You are on page 1of 19

Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


journal homepage: www.elsevier.com/locate/rser

Life cycle environmental impacts of wind energy technologies: A review of T


simplified models and harmonization of the results
Barbara Mendeckaa, Lidia Lombardib,∗
a
University of Florence, via S. Marta 3, 50139 Florence, Italy
b
Niccolò Cusano University, via don Carlo Gnocchi 3, 00166 Rome, Italy

ARTICLE INFO ABSTRACT

Keywords: Nowadays, wind energy is taking on primary role within renewable energies scenario. However, impact of wind
Systematic review energy industry on the environment still requires to be fully understood and better quantified. This study pro-
Wind turbines vides an updated review of Life Cycle Assessment (LCA) studies of electricity produced from onshore and off-
Wind energy shore wind turbines (WTs). Special emphasis is put on results harmonization and simplified LCA models existing
Life cycle assessment
in the literature. The synthesis of the results is performed for wide range of WTs capacities, providing an ex-
Environmental impact
Onshore
haustive and general frame of the environmental impacts of WTs systems. Moreover, new simplified LCA models,
Offshore which make use of a non-linear regression, were developed in this work for the following impact categories:
Acidification Potential (AP), Eutrophication Potential (EP), Global Warming Potential (GWP) and Cumulative
Energy Demand (CED) and for onshore (1–5000 kW) and offshore (500–8000 kW) WTs. Nonlinear data fitting
models are provided with a sufficiently high correlation coefficient for total life cycle impacts. Moreover, the
proposed simplified LCA models predict the final results with acceptable uncertainty. This indicates that the one-
term power series describes the behavior of the impact indicators accurately, providing a useful correlation to
estimate the life cycle environmental performance for a specific turbine model with a given nominal power.
Furthermore, obtained simplified LCA models were generalized for different site-specific wind conditions, i.e.
wind speeds and wind classes. By analyzing all the considered impact indicators for electricity generation, we
notice that the highest values of life cycle impacts of electricity, for a particular WT correspond to the highest
wind velocities. This is particularly valid for low nominal power turbines, which seem to be significantly affected
by wind conditions. The trends exhibit an asymptotic behavior, indicating that, on the contrary, wind conditions
are a minor contributor to the environmental impact of large-scale systems.

1. Introduction 18.8 GW [1].


To date, the European Union (EU) contribution accounts for
Nowadays, wind energy technologies are taking on a primary role 168.7 GW of total net installed WTs capacity with 85% of total installed
within the renewable energies scenario. They represent a mature and worldwide offshore capacity. In 2017, the EU connected 15.6 GW of
valid solution toward the fulfillment of the world objectives on sus- new WTs capacity and produced a total of 336 TWh covering 11.6% of
tainable development. Wind turbines (WTs) industry development was total electric demand [2]. In the same year, the average size of all newly
evolving quickly in the past years, and it is predicted that the trend of installed onshore wind turbines ranged between 1.4 and 2.9 MW. Off-
progress will continue. According to the International Energy Agency shore wind turbines capacity was generally higher varying from 2.5 to
(IEA), only in 2017, 53 GW of new WTs capacity was added, reaching 7 MW. Moreover, during the past few years, substantial technological
the total net installed capacity of wind energy around 539.1 GW and progress has been observed, especially for offshore applications. The
covering the 5.6% of electric demand worldwide. Moreover, the off- average land-based wind capacity factors increased from 22.5% to
shore wind power market continued to grow in all the developed 25.5%, between 2011 and 2017. Offshore capacity factors also im-
countries, and its potential attracts the interest of industries and poli- proved, rising from 37.1% to 45.6% [3].
tical leaders. About 4.3 GW of new offshore WTs capacity was in- Large scale wind energy systems are considered as the most af-
tegrated to the grid in 2017, reaching the total cumulative capacity of fordable among all of the other known renewables [4,5]. However,


Corresponding author.
E-mail address: lombardi.lidia@gmail.com (L. Lombardi).

https://doi.org/10.1016/j.rser.2019.05.019
Received 18 December 2018; Received in revised form 29 April 2019; Accepted 7 May 2019
Available online 24 May 2019
1364-0321/ © 2019 Elsevier Ltd. All rights reserved.
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Nomenclature HH Human Health


HTTP Human Toxicity Potential
AP Acidification Potential IEA International Energy Agency
CED Cumulative Energy Demand IEC International Electrotechnical Commission
CML Centre for Environmental Studies, Leiden, Netherlands ILCD International Reference Life Cycle Data
CV Coefficient of variation JRC Joint Research Centre
EI Environmental impact TEC Life Cycle Thermoecological Cost
ELCD European Reference Life Cycle Database LCA Life Cycle Assessment
EP Eutrophication Potential LCI Life Cycle Inventory
EPBT Energy Pay Back Time LCIA Life Cycle Impact Assessment
EU European Union MAETP Marine Aquatic EcoToxicity Potential
FAETP Freshwater Aquatic EcoToxicity Potential NR Non-Renewable
F Characterization Factor ODP Ozone Depletion Potential
FU Functional Unit TETP Terrestrial EcoToxicity Potential
GHG Greenhouse Gases TRL Technology Readiness Level
GSA Global Sensitivity Analysis VAWTs Vertical Axis Wind Turbines
GWP Global Warming Potential WTs Wind Turbines
HAWTs Horizontal Axis Wind Turbines

