD of £150/MWh, negotiated through the UK
government. A worst-case scenario would force the project to compete with offshore wind
installations, thus a CfD of approximately £60/MWh would be necessary to compete in the CfD
auctions. Alternatively, a best-case scenario would entail ring-fencing for marine energy projects; in
this instance, there is potential for the project to secure a CfD for £300 (the strike price for the
previous round of CfD auctions, in which there were no bids from marine energy generation
projects).

6.6.2.4 Loan
Loan capital is predicted to match CAPEX flex, while loan period is adjudged to equal project life.
However, the offered loan rate can be flexed independently. A best guess loan rate of 6% is
assumed, given the long life but relatively high risk of the project. This has been flexed to 7.5% as a
worst-case value, should potential investors deem the project as a significantly high-risk investment.
A best-case value of 4.5% is based on the assumption of potential government backing, grants etc.
reducing the risk perceived by a potential financial backer.

6.6.2.5 Project life


The Scotrenewables SR2000 device is designed for an operating lifespan of 20 years. A worst case of
15 years is assumed, a scenario in which the consistently high current velocities accelerate wear on
the turbines and drivetrain. The best and worst cases depend on whether the project is net
profitable, or making a loss.

6.6.2.6 Corporation tax


Corporation tax is currently set at 19%, and hence, this value is taken for the best guess. This is the
lowest rate seen for decades (Figure 6-4), and as such it is not expected to fall significantly lower,
over the project life of the installation; therefore, a best-case tax rate of 15% has been assumed. As
a worst case, a return to the 30% corporation tax rate of 1997-2008 has been used.

Figure 6-4: Graph illustrating the UK corporation tax rate from 1981 to 2018. The red area indicates the chosen
potential range of corporate tax rate for the subsequent modelling. Modified from tradingeconomics.com
107
6.6.2.7 Inflation
Rate of inflation, which primarily impacts the OPEX and discount rate estimate of the simplified cash
flow model, has been estimated based on historic data for the UK inflation rate. The Bank of England
target rate, set by the incumbent government, is set to 2%, and as such is the ‘best guess’ figure
used for the analysis. As illustrated in Figure 6-5, inflation is almost entirely maintained between 1%
and 3% over significant (multi-year) time periods; hence, these values have been utilised for best
and worst-case inflation rates respectively.

It should be noted that in this case a higher inflation is detrimental, due to the impact of increasing
costs and the fixed price for produced energy. This may not always be the case, if the energy
agreement encompasses inflation over time.

Figure 6-5: Graph illustrating the UK corporation tax rate from 1981 to 2018. The red area indicates the chosen potential
range of corporate tax rate for the subsequent modelling. Modified from tradingeconomics.com

By flexing each parameter in isolation, the NPV and IRR of the project is calculated for each
parameter flex. This produces the following analysis:

Figure 6-6: Subjective sensitivity analysis for a single device installation, with steeper gradients indicating a
greater impact on NPV.

108
Figure 6-7: Subjective sensitivity analysis for a three-device installation, with steeper gradients indicating a
greater impact on NPV.

Here, a clearer indication of the most sensitive parameters is ascertained. For a single device, the
best and worst-case price and capital expenditure have a far greater impact on the success of a
project than the remaining factors. Interestingly, loan rate has the third greatest influence on the
profitability or otherwise of a single device; this is due to the link between capital expenditure,
which is extremely high for the 2 MW installation, and the interest paid on the loaned capital. The
remaining parameters have a much smaller impact on NPV, with worst-case to best-case ranges at
less than 15% of the base case value.

For three devices, it is clear that the price per MWh is overwhelmingly dominant. This is expected,
given the massive potential range of CfD values (from competing with offshore wind at £60/MWh, to
a ring-fenced subsidy of £300/MWh). Given the installation produces three times the output of a
single device, the impact of price is amplified, while the lower CAPEX/MW dampens the impact of
capital expenditure simultaneously. That said, CAPEX, OPEX and loan rate all still have a significant
impact on NPV; if a CfD were to be secured, these parameters would need to be attended to.

6.6.3 Discussion
It is clear that large ranges in price and capital expenditure are the dominant risks to the confidence
of either project going forward. As such, it is imperative to reduce the ranges of these parameters.
The range of potential prices could be reduced via preliminary discussions with government entities,
and sounding out possible power purchase agreements from non-grid parties. Capital expenditure
confidence should increase as the study continues, as increasing amount of both secondary and
primary data are collected. In addition, grant funding has the potential to drastically reduce the level
of CAPEX (and associated loan or other funding) required for an installation to go ahead. While price
and CAPEX require the majority of attention, it is important not to ignore the potential impact of
operational expenditure, which would continue to increase in importance as a deployment increases
capacity.

6.7 Scenario Analysis


The NPV and IRR of the two proposed scales of installation are calculated for best guess, absolute
best-case, and absolute worst-case scenarios. The range between these predictions helps to quantify
the risk associated with the project extremes. Given the high-level nature of this initial study, and
general uncertainty surrounding the tidal energy industry quantified within the sensitivity analysis
above, a large range is expected.

109
6.7.1 Standard scenario analysis
Table 6-3: Scenario analysis for both installation sizes, defining the absolute minimum and maximum project net present
values, given the parameter ranges defined during the preceding subjective sensitivity analysis.

Installation size Extreme worst case Base case Extreme best case Range
2 MW (1 device) £-23,361,214 £-13,097,789 £704,799 £24,660,013
6 MW (3 devices) £-38,238,537 £-15,425,228 £16,557,704 £54,796,241

6.7.2 Fixed price scenario analysis


As identified by the sensitivity analysis, the hypothetical agreed CfD price has an overwhelming
impact on the profitability of the proposed projects. Given that an installation would not go ahead
without an agreed price per unit energy produced, a scenario analysis has been performed with a
fixed price of £150/MWh, to give a more representative indication of project risk, should a power
purchase agreement be successfully negotiated.

Table 6-4: Scenario analysis for both installation sizes, defining the minimum and maximum project net present values in
the scenario of a fixed £150/MWh CfD, given the parameter ranges defined during the preceding subjective sensitivity
analysis.

Installation size Extreme worst case Base case Extreme best case Range
2 MW (1 device) £-20,491,585 £-13,097,789 £-6,420,345 £14,071,240
6 MW (3 devices) £-29,629,650 £-15,425,228 £-2,302,080 £27,327,570

6.7.3 Discussion
At such a high level of study, the ranges involved between absolute worst and best-case net present
values are monumental, exacerbated by the lack of confidence in the estimation of many
parameters. While both extreme best cases are profitable, the chances of such and occurrence are
so infinitesimally small that it would be foolish to predicate any decisions upon it. The noteworthy
outcome is the reduction in range when the price of sold energy is fixed at £150/MWh. The range is
more-or-less halved for both deployment scenarios, and while the level of risk is still extremely high,
there is evidence that further investigation can surely add significant confidence to any study going
forward.

6.8 Monte Carlo Analysis


6.8.1 Methodology
An initial Monte Carlo analysis was performed, to assess the possible outcomes of the parameter
ranges detailed within section 6.6.2. For the analysis to be performed, the following assumptions
were made:

- CAPEX; OPEX; Gross output; Loss factor; and inflation rate are normally distributed, with
worst/best case values at three times the standard deviation from the mean. This gives a
99.7% probability of each parameter falling within the worst case-best case range.
- Price and corporation tax are uniformly sampled from worst case, ‘bad case’, base case,
‘good case’ and best-case values. The ‘bad’ and ‘good’ cases represent the mean of the base
case and the corresponding worst or best case.
- Three possible project lifespans are uniformly simulated: 15, 20, and 25 years.
- Net output, revenue, loan capital and loan period are not sampled, as they are values
derived from other sampled values listed above, and are linked accordingly in the model. For
example, the loan capital will always match the random sample of CAPEX.

