You are on page 1of 11

Energy 269 (2023) 126655

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Evolution of offshore wind resources in Northern Europe under


climate change
A. Martinez a, L. Murphy a, G. Iglesias a, b, *
a
MaREI, Environmental Research Institute & School of Engineering, University College Cork, College Road, Cork, Ireland
b
University of Plymouth, School of Engineering, Marine Building, Drake Circus, Plymouth, PL4 8AA, United Kingdom

A R T I C L E I N F O A B S T R A C T

Handling Editor: Jesse L. The Climate-change impacts on offshore wind resources in Northern Europe over the 21st century are investigated
based on the most up-to-date narratives of societal development and land use, the Shared Socioeconomic
Keywords: Pathways (SSPs). Three scenarios are considered: a high-emissions (pessimistic) scenario, SSP5-8.5; an inter­
Offshore wind energy mediate scenario, SSP2-4.5, in which current emissions do not vary notably; and, for the first time in this type of
Wind power
work, a low-emissions (optimistic) scenario, SSP1-1.9, representing the fulfilment of the ambitious goals of the
Offshore renewable energy
EU’s Green Deal and the Paris Agreement. A multi-model ensemble is constructed with the global climate models
Renewable energy
Climate change that are found to best reproduce the wind climate in Northern Europe. The results anticipate an overall decline in
Multi-model ensemble wind power density, especially in the high-emissions scenario and in certain regions (up to 30% off Western
Shared socioeconomic pathway Ireland), which should be taken into account in planning future offshore wind deployments. As an exception,
slight increases (around 10%) are projected in certain areas of the Baltic Sea. The general decline is less pro­
nounced in the low-emissions scenario. Indeed, the results prove that reducing emissions as advocated by current
climate objectives would not only weaken the declining trend but also lead to a more stable resource.

1. Introduction turbine size, noise nuisance and visual impacts [10–12]. Indeed,
offshore energies have arisen as a great alternative in the context of
Cutting carbon emissions to mitigate climate change and reducing climate change, e.g., wave [13,14] and tidal [13,15,16], of particular
the over-dependence on fossil fuels are listed as a priority in a great interest in achieving energy independence in small islands [17].
number of countries throughout the world. As a result, investments in Particularly, offshore wind offers the unique opportunity of combined
the energy sector are shifting towards renewable energies; among all the exploitation with other renewable energies in co-located structures [18,
renewables, wind energy is experiencing remarkable growth in recent 19], such as wind-wave [20,21] or solar-wind [22,23]. Globally, a re­
years. Globally, the wind energy sector has undergone two record years cord 21.1 GW of new offshore wind capacity was installed in 2021,
in new capacity installations – a total of 93 GW in 2020 [1] and 94 GW in leading to a total of 57.2 GW [2]. Particularly, Europe has been one of
2021 [2]. However, these reports also emphasise the need to quadruple the strongest markets for this technology, accounting for more than 45%
the current growth rate to stay on course for the ambitious goals of the of the global installations – vastly dominated by bottom-fixed in­
Paris Agreement [3] and the European Union Green Deal [4], i.e., stallations in countries bordering the North Sea [24], i.e., the UK, Ger­
net-zero emissions by 2050. many, the Netherlands and Denmark. Furthermore, Northern Atlantic
However, onshore wind presents several drawbacks, e.g., the scarcity Europe has been found to be of particular interest for future floating
and high prices of available land, especially in developed countries, offshore wind farms [25], and a great number of new installations are
noise nuisance and visual impacts. Consequently, the energy sector has expected in the coming years [26].
found new technologies to harvest wind energy in offshore areas. With Importantly, the economic viability and profitability of an offshore
wind turbines mounted on bottom-fixed platforms (jackets [5] or wind farm are determined by the amount of energy that it can produce,
monopiles [6]) or floating platforms (tension leg [7], spar buoy [8] or and thus by the available wind resource [27]. Wind power density being
semi-submersible [9]), the possibility of exploiting the stronger and proportional to the wind speed cubed, it is highly sensitive to variations
steadier offshore winds is combined with the fewer constraints on in the wind climate [28]. Therefore, the effects of climate change on

* Corresponding author. MaREI, Environmental Research Institute & School of Engineering, University College Cork, College Road, Cork, Ireland.
E-mail address: gregorio.iglesias@ucc.ie (G. Iglesias).

https://doi.org/10.1016/j.energy.2023.126655
Received 14 December 2022; Accepted 7 January 2023
Available online 9 January 2023
0360-5442/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
A. Martinez et al. Energy 269 (2023) 126655

Fig. 1. Limits of the study area and its regions.

Table 1
GCMs considered in this work.
Model Centre Resolution Reference
(lat × lon)

FGOALS-f3- Chinese Academy of Sciences 1◦ × 1.25◦ [46]


L (China)
CanESM5 Canadian Centre for Climate 1.775◦ × [47]
Modelling and Analysis (Canada) 2.1825◦
MPI-ESM1- Max Planck Institute for 0.93◦ × [48]
2-HR Meteorology (Germany) 0.9375◦
MPI-ESM1- Max Planck Institute for 1.85◦ × 1.875◦ [49–51]
2-LR Meteorology (Germany)
EC-Earth3 EC-EARTH-CONSORTIUM (Europe) 0.7◦ × 0.7031◦ [52]
IPSL- Institut Pierre Simon Laplace 1.2676◦ × 2.5◦ [53]
CM6A-LR (France) Fig. 2. Normalised bias of the historical data of the multi-model ensemble
MIROC6 JAMSTEC (Japan Agency for 1.4◦ × 1.4063◦ [54] relative to the ERA-5 time series.
Marine-Earth Science and
Technology) (Japan)
MRI-ESM2- Meteorological Research Institute 1.12◦ × 1.125◦ [55]
0 (Japan)
GFDL- NOAA-GFDL (USA) 1◦ × 1.25◦ [56]
ESM4

Table 2
Number of points in the grid statistically similar to their ERA-5 counterparts
(percentage of the total number of points).
Model Number of grid points statistically similar