according to recent predictions, in the EU around 50% of the cumula- and Kendall [16], who provided an exhaustive study on LCAs in-
tive large WTs capacity will reach the end of its operational life by 2030 vestigating life cycle greenhouse gases (GHG) emissions of wind power
[6]. This implies that repowering market will start to emerge [7]. The in a vast study region. In another review study, Raadal et al. [17,18]
repowering process consists in substituting the exploited wind turbines reported the total GHG emissions in the range of 4.6–55.6 g CO2 kWh−1
with new ones at the same place and thus increasing the power output for onshore WTs capacity between 30 and 3000 kW, and of 18.0–31.4 g
capacity of the wind farm [8]. It should also be mentioned that in the CO2 kWh−1 for offshore WTs capacities between 500 and 5000 kW.
context of the sustainable development of energy systems even small Similarly, Lenzen and Munksgaard [19] reported different values of
and medium-size onshore WTs are acquiring significant importance. In GHG emission intensities: from 7.9 to 123.7 g CO2 kWh−1 across the
particular, small wind turbine technology, below 200 kW, is developing onshore WT power capacity ratings between 0.3 and 3000 kW. Arvesen
rapidly, becoming essential to the private market [9]. The rise of wind and Hertwic [20] conducted an extensive, critical review of the en-
power generation requires, therefore, more efficient WTs and their vironmental impacts of on and offshore WTs with capacity ranging from
environmental concerns are still a priority. 1 to 5000 kW. They reported a wide range of variability of results also
The environmental assessment of wind energy should be assessed for different environmental stressors, in particular, CO2, CH4, CO, NH3,
from the entire life cycle perspective. Such an approach allows to NMVOC, N2O, NOx, particulates, SO2 and Cumulative Energy Demand
compare the environmental impacts of different renewable and non – (CED). Furthermore, they observed that several authors provide dif-
renewable energy sources utilized in different energy systems fulfilling ferent uncertainties for WTs scales among each other. In particular,
the same function. It accounts for the environmental burdens occurring some of the results for parameters such as CED, GHG or SO2 are char-
at different life cycle stages – from the extraction of raw materials to acterized by higher values of uncertainty for small (< 100 kW) and
production, transport, use, and end-of-life [10,11]. The Life Cycle As- medium (100–1000 kW) size WTs, than for the largest sizes. An esti-
sessment (LCA) approach provides quantitative results, which may be mation of this set of pollutants is of great importance in order to pro-
prerequisite to aid the decision makers. The general methodological vide a proper impact assessment concerning global environmental
framework and standards for LCA are defined by ISO 14040 and ISO concerns such as climate change, acidification, eutrophication or pho-
14044 series [12,13]. Moreover, other guidelines are available for dif- tochemical oxidation.
ferent product types and regions of the world as those presented by the The discrepancy of the results reported in the literature is caused
International Reference Life Cycle Data (ILCD) System [14]. Current mainly by input data inconsistencies and differences in assumptions
methodological LCA standards provide four steps approach, which are: regarding site-specific wind conditions, wind turbine sizes and design.
(1) goal and scope definition, (2) inventory assessment, (3) impact as- Another factor affecting the final result of LCA may be the difference in
sessment and (4) interpretation. However, the abovementioned fra- the modeling approach, which may adopt different pathways to eval-
meworks leave the practitioner a wide margin of flexibility in how to uate the same environmental impact. This factor is particularly im-
perform an LCA of a particular product or process. While this flexibility portant for the impact categories representing toxicity, but also for
is necessary for responding to the large variety of concerns addressed, those regarding climate change or acidification [21]. Overall, such a
further generalization is needed to support the consistency of the spe- variability represents a further limitation for the decision making pro-
cific assessment. Another issue concerns the data quality and avail- cess, since it does not allow unequivocal identification of the environ-
ability. Gathering detailed Life Cycle Inventory (LCI), which represents mental effects produced by a considered system. Everything considered,
the core of every LCA, is a complex task indeed. In real energy appli- detailed LCA analysis results in a rather time-consuming and a so-
cations, it is rarely possible to include each and all of the input and phisticated approach, which may lead the method, in some cases, to be
output flows within different life cycle stages [15]. Furthermore, the even impractical. For these reasons, it is often reasonably acceptable to
uncertainty and subjectivity of the results propagate moving to the part simplify the LCA models by adopting estimations and approximations
of the Life Cycle Impact Assessment (LCIA). or harmonizing the LCA results instead of specifying the detailed in-
The number of LCA studies on environmental impacts of wind tur- formation required by a given process.
bines is significant, with the reported environmental profiles of WTs The goal of this work is twofold. Firstly, we provide a thorough and
varying considerably, according to the technology of the considered updated review of the currently available in the literature studies
power plant, for instance, the site location or the temporal analysis concerning LCA applied to wind energy. The synthesis of the results is
resolution. Systematic reviews published in areas related to WTs en- performed on wind turbines of a wide range of capacities, thus pro-
vironmental impacts confirm this information, such as the one of Price viding an exhaustive and general frame of the environmental impacts of

463
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

the wind energy systems. Secondly, we present the potential of sim- harmonize the LCA results for the various renewable technologies. The
plified LCA models to provide reliable and quantitative predictions of harmonization formula they proposed for normalizing environmental
the environmental impact of wind energy, emphasizing benefits and effects of wind turbines was based on a linear approximation and
drawbacks of this methodology in comparison with a detailed LCA. considered two harmonization parameters, i.e. capacity factor and
Some of the approaches presenting simplified LCA models and har- lifetime. Dolan and Heath [25] proposed a more complex harmoniza-
monization of results as an alternative to detailed LCA of WTs have tion procedure for GWP by manipulating two parameters, such as
already been proposed in the literature. We discuss them in detail in lifetime and capacity factor, as well as the GWP methodology and the
Section 2 of this manuscript. Differently, from the previous works implemented system boundary, consecutively. Specifically, they per-
available in the literature, we focus on the regressions based on a re- formed the harmonization of 107 onshore and 16 offshore GHG esti-
presentative set of detailed LCA results derived from a systematic and mates. Their findings suggest that the harmonization process either by
updated literature review for a wider range of onshore and offshore applying one or all the parameters, increased the precision of life cycle
WTs capacities (from 1 kW up to 5000 kW for onshore WTs and from GHG emission estimates in the literature while having little effect on
500 to 8000 kW for offshore WTs). the published central tendency. Moreover, they found the capacity
The remaining part of the paper is organized as follows. Section 2 factor to have the most significant effect on reducing the variability of
provides background information and review on the simplified LCA the results. Carbajales-Dale [26] presented the results of a meta-ana-
models and harmonization procedures available for the wind power lysis and harmonization of CED and energy payback time of wind tur-
industry. Section 3 provides a systematic review with a detailed ana- bines. The harmonization procedure involved the capacity factor and
lysis of the life cycle environmental impacts of wind energy and lifetime. From that study, it emerged that, by harmonizing CED by one
methodological aspects concerning simplified LCA models. Section 4 parameter, i.e. capacity factor, the interquartile range increases. On the
presents the proposed regression models and the validity assessment of contrary, after subsequent harmonization of turbine lifetime, the un-
the proposed approach as well as limitations and practical implications certainty decreases to a value which is still higher compared to the
of this work. Finally, Section 5 presents concluding remarks and re- unharmonized ones. Table 1 summarizes the harmonization procedures
commendations. available for wind power.