110
The simulation was run for 3,000 iterations using Microsoft Excel software.

6.9 Results and discussion

Figure 6-8: Monte Carlo analysis, resulting from 3000 simulations, illustrating a range of probable outcomes for a single
device scenario.

Figure 6-9: Monte Carlo analysis, resulting from 3000 simulations, illustrating a range of probable outcomes for a
three-device scenario. The 86th percentile is indicated as the percentile in which the NPV becomes positive.

The results of the Monte Carlo analyses for both scales of installation are illustrated in Figures 6-8
and 6-9. It is clear immediately that a single device installation will essentially never be profitable.
This was confirmed in the scenario analysis, where even the absolute best-case scenario had a NPV
of less than £1m. While there is certainly potential to reduction of capital expenditure, it is difficult
to imagine a scenario where deployment of a single device in the centre of the Pentland Firth is of
positive economic value.

111
However, the analysis for a three-device installation indicates a more optimistic outlook. Over the
course of 3000 randomly sampled simulations, over 14% (425/3000) elicited a positive NPV. While
these almost certainly represent scenarios where the price was sampled at £300/MWh, it is
evidence that there is potential for economically successful floating tidal stream installations within
the Pentland Firth. That said, the vast majority of simulations produced a negative result, with a
mean NPV of £-13,933,712, and a median of £-16,037,438. While a minor proportion of iterations
are profitable, it is clear that a significant amount of study is still necessary to define a genuinely
profitable opportunity for floating tidal stream energy in the Pentland Firth.

6.10 Levelised Cost of Energy


A levelised cost of energy (LCOE) has been calculated using Microsoft Excel ‘Solver’ software,
calculating the price per MWh required for the NPV of the installation to be equal to zero.
Essentially, the LCOE is the minimum price of energy required for the project to break even.

The LCOE for a single device installation is 482.08 £/MWh. For a three-device installation, the LCOE is
280.14 £/MWh

The levelised cost for a single device is prohibitively high, and well beyond even optimistic
suggestions for recommended strike prices for tidal energy. Astariz, Vazquez and Iglesias (2015)
detailed a recommended ‘catalysing’ strike price of £280-300/MWh, while the World Energy Council
indicates top-end LCOEs of around £335/MWh for developed, but pre-commercial arrays. However,
the three-turbine array produces energy at a cost of approximately £280/MWh. While this is not
competitive in a market amongst more developed forms of energy generation, it is certainly within
the range of a competitive LCOE in the tidal range industry, especially when considering the
significant scope for cost reduction.

6.11 Payback period


As the base case scenario used in this analysis does not produce a positive net present value, the
payback period cannot be defined, i.e. the payback period is beyond the planned lifespan of the
installation.

6.12 Scaling
A basic scaling has been applied to the installation, to simulate LCOE up to 20 MW scale. The
assumptions are as follows:

- Device cost, mooring costs, operational expenditure and output are assumed to scale
linearly.
- Power distribution costs are assumed to increase by 10% for every 2 MW device added to
the array, due to the fact that the increasing capacity of any laid cable would not affect the
cost in a linear fashion.
- Installation and project management, segments of capital expenditure, remain at a fixed
percentage of the total CAPEX.

112
Figure 6-10: Levelised costs of energy at scales between 2 and 20 MW (1 to 10 devices), indicating a steady
decrease in LCOE at increasing scale.

As illustrated in Figure 6-10, levelised cost of energy is expected to reduce significantly after the
installation of more than a single turbine. While increasing capacity is accompanied by a
decrease in levelised costs, there is a definite limit below which the cost cannot decrease, at
around £230/MWh.

This will in part be due to the sheer volume of steel required in the fabrication of the SR2000
device, but also a result of the mooring systems, power transmission etc. being optimised for
this particular, small-scale study. At scales of tens or hundreds of devices, additional
considerations such as floating high voltage substations, grid strengthening, proximal O&M
stations etc. would have to be considered, and are beyond the scope of this particular economic
assessment. The basic economy of scale theory would suggest that increased scales would
induce lower, more competitive levelised costs of energy; Scotrenewables project a possible
£120/MWh 10 MW array of SR2000 devices.

113
6.13 Conclusions
- The capital costs of installing a single device in the weakly-gridded Pentland Firth region are
prohibitively expensive, and near impossible to undertake with economically valuable
outcomes. This is reduced somewhat when three devices are deployed.

- ‘Best guess’ discounted cash flow models indicate that in their present state, both
installations would make significant losses, with multi-million-pound negative net present
values.

- Sale price and capital costs have the greatest impact on the net present values of both
installations, with sale price an overwhelmingly dominant factor in the success or failure of
the proposed 3 MW installation.

- Both projects offer very low levels of confidence, and as such large ranges of potential
outcomes, from £30m losses to multi-million profits.

- Increasing the scale of an installation, even if only to three devices, introduces the possibility
of profitability, estimated at 14% of scenarios. However, this is heavily dependent on
favourable CfD negotiation with the UK government.

- Levelised costs of energy for a single device is decidedly infeasible, with a value of
£482/MWh. A deployment of three devices is more in line with typical LCOEs in the tidal
industry, with a production cost of £280/MWh.

- Increasing scale of installation does decrease cost of energy, however, the pure material
costs for the SR2000 device prevent major reductions, at small to medium scales.

6.14 Recommendations for future work


- Primary data collection, alongside discourse with potential suppliers, installers, ports etc. are
imperative for reducing uncertainty in all constituent cost elements.

- This would allow a more confident, defensible levelised cost of energy to be ascertained, an
important basis upon which CfD negotiations and potential investment would be based.

- Further study into the possibility of increasing the installation scale to ten devices and
beyond, utilising the lower per unit cost, and potential for more cost-efficient mooring and
power configurations, installation and O&M strategies etc.

- Long term, with an aim to produce energy at competitive (sub-£100) LCOEs, the SR2000
device itself would have to be refined, as the volume of steel structure begets an unyielding
base cap on any LCOE reductions.

114
7 Project Conclusions
The scoping study presented in this document covers all relevant areas including the resource and
wave climate across the site and other important site information, turbine selection and mooring
design, installation and O&M procedures. The individual tasks are concluded by culminating in an
economic assessment.

The site assessment showed the sea bed at the site is rocky, uneven, and primarily consisting of old
red sandstone. The tidal resource is strong across the Pentland firth, and in places the tidal flow
would have frequency exceeded the cut-out speed of the device. The wave climate reduces in
magnitude towards the east, increasing potential power consumption and installation windows.
Water depth, Underwater obstacles and shipwrecks, protected areas, existing marine licences and
the shipping lanes were also considered when choosing the site; allowing the selection of the
following section following location: 58.698, -2.966.

Having selected a site, a resource assessment of the specific site was undertaken in order to
calculate the AEP, Data from the FoW was used to predict the available extracted power in the
Pentland Firth. Both locations have a very similar tidal resource and therefor no correction factor
was applied. The results suggest an annual average extractable power within a region of 0.587 MW
and 0.88 MW per device, and an annual extractable energy yield of between 5.15 GWh and 7.72
GWh per a device. Due to the methodology used it is likely that this is an under estimation, which
was confirmed in comparison with the UK renewable Atlas. The flow is bi-directional, flowing
through the channel in both directions and has a sheer profile similar in shape but larger in
magnitude to that in A-11.