MRI-ESM2-0 92%
MIROC6 88%
MPI-ESM1-2-LR 86%
EC-Earth3 84%
FGOALS-f3-L 82%
CanESM5 74% Fig. 3. Mean power density (baseline) (Wm-2).
MPI-ESM1-2-HR 72%
IPSL-CM6A-LR 69%
GFDL-ESM4 51% have been initially studied based on scenarios A2 and B2 of cumulative
greenhouse gas (GHG) emissions [31], focusing on the long-term evo­
lution [32]. Later analysis using different regional climate models have
atmospheric patterns may be expected to have significant repercussions been used to evaluate the future wind resource in Europe [33] and its
on wind energy resources [29]. Indeed, recent studies have shown that effects on energy production [34]; however, these studies strongly differ
offshore wind resources are poised to decrease in Northern Europe as a in areas of the North Sea – it appears that these climate predictions are
result of climate change [30]. It follows that to plan new offshore wind subject to great uncertainty. Recently, the 6th phase of the Coupled
projects and operate current installations, a characterisation of future Model Intercomparison Project (CMIP6) [35] has released an extensive
offshore wind energy resources is much needed. number of global climate models (GCMs). Among its activities, the
The effects of climate change on the wind climate in Northern Europe ScenarioMIP develops climate projections by means of the most novel

2
A. Martinez et al. Energy 269 (2023) 126655

Fig. 4. Continuous evolution (%) of the mean power density for climate-change scenarios SSP1-1.9 (left column), SSP2-4.5 (centre column) and SSP5-8.5
(right column).

resource in North America [37], China [38], the UK [39], the North­
western Passage [40] and Europe [41]. In the latter, a large-scale study
predicted relevant reductions in wind power density in the offshore
Atlantic, which were found to depend significantly on the scenario
considered.
Large-scale studies may miss the special characteristics of a specific
area. Particularly, previous studies showed that the performance of the
GCMs is inferior in mountainous areas due to their coarse resolution.
Similarly, GCMs may struggle to model the offshore climate if their
validation was dominated by continental regions. In this sense, an
offshore specific approach is well-justified, and studies on the evolution
of the offshore wind resource have been undertaken for the USA [42]
and, more recently, China [43].
In this work, the evolution of the offshore wind energy resource in
Northern Atlantic Europe in the 21st century is studied by considering
Fig. 5. Coefficient of variation (COV), baseline.
the most novel and complex scenarios of climate change, the SSPs. Three
climate change scenarios are considered: a high-emissions scenario, an
and complex scenarios of climate change, the shared socioeconomic intermediate scenario in which current policies persist and, for the first
pathways (SSPs) [36]. These updated scenarios consist of different time, a low-emissions scenario. Particularly, the low-emissions scenario
narratives of global socio-economic development, employed by the represents a pathway in which the EU’s Green Deal and Paris Agreement
CMIP6 as a means to reach a certain radiative forcing. These models goals are achieved, with the atmospheric concentration of GHG (and
have been recently employed to evaluate future changes in the wind

3
A. Martinez et al. Energy 269 (2023) 126655

Fig. 6. Continuous evolution (%) of the coefficient of variation (COV) for climate-change scenarios SSP1-1.9 (left column), SSP2-4.5 (centre column) and SSP5-8.5
(right column).

thus global temperature) declining in the second half of the 21st century. 3. Materials and methods
Hence, it is of particular interest to evaluate changes in the wind climate
in this scenario, which has not been undertaken so far. 3.1. Global climate models
Global Climate Models from the CMIP6 activities are employed and
validated in the specific area of study, and those that perform best for the The climate projections used in this work, which derive from the
past wind climate are used to form a multi-model ensemble (MME). In activities of the CMIP6, employ the most novel and complex scenarios of
this manner, the individual uncertainties of the GCMs in the region are climate change, the shared socioeconomic pathways (SSPs) [45]. In this
mitigated [44]. The evolution of the offshore wind energy resource is work, three scenarios are considered: SSP1-1.9, a low-emissions scenario
evaluated by studying changes in mean wind energy and its overall and in which climate targets of the EU’s Green Deal and Paris Agreement are
intra-annual (seasonal) variability. achieved, resulting in net-zero emissions by 2050 and ~1.4 ◦ C of global
warming by 2100 (compared to current values); SSP2-4.5, a moderate
2. Study area scenario in which current energy policies continue without substantial
deviation; and SSP5-8.5, a high-emissions scenario resulting from an
The study area is within the ranges of latitude (48◦ N, 66◦ N) and intensive energy consumption from fossil fuel sources.
longitude (13◦ W, 31◦ E), encompassing the offshore regions of Northern Data from all the GCMs involved in the ScenarioMIP activity of the
Europe (Fig. 1). Hereinafter, these regions are separated and referred to CMIP6 providing data on the wind climate for the three scenarios are
as: the Norwegian Sea, Atlantic Ocean, Irish/Scottish Sea, Celtic Sea, considered. Moreover, given the relatively small study area, GCMs with
North Sea, English Channel and Baltic Sea (Fig. 1). The vast majority of a lower resolution than 2.5◦ in latitude or longitude are considered unfit
the current European offshore wind capacity is located in these regions, – very low resolutions are not able to capture the particularities of the
which are expected to accommodate a great number of new installations coastline of Denmark, the UK and Ireland. As a result, a total of 9 GCMs
in the future. are taken into consideration for the study (Table 1). Notably, the MPI-

4
A. Martinez et al. Energy 269 (2023) 126655

Fig. 7. Seasonal mean power density (baseline) (Wm-2) – MAM (March-April-May), JJA (June-July-August), SON (September-October-November) and DJF
(December-January-February).