2. Review of simplified LCA models and harmonization of the


2.2. Simplified LCA models
results for wind energy
Several simplified LCA approaches for wind energy has been already
In this section, we present the literature background concerning
established. Padey et al. [27,28] defined and developed a simplified
harmonization procedures and simplified LCA models applied to the
model estimating GHG emissions of wind energy systems. The metho-
wind power system. LCA harmonization procedure can reduce the
dological framework they proposed involves the meta-analysis of pub-
variability of the impacts and increase the consistency of the results.
lished GHG data. In particular, in Ref. [27], a regression model has been
This procedure is one of the meta-analytical methods involving sys-
established with the use of detailed LCA results from 17 industrial on-
tematic review [22]. In the LCA context, this procedure allows to
shore wind turbines with capacities between 800 kW and 4500 kW in-
identify sensitive parameters, inconsistent methods, and assumptions
stalled and manufactured in Europe. The simplified model was ex-
and then to decrease uncertainty by adjusting the results to alternative
pressed as a function of two independent variables, i.e., wind speed and
conditions [23]. Simplified LCA models instead, provide a more
wind turbine lifetime.
straightforward and efficient way of estimating environmental impacts.
The same authors proposed a generic methodology providing the
They facilitate obtaining of the results from a comprehensive, detailed
simplified GHG performance for medium size (over 500 kW) WTs em-
LCA of any product or process, by considering technical, temporal and
ploying Global Sensitivity Analysis approach and Monte Carlo simula-
geographical variability of the parameters of the considered model.
tion [28]. This strategy allowed to analyze a hierarchy of the main
parameters and to select those that contributed to the overall un-
2.1. LCA data harmonization certainty of results. In order to create an LCI of considered WTs, they
applied scaling functions handling a relation between the amount of
Some of the approaches harmonization of results as an alternative to construction materials required for WT and its nominal power. Finally,
detailed LCA of WTs have already been proposed in the literature. they defined the simplified regression function of GHG basing on the
Asdrubali et a [24]. presented a simple methodological approach to capacity factor and the wind turbine lifetime as the most significant

Table 1
Overview of the harmonization procedures for wind power.
Harmonization parameter Reference parameters values Harmonization formula for environmental impact (EI ) Ref.

Capacity factor, CF Onshore turbines: 35% CFpub LTpub [24]


EIharm = EIpub
Offshore turbines: 45% CFref LTref
Lifetime, LT 20 years
Capacity factor, CF Onshore turbines: 30% CFpub [25]
EIharm = EIpub
Offshore turbines: 40% CFref

Lifetime, LT 20 years LTpub


EIharm = EIpub
LTref
Characterization factors for GHG According to the Intergovernmental Panel on Climate Change (IPCC) Fpub, i mpub, i
EIharm = EIpub
emissions, F methodology for 100 years horizon (25 g CO2-eq/g methane (CH4) and 298 g Fref , i
CO2-eq/g nitrous oxide (N2O).
System boundary Disaggregated GHG emission for upstream (construction), on-going EIharm = EIpub,up + EIref ,on + EIref , down
(operational) and downstream (end-of-life) stages add- on values EI:
Onshore turbines: 1.3/0.44 g CO2-eq/kWh (on-going/downstream)
Offshore turbines: 0.50/0.56 g CO2-eq/kWh (on-going/downstream)
Capacity factor, CF All turbines: 25% CFpub LTpub [26]
EIharm = EIpub
Lifetime, LT 25 years CFref LTref

464
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

parameters identified. The results obtained from the model were com- Publications were screened applying the following criteria: (1) the
pared with the results presented for 2 MW WT by Guezuraga et al. [29] publication should use referenced data; (2) the publication should de-
and Ardente et al. [30]. In this comparison, only small deviations were scribe the assumptions quantitatively; (3) the publication should pre-
observed. sent full LCIA assessment results including the results of classification
In previous work by the authors, Mendecka et al. [31] proposed a and characterization steps at the midpoint level and/or report LCI for
simplified regression model for estimation of life cycle Thermo- the cases where the impacts are reported only after normalization,
ecological Cost (LC - TEC). The parameters taken into consideration grouping and weighting (endpoint or single score level). The final re-
were the nominal power and wind conditions i.e. wind speed and wind view included 37 peer-reviewed articles, 33 scientific and industrial
variability. The obtained correlation models provided estimates of the reports, and one database, which lead to a complete set of 148 WTs LCA
final value with sufficient accuracy comparing to detailed LCA. Another case studies published between 2000 and 2018. Table 2 presents an
publication, the one of Caduff [32], presents an empirical regression overview of the considered studies.
model for estimating the midpoint indicators from ReCiPe metho- Afterward, we carried out a detailed analysis of each study con-
dology. The independent values in the model were the hub height and sidering the criteria related to technological, geographical and metho-
rotor diameter. The model coefficients were derived from analysis of 12 dological aspects. In particular, we focused on size and type of tech-
previously published case studies of onshore WTs, with a range of ca- nology (onshore, offshore, HAWT, VAWT), operating time of wind
pacities from 30 to 3000 kW. This analysis was conducted via the im- turbine and plant components, capacity factors and electricity outputs,
plementation of reduced LCA approach. Moreover, the reported model system boundary, the methodology used and results of LCIA. The re-
did not consider the geographical variability. The model was tuned presentative studies are analyzed below. Raw data used for this analysis
assuming wind speed of 5 m/s at 10 m height and a wind shear gradient are presented in Supplementary Material.
of 1/7. This indicates that the scaling factors used in that model are Among the considered case studies, we found that the most ana-
valid only for a specific location and wind regime. lyzed case is the horizontal WT, which accounts for 109 studies for
The reported studies show the possibility of decreasing the use of onshore and 30 for offshore technology. The remaining 9 studies con-
detailed LCAs in favour of simplified models. However, evaluation cerned the vertical WT technology, of which 7 were dedicated to on-
methodologies based on a single indicator, i.e. the GHG emissions, shore applications, while only 2 dealt with offshore applications. The
present some advantages with respect to full LCA approaches, which are major part, 95 over the 148 of the published case studies, reports de-
inherent to their simplicity. This may lead to undesired over- tailed LCI. For remained 53 case studies only LCIA at midpoint level
simplifications and has been demonstrated by several authors, as shown was available. However, they were included in further analysis. As far
in the review study of Turconi et al. [33]. Wind energy technology is as the range of power is concerned, the onshore technology was con-
also related to water, ground, wildlife and landscape impacts. There- sidered for nameplate capacity ranging between 1 kW and 5000 kW,
fore, evaluating only the global CO2 emissions cannot be sufficient to while the offshore aWTs power varied between 500 and 8000 kW. In
describe its total impact adequately. next Fig. 1, we report the distribution of the WTs LCA case studies in
terms of size and technology analyzed.
It can be noted that the medium-to-large scale (e) WTs attract the
3. Material and methods greatest interest. In fact, at least 51 case studies reported, of which 43
were subjected to onshore and 8 to offshore applications were reported
3.1. Systematic review and meta-analysis of LCA case studies included for technologies with power ranging from 1000 to 2500 kW. Moreover,
we observed a continuing growth of interest in analyzing WTs tech-
This section provides a systematic review of environmental impacts nology in this power range that is evidenced by the fact of a growing
of electricity produced from onshore and offshore wind power systems. number of case studies reported. This was indeed expected, since this is
A literature review was carried out by considering peer-reviewed arti- the average size of all newly installed turbines in 2017, as mentioned in
cles, industry reports and databases, with a focus on the newest data. the Introduction. Also, large scale WTs characterized by higher capa-
When referring to environmental impacts, we have taken into con- cities are extensively studied in the literature. In particular, among the
sideration midpoint oriented impact categories. Impacts at the midpoint WTs with nominal power above 5000 kW (g), at least 17 studies dealt
level are modeled quite accurately and are in line with the recent en- with offshore applications, while only 3 were dedicated to onshore
ergy and environmental policies. The examples of midpoints are applications. This information depicts well the current research sce-
Acidification Potential (AP), Eutrophication Potential (EP), Global nario, in which the offshore applications for large size WTs are emer-
Warming Potential (GWP) and Cumulative Energy Demand (CED), ging as a very attractive technology. This analysis also indicated that
Photochemical Ozone Creation Potential (POCP), Human Toxicity micro, small and medium scales WTs still require further investigation.
(cancer and non-cancer related), Eco Toxicity or Land Use [14]. Con- In particular, only a few works in literature has been devoted to small
versely, endpoint categories such as damages to human health, eco- and medium scale WTs technology (b and c).
system diversity or resource scarcity are more understandable but have From our survey, we also observed that the overall trend of interest,
not sufficient transparency for detailed analysis, therefore were ex- in terms of capacity factor, follows the one for the nominal capacity. In
cluded from this review.