The power curve for Cornett’s generic device was used to calculate the AEP figures from the raw
tidal data. peak current speed is not expected to typically exceed 4 ms -1 which is below the cut-out
speed for the SR2000, while the mean current speeds are within a high efficiency area of the generic
device, suggesting that the majority of the flow regime is extractable. This report recommends that
the results from this section are considered a highly conservative estimate.

The extreme wave was calculated to inform the mooring design. The resulting wave was 16 meters
in height at the site, however significant uncertainty was present here as modelled data and
measured data had an error of ≈20%.

A preliminary design of a mooring system was developed to allow for cost estimation. The design
followed DNV standards; it was designed for the 100 year wave with FOS of 3. The design utilised a
taught mooring and features a 2 part rope comprising of a short section of chain at the anchor
points, with the remaining mooring line made up of synthetic rope. For the cost assessment, the
outputs from a study utilising industry standard tool, OrcaFlex were used. A total length of the 370m
and a diameter of 0.25m. This mooring design resulted in a total packing density of 64MW/km2,
which was found to be close comparison to similar figures generated by the device manufacturer.
The anchor was designed in consultation with the bathymetry, which due to its hard nature and
unevenness eliminated gravity anchors, and suction or drilled piles. Therefore a single drilled pin pile
of 20 meters in length is suggested for each mooring line.

A preliminary study of the electrical connection was undertaken to assess the costs and other
important factors. The length of cable is dependent on the connection point assigned by the DNO

115
however, it is unlikely to be more than 10 km and therefore the cost would not exceed 20 million
pounds. The local 132 KV transmission line is constrained, and therefore the cost of grid connection
is highly dependent on whether this line is upgraded by the DNO to meet the local demand. If this is
the case the onshore electrical connection would cost 6.5 million pounds. Alternatively, if this is not
the case and the entire upgrade is paid for by the developer, this would cost in excess of 30 million
pounds. It has been recommended that directional drilling is used to bring the offshore cables
onshore, and that they are sized to accommodate a future increase in generating capacity.

Based on the above simulation it is seen that summer installation offers benefits due to better
weather conditions. Even though winter installation is cheaper than summer installation, there are
tremendous risks involved with it. Delays in tasks could have knock on effects which in turn may
increase the installation costs.

Analysis of reliability data was conducted to provide a optimisation of O&M strategies based on the
device specific subsystems for tidal stream arrays. Due to lack of data availability, surrogate data
from the offshore wind industry is used to perform this analysis for the SR2000. The results suggest
that 80% of failures will be minor and repairable with a small RIB vessel. A Mermaid analysis was
conducted; it suggested that minor failures are repaired during summer where possible as there is
better site access availability. Larger failures that require the device to be towed to shore are
predicted to occur 9 times throughout device lifetime. Due to the SR2000 design these operations
will not require heavy life vessels, thus significantly lowering the O&M cost. It is recommended that
maintenance and inspection is carried out quarterly to minimise the risk of a major failure. The
Mermaid analysis was also used to assess the impact of weather conditions on the offshore
operations throughout the year. The results show that any major replacement required from
January to March is likely to double the operation cost due to small available weather windows.
Minor repairs are less affected by the adverse weather conditions due to their smaller operation
time offshore.

An economic analysis was performed utilizing cost data from predeceasing parts of the report. This
analysis concluded extensively that the installation of an SR2000 in the Pentland Firth would make
significant loss in the current economic landscape. This is magnified when only a single device is
installed due to the required distribution Grid Upgrade. A sensitivity study showed that sale price
and capital cost have the most impact on NPV, with the latter being more dominant. In the best
case, where three devices are installed, a Monte Carlo analysis estimated a 14% chance that the
project would turn a profit, while potential losses amount to 30 million in the worst case. The LCOE
was estimated at £280 for the installation of three devices. Finally, the results show that while this
value can be reduced a little by increasing the number of devices, the material cost of the device
limits the impact this has. Therefore, the final conclusion that may be drawn from this study is that
unless a CFD strike price or similar financial assistance be implemented in the near future it would be
unwise to advance this project in the current economic climate.

116
8 References
 ABPmer (2018) UK Renewable Atlas. Available at: http://www.renewables-atlas.info/explore-the-atlas/
(Accessed: 14 March 2018).
 Adcock, T. A. A. et al. (2013) ‘The available power from tidal stream turbines in the Pentland Firth’,
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 469(2157), pp.
20130072–20130072. doi: 10.1098/rspa.2013.0072.
 Alexander, 2003. Application of Monte Carlo simulation to system reliability analysis. 20th Int. Pump User
Symp., 91-94.
 Allan, G. et al. (2011) ‘Levelised costs of Wave and Tidal energy in the UK: Cost competitiveness and the
importance of “banded” Renewables Obligation Certificates’, Energy Policy. Elsevier, 39(1), pp. 23–39. doi:
10.1016/J.ENPOL.2010.08.029.
 All-energy.co.uk. (2018). [online] Available at: https://www.all-energy.co.uk/RXUK/RXUK_All-
Energy/2016/Presentations%202016%20Day%202/UK%20Marine%202/Michael%20McSherry.pdf?v=6359
96132146582515 [Accessed 20 Mar. 2018].
 Astariz, S., Vazquez, A. and Iglesias, G. (2015) ‘Evaluation and comparison of the levelized cost of tidal,
wave, and offshore wind energy’, Journal of Renewable and Sustainable Energy, 7(5), p. 53112. doi:
10.1063/1.4932154.
 Bank of England (2018) Bank of England. Available at: https://www.bankofengland.co.uk/monetary-
policy/inflation (Accessed: 21 March 2018).
 Boon, J., Green, M. and Suh, K. (1996). Bimodal wave spectra in lower Chesapeake Bay, sea bed energetics
and sediment transport during winter storms. Continental Shelf Research, 16(15), pp.1965-1988.
 Boronowski S.1, Monahan K.2 and van Kooten G.C.3. The Economics of Tidal Stream Power [Internet].
1University of Victoria, Department of Mechanical Engineering, Victoria, Canada 2 University of Victoria,
Department of Economics, Victoria, Canada 3 University of Victoria, Department of Economics, Victoria,
Canada; 2012. Available from: https://ageconsearch.umn.edu/bitstream/44260/2/403.pdf
 Bruce, D. and Nicholson, P. (1998). The Practice and Application Of Pin piling. [online] Available at:
http://www.geosystemsbruce.com/v20/biblio/033%20The%20Practice%20and%20Application%20of%20P
in%20Piling.pdf [Accessed 20 Mar. 2018].
 Bryden, I. G. et al. (2007) ‘Tidal current resource assessment’, Proceedings of the Institution of Mechanical
Engineers, Part A: Journal of Power and Energy, 221(2), pp. 125–135. doi: 10.1243/09576509JPE238.
 Cahill, B. and Lewis, T. (2013). Wave energy resource characterisation of the Atlantic Marine Energy Test
Site. International Journal of Marine Energy, 1, pp.3-15.
 Carbon Trust (2005) ‘Capital, operational and maintenance costs’. Available at:
http://www.carbontrust.co.uk/emerging-technologies/current-focus-areas/marine-energy-
accelerator/pages/default.aspx (Accessed: 21 March 2018).
 Carbon Trust (2006) ‘Cost estimation methodology’, Carbon, (May). Available at:
http://www.carbontrust.co.uk/emerging-technologies/current-focus-areas/marine-energy-
accelerator/pages/default.aspx (Accessed: 21 March 2018).
 Carbon Trust (2018) Marine energy (wave & tidal stream energy) - Carbon Trust. Available at:
https://www.carbontrust.com/resources/reports/technology/marine-energy/ (Accessed: 21 March 2018).
 Carrol, J., McDonald, A., McMillan, D. (2015). Failure rate, repair time and unscheduled O&M cost analysis
of offshore wind turbines. Wind Energy, 16(6), pp.1107-1119.
 Catapult Offshore Renewable Energy (2014) ‘Financing solutions for wave and tidal energy’, (November).
Available at: https://ore.catapult.org.uk/documents/10619/150618/pdf/6d8ad914-ecc4-43e1-b4a9-
55082d77fb04 (Accessed: 21 March 2018).
 Codecogs.com. (2018). JONSWAP - Spectra - Waves - Fluid Mechanics - Engineering in C, C++. [online]
Available at: http://www.codecogs.com/library/engineering/fluid_mechanics/waves/spectra/jonswap.php
[Accessed 14 Mar. 2018].