ESM1-2-HR does not consider the SSP1-1.9, unlike the low-resolution significance level of 5% – by eliminating the seasonal bias, the K–S test is
model (MPI-ESM1-2-HR). Nonetheless, both GCMs are studied, with shown to detect distributional differences in higher-order moments
the possibility of using the high-resolution model as a substitute for the [64]. The number of points in the GCM grid that are statistically similar
SSP2-4.5 and SSP5-8.5 scenarios. Analysing the performance of both to the ERA-5 time series is calculated as a percentage of the total number
models would also illustrate differences between high- and low- of points (Table 2).
resolution models.
Historical data of the above-mentioned GCMs are validated, and the
GCMs that are considered to best represent the wind climate in the re­ 3.3. Selection of GCMs
gion of study are selected to form a multi-model ensemble (MME). With
this approach, individual biases and uncertainties of the GCMs are To study the evolution of the wind resource, wind projections from
significantly mitigated [57]. different GCMs are combined into a multi-model ensemble and
compared against historical data of the same multi-model ensemble. The
GCMs that are found to best reproduce historical data are selected for the
3.2. Validation of GCMs MME; how well a GCM performs is measured by the number of grid
points that are statistically similar to their reanalysis data counterparts
Historical data of the GCMs are compared and validated against the (Table 2). In this work, the five models that present more than 80% of
ERA-5 reanalysis database [58] produced by the European Centre for statistically similar points are selected to form the MME, namely: MRI-
Medium-Range Weather Forecasts (ECMWF) and accessible through the ESM2-0, MPI-ESM1-2-LR, MIROC6, FGOALS-f3-L, EC-Earth3.
Copernicus Climate Change Service (C3S) Climate Data Store [59]. The Notably, the number of points of the GCMs that are statistically
ERA-5 is the most recognised database in reanalysis products, it has been similar to the reanalysis data is significantly higher than in studies
employed extensively for validation in previous works, e.g., Europe including onshore regions [37]. Moreover, even though the GFDL-ESM4
[30], Africa [60] or the Mediterranean [61], and is the official validation was selected for its accuracy in previous studies in Europe [41], in the
dataset for the CMIP downscaling initiatives [62]. Data on near-surface area of study it performs poorly in comparison. These facts reiterate the
historical (2000–2014) wind speed are thus compared with their ERA-5 importance of an offshore-specific approach, which may be expected to
counterparts. lead to better results.
The validation is conducted to determine distributional differences To quantify errors that may be expected in the MME, the normalised
using the two-sample Kolmogorov-Smirnov test (K–S test), which ex­ bias, i.e., the bias normalised against the standard deviation, of the
amines the hypothesis of two groups of data belonging to the same historical (2000–2014) MME against the ERA-5 in the same period is
distribution. In the following analysis, the evolution of the wind calculated (Fig. 2). It is clear that the MME performs better offshore than
resource is quantified by comparing wind projections of a certain GCM nearshore.
against historical data belonging to the same GCM. Hence, for the scope The evolution of the wind energy resource in the 21st century is
of this paper, it is more relevant to test the similarity of the distributions studied by separating the time frame into 15-year intervals: 2025–2039,
rather than the biases. 2040–2054, 2055–2069, 2070–2084 and 2085–2099. Importantly, the
Given that all the GCMs (Table 1) and ERA-5 present different res­ scenario with low emissions, SSP1-1.9, represents a pathway in which
olutions, before the validation a remapping of all the data is conducted. atmospheric concentration of GHG and global temperature increase by
The data are remapped into a regular 1◦ × 1◦ following a first-order 2050, but later decline as a result of climate change policies. Hence, it is
conservative remapping, maintaining the flux integrals [63]. The of great interest to evaluate the effects of this change in trends in global
two-sample K–S test is applied to the unbiased time-series data of the conditions on the wind resource.
GCMs against the unbiased time-series data of the ERA-5 with a In the following, the evolution of the available wind resource is

5
A. Martinez et al. Energy 269 (2023) 126655

Fig. 8. Continuous evolution (%) of the MAM (March-April-May) mean power density for climate-change scenarios SSP1-1.9 (left column), SSP2-4.5 (centre column)
and SSP5-8.5 (right column).

measured by comparing the wind resource of the MME in a certain time 4.1. Mean wind power density
period against the historical (2000–2014) wind resource of the same
MME, hereinafter referred to as the Baseline. Variations in the different The mean power density of the multi-model ensemble in the three
variables characterising the wind resource are quantified as a percent­ climate change scenarios is computed and compared with the baseline
age of change to the Baseline. values (Fig. 3). The baseline shows that the greatest wind energy
The available wind resource is measured by the wind power density resource is concentrated in the open Atlantic areas off western Ireland
(P), obtained from the wind speed (U) as (~1000 Wm-2), where strong westerlies prevail. For this reason, this
area is considered to have great potential for future offshore wind
1
P = ρU 3 , (1) development [65]. Conversely, the weakest resource occurs in the more
2
sheltered areas of the Baltic Sea (~100 Wm-2), where many
where ρ is the air density, taken to be 1.225 kgm-3. Since P is propor­ bottom-fixed wind turbines are installed in its relatively shallow waters
tional to U cubed in Eq. (1), small variations in wind speeds are [66]. Nonetheless, the vast majority of current offshore wind in­
amplified in the available wind energy resource. stallations are allocated in the North Sea [67], where relatively high
values of mean power density (~700 Wm-2) occur. This region is ex­
4. Results and discussion pected to develop intensively offshore wind in the coming years and
therefore is of special interest in the subsequent study [68].
The multi-model ensemble is constructed as indicated in section 2. Importantly, the evolution of the resource shows a generalised
Five 15-year periods are considered: 2025–2039, 2040–2054, decline in the available wind power density in the second half of the 21st
2055–2069, 2070–2084 and 2085–2099. Changes in the wind energy century regardless of the climate change scenario (Fig. 4). This decline is
resource predicted by the MME are measured in comparison with his­ greater in the scenario with the highest GHG emissions, SSP5-8.5, with
torical data (2000–2014) of the same MME. reductions of up to 30%. The evolution of the wind energy resource does