Table 2
Overview of the studies included in the literature review.
Year of publication Number of case studies References

Article Report Database

2018 - 32 6 – Articles: [8,31,34–37], Reports [38–43]:


2016–2018 9 11 8 Articles: [44–48], Reports: [49–58], Databases [59]:
2014–2016 15 7 – Articles: [18,60–64], Reports [65–71]:
2012–2014 18 3 – Articles: [29,72–77], Reports [78–80]:
2010–2012 13 5 – Articles: [81–85], Reports [86–90]:
- 2010 17 4 – Articles: [30,91–99], Reports [100–102]:
TOTAL 104 36 8 Articles: 38, Reports: 34, Database: 1

465
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

(manufacturing, construction) as well as downstream (dismantling)


processes.
The majority of LCA studies highlights the particular importance of
the upstream stage. However, some potential system-level adverse ef-
fects, which are connected with the wind energy system integration to
the existing grid, are also significant. As an example, recent assessments
[36,103] quantify the additional emission of part load and start up/shut
down the conventional power plants with increasing penetration of
wind power, and state that this impact should be associated with the
operation stage of WTs [36]. Nevertheless, these impacts are not typi-
cally considered in LCA studies, and for the matter of consistency, they
are not discussed in this study.
The environmental impacts associated with WTs are evaluated using
a great variety of methodologies, within the LCA framework, in lit-
erature. Few studies applied more than one methodology to assess the
environmental performance of the WTs. In Table 3 we report an over-
view of the methods adopted in the representative studies.
The GHG emission analysis seems to be the approach of most in-
terest. GHG emissions still represent a major environmental concern.
This is the reason why this indicator is widely used for the evaluation of
the environmental performances of an energy system. Further, another
significant amount of studies propose to analyze the environmental
Fig. 1. WTs case studies distribution in terms of WTs size and technology. performances of WTs power plants through a single elementary air
Onshore site WTs are indicated with “On”, while “Off” stands for the offshore emission approach, such as CO2, SO2 or NOx, for instance. In general,
site. The corresponding overall number of analyzed works is reported for each the analysis of a particular air emission component is particularly
case study. Letters “a) - g)” refer to the nameplate capacity range, with: a) micro useful, since it allows a direct comparison of the environmental per-
WTs (below 20 kW); b) small WTs (20–100 kW; c) medium I WTs formances among different typologies of electricity generation systems.
(100–300 kW); d) medium II WTs (500 kW–1000 kW); e) large I WTs Moreover, these emissions can provide a reasonable approximation of
(1000–2500 kW); f) large II WTs (2500–5000 kW) and g) large III WTs (above environmental impacts such as GWP, AP or EP [103]. Other interesting
5000 kW). concepts are those related to the CED and the Energy Payback Time
(EPBT). Several studies, among those taken into consideration by the
other words, WTs characterized by relatively high capacity factor were present review, make use of methodological approaches based on these
the most studied, since they are typically those of medium-to-large and two indicators. In fact, CED is one of the standard impact indicators
larger size. This is in line with the general expectation. Another aspect which requires fewer data in the inventory analysis comparing to other
concerns the values for the capacity factors reported in the literature. indicators involving emission data and specific impact assessment
Fig. 2 presents the variability of capacity factors of WTs as reported in procedures [104]. Furthermore, considering the energy performance
the literature. from a cradle to gate perspective provides useful information and gives
We found out that, for medium size WTs, offshore applications were better insight into the overall efficiency of the renewables-based energy
characterized by lower average capacity factor values concerning the system than conventional energy analysis. Also, we observe that other
onshore ones, while for larger sizes, offshore applications had a gen- LCIA strategies as those developed at the Centre of Environmental
erally higher capacity factor. In particular, across the whole range of Science of Leiden University (CML) are quite common in literature. As
WTs sizes, capacity factors for onshore applications ranged between shown by Martinez et al. [61], CML represents one of the most robust
0.02 and 0.56, with a median value of 0.335. Offshore installations methods and provides a useful and flexible tool for evaluating en-
showed a minimum capacity factor equal to 0.16, a maximum of 0.59 vironmental performances. Furthermore, the literature analysis shows
and median value equal to 0.295. No comparison can be made for micro that the life cycle LC - TEC is a method of great success, which is applied
and small installations since there are no case studies concerning off- in a vast number of case studies. This strategy makes use of the concept
shore applications for these classes. of exergy to measure the quality of a given natural resource and as-
Regarding the methodological aspects, we report 143 up to 148 case sumes a balance boundary which encloses the resources extraction
studies assuming an operating time for the WTs of 20 years. This re- process. By means of the LC-TEC approach, it is then possible to assess
presents the most significant share of papers. Among these, in 115 the consumption of non-renewable natural sources considering both
studies the onshore installations were analyzed, while the remaining 28 their different quality and the quality of the energy carriers. Besides,
focused on offshore WTs. The assumption of an operating time equal to the LC-TEC takes into account the environmental losses caused by the
25 years was applied in fewer case studies presented in the literature. rejection of harmful substances to the natural environment [105,106].
We report only 8 studies over 148 considering such an approach. Of Finally, among the other methods proposed in the literature, few others
these, 3 case studies performed in Ref. [76] are dedicated to onshore are attracting a growing interest within the scientific community. We
applications, while 5 case studies presented in Refs. [37,42,43,85] focus mention IMPACT 2002, Ecoindicator 99/95 and ReCiPe, as the most
on offshore WT technologies. used. These methodologies consider a different set of midpoint cate-
The functional unit (FU) for the compared studies was always 1 kWh gories to evaluate the sustainability of the product or process and adopt
of electricity produced. The analyzed studies included both upstream the characterization factors at the endpoint level for the categories such
(raw material extraction, manufacturing, wind turbine construction) as Human Health, Climate Change, Resources or Ecosystem Quality. We
operational (power generation, operation, and maintenance) and noted that studies involving Ecoindicator 99/95 and ReCiPe assessment
downstream (decommissioning and end of life scenarios) processes. The methods [44,62,75,83,92] report the LCA results for WTs as endpoints
studies differ mainly in the assumptions on input and output flows or normalized weighted single scores rather than the midpoint or LCI
concerning upstream and downstream processes and their associated results. Such an approach is simple and clear. Therefore it is often
emissions. One of the primary sources of inconsistency among the chosen in decision making support [107]. However, it is a matter of fact
studies regards the impacts of electricity mix used in the upstream that midpoint modeling, as well as reporting LCI results, make the LCA