117
 Colby, J. (2012) ‘Marine Renewables’. Available at: http://www.iec.ch/etech/2012/etech_1112/tech-
2.htm.
 Cornett, A. (2006) ‘OES-IA Guidance on Assessing Tidal Current Energy Resources’, Energy, pp. 1–7.
 Cto, M. H. (2012) ‘Technology Update – May 2012’, (May).
 Delorm, T.M., 2014. Tidal stream devices: reliability prediction models during their conceptual &
development phases. Durham Thesis, Durham University. Available at: Durham E-Theses Online:
<http://etheses.dur.ac.uk/9482/>.
 Design and analysis of station turbine systems for floating structures. American petroleum industry; 2015.
 DNV (2010) ‘Offshore Standard DNV-OS-E-301 - Position Mooring’, Det Norske Veritas, (DNV-OS-E301), p.
100.
 DNV (2016) ‘RP-0360 Subsea power cables in shallow water’, (June).
 DNV C205, G. (2010) ‘Environmental conditions and environmental testing’, Proceedings of Intelec 93:
15th International Telecommunications Energy Conference, 2(October), pp. 92–99. doi:
10.1109/INTLEC.1993.388591.
 Draper, S. et al. (2014) ‘Estimate of the tidal stream power resource of the Pentland Firth’, Renewable
Energy. Elsevier Ltd, 63, pp. 650–657. doi: 10.1016/j.renene.2013.10.015.
 Easton, M. C., Woolf, D. K. and Bowyer, P. A. (2012) ‘The dynamics of an energetic tidal channel, the
Pentland Firth, Scotland’, Continental Shelf Research. Elsevier, 48, pp. 50–60. doi:
10.1016/j.csr.2012.08.009.
 EMEC, 2018. Scotrenewables Tidal Power. [online] Available at: < http://www.emec.org.uk/about-us/our-
tidal-clients/scotrenewables/> [Accessed 9 March 2018].
 Emec.org.uk. (2018). Blog: Extreme weather generates 14m wave at Billia Croo : EMEC: European Marine
Energy Centre. [online] Available at: http://www.emec.org.uk/blog-extreme-weather-generates-14m-
wave-at-billia-croo/ [Accessed 13 Jan. 2013]
 Ernst & Young LLP (2010) ‘Cost of and financial support for wave, tidal stream and tidal range generation
in the UK’. Available at:
http://webarchive.nationalarchives.gov.uk/20121205081857/http://www.decc.gov.uk/assets/decc/what
we do/uk energy supply/energy mix/renewable energy/explained/wave_tidal/798-cost-of-and-finacial-
support-for-wave-tidal-strea.pdf (Accessed: 21 March 2018).
 Faraci, V. (2006). Calculating failure rates of series/parallel networks. Journal of the System Reliability
Centre.
 Foundations and Moorings for Tidal Stream Systems. Carbon Trust [Internet]. 2009 Sep; Available from:
https://www.carbontrust.com/media/170419/foundations-and-moorings-tidal-systems.pdf
 GL Garrad Hassan (2013) ‘A guide to UK offshore wind operations and maintenance’, Scottish Enterprise
and The Crown Estate, p. 42. Available at: https://www.thecrownestate.co.uk/media/5419/ei-km-in-om-
om-062013-guide-to-uk-offshore-wind-operations-and-maintenance.pdf (Accessed: 21 March 2018).
 Gōda, Y. (2010). Random seas and design of maritime structures. Singapore: World Scientific, p.380.
 Grubb, M. and Newbery, D. (2018) UK Electricity Market Reform and the Energy Transition: Emerging
Lessons. Available at:
https://nicholasinstitute.duke.edu/sites/default/files/UK_Electricity_Market_Reform_and_the_Energy_Tr
ansition_Michael_Grubb.pdf (Accessed: 21 March 2018).
 Hamilton, M. (2012) Technology Update – May 2012. Available at: http://www.all-
energy.co.uk/__novadocuments/14981 (Accessed: 21 March 2018).
 Harnois, V. et al. (2010) ‘Numerical model validation for mooring systems: Method and application for
wave energy converters’, Renewable Energy, 75, pp. 869–887. doi: 10.1016/j.renene.2014.10.063.
 Hoffman, M., 2011. A review of decision support models for offshore wind farms with an emphasis on
operation and maintenance strategies. Wind Engineering, 35, pp.1-16.
 Installing the SABELLA D10 Tidal turbine [Internet]. Available from:
https://www.youtube.com/watch?v=zo47IesiC3c

118
 Joe Barnaby, Stuart Campbell, Thomas Clyde, Shane Crowley, Alexandros Drymonakos, Benjamin
Fisher,Francesca Ford, Jesse Gyan, Blair Gordon, Kerry Hayes, Hamish Kerr, Seumas MacKenzie, Harry
Newton, Edward Olivier, Shaun Rafferty, Adam Roberts, Giovanni, Rosato, Elizabeth Rudd, Matthew
Rundle, Olusesi Tajudeen, Joseph Wellard, Richard Wheal, and Thomas van Lanschot. Offshore Renewable
Energy for Guernsey. 2012 Dec; Available from:
http://www.guernseyrenewableenergy.com/documents/managed/Offshore%20Renewable%20Energy%2
0for%20Guernsey.pdf
 Karimirad, M. et al. (2013) ‘Applicability of offshore mooring and foundation technologies for marine
renewable energy ( MRE ) device arrays’, (Carrell 2011), pp. 1–8.
 Lawrence, J., . Kofoed-Hansen, H. and Chevalier, C. (2013). High-resolution metocean modelling at EMEC’s
(UK) marine energy test sites. [online] pp.2,3,5. Available at:
http://www.homepages.ed.ac.uk/shs/Wave%20Energy/EWTEC%202009/EWTEC%202009%20(D)/papers/
272.pdf [Accessed 12 Mar. 2018].
 Lcicg (2012) ‘Carbon Innovation Coordination Group Technology Innovation Needs Assessment ( TINA )
Marine Energy Summary Report’, Energy, (August). Available at:
https://www.carbontrust.com/media/168547/tina-marine-energy-summary-report.pdf (Accessed: 21
March 2018).
 Lewis, M. et al. (2017) ‘Characteristics of the velocity profile at tidal-stream energy sites’, Renewable
Energy. Elsevier Ltd, 114, pp. 258–272. doi: 10.1016/j.renene.2017.03.096.
 Lewis, M., Neill, S., Hashemi, M. and Reza, M. (2014). Realistic wave conditions and their influence on
quantifying the tidal stream energy resource. Applied Energy, 136, pp.495-508.
 M. Willis, I. Masters, S. Thomas, R. Gallie, J. Loman, A. Cook, R. Ahmadian, R. Falconer, B. Lin, G. Gao, M.
Cross, N. Croft, A. Williams, M. Muhasilovic, I. Horsfall, R. Fidler, C. Wooldridge, I. Fryett, P. Evans,, T.
O’Doherty, D. O’Doherty and A. Mason-Jones. Tidal turbine deployment in the Bristol Channel: a case
study. 2010 Dec;
 Marine.gov.scot. (2018). Pentland Firth Bathymetry 2009 | Marine Scotland Information. [online]
Available at: http://marine.gov.scot/information/pentland-firth-bathymetry-2009 [Accessed 20 Mar.
2018].
 Marinetraffic.com. (2018). MarineTraffic: Global Ship Tracking Intelligence | AIS Marine Traffic. [online]
Available at: https://www.marinetraffic.com/en/ais/home/centerx:-3.1/centery:58.7/zoom:9 [Accessed
14 Mar. 2018].
 Martin-Short, R. et al. (2015) ‘Tidal resource extraction in the Pentland Firth, UK: Potential impacts on flow
regime and sediment transport in the Inner Sound of Stroma’, Renewable Energy. Elsevier Ltd, 76, pp.
596–607. doi: 10.1016/j.renene.2014.11.079.
 Martin-Short, R., Hill, J., Kramer, S., Avdis, A., Allison, P. and Piggott, M. (2015). Tidal resource extraction in
the Pentland Firth, UK: Potential impacts on flow regime and sediment transport in the Inner Sound of
Stroma. Renewable Energy, 76, pp.596-607.
 Mathiesen, M., Goda, Y., Hawkes, P., Mansard, E., Martín, M., Peltier, E., Thompson, E. and Van Vledder,
G. (1994). Recommended practice for extreme wave analysis. Journal of Hydraulic Research, 32(6),
pp.803-814.
 Meygen Ltd (2017) ‘Lessons Learnt from MeyGen Phase 1a Part 1/3: Design Phase Lessons Learnt from
MeyGen Phase 1a Part 1/3: Design Phase LESSONS LEARNT FROM MEYGEN PHASE 1A. PART 1/3: DESIGN
PHASE CONTENTS’. Available at: https://tethys.pnnl.gov/sites/default/files/publications/MeyGen-2017-
Part1.pdf (Accessed: 21 March 2018).
 Mygen Tidal Energy project phase 1 [Internet]. Mygen; 2010. Available from:
https://www.iema.net/assets/nts/Xodus/Meygen_Tidal_Energy_Project_NTS_August_2011.pdf
 National Grid (2015) ‘Electricity Ten Year Statement 2015’, UK electricity transmission, (November), pp. 1–
145. doi: 10.1016/B978-0-7506-7766-0.50021-X.