6
A. Martinez et al. Energy 269 (2023) 126655

Fig. 9. Continuous evolution (%) of the JJA (June-July-August) mean power density for climate-change scenarios SSP1-1.9 (left column), SSP2-4.5 (centre column)
and SSP5-8.5 (right column).

not show clear trends up until mid-century, with increases and decreases pronounced changes in the SSP5-8.5. The climate projections anticipate
spread in the area of study without following any particular pattern. a pronounced decrease of up to 30% in areas off Western Ireland which
Conversely, from mid-century onwards tendencies in the evolution of progressively extends eastwards, leading to a generalised drop of ~10%
the mean wind power density become apparent – given a certain sce­ in the North Sea. On the contrary, localised increases of up to 10% are
nario, changes in the mid-century are maintained or exacerbated to­ predicted in the nearshore regions of the Gulf of Riga, the Gulf of Finland
wards 2100. and the Gulf of Bothnia. Notwithstanding, no significant changes are
The scenario with the lowest GHG emissions, SSP1-1.9, shows the predicted in most of the Baltic Sea.
least changes relative to the baseline. Interestingly, this climate-change
scenario represents a pathway in which global temperature returns to 4.2. Variability
current values after a short increase. However, this return is not re­
flected in the evolution of the resource. Even though all the climate The variability of the resource has a large bearing on wind farm
targets of the EU’s Green Deal and Paris Agreement are achieved, an planning and project funding, as investors often prefer a stable revenue.
overall decline in wind power density is projected in all the regions For that reason, the temporal variability of the resource and its evolution
studied, and in particular in the English Channel and nearshore regions are of great interest to the future of the offshore wind industry.
of Ireland (~20%). More generally, this scenario predicts a slight in­ The temporal variability of the resource is studied using the coeffi­
crease in wind power density in the North Sea and western regions of the cient of variation (COV) of wind power density [69], computed for each
Norwegian Sea in the first half of the century, which later evolves to a 15-year period. The COV can be defined as the ratio of the standard
slight decline towards 2100. deviation (σ ) to the mean value (x) of a statistical sample,
The other two scenarios predict a similar evolution – except for a
σ
slight increase in the North Sea in SSP2-4.5 in the first half of the cen­ COV = . (2)
tury. Trends in wind power density are parallel, but with more x
The COV is computed for every period and compared against the

7
A. Martinez et al. Energy 269 (2023) 126655

Fig. 10. Continuous evolution (%) of the SON (September-October-November) mean power density for climate change scenarios SSP1-1.9 (left column), SSP2-4.5
(centre column) and SSP5-8.5 (right column).

COV of the baseline to quantify changes in the temporal variability. 4.3. Intra-annual variability (seasonality)
The COV in the baseline (Fig. 5) shows relatively similar values in the
overall area of study. The largest values are concentrated in the southern The evolution of the intra-annual variability of the resource is ana­
regions of the North Sea around the UK and the Netherlands and in the lysed in the three climate-change scenarios through the evolution of the
nearshore regions of the Norwegian Sea, where the presence of land­ seasonal mean power density. The year is divided into four three-month
masses induces more variability. The lack of obstacles in the open periods – MAM (March-April-May), JJA (June-July-August), SON
Atlantic regions, on the contrary, produces the most stable resource. (September-October-November) and DJF (December-January-
The evolution of the COV values (Fig. 6) shows that major changes February).
are projected exclusively in the scenario with the highest emissions, The baseline values (Fig. 7) show that the most energetic winds occur
SSP5-8.5. In this scenario, an overall increase of the variability is in DJF, regardless of location. The greatest values of mean wind power
anticipated, which is of greater magnitude (up to 15% of the COV density (up to 1200 Wm-2) occur in open areas of the Atlantic Ocean, off
values) in the Celtic Sea and nearshore regions of the North Sea. The West Ireland. On the contrary, the resource is weakest in JJA, especially
only exception to this tendency is found in the open Atlantic Ocean, in the Baltic Sea (<100 Wm-2). With this pattern, increases in the pro­
which shows negligible changes. jected mean wind power density in DJF and/or decreases in JJA result in
Importantly, the other scenarios anticipate little to no significant augmented variability and, conversely, decreases in DJF and/or in­
changes in the variability – generally in the range of ±5%. Nonetheless, creases in JJA lead to a more stable resource.
SSP1-1.9 predicts a general decrease in the variability up to 10% in the The climate projections (Figs. 8–11) anticipate different trends for
English Channel and adjacent regions. the climate change scenarios. The climate change scenario with the
highest emissions, SSP5-8.5, shows a well-spread decrease in wind
power density that is persistent throughout the year, of greater magni­
tude off West Ireland in JJA (up to 30%). An exception to this trend is

8
A. Martinez et al. Energy 269 (2023) 126655

Fig. 11. Continuous evolution (%) of the DJF (December-January-February) mean power density for climate-change scenarios SSP1-1.9 (left column), SSP2-4.5
(centre column) and SSP5-8.5 (right column).