466
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 2. Capacity factors reported in the literature for onshore and offshore WTs technology and different nameplate capacity ranges. Onshore site WTs are indicated
with “On”, while “Off” stands for the offshore site. The corresponding numbers refer to the case studies sample size. Box is constrained by first and third quartile;
horizontal red bar within the box is the median value, while upper and lower whiskers are maximum and minimum values. (For interpretation of the references to
colour in this figure legend, the reader is referred to the Web version of this article.)

study more transparent. For this reason, even though the mentioned charts presented in Figs. 4–7 describe the distribution of the considered
works are still crucial for this research, as they report a material in- indicators for the environmental performances. Each Figure is divided
ventory of WTs, we exclude endpoint and single score results collected into seven subplots, each of them referring to a specific range of
in the literature from the present analysis. Fig. 3 presents a number of nameplate capacity for the WT. This allows to perform a comprehensive
case studies analyzing particular midpoints and LCI results. comparison among different cases. Further, we report on the x-axis of
With more than 80 case studies available, the midpoint impacts each subplot the number of case studies considering the given indicator,
GWP, AP and EP are the most considered in the literature. It follows the by distinguishing between onshore and offshore applications. This also
non-renewable (NR) CED midpoint impact, which appears in a wide serves to provide a better quantitative representation of the current
number of works as well. We, therefore, select the mentioned in- research trend.
dicators, and we analyze them in details, by providing statistical data In Fig. 4 the GWP environmental performance indicator is reported.
about their resulting values from the different works. The following As it was also mentioned before, offshore applications are taken into

467
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Table 3
Overview of the impact assessment methods included in the case studies reviewed.
LCIA Nr of case studies References

GHG emissions 58 [18,29,34,37,41–43,45,48,59,63,64,71,74,76,77,81,82,90–92,94–96]


Single emissions: CO2, SO2, NOx, NH3, N2O, PM, NMVOC 48 [29,30,34,37,39,40,47,48,55,56,58–60,64,71,81,102]
CED, EPBT 48 [18,29,38,44,48–54,59,61,68–70,76,78,79,82,86,87,89,98,100]
CML 48 [8,35,38–40,45,49–59,61,65–70,72,73,78–80,86–89,93,102]
LC-TEC 24 [31,36,45]
IMPACT2002+ 20 [48,59,61,84,94]
Ecoindicator95/99 16 [44,46,59,61,75,83,99]
ReCiPe midpoint 7 [59,62,85]
Other: UMIP, TRACI, EPID, ILCD, DIP, EPS 4 [61,62,71,100]

median values decrease with the increase of the nominal WT capacity,


as expected. This is true for both onshore and offshore installations.
Finally, we observe that results, in terms of GWP, are comparable be-
tween onshore and offshore systems among all the cases of large-size
power turbines.
Similar considerations to those as for GWP can be applied to the
other considered indicators. In particular, we show the AP in Fig. 5. For
this midpoint impact, the overall median values are 0.057 and 0.084 g/
kWh for onshore and offshore applications, respectively. Also, we re-
port that, overall, AP values range between 0.012 and 4.100 g/kWh in
the case of onshore applications, while the offshore ones exhibit a range
of variability which is comprised between 0.027 and 0.215 g/kWh.
The number of estimates for offshore WTs is still very limited.
Therefore the range of variability of AP across all the nominal capa-
cities does not provide an exhaustive scenario. Despite this, we observe
that offshore wind power systems show comparable or slightly higher
AP than onshore systems for large-size installations with capacity lower
than 5000 kW. For higher nominal capacities, the offshore WTs seem to
be characterized by significantly higher values of AP.
In Fig. 6 we present results for the EP impact indicator. The litera-
ture survey leads to median values of 0.007 and 0.005 g/kWh for on-
Fig. 3. Number of WTs case studies considered in terms of environmental
shore and offshore systems, respectively. The ranges of variability are
stressors and methodology used in LCA. GWP100 – Global Warming Potential
from 0.001 to 0.380 g/kWh for onshore applications and from 0.001 to
(100 years' time horizon); AP – Acidification Potential, EP – Eutrophication
Potential, ODP – Ozone Layer Depletion, TETP – Terrestrial EcoToxicity, HTTP
0.009 g/kWh for offshore ones. The general trends observed in the
– Human Toxicity Potential, FAETP – Freshwater Aquatic EcoToxicity, MAETP - previously analyzed indicators are also found here. Micro-scale power
Marine Aquatic EcoToxicity, CED – Cumulative Energy Demand, TEC – systems results are characterized by relatively high uncertainty. Also,
Thermoecological Cost, GHG – Greenhouse Gases; elementary emissions: CO2, for onshore installations, median values for EP tend to decrease as the
NH3, NMVOC, N2O, NOx, PM, SO2. nameplate capacity is increased, while for offshore applications the
available works report roughly similar results.
consideration only for large-size cases, namely for power above Finally, in Fig. 7, the CED impact indicator is shown. The overall
1000 kW. Specifically, only 6 case-studies in total are available for median values across all of the considered power plant sizes are 0.130
GWP. The overall median values, for onshore and offshore are 9.700 and 0.103 MJ/kWh for onshore and offshore systems, respectively. The
and 9.990 g/kWh, respectively. For onshore applications, in particular, minimum value for onshore WT applications is 0.030 MJ/kWh, while
the range of variability for the GWP is significantly high, that is be- its maximum value is 9.200 MJ/kWh. As far as the offshore systems are
tween 4.8 and 560.0 g/kWh. This indicates a high dispersion of results, concerned, minimum and maximum values are 0.078 and 0.196 MJ/
especially for micro-size cases, where different base assumptions lead to kWh, respectively. In particular, we observe that for large-scale WTs,
very different estimations. The 50% of the results published for WTs the CED indicator estimation for offshore cases is comparable to the one
below 20 kW report values for GWP varying between 72 and 189 g/ related to onshore systems.
kWh. However, in the same power range, 4 over 14 estimates of
equivalent CO2 emissions were higher than 200 g CO2 eq/kWh, with two
3.2. Simplified LCA models for AP, EP, GWP and CED impacts
of them referring to micro VAWTs of nameplate capacity below 1 kW
[45,64]. The remaining cases were studied by Amor et al. [84], who
Next step of this study was to generate simplified and harmonized
reported the highest values of GWP for both micro and small-scale WTs
LCA models. Simplified LCA models were established for AP, EP, GWP
(560 g/kWh and 240 g/kWh). In particular, Amor and co-authors took
and CED impacts on the basis of the detailed LCA of onshore wind
into account a total of three HAWTs with power rate of 1 kW, 10 kW
turbines with capacities between 1 and 5000 kW, and offshore wind
and 30 kW, and they assumed different electricity outputs depending on
turbines with capacity ranging from 500 to 8000 kW. Detailed LCA of
the site wind conditions.
wind turbines followed the ISO 14040 and ISO 14044 guidelines
On the contrary, there is only a slight variation in the results for the
[12,13]. The function of the wind turbine is to produce electricity. The
offshore cases, with minimum and maximum values of 7.767 and
FU is defined as 1 kWh of electricity generated by WTs. An attributional
32.000 g/kWh, respectively. This is due to the low amount of works
approach was applied to LCI modeling. Fig. 8 presents the boundary of
available. More case studies are indeed required before judgment in this
the system.
concern can be made. In addition, we notice that, in general, GWP
All the relevant flows were considered from cradle to grave