119
 Neill, S. P. et al. (2017) ‘The wave and tidal resource of Scotland’, Renewable Energy, 114, pp. 3–17. doi:
10.1016/j.renene.2017.03.027.
 Neill, S., Vögler, A., Goward-Brown, A., Baston, S., Lewis, M., Gillibrand, P., Waldman, S. and Woolf, D.
(2017). The wave and tidal resource of Scotland. Renewable Energy, 114, pp.3-17.
 Orkney 3. EMEC Tidal Power | Fully Charged [Internet]. Available from:
https://www.youtube.com/watch?v=YEQQl-qpkCc&t=323s
 Parsons Brinckerhoff (2012) ‘Pentland Firth and Orkney Waters Onshore Infrastructure Information Note’,
The Crown Estate.
 Paterson, J., Amico, F.D., Thies, P.R., Kurt, R.E., Harrison, G. (2018). Offshore wind installation vessels – A
comparative assessment for UK offshore rounds 1 and 2. Ocean Engineering, 148, pp.637-649.
 Planning, I. and Forms, P. (2014) ‘Cable Details and Grid Connection Statement’, 6(March).
 Pye, D. (2014) ‘Subsea Connectors in Marine Renewable Energy’, Icoe, (September).
 Research.dnv.com. (2018). DNV Ocean *283*# ~ JONSWAP Spectrum.. [online] Available at:
http://research.dnv.com/hci/ocean/bk/c/a28/s3.htm [Accessed 13 Mar. 2018].
 Ridge, I. M. L., Banfield, S. J. and Mackay, J. (2010) ‘Nylon fibre rope moorings for wave energy
converters’, in OCEANS 2010 MTS/IEEE SEATTLE. IEEE, pp. 1–10. doi: 10.1109/OCEANS.2010.5663836.
 Rinaldi, G., Thies P.R., Johanning, L., Walker, R. (2017). A decision support model to optimise the operation
and maintenance strategies of an offshore renewable energy farm. Ocean Engineering, 145, pp. 250-262.
 Rinaldi, G., Thies, P.R., Johanning, L., Walker, R.T. (2016). A Computational Tool for the Pro-Active
Management of Offshore Farms. University of Exeter & Mojo Maritime Ltd.
 Saruwatari, A., Ingram, D. and Cradden, L. (2013). Wave-current interaction effects on marine energy
converters. Ocean Engineering, 73, pp.106-118.
 Scot Renewables, 2018. SR2000. [online] Available at: <http://www.scotrenewables.com/technology-
development/sr2000> [Accessed 25 February 2018].
 Scotrenewables (2018) SR2000. Available at: http://www.scotrenewables.com/technology-
development/sr2000 (Accessed: 21 March 2018).
 Scotrenewables (n.d). Enabling Low Cost Maintenance. [online] Available at: <https://www.sut.org/wp-
content/uploads/2017/10/SUT_171102-Presentation.pdf> [Accessed 15 March 2018].
 Scotrenewables Tidal Power complete deployment of advanced modular anchoring system. 2017 Mar;
Available from: http://www.scotrenewables.com/news/89-scotrenewables-tidal-power-ltd-srtp-complete-
deployment-of-advanced-modular-anchoring-system
 Scotrenewables, 2012. Technology Update – May 2012. [online] Available at: <http://www.all-
energy.co.uk/__novadocuments/14981> [Accessed 5 March 2018].
 Scott, N. C. et al. (2009) ‘Pentland Firth Tidal Energy Project Grid Options Study Prepared for : Highlands
and Islands Enterprise’, (January), p. 135.
 Scottish and Southern energy networks. (2018). Generation availability map. [online] Available at:
https://www.ssepd.co.uk/GenerationAvailabilityMap/?mapareaid=2 [Accessed 14 Mar. 2018].
 Scottish Enterprise (2014) Renewable Energy Investment Fund Q&A. Available at: http://www.scottish-
enterprise.com/services/attract-investment/renewable-energy-investment-fund/overview (Accessed: 21
March 2018).
 Scottish Government (2014) £6 million for Scotland’s wave and tidal industry, Scottish Government.
Available at: https://news.gov.scot/news/6-million-for-scotlands-wave-and-tidal-industry (Accessed: 21
March 2018).
 Smith, H. (2016) ‘Tidal Anaylisis workshop’. University of Exeter.
 SR2000 Tidal turbine launch [Internet]. Available from: https://www.youtube.com/watch?v=AXXibP2CI5o
 SR2000-1 turbine connection [Internet]. Available from:
https://www.youtube.com/watch?v=bZIRVKHWEY8&t=67s