found in the Baltic Sea, where the wind energy resource is predicted to 5. Conclusions
increase in DJF and MAM. Being the strongest decreases in JJA and the
increases in DJF, the intra-annual variability of the resource is poised to In this work, the evolution of wind energy resources in offshore re­
increase, which is also reflected in the evolution of the COV in SSP5-8.5 gions of Northern Europe was studied considering the novel scenarios of
(Fig. 6). climate change — the shared socioeconomic pathways (SSPs). In
In climate scenario SSP1-1.9, the projections show two opposite particular, three climate-change scenarios were taken into account: a
trends throughout the year. On the one hand, increases in wind power high-emissions scenario, SSP5-8.5; an intermediate scenario, SSP2-4.5,
density (up to 20%) are anticipated in JJA and, to a lesser extent, MAM – and a low-emissions scenario, SSP1-1.9. Climate projections of global
of greater significance in the near- and mid-term future. On the other climate models (GCMs) participating in the CMIP6 activities were
hand, a decrease in wind power density is predicted for the rest of the considered and validated, and a multi-model ensemble was constructed
year, most notably in the long-term future. Being of similar values but using the GCMs that were found to best simulate a historical period
opposite signs, these two trends nullify changes in the mean wind power (2000–2014) for the study area.
density values (Fig. 4). However, in stark contrast with SSP5-8.5, trends The validation of the GCMs performed in this work showed that an
in SSP1-1.9 lead to a decline of the intra-annual variability, which is also offshore-specific approach leads to a selection of much more statistically
reflected in the evolution of the COV. This highlights the importance of accurate models and thus to more reliable results. Due to their relatively
studying the evolution of intra-annual variability. coarse resolution, the GCMs present bigger uncertainties in onshore
Finally, climate-change scenario SSP2-4.5 predicts slight regional areas, which would influence the validation if included. In other words,
increases in MAM and JJA in the Baltic Sea, more notable in the near- the differences between onshore and offshore areas regarding uncer­
and mid-term future. In other regions, general decreases are anticipated. tainty levels in the modelling of the wind regime make an offshore-
specific approach all the more relevant, even though the multi-model
ensemble helps to reduce individual uncertainties.

9
A. Martinez et al. Energy 269 (2023) 126655

Under climate change the greatest changes are anticipated in the supporting the WCRP and ESGF. The authors are also grateful to the
scenario with the highest emissions, SSP5-8.5, in the form of a general European Centre for Medium-Range Weather Forecasts (ECMWF) for
reduction in wind power density, which is particularly pronounced off producing and making available their reanalysis products.
West Ireland (of up to 30%). At the other extreme, in the Baltic Sea, the ERA-5 data on wind speed were downloaded from the Copernicus
exception to this general trend is found in the form of local nearshore Climate Change Service (C3S) Climate Data Store. The results contain
increases (up to 10%) in the Gulfs of Riga, Finland and Bothnia. modified Copernicus Climate Change Service information 2020. Neither
Importantly, in this study climate-change scenario SSP1-1.9 was the European Commission nor ECMWF is responsible for any use that
considered for the first time in this type of analysis. The scenario with may be made of the Copernicus information or data it contains.
the lowest emissions, SSP1-1.9 represents a pathway in which global
temperature returns to current values after a short period of increased References
values. However, this return to current temperature values is not
replicated in the wind climate projections. Instead, a decrease in wind [1] Global Wind Energy Council. GWEC| global wind report 2021. 2021.
power density is anticipated in the entire study area, and in particular in [2] Global Wind Energy Council. GWEC| global wind report 2022. 2022.
[3] United Nations. The Paris agreement. What is the Paris agreement?. https://unfccc.
the English Channel and nearshore regions of Ireland (~15%). int/process-and-meetings/the-paris-agreement/the-paris-agreement;. [Accessed 30
SSP1-1.9 does lead to lesser changes in the mean wind power density September 2022]. Accessed.
relative to the baseline than the other scenarios considering higher [4] European Union. A clean energy transition. https://ec.europa.eu/info/strategy/
priorities-2019-2024/european-green-deal/energy-and-green-deal_en; 2021.
levels of carbon emissions. This difference between scenarios is further [Accessed 17 September 2021]. Accessed.
corroborated when studying the temporal variability of the wind [5] Perez-Collazo C, Greaves D, Iglesias G. A novel hybrid wind-wave energy converter
resource. While the temporal variability is predicted to increase to a for jacket-frame substructures. Energies 2018;11(3):637.
[6] Negro V, López-Gutiérrez J-S, Esteban MD, Alberdi P, Imaz M, Serraclara J-M.
large extent in the high-emissions scenario, the low-emissions scenario Monopiles in offshore wind: preliminary estimate of main dimensions. Ocean Eng
shows a decrease in most of the study area. The intermediate scenario, 2017;133:253–61.
conversely, does not show significant changes. [7] Sclavounos P, Lee S, DiPietro J, Potenza G, Caramuscio P, De Michele G. Floating
offshore wind turbines: tension leg platform and taught leg buoy concepts
Finally, the importance of studying the intra-annual variability is
supporting 3-5 MW wind turbines. In: European wind energy conference EWEC;
demonstrated. Changes in the available wind resource of similar mag­ 2010.
nitudes but of different signs occurring in different periods of the year [8] Tomasicchio GR, D’Alessandro F, Avossa AM, Riefolo L, Musci E, Ricciardelli F,