468
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 4. Global Warming Potential impact indicator results reported in the literature for different types and nameplate capacities range of WTs. Onshore site WTs are
indicated with “On”, while “Off” stands for the offshore site. The corresponding numbers refer to the case studies sample size. Box is constrained by first and third
quartile; horizontal red bar within the box is the median value, while upper and lower whiskers are maximum and minimum values. (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)

perspective. In order to provide harmonized results, analogous LCI and Inventory analysis was developed using both primary and secondary
LCIA modeling approaches for system boundary were considered for all data. Primary data for the main processes in the foreground system, i.e.
analyzed cases. Thus, both upstream (raw material extraction, manu- input energy and materials flows at the construction, maintenance, and
facturing, wind turbine construction) operational (power generation, transportation stages were gathered from the literature. Detailed in-
operation and maintenance) and downstream (decommissioning and formation is reported in our previous work [18]. Assumptions regarding
end of life scenarios) processes were considered. Moreover, the fore- end – of – life and transportation phases are presented in Supplemen-
ground and the background systems are included in the system tary Material. Secondary data such as inventory for electricity or con-
boundary. Foreground system refers to construction and operation of struction material production were retrieved from the European Re-
the facility fulfilling the FU, i.e. wind power plant. Background system ference Life Cycle Database (ELCD) and the Ecoinvent database version
considers all the processes of production of entering flows and dealing 3.2. For consistency average EU mixes regarding electricity, materials
with the exiting flows. Operating lifetime for WTs was assumed to be 20 and other resources were considered.
years. The considered environmental impacts were modeled as a sum of

469
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 5. Acidification Potential impact indicator results reported in the literature for different types and nameplate capacities range of WTs. Onshore site WTs are
indicated with “On”, while “Off” stands for the offshore site. The corresponding numbers refer to the case studies sample size. Box is constrained by first and third
quartile; horizontal red bar within the box is the median value, while upper and lower whiskers are maximum and minimum values. (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)

cumulative equivalent emission or energy of all resources that are re- energy carrier (MJ or MJ/kWh); EIj– environmental impact per MJ of
quired to provide a process or a product and fulfill the FU. We used energy carrier j (kg/MJ); rsubs– direct emissions in foreground system
SimaPro software in combination with the Ecoinvent database version (kg or kg/kWh); Fsubs – characterization factor for direct emission and
3.2 to calculate the equivalent emissions. Generic life cycle environ- environmental impact used (kg/kg).
mental impact, EILC , is defined as follows (1). Environmental performance of electricity produced by WTs depends
on specific site conditions that determine the energy production of the
EILC = mi EIi + ej EI j + mk EIk + rsubs Fsubs system. Annual electricity generated by WT was calculated using the
i j k subs (1) power curves provided by manufactures [108] and Weibull wind speed
distribution at the site considering wind speed classes reported in In-
where: mi – mass of material resource I (kg or kg/kWh); EIi – en- ternational Electrotechnical Commission (IEC). IEC standards [109]
vironmental impact per kg of substance i (kg/kg); mk – mass of waste k provides wind turbine classes that are determined by two parameters–-
produced in the process (kg or kg/kWh); EIk – environmental impact the average wind speed and turbulence, as presented in Table 4, and are
per treatment of kg of waste k (kg/kg); ej – amount of energy from

470
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 6. Eutrophication impact indicator results reported in the literature for different types and nameplate capacities range of WTs. Onshore site WTs are indicated
with “On”, while “Off” stands for the offshore site. The corresponding numbers refer to the case studies sample size. Box is constrained by first and third quartile;
horizontal red bar within the box is the median value, while upper and lower whiskers are maximum and minimum values. (For interpretation of the references to
colour in this figure legend, the reader is referred to the Web version of this article.)

commonly used for design purposes. characterized by the cut-off velocity, meaning that over a certain value
A specific site can be characterized by two parameters: the shape of wind velocity the WT is shut down for safety reasons. The cut-off
factor k and the scale factor c. The Weibull probability density function velocity increases with the WT size.
(PDF) is given by equation (2): The final environmental impact of electricity generated in the
considered system is calculated as follows (3):
k
v
W (v ) = 1 exp EILC
c (2) EI =
LT × Annual Electricity (3)

The scale factor c has units of speed, and it is directly proportional Generation of simplified LCA models of selected impacts of elec-
to the mean wind speed at the site. The shape factor k is a non-di- tricity produced by WTs was based by considering a detailed LCI of 25
mensional parameter, and it is inversely related to the variance of the different WTs, as well as literature data. As far as literature data is
average wind speed [110]. By applying the wind speed distribution of concerned, we considered only those studies which resulting impact
manufacturers power curves annual electricity production is estimated. indicator values were close to the averaged-value. In particular, we
It is important to remind that manufacturer's power curves are considered data points falling within the range of the first and third

471
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 7. Cumulative Energy Demand impact indicator results reported in the literature for different types and nameplate capacities range of WTs. Onshore site WTs are
indicated with “On”, while “Off” stands for the offshore site. The corresponding numbers refer to the case studies sample size. Box is constrained by first and third
quartile; horizontal red bar within the box is the median value, while upper and lower whiskers are maximum and minimum values. (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)

quantile, as shown in Figs. 4–7. data points.