120
 SR250 Full Power October 2013 [Internet]. 2013. Available from:
https://www.youtube.com/watch?v=oYuCH42atCY
 Standard, O. (2003) ‘Electrical Installations Foreword’, (January).
 Starling, M. and Scott, A. (2009) ‘Foundations and Moorings for Tidal Stream Systems’, Carbon Trust,
(September). Available at: http://www.carbontrust.com/media/170419/foundations-and-moorings-tidal-
systems.pdf.
 Starling, M., Scott, A. and Parkinson, R. (2009). Foundations and Moorings for Tidal Stream Systems.
[online] The Carbon Trust. Available at: https://www.carbontrust.com/media/170419/foundations-and-
moorings-tidal-systems.pdf [Accessed 20 Mar. 2018].
 Stegman, A. et al. (2017) ‘Exploring Marine Energy Potential in the UK Using a Whole Systems Modelling
Approach’, Energies. Multidisciplinary Digital Publishing Institute, 10(12), p. 1251. doi:
10.3390/en10091251.
 Taaffe, D. and Manager, P. (2016) ‘MeyGen Tidal Energy Project Phase 1a – Progress Update’, (May).
 Technology, N. G. and Hamilton, T. M. (no date) ‘Scotrenewables Floating Tidal Energy Next Generation
Technology … Today Mark Hamilton - CTO Scotrenewables – Founded 2002 in Orkney’.
 The Crown Estate, 2013. A Guide to UK Offshore Wind Operations and Maintenace. [online] Available at:
<https://www.thecrownestate.co.uk/media/5419/ei-km-in-om-om-062013-guide-to-uk-offshore-wind-
operations-and-maintenance.pdf>.
 Thies P.R., Smith, H.G., Johanning, L. (2012). Addressing failure rate uncertainties of marine energy
converters. Renewable Energy, 44, pp.359-367.
 Tidal Energy Today (2017) MeyGen aims at half-priced tidal CfD | Tidal Energy Today, Tidal Energy Today.
Available at: https://tidalenergytoday.com/2017/09/20/meygen-aims-at-half-priced-tidal-cfd/ (Accessed:
21 March 2018).
 Tidal Energy Today (2018) Scottish National Investment Bank to benefit tidal & wave? | Tidal Energy
Today, Tidal Energy Today. Available at: https://tidalenergytoday.com/2018/02/28/scottish-national-
investment-bank-to-benefit-tidal-wave/ (Accessed: 21 March 2018).
 Vazquez, A. and Iglesias, G. (2015) ‘LCOE (levelised cost of energy) mapping: A new geospatial tool for tidal
stream energy’, Energy. Pergamon, 91, pp. 192–201. doi: 10.1016/J.ENERGY.2015.08.012.
 Weller, S. D. et al. (2015) ‘Synthetic mooring ropes for marine renewable energy applications’, Renewable
Energy, pp. 1268–1278. doi: 10.1016/j.renene.2015.03.058.
 Whiterow, P. (2017) Atlantis Resources Ltd confident of CfD deal for MeyGen tidal power project,
Proactive Investors. Available at: http://www.proactiveinvestors.co.uk/companies/news/185900/atlantis-
resources-confident-of-cfd-deal-for-meygen-tidal-power-project-185900.html (Accessed: 21 March 2018).
 Wikiwaves.org. (2018). Ocean-Wave Spectra - WikiWaves. [online] Available at:
https://www.wikiwaves.org/Ocean-Wave_Spectra [Accessed 13 Mar. 2018
 World Energy Council (2016) ‘World Energy Resources: Marine Energy 2016’, p. 79. doi:
http://www.worldenergy.org/wp-content/uploads/2013/09/Complete_WER_2013_Survey.pdf.
 www.gov.uk (2018) Corporation Tax Rates and Reliefs. Available at: https://www.gov.uk/corporation-tax-
rates (Accessed: 21 March 2018).

121
9 Appendix A

Table 9-1: Failure rate sub-assembly breakdown adapted from Delorm (2014).

Sub-System Sub-Assembly failure rate (λ)


(failures/year)
Drive Train Rotor blades (fixed pitch) 0.874
Hub 0.25
Shaft seal 0.123
Main shaft, main bearing, couplings 0.11
Brake system 0.459
PM Generator 0.46
Rectifier AC/DC 0.051
Inverter DC/AC and control system 0.366
Generator circuit breaker 0.175
Electrical System Generator transmission cable 0.01
Transformer and cooling system 0.081
400V/11kV)
Main circuit breaker 0.175
Isolator switch 0.037
LV DC Uninterruptible Electrical LV load-break switch 0.01
Supply
Converter AC/DC 0.015
LV circuit breaker 0.011
Battery 0.098
LV DC cables 0.02
Grid Connection sub-sea connector 0.018
Umbilical electrical cable 0.222
Umbolical fibre optic cable 0.052
Ancillary System Heat exchanger 0.07
Ventilation system 0.105
Control and Management Programmable controller 0.63
Turbine Controller 0.328
Environment controller 0.117
SCADA 0.754
Corrosion Protection Corrosion control system 0.115
Hydraulic system Hydraulic cylinders 0.079
Hydraulic power pack 0.039
Moving Structures Nacelle Structure 0.018
Retractable leg 0.018
Floatation Tube 0.001
Fixed Structure Buoy 0.05
Mooring system 0.171
Yoke system 0.05

122
Table 8.2: Failure rate correlation Delorm (2014) and Carrol et al. (2015).

Category Delorm (2014) Carrol et al.


failure rate (2015) failure
(failures/year) rate
(failures/year)
hub 0.25 0.221
blades 0.874 0.467
generator 0.635 0.901
main circuit breaker 0.175 0.382
main shaft, seals, 0.233 0.221
bearing
brake system 0.459 0.377
DC/DC converter 0.417 0.162
transformer 0.081 0.056
electrical components 0.376 0.414
controls 1.075 0.754
structure and mooring 0.323 0.181
ancillary system 0.175 0.197

Table 9-2: Failure rates split into categories: major replacement, major repair and minor repair.

percentage likelihood of failure from each repair category (Carrol et


al.)
Category major major minor
replacement repair repair
hub 0.5 17.2 82.4
blades 0.2 2.1 97.6
generator 10.5 35.6 53.8
main circuit breaker 0.5 14.1 85.3
main shaft, seals, 0.5 17.2 82.4
bearing
brake system 0.0 1.1 98.9
DC/DC converter 3.1 50.0 46.9
transformer 1.8 5.4 92.9
electrical components 0.5 13.0 86.5
controls 0.1 16.8 83.1
structure and mooring 0.0 49.2 50.8
ancillary system 0.0 3.6 96.4

123
Table 9-3: Failure rate over lifetime (20 years)

failure rate LIFETIME per category (=failure rate *20 * percentage


likelihood of failure from repair category)
Category major major minor
replacement repair repair
hub 0.023 0.860 4.118
blades 0.037 0.374 17.068
generator 1.339 4.525 6.836
main circuit breaker 0.018 0.495 2.987
main shaft, seals, 0.021 0.801 3.838
bearing
brake system 0.000 0.097 9.083
DC/DC converter 0.257 4.170 3.913
transformer 0.029 0.087 1.504
electrical components 0.036 0.981 6.503
controls 0.022 3.603 17.870
structure and mooring 0.000 3.176 3.284
ancillary system 0.000 0.124 3.376

Table 9-4: MTTR

MTTR (Carrol et al., 2015)


Category major major minor
replacement repair repair
hub 298 40 10
blades 288 21 9
generator 81 24 7
main circuit breaker 150 19 4
main shaft, seals, 298 40 10
bearing
brake system 0 7 2
DC/DC converter 57 14 7
transformer 1 26 7
electrical components 18 14 5
controls 12 14 8
structure and mooring 0 2 5
ancillary system 0 14 5

124
Table 9-5: MTTR over 20-year lifetime

MTTR (Time taken to repair unexpected failures over lifetime)


technician time only = failure rate LIFETIME per category * MTTR
Category major major minor
replacement repair repair
hub 6.7 34.4 41.2
blades 10.8 7.9 153.6
generator 108.5 108.6 47.9
main circuit breaker 2.7 9.4 11.9
main shaft, seals, 6.3 32.1 38.4
bearing
brake system 0.0 0.7 18.2
DC/DC converter 14.7 58.4 27.4
transformer 0.0 2.3 10.5
electrical components 0.7 13.7 32.5
controls 0.3 50.4 143.0
structure and mooring 0.0 6.4 16.4
ancillary system 0.0 1.7 16.9