Vicinanza D. Experimental modelling of the dynamic behaviour of a spar buoy
may cancel out and therefore are not reflected in the mean wind power
wind turbine. Renew Energy 2018;127:412–32.
density evolution. Notwithstanding, they may lead to great changes in [9] Lopez-Pavon C, Souto-Iglesias A. Hydrodynamic coefficients and pressure loads on
the variability. This is the case of the projections in SSP1-1.9, where heave plates for semi-submersible floating offshore wind turbines: a comparative
increases in JJA and decreases in DJF lead to a decrease in temporal and analysis using large scale models. Renew Energy 2015;81:864–81.
[10] Vasconcelos RMd, Silva LLC, González MOA, Santiso AM, de Melo DC.
intra-annual variability. Environmental licensing for offshore wind farms: guidelines and policy
The results of this study are relevant in assessing the future viability implications for new markets. Energy Pol 2022;171:113248.
of wind power projects. In the case of Western Ireland, for instance, the [11] Ramos V, Giannini G, Calheiros-Cabral T, Rosa-Santos P, Taveira-Pinto F. Legal
framework of marine renewable energy: a review for the Atlantic region of Europe.
wind energy resource is among the most substantial of the study area in Renew Sustain Energy Rev 2021;137:110608.
the baseline situation; however, it will be affected by the largest pro­ [12] Lee M-K, Nam J, Kim M. Valuing the public preference for offshore wind energy:
jected reductions in wind power density, of up to 30%. the case study in South Korea. Energy 2023;263:125827.
[13] López I, Carballo R, Taveira-Pinto F, Iglesias G. Sensitivity of OWC performance to
As indicated, the results predict significant changes in available wind air compressibility. Renew Energy 2020;145:1334–47.
energy resources — generally, in the form of a decline. A ray of hope, [14] Calheiros-Cabral T, Clemente D, Rosa-Santos P, Taveira-Pinto F, Ramos V,
however, is also to be found in the results, more specifically in the Morais T, Cestaro H. Evaluation of the annual electricity production of a hybrid
breakwater-integrated wave energy converter. Energy 2020;213:118845.
comparison between the three climate-change scenarios, which implies
[15] Fouz D, Carballo R, López I, Iglesias G. Tidal stream energy potential in the
that reducing carbon emissions by adhering to current climate objectives Shannon Estuary. Renew Energy 2022;185:61–74.
would weaken the decline and produce more stable resources. Conse­ [16] Fernández-Jiménez A, Álvarez-Álvarez E, López M, Fouz M, López I, Gharib-
Yosry A, Claus R, Carballo R. Power performance assessment of vertical-Axis tidal
quently, both policy-makers and developers should consider the evolu­
turbines using an experimental test rig. Energies 2021;14(20):6686.
tion of wind resources in planning the future energy mix and offshore [17] Veigas M, Ramos V, Iglesias G. A wave farm for an island: detailed effects on the
wind farms in Europe. nearshore wave climate. Energy 2014;69:801–12.
[18] Ramos V, Giannini G, Calheiros-Cabral T, López M, Rosa-Santos P, Taveira-Pinto F.
Assessing the effectiveness of a novel WEC concept as a Co-located solution for
Declaration of competing interest offshore wind farms. J Mar Sci Eng 2022;10(2):267.
[19] Astariz S, Abanades J, Perez-Collazo C, Iglesias G. Improving wind farm
The authors declare that they have no known competing financial accessibility for operation & maintenance through a co-located wave farm:
influence of layout and wave climate. Energy Convers Manag 2015;95:229–41.
interests or personal relationships that could have appeared to influence [20] Astariz S, Perez-Collazo C, Abanades J, Iglesias G. Co-located wind-wave farm
the work reported in this paper. synergies (Operation & Maintenance): a case study. Energy Convers Manag 2015;
91:63–75.
[21] Wen Y, Kamranzad B, Lin P. Joint exploitation potential of offshore wind and wave
Data availability energy along the south and southeast coasts of China. Energy 2022;249:123710.
[22] Yunna W, Geng S. Multi-criteria decision making on selection of solar–wind hybrid
The authors do not have permission to share data. power station location: a case of China. Energy Convers Manag 2014;81:527–33.
[23] de Souza Nascimento MM, Shadman M, Silva C, de Freitas Assad LP, Estefen SF,
Landau L. Offshore wind and solar complementarity in Brazil: a theoretical and
Acknowledgements technical potential assessment. Energy Convers Manag 2022;270:116194.
[24] Durakovic G, del Granado PC, Tomasgard A. Powering Europe with North Sea
offshore wind: the impact of hydrogen investments on grid infrastructure and
This research was supported by European Union’s Horizon 2020 power prices. Energy 2023;263:125654.
European Green Deal Research and Innovation Program (H2020-LC-GD- [25] Martinez A, Iglesias G. Mapping of the levelised cost of energy for floating offshore
2020-4), grant No. 101037643 – ILIAD (Integrated Digital Framework wind in the European Atlantic. Renew Sustain Energy Rev 2022;154:111889.
[26] Buljan A. Denmark could build its first and world’s largest floating wind farm at
for Comprehensive Maritime Data and Information Services).
bornholm. ; 23 September 2022. https://www.offshorewind.biz/2022/09/23/den
The authors are grateful to the World Climate Research Programme mark-could-build-its-first-and-worlds-largest-floating-wind-farm-at-bornholm/?
(WCRP), which is responsible for the CMIP6, and to the groups utm_source=offshorewind&utm_medium=email&utm_campaign=newsletter
participating in the ScenarioMIP activities, for producing and making _2022-09-26.
[27] Martinez A, Iglesias G. Multi-parameter analysis and mapping of the levelised cost
their model outputs available; the Earth System Grid Federation (ESGF), of energy from floating offshore wind in the Mediterranean Sea. Energy Convers
for storing and making the data accessible; and the funding agencies Manag 2021;243:114416.