The reference simplified model of environmental impact is based on
the one-term power series as it describes well the correlation between 4. Results and discussion
the environmental impacts and the turbine capacity [31].
4.1. Synthesis and generalization of the environmental impacts
EI = C1 × PC2 (4)
We consider the installed capacity of WTs as the main parameter in Figs. 9–12 show the total LC-environmental impacts EILC of the
the model, the fitting of data was performed for different wind condi- analyzed WTs, as a function of the nominal turbine power. The results
tions. We applied the nonlinear regression model based on least-squares are presented in logarithmic scale. Each Figure refers to a particular
for estimation of function parameters C1 and C2. The process started impact indicator and report fitting model results (line), as well as em-
with initial estimates and then iteratively converged on parameter es- pirical data coming from detailed LCA (circles), for either onshore and
timates that provide the best fit of the underlying model to the actual offshore systems. Overall, the total LC-impacts exhibits a qualitatively

472
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 8. System boundary in the LCA of the wind turbine.

Table 4
Wind classes according to IEC standards [109].
Wind class Annual average speed (m/ Extreme 50 – year speed (m/
s) s)

IEC I high wind 10 70.0


IEC II medium wind 8.5 59.5
IEC III low wind 7.5 52.5
IEC IV 6.0 42.0

similar trend for onshore and offshore WTs. This is especially true for
GWP, EP and CED indicators. The power law behavior is mainly due to
the nonlinear scaling factor existing in the amount of materials required
for the construction of the turbines. Data fitting provides a sufficiently
high correlation coefficient (R2 is between 0.9460 and 0.9874 for on-
shore and between 0.8710 and 0.9845 for offshore). This indicates that
the one-term power series describes the behavior of the impact in-
dicators accurately, thus providing a useful correlation to estimate the
global LC-environmental performance indicator for a specific turbine
model with a given nominal power. From Figs. 9–12, we observe that
differences of the environmental impact of WTs with different nominal
power are much greater in the case of offshore applications, for all of
the considered indicators. In fact, we notice that, within the range of
large-scale systems, even slight variations in nominal WT power cor-
respond to significant variations in the values of considered impact
indicators. Figs. 9–12 highlight also the higher general impact of off- Fig. 9. LC-GWP as a function of nominal turbine power.
shore WTs with respect to the onshore ones. Values for GWP, AP, EP,
and CED are higher for offshore applications in the whole range of
onshore transmission system. Moreover, the construction of an offshore
nominal power considered. These results may be justified by observing installation presents a greater operational complexity as well as several
that offshore installations require, in general, a more complex structure
technical challenges. This increasing level of difficulty unavoidably
than those at onshore sites. Offshore WTs need to be connected to the leads to a higher environmental impact, comparing to the onshore

473
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 12. LC-CED as a function of nominal turbine power.

We perform Monte Carlo simulation that allows propagating the


Fig. 10. LC-AP as a function of nominal turbine power.
parameter uncertainties and estimating the overall uncertainty of the
result. The random values from the probability distributions were es-
timated at 10 000 runs. The forecast distribution of the result was ob-
tained and analyzed using the coefficient of variation (CV), which re-
present the ratio of the standard deviation (St.Dev) and mean values.
Tables 6 and 7 report results of uncertainty analysis for onshore and
offshore WTs.
Different CV values are observed for different WTs capacities for
both onshore and offshore applications. Generally, the highest variation
was observed for EP impact, while the lowest for CED impact for both
onshore and offshore WTs datasets. The average CV was 24.5, 21.7,
26.9 and 19.6 for GWP, AP, EP and CED of onshore WTs. The average
uncertainty was slightly higher for offshore wind power systems (26.0,
27.3, 28.6 and 21.3 for GWP, AP, EP and CED, respectively). Such
uncertainty is acceptable for preliminary evaluation of the environ-
mental performance of wind energy systems.
Environmental impacts of electricity produced by specific WTs may
also be affected by specific wind conditions. Figs. 13–16 present the
environmental performance of wind energy under different wind con-
ditions. Results are provided for onshore installations only since the
obtained estimations for offshore cases were found to be too dispersed
to the point that no suitable fitting could be applied. This emphasizes
the need for data regarding the offshore applications that, as discussed
before, are available in a small amount.
Fig. 11. LC-EP as a function of nominal turbine power. Data fitting for wind conditions analysis has been performed by
means of the one-term power series model. Table 8 presents fitting
parameters obtained for onshore WTs for different IEC wind speed class.
installations.
In fact, also, in this case, this mathematical law was found to be the
As a next step, we assess the reliability of the simplified LCA model.
most appropriate to describe the trends for the LC-environmental im-
Background process LCI data are a major source of the results un-
pact of electricity produced in wind power systems.
certainty in the LCA. In order to investigate the distribution of the
From Figs. 13–16, we observe that the impact of the LC-environ-
possible outputs and estimate the results uncertainty, deterministic
mental of electricity decreases when higher nominal power is con-
input secondary data are transformed into probabilistic data. We con-
sidered, as expected. The total environmental impact is higher at the
sider main input parameters such as impact of steel, aluminum and
larger WTs capacity scales. However, WTs of higher nominal power are
concrete production processes as a variables. It was assumed that all the
generally more efficient, thus leading to LC-impacts of electricity lower
considered parameters are uniformly distributed. Maximum and
than the one obtained at smaller scales. In particular, by analyzing all
minimum values for the uniform probability distribution are based on
the considered impact indicators for a given WT capacity, we notice
the data reported in the Ecoinvent database. The range of variation of
that, in general, the highest values of LC-impacts of electricity corre-
input parameters was chosen on the basis of the data regarding average
spond to the highest wind velocities. This is particularly valid for low
World and European production mixes and presented in Table 5.
nominal power turbines, which seem to be significantly affected by

474
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Table 5
Selected parameters and their uncertainty distribution.
LCI parameter Uncertainty distribution Reference Ecoinvent records

Chromium steel, hot rolled, kg GWP Uniform (0.184–0.196) kg CO2 eq Chromium steel, hot rolled {RER}\{GLO}|, market for, Alloc Rec, S
AP Uniform (0.550–0.621) kg SO2 eq
EP Uniform (0.191–0.196) kg PO4 eq
CED Uniform (1.045–1.226) MJ
Aluminum, kg GWP Uniform (7.259–9.379) kg CO2 eq Aluminum, wrought alloy {RER}/{GLO}| market for | Alloc Rec, S
AP Uniform (75.990–81.000) kg SO2 eq
EP Uniform (11.943–20.888) kg PO4 eq
CED Uniform (76.162–81.572) MJ
Concrete, kg GWP Uniform (4.495–4.790) kg CO2 eq Concrete block {RER}/{GLO}| market for | Alloc Rec, S
AP Uniform (28.493–29.865) kg SO2 eq
EP Uniform (8.359–8.859) kg PO4 eq
CED Uniform (44.105–47.451) MJ