Table 9-6: MTTR over 20-year lifetime

Av. Material cost per failure (euros)


Category major major minor
replacement repair repair
hub 95000 1500 160
blades 90000 1500 170
generator 60000 3500 160
main circuit 13500 2300 260
breaker
main shaft, 95000 1500 160
seals, bearing
brake system 0 2400 130
DC/DC converter 13000 5300 240
transformer 70000 2300 95
electrical 12000 2000 100
components
controls 13000 2000 200
structure and 0 1100 140
mooring
ancillary system 0 1300 465

125
Table 9-7: Average material cost over lifetime

Exchange Rate 0.88


Av cost (EUROS)over LIFETIME for repair (material
cost only) = failure rate LIFETIME per category * Av.
Cost per failure
Category major major repair minor repair
replacement
hub 1891 1135 580
blades 2964 494 2553
generator 70703 13936 963
main circuit 218 1001 683
breaker
main shaft, 1763 1058 540
seals,
bearing
brake 0 206 1039
system
DC/DC 2945 19449 826
converter
transformer 1782 176 126
electrical 384 1726 572
components
controls 246 6342 3145
structure 0 3075 405
and
mooring
ancillary 0 142 1381
system

Totals 82895 48739 12814

126
10 Appendix B

Table 10-1 Winter cost data (installation)

Name Unwea Min Max Mea Med P10 P25 P75 P90 P75 Max
thered n ian - - Min
P25
drillship Overall Cost (GBP) 128259 2257 8459 5650 5145 2417 3858 7984 8419 4125 6201
.4 59.4 09.4 99.4 84.4 59.4 71.9 09.4 09.4 37.5 50
drillship Day Rate (GBP) 122500 1400 1575 1452 1400 1400 1400 1575 1575 1750 1750
00 00 50 00 00 00 00 00 0 0
drillship Standby Rate (GBP) 0 8000 7000 4140 3600 9600 2400 6525 6960 4125 6200
0 00 00 00 0 00 00 00 00 00
drillship Port Fees (GBP) 5000 5000 5000 5000 5000 5000 5000 5000 5000 0 0
drillship Accommodation Fees (GBP) 450 450 600 540 600 450 450 600 600 150 150
drillship Fuel Fees (GBP) 309.4 309. 309. 309. 309. 309. 309. 309. 309. 0 0
4 4 4 4 4 4 4 4
drillship Working Offshore Duration 3 3 3 3 3 3 3 3 3 0 0
(Days)
drillship Waiting Offshore Duration 0 0 0.23 0.02 0 0 0 0 0.21 0 0.239
(Days) 9583 3958 5625 583
drillship Working Port Duration (Days) 2 2 2 2 2 2 2 2 2 0 0
drillship Waiting Port Duration (Days) 0 9.06 70.9 43 37.9 10.6 25.5 66.4 70.5 40.9 61.87
25 375 4271 6875 7396 7396 7396 5
drillship Transit Duration (Days) 0.8333 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0 0
33 3333 3333 3333 3333 3333 3333 3333 3333
drillship Offshore Arrivals 1 1 1 1 1 1 1 1 1 0 0
drillship Transits 2 2 2 2 2 2 2 2 2 0 0
drillship Port Calls at scrabster 2 2 2 2 2 2 2 2 2 0 0
drillship Port Calls at substation 0 0 0 0 0 0 0 0 0 0 0

cable laying vessel Overall Cost (GBP) 198859 2563 9365 6754 7452 2703 4263 8888 9325 4625 6801
.4 59.4 09.4 34.4 59.4 59.4 59.4 96.9 09.4 37.5 50
cable laying vessel Day Rate (GBP) 192500 1925 2275 2100 2100 1942 2100 2100 2257 0 3500
00 00 00 00 50 00 00 50 0
cable laying vessel Standby Rate (GBP) 0 4000 7200 4590 5200 5400 2100 6725 7160 4625 6800
0 00 00 00 0 00 00 00 00 00
cable laying vessel Port Fees (GBP) 5000 5000 5000 5000 5000 5000 5000 5000 5000 0 0
cable laying vessel Accommodation Fees 1050 1050 1200 1125 1125 1050 1050 1200 1200 150 150
(GBP)
cable laying vessel Fuel Fees (GBP) 309.4 309. 309. 309. 309. 309. 309. 309. 309. 0 0
4 4 4 4 4 4 4 4
cable laying vessel Working Offshore 7 7 7 7 7 7 7 7 7 0 0
Duration (Days)
cable laying vessel Waiting Offshore 0 0 0 0 0 0 0 0 0 0 0
Duration (Days)
cable laying vessel Working Port 2 2 2 2 2 2 2 2 2 0 0
Duration (Days)
cable laying vessel Waiting Port 0 4.94 73.6 47.1 53.9 6.34 21.8 68.3 73.1 46.4 68.65
Duration (Days) 7917 0417 3958 4271 6875 9063 099 6771 1927 625
cable laying vessel Transit Duration 0.8333 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0 0
(Days) 33 3333 3333 3333 3333 3333 3333 3333 3333
cable laying vessel Offshore Arrivals 1 1 1 1 1 1 1 1 1 0 0
cable laying vessel Transits 2 2 2 2 2 2 2 2 2 0 0
cable laying vessel Port Calls at scrabster 2 2 2 2 2 2 2 2 2 0 0
cable laying vessel Port Calls at 0 0 0 0 0 0 0 0 0 0 0
substation

Renewable service vessel/ multi cat vessel 411631 4270 1381 9601 1064 4390 5615 1241 1372 6800 9545
Overall Cost (GBP) 68.3 631 60.3 206 10.2 95 631 131 35.9 62.7
Renewable service vessel/ multi cat vessel 275000 2500 3500 2850 2750 2500 2687 3062 3475 3750 1000
Day Rate (GBP) 00 00 00 00 00 50 50 00 0 00

127
Renewable service vessel/ multi cat vessel 120000 4500 1050 6450 7575 6600 2775 9375 1045 6600 1005
Standby Rate (GBP) 0 000 00 00 0 00 00 500 00 000
Renewable service vessel/ multi cat vessel 15000 1500 3000 2850 3000 1650 3000 3000 3000 0 1500
Port Fees (GBP) 0 0 0 0 0 0 0 0 0
Renewable service vessel/ multi cat vessel 1200 1050 1350 1215 1200 1065 1200 1237 1350 37.5 300
Accommodation Fees (GBP) .5
Renewable service vessel/ multi cat vessel 430.95 287. 718. 445. 430. 301. 430. 430. 689. 0 430.9
Fuel Fees (GBP) 3 25 315 95 665 95 95 52 5
Renewable service vessel/ multi cat vessel 7.25 7.25 7.25 7.25 7.25 7.25 7.25 7.25 7.25 0 0
Working Offshore Duration (Days)
Renewable service vessel/ multi cat vessel 0.4166 0 0.62 0.11 0 0 0 0.20 0.59 0.20 0.625
Waiting Offshore Duration (Days) 67 5 7708 0521 2708 0521
Renewable service vessel/ multi cat vessel 0.0625 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0 0
Working Port Duration (Days) 25 25 25 25 25 25 25 25
Renewable service vessel/ multi cat vessel 9.2187 6.19 72.6 44.6 51.4 7.42 19.0 63.5 72.1 44.4 66.41
Waiting Port Duration (Days) 5 7917 1458 0104 4271 7083 5208 2344 125 7135 667
Renewable service vessel/ multi cat vessel 1.0937 0.72 1.82 1.13 1.09 0.76 1.09 1.09 1.75 0 1.093
Transit Duration (Days) 5 9167 2917 0208 375 5625 375 375 75
Renewable service vessel/ multi cat vessel 2 1 3 2 2 1.1 2 2 2.9 0 2
Offshore Arrivals
Renewable service vessel/ multi cat vessel 3 2 5 3.1 3 2.1 3 3 4.8 0 3
Transits
Renewable service vessel/ multi cat vessel 2 2 3 2.1 2 2 2 2 2.9 0 1
Port Calls at scrabster
Renewable service vessel/ multi cat vessel 0 0 0 0 0 0 0 0 0 0 0
Port Calls at substation