10
A. Martinez et al. Energy 269 (2023) 126655

[28] Tian Q, Huang G, Hu K, Niyogi D. Observed and global climate model based ESM1.2-LR model output prepared for CMIP6 ScenarioMIP ssp 119. Earth System
changes in wind power potential over the Northern Hemisphere during Grid Federation, http://cera-www.dkrz.de/WDCC/meta/CMIP6/CMIP6.
1979–2016. Energy 2019;167:1224–35. ScenarioMIP.MPI-M.MPI-ESM1-2-LR.ssp119; 2018.
[29] Gao Y, Ma S, Wang T. The impact of climate change on wind power abundance and [50] Wieners K-H, Giorgetta M, Jungclaus J, Reick C, Esch M, Bittner M, Gayler V,
variability in China. Energy 2019;189:116215. Haak H, de Vrese P, Raddatz T, Mauritsen T, von Storch J-S, Behrens J, Brovkin V,
[30] Carvalho D, Rocha A, Gómez-Gesteira M, Silva Santos C. Potential impacts of Claussen M, Crueger T, Fast I, Fiedler S, Hagemann S, Hohenegger C, Jahns T,
climate change on European wind energy resource under the CMIP5 future climate Kloster S, Kinne S, Lasslop G, Kornblueh L, Marotzke J, Matei D, Meraner K,
projections. Renew Energy 2017;101:29–40. Mikolajewicz U, Modali K, Müller W, Nabel J, Notz D, Peters K, Pincus R,
[31] Räisänen J, Hansson U, Ullerstig A, Döscher R, Graham L, Jones C, Meier H, Pohlmann H, Pongratz J, Rast S, Schmidt H, Schnur R, Schulzweida U, Six K,
Samuelsson P, Willén U. European climate in the late twenty-first century: regional Stevens B, Voigt A, Roeckner E. MPI-M MPI-ESM1.2-LR model output prepared for
simulations with two driving global models and two forcing scenarios. Clim Dynam CMIP6 ScenarioMIP ssp 245. Earth System Grid Federation, https://doi.org/10.22
2004;22(1):13–31. 033/ESGF/CMIP6.6693; 2019.
[32] Pryor S, Barthelmie R, Kjellström E. Potential climate change impact on wind [51] Wieners K-H, Giorgetta M, Jungclaus J, Reick C, Esch M, Bittner M, Gayler V,
energy resources in northern Europe: analyses using a regional climate model. Clim Haak H, de Vrese P, Raddatz T, Mauritsen T, von Storch J-S, Behrens J, Brovkin V,
Dynam 2005;25(7–8):815–35. Claussen M, Crueger T, Fast I, Fiedler S, Hagemann S, Hohenegger C, Jahns T,
[33] Hueging H, Haas R, Born K, Jacob D, Pinto JG. Regional changes in wind energy Kloster S, Kinne S, Lasslop G, Kornblueh L, Marotzke J, Matei D, Meraner K,
potential over Europe using regional climate model ensemble projections. J Appl Mikolajewicz U, Modali K, Müller W, Nabel J, Notz D, Peters K, Pincus R,
Meteorol Climatol 2013;52(4):903–17. Pohlmann H, Pongratz J, Rast S, Schmidt H, Schnur R, Schulzweida U, Six K,
[34] Tobin I, Vautard R, Balog I, Bréon F-M, Jerez S, Ruti PM, Thais F, Vrac M, Yiou P. Stevens B, Voigt A, Roeckner E. MPI-M MPI-ESM1.2-LR model output prepared for
Assessing climate change impacts on European wind energy from ENSEMBLES CMIP6 ScenarioMIP ssp 585. Earth System Grid Federation, https://doi.org/10.22
high-resolution climate projections. Climatic Change 2015;128(1–2):99–112. 033/ESGF/CMIP6.6705; 2019.
[35] Eyring V, Bony S, Meehl GA, Senior CA, Stevens B, Stouffer RJ, Taylor KE. [52] EC-Earth Consortium. EC-Earth-Consortium EC-Earth3 model output prepared for
Overview of the coupled model Intercomparison project phase 6 (CMIP6) CMIP6 ScenarioMIP. Earth Syst Grid Fed 2019. https://doi.org/10.22033/ESGF/
experimental design and organization. Geosci Model Dev (GMD) 2016;9(5): CMIP6.251.
1937–58. [53] Boucher O, Denvil S, Caubel A, Foujols MA. IPSL IPSL-CM6A-LR model output
[36] O’Neill BC, Kriegler E, Ebi KL, Kemp-Benedict E, Riahi K, Rothman DS, van prepared for CMIP6 ScenarioMIP. Earth Syst Grid Fed 2019. https://doi.org/
Ruijven BJ, van Vuuren DP, Birkmann J, Kok K. The roads ahead: narratives for 10.22033/ESGF/CMIP6.1532.
shared socioeconomic pathways describing world futures in the 21st century. [54] Shiogama H, Abe M, Tatebe H. MIROC MIROC6 model output prepared for CMIP6
Global Environ Change 2017;42:169–80. ScenarioMIP. Earth Syst Grid Fed 2019. https://doi.org/10.22033/ESGF/
[37] Martinez A, Iglesias G. Climate change impacts on wind energy resources in North CMIP6.898.
America based on the CMIP6 projections. Sci Total Environ 2022;806:150580. [55] Yukimoto S, Koshiro T, Kawai H, Oshima N, Yoshida K, Urakawa S, Tsujino H,
[38] Wu J, Shi Y, Xu Y. Evaluation and projection of surface wind speed over China Deushi M, Tanaka T, Hosaka M, Yoshimura H, Shindo E, Mizuta R, Ishii M,
based on CMIP6 GCMs. J Geophys Res Atmos 2020;125(22). e2020JD033611. Obata A, Adachi Y. MRI MRI-ESM2.0 model output prepared for CMIP6
[39] Moradian S, Akbari M, Iglesias G. Optimized hybrid ensemble technique for CMIP6 ScenarioMIP. Earth Syst Grid Fed 2019. https://doi.org/10.22033/ESGF/
wind data projections under different climate-change scenarios. Case study: United CMIP6.638.
Kingdom, vol. 826. Science of The Total Environment; 2022, 154124. [56] John JG, Blanton C, McHugh C, Radhakrishnan A, Rand K, Vahlenkamp H,
[40] Qian H, Zhang R. Future changes in wind energy resource over the Northwest Wilson C, Zadeh NT, Gauthier PPG, Dunne JP, Dussin R, Horowitz LW, Lin P,
Passage based on the CMIP6 climate projections. Int J Energy Res 2021;45(1): Malyshev S, Naik V, Ploshay J, Silvers L, Stock C, Winton M, Zeng Y. NOAA-GFDL
920–37. GFDL-ESM4 model output prepared for CMIP6 ScenarioMIP. Earth Syst Grid Fed
[41] Martinez A, Iglesias G. Wind resource evolution in Europe under different scenarios 2018. https://doi.org/10.22033/ESGF/CMIP6.1414.
of climate change characterised by the novel Shared Socioeconomic Pathways. [57] Moradian S, Torabi Haghighi A, Asadi M, Mirbagheri SA. Future changes in
Energy Convers Manag 2021;234:113961. precipitation over Northern Europe based on a multi-model ensemble from CMIP6:
[42] Costoya X, DeCastro M, Carvalho D, Gómez-Gesteira M. On the suitability of Focus on Tana river basin. Water Resources Management; 2022.
offshore wind energy resource in the United States of America for the 21st century. [58] Hersbach H, Dee D. ERA5 reanalysis is in production. ECMWF Newsl 2016;147(7):
Appl Energy 2020;262:114537. 5–6.
[43] Zhang S, Li X. Future projections of offshore wind energy resources in China using [59] Hersbach H, Bell B, Berrisford P, Biavati G, Horányi A, Muñoz Sabater J, Nicolas J,
CMIP6 simulations and a deep learning-based downscaling method. Energy 2021; Peubey C, Radu R, Rozum I, Schepers D, Simmons A, Soci C, Dee D, Thépaut J-N.
217:119321. ERA5 hourly data on single levels from 1979 to present. Copernicus Climate
[44] Yazdandoost F, Moradian S, Zakipour M, Izadi A, Bavandpour M. Improving the Change Service (C3S) Climate Data Store (CDS). 2018. https://doi.org/10.24381/
precipitation forecasts of the North-American multi model ensemble (NMME) over cds.adbb2d47. Accessed 23rd March 2022.
Sistan basin. J Hydrol 2020;590:125263. [60] Dosio A, Panitz H-J, Schubert-Frisius M, Lüthi D. Dynamical downscaling of CMIP5
[45] Riahi K, Van Vuuren DP, Kriegler E, Edmonds J, O’neill BC, Fujimori S, Bauer N, global circulation models over CORDEX-Africa with COSMO-CLM: evaluation over
Calvin K, Dellink R, Fricko O. The shared socioeconomic pathways and their the present climate and analysis of the added value. Clim Dynam 2015;44(9–10):
energy, land use, and greenhouse gas emissions implications: an overview. Global 2637–61.
Environ Change 2017;42:153–68. [61] Bloom A, Kotroni V, Lagouvardos K. Climate change impact of wind energy
[46] Yu Y. CAS FGOALS-f3-L model output prepared for CMIP6 ScenarioMIP. Earth Syst availability in the Eastern Mediterranean using the regional climate model PRECIS.
Grid Fed 2019. https://doi.org/10.22033/ESGF/CMIP6.2046. Nat Hazards Earth Syst Sci 2008;8(6).
[47] Swart NC, Cole JNS, Kharin VV, Lazare M, Scinocca JF, Gillett NP, Anstey J, [62] Gutowski WJ, Giorgi F, Timbal B, Frigon A, Jacob D, Kang H-S, Raghavan K, Lee B,
Arora V, Christian JR, Jiao Y, Lee WG, Majaess F, Saenko OA, Seiler C, Seinen C, Lennard C, Nikulin G. WCRP coordinated regional downscaling experiment
Shao A, Solheim L, von Salzen K, Yang D, Winter B, Sigmond M. CCCma CanESM5 (CORDEX): a diagnostic MIP for CMIP6. 2016.
model output prepared for CMIP6 ScenarioMIP. Earth Syst Grid Fed 2019. https:// [63] Jones PW. First-and second-order conservative remapping schemes for grids in
doi.org/10.22033/ESGF/CMIP6.1317. spherical coordinates. Mon Weather Rev 1999;127(9):2204–10.
[48] Schupfner M, Wieners K-H, Wachsmann F, Steger C, Bittner M, Jungclaus J, Früh B, [64] Brands S, Herrera S, Fernández J, Gutiérrez JM. How well do CMIP5 Earth System
Pankatz K, Giorgetta M, Reick C, Legutke S, Esch M, Gayler V, Haak H, de Vrese P, Models simulate present climate conditions in Europe and Africa? Clim Dynam
Raddatz T, Mauritsen T, von Storch J-S, Behrens J, Brovkin V, Claussen M, 2013;41(3–4):803–17.
Crueger T, Fast I, Fiedler S, Hagemann S, Hohenegger C, Jahns T, Kloster S, [65] Martinez A, Iglesias G. Site selection of floating offshore wind through the levelised
Kinne S, Lasslop G, Kornblueh L, Marotzke J, Matei D, Meraner K, Mikolajewicz U, cost of energy: a case study in Ireland. Energy Convers Manag 2022;266:115802.
Modali K, Müller W, Nabel J, Notz D, Peters K, Pincus R, Pohlmann H, Pongratz J, [66] Wikipedia. List of offshore wind farms in the Baltic Sea. https://en.wikipedia.org/
Rast S, Schmidt H, Schnur R, Schulzweida U, Six K, Stevens B, Voigt A, Roeckner E. wiki/List_of_offshore_wind_farms_in_the_Baltic_Sea;. [Accessed 7 October 2022].
DKRZ MPI-ESM1.2-HR model output prepared for CMIP6 ScenarioMIP. Earth Syst Accessed.
Grid Fed 2019. https://doi.org/10.22033/ESGF/CMIP6.2450. [67] Wikipedia. List of offshore wind farms in the North Sea. https://en.wikipedia.
[49] Milinsky S, Li H, Brune S, Wachsmann F, Wieners K-H, Giorgetta M, Jungclaus J, org/wiki/List_of_offshore_wind_farms_in_the_North_Sea;. [Accessed 7 October
Reick C, Esch M, Gayler V, Haak H, de Vrese P, Raddatz T, Mauritsen T, von 2022]. Accessed.
Storch J-S, Behrens J, Brovkin V, Claussen M, Crueger T, Fast I, Fiedler S, [68] North Sea offshore wind to help repower the EU. Wind Europe; 22 May 2022. https
Hagemann S, Hohenegger C, Jahns T, Kloster S, Kinne S, Lasslop G, Kornblueh L, ://windeurope.org/newsroom/press-releases/north-sea-offshore-wind-to-help-re
Marotzke J, Matei D, Meraner K, Mikolajewicz U, Modali K, Müller W, Nabel J, power-the-eu/.
Notz D, Peters-von Gehlen K, Pincus R, Pohlmann H, Pongratz J, Rast S, Schmidt H, [69] Walpole RE, Myers RH, Myers SL, Ye K. Probability and statistics for engineers and
Schnur R, Schulzweida U, Six K, Stevens B, Voigt A, Roeckner E. MPI-M MPI- scientists, vol. 5. New York: Macmillan; 1993.

11

You might also like