All Vessels Overall Cost (GBP) 7387 1069 2873 2200 2552 1079 1399 2744 2872 1345 1804
187 900 694 725 129 451 675 415 223 713
49.8
All Vessels Day Rate (GBP) 5900 6000 7000 6402 6337 6017 6231 6562 6975 3312 1000
00 00 50 50 50 25 50 00 5 00
00
All Vessels Standby Rate (GBP) 1200 3250 2180 1518 1875 3440 7400 2050 2176 1310 1855
00 000 000 000 00 00 000 000 000 000
00
All Vessels Port Fees (GBP) 2500 2500 4000 3850 4000 2650 4000 4000 4000 0 1500
0 0 0 0 0 0 0 0 0
0
All Vessels Accommodation Fees (GBP) 2700 2700 3150 2880 2850 2700 2700 3000 3135 300 450
All Vessels Fuel Fees (GBP) 1049. 906. 1337 1064 1049 920. 1049 1049 1308 0 430.9
1 .05 .115 .75 465 .75 .75 .32 5
75

128
Table 10-2 Summer cost data (installation)

Name Unwea Min Max Mea Med P10 P25 P75 P90 P75 Max
thered n ian - - Min
P25
cable laying vessel Overall Cost (GBP) 281709 2817 4667 3462 2817 2817 2817 4329 4652 1512 1850
.4 09.4 09.4 09.4 09.4 09.4 09.4 59.4 09.4 50 00
cable laying vessel Day Rate (GBP) 275000 2750 3250 2900 2750 2750 2750 3062 3250 3125 5000
00 00 00 00 00 00 50 00 0 0
cable laying vessel Standby Rate (GBP) 0 0 1350 4950 0 0 0 1200 1335 1200 1350
00 0 00 00 00 00
cable laying vessel Port Fees (GBP) 5000 5000 5000 5000 5000 5000 5000 5000 5000 0 0
cable laying vessel Accommodation Fees 1400 1400 1400 1400 1400 1400 1400 1400 1400 0 0
(GBP)
cable laying vessel Fuel Fees (GBP) 309.4 309. 309. 309. 309. 309. 309. 309. 309. 0 0
4 4 4 4 4 4 4 4
cable laying vessel Working Offshore 7 7 7 7 7 7 7 7 7 0 0
Duration (Days)
cable laying vessel Waiting Offshore 0 0 0 0 0 0 0 0 0 0 0
Duration (Days)
cable laying vessel Working Port 2 2 2 2 2 2 2 2 2 0 0
Duration (Days)
cable laying vessel Waiting Port 0 0 11.6 4.02 0 0 0 9.40 11.5 9.40 11.61
Duration (Days) 1458 2917 1042 1667 1042 458
cable laying vessel Transit Duration 0.8333 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0 0
(Days) 33 3333 3333 3333 3333 3333 3333 3333 3333
cable laying vessel Offshore Arrivals 1 1 1 1 1 1 1 1 1 0 0
cable laying vessel Transits 2 2 2 2 2 2 2 2 2 0 0
cable laying vessel Port Calls at scrabster 2 2 2 2 2 2 2 2 2 0 0
cable laying vessel Port Calls at 0 0 0 0 0 0 0 0 0 0 0
substation

Renewable service vessel/ multi cat vessel 417031 4170 5716 4798 4469 4170 4170 5716 5716 1546 1546
Overall Cost (GBP) 31 87.3 73.5 31 31 31 87.3 87.3 56.4 56.4
Renewable service vessel/ multi cat vessel 385000 3150 4200 3605 3850 3150 3150 3850 4165 7000 1050
Day Rate (GBP) 00 00 00 00 00 00 00 00 0 00
Renewable service vessel/ multi cat vessel 0 0 2250 8750 1250 0 0 2250 2250 2250 2250
Standby Rate (GBP) 00 0 0 00 00 00 00
Renewable service vessel/ multi cat vessel 30000 3000 3000 3000 3000 3000 3000 3000 3000 0 0
Port Fees (GBP) 0 0 0 0 0 0 0 0
Renewable service vessel/ multi cat vessel 1600 1400 1600 1500 1500 1400 1400 1600 1600 200 200
Accommodation Fees (GBP)
Renewable service vessel/ multi cat vessel 430.95 287. 430. 373. 430. 287. 287. 430. 430. 143. 143.6
Fuel Fees (GBP) 3 95 49 95 3 3 95 95 65 5
Renewable service vessel/ multi cat vessel 7.25 7 7.25 7.15 7.25 7 7 7.25 7.25 0.25 0.25
Working Offshore Duration (Days)

Renewable service vessel/ multi cat vessel 0.4166 0 0.48 0.25 0.41 0 0 0.41 0.48 0.41 0.489
Waiting Offshore Duration (Days) 67 9583 7292 6667 6667 2292 6667 583
Renewable service vessel/ multi cat vessel 0.0625 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0 0
Working Port Duration (Days) 25 25 25 25 25 25 25 25
Renewable service vessel/ multi cat vessel 1.2187 1.21 9 4.26 2.13 1.21 1.21 9 9 7.78 7.781
Waiting Port Duration (Days) 5 875 1458 5417 875 875 125 25
Renewable service vessel/ multi cat vessel 1.0937 0.72 1.09 0.94 1.09 0.72 0.72 1.09 1.09 0.36 0.364
Transit Duration (Days) 5 9167 375 7917 375 9167 9167 375 375 4583 583
Renewable service vessel/ multi cat vessel 2 1 2 1.6 2 1 1 2 2 1 1
Offshore Arrivals
Renewable service vessel/ multi cat vessel 3 2 3 2.6 3 2 2 3 3 1 1
Transits
Renewable service vessel/ multi cat vessel 2 2 2 2 2 2 2 2 2 0 0
Port Calls at scrabster
Renewable service vessel/ multi cat vessel 0 0 0 0 0 0 0 0 0 0 0
Port Calls at substation

All Vessels Overall Cost (GBP) 6987 6987 1023 8260 7286 6987 6987 9983 1020 2996 3246
40.4 397 82.9 40.4 40.4 40.4 96.7 897 56.4 56.4
40.4

129
All Vessels Day Rate (GBP) 6600 6150 6950 6505 6600 6150 6337 6600 6915 2625 8000
00 00 00 00 00 50 00 00 0 0
00
All Vessels Standby Rate (GBP) 0 0 3450 1370 1250 0 0 3450 3450 3450 3450
00 00 0 00 00 00 00
All Vessels Port Fees (GBP) 3500 3500 3500 3500 3500 3500 3500 3500 3500 0 0
0 0 0 0 0 0 0 0
0
All Vessels Accommodation Fees (GBP) 3000 2800 3000 2900 2900 2800 2800 3000 3000 200 200
All Vessels Fuel Fees (GBP) 740.3 596. 740. 682. 740. 596. 596. 740. 740. 143. 143.6
7 35 89 35 7 7 35 35 65 5
5

130

You might also like