Table 6
Uncertainty analysis for onshore WTs.
Onshore wind turbines model and capacity GWP AP EP CED

kg CO2 eq kg SO2 eq kg PO4 eq MJ

Bergey BWCXL1 1 kW Mean 2740.00 17.86 2.56 31862.13


St.Dev 849.40 4.91 0.87 7930.65
CV, % 31.00 27.50 34.10 24.89
Westwind 3 kW Mean 6600.00 40.42 5.79 106312.94
St.Dev 1932.00 10.50 1.86 24987.42
CV, % 29.27 25.97 32.20 23.50
Endurance S-343 5 kW Mean 8380.00 52.50 8.46 127307.02
St.Dev 2133.09 11.85 2.37 26018.95
CV, % 25.45 22.58 28.00 20.44
Jacobs. C&F Green Energy 20 kW Mean 25000.00 137.61 23.65 283966.62
St.Dev 4977.27 24.30 5.18 45393.20
CV, % 19.91 17.66 21.90 15.99
Pitchwind 30 kW Mean 29800.00 186.13 31.94 382465.10
St.Dev 7314.55 40.53 8.62 75376.33
CV, % 24.55 21.77 27.00 19.71
Northern Power 100 kW Mean 104000.00 598.36 78.03 1289273.03
St.Dev 27796.36 141.87 22.94 276676.11
CV, % 26.73 23.71 29.40 21.46
Vestas V39-500 kW Mean 210000.00 1460.42 257.48 2723695.11
St.Dev 55936.36 345.08 75.44 582512.90
CV, % 26.64 23.63 29.30 21.39
Tacke 600e. Vestas V42 600 kW Mean 363000.00 2444.81 294.77 5134528.30
St.Dev 85833.00 512.82 76.67 974810.81
CV, % 23.65 20.98 26.01 18.99
Vestas V47 660 kW Mean 266000.00 1733.11 316.36 3711506.24
St.Dev 59003.64 341.03 77.19 661027.39
CV, % 22.18 19.68 24.40 17.81
WindEnergyLebanon 750 kW Mean 340000.00 2099.78 347.83 4938302.80
St.Dev 86730.91 475.16 97.60 1011450.92
CV, % 25.51 22.63 28.06 20.48
Vestas V52 850 kW Mean 345000.00 2321.10 381.67 4542079.73
St.Dev 75304.09 449.43 91.64 796024.34
CV, % 21.83 19.36 24.01 17.53
SuzlonS64 1250 kW Mean 411000.00 2700.43 508.08 5817803.83
St.Dev 102563.18 597.80 139.47 1165684.05
CV, % 24.95 22.14 27.45 20.04
Leitwind LTW80 1.8 MW Mean 620000.00 4392.43 665.90 9608654.81
St.Dev 150941.82 948.62 178.33 1878246.54
CV, % 24.35 21.60 26.78 19.55
Gamesa G80 2.0 MW Mean 854555.52 4892.22 720.03 10805973.32
St.Dev 172387.15 875.47 159.77 1750252.17
CV, % 20.17 17.90 22.19 16.20
Vestas 2 MW Mean 741000.00 3351.26 720.03 10805973.32
St.Dev 155407.91 623.50 166.11 1819662.81
CV, % 20.97 18.60 23.07 16.84
Vestas V90 3 MW Mean 929000.00 6093.93 972.69 12349020.46
St.Dev 216794.82 1261.54 249.69 2313863.91
CV, % 23.34 20.70 25.67 18.74
Gamesa G128 5 MW Mean 2679686.52 11081.20 1314.01 16978991.09
St.Dev 680396.77 2495.95 367.00 3461483.37
CV, % 25.39 22.52 27.93 20.39

475
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Table 7
Uncertainty analysis for offshore WTs.
Offshore wind GWP AP EP CED
turbines model
and capacity kg CO2 eq kg SO2 eq kg PO4 eq MJ

Vestas V39- Mean 255795.16 670.99 149.39 3313572.00


500 kW St.Dev 69320.49 190.50 44.53 858935.49
CV, % 27.10 28.39 29.81 25.92
Vestas 80 m Mean 1128855.56 1111.88 584.87 14491012.95
2 MW St.Dev 251426.92 259.44 143.29 3087215.80
CV, % 22.27 23.33 24.50 21.30
Enercon E82 Mean 1311111.70 2084.20 671.15 16815216.71
2300 kW St.Dev 369614.31 615.53 208.12 4534259.74
CV, % 28.19 29.53 31.01 26.97
General Electric Mean 1433571.33 3032.07 728.56 18375741.87
2.5xl St.Dev 374031.79 828.77 209.10 4585946.01
2500 kW CV, % 26.09 27.33 28.70 24.96
Vestas 112 m Mean 1742667.39 6882.07 871.81 22311122.05
3 MW St.Dev 481609.90 1992.52 265.03 5897896.61
CV, % 27.64 28.95 30.40 26.43
General Electric Mean 2118408.46 15620.61 1043.22 27089309.96
3600 kW St.Dev 537305.42 4150.62 291.06 6572102.16
CV, % 25.36 26.57 27.90 24.26
Repower 5 MW Mean 3011565.62 68406.18 1441.56 38427846.57
St.Dev 761104.77 18111.35 400.75 9289514.21
CV, % 25.27 26.48 27.80 24.17
Vestas V164 Mean 4317982.16 120492.04 2007.68 54976535.70
7 MW St.Dev 1169780.62 34196.79 598.29 14246093.60 Fig. 14. LC-AP of electricity as a function of nominal turbine power, for dif-
CV, % 27.09 28.38 29.80 25.91 ferent wind speed.
Vestas V164 Mean 4981778.57 165936.59 2289.75 63372534.63
8 MW St.Dev 1259031.31 43933.69 636.55 15319621.41
CV, % 25.27 26.48 27.80 24.17

Fig. 15. LC-EP of electricity as a function of nominal turbine power, for dif-
ferent wind speed.
Fig. 13. LC-GWP of electricity as a function of nominal turbine power, for
different wind speed. indicators and different ranges of nominal power. These results can be
explained by considering the variability of the capacity factor with the
wind conditions. The trends exhibit an asymptotic behavior, indicating wind conditions. Therefore, we may conclude that the capacity factor
that, on the contrary, wind conditions are a minor contributor to the plays an essential role in the final estimation of the environmental
environmental impact of large-scale systems. performance of electricity produced.
Figs. 13–16 also reveal another interesting aspect. By comparing the Finally, in Table 9, we report a comparison between the values re-
four considered impact indicators among each other, we note that, for a ported by manufacturers and those as derived from our simplified LCA
given WT model, there is no linear correlation between the wind ve- models, for all of the impact indicators considered.
locity and the value for environmental impacts. For instance, from In Table 5 only a few selected cases reporting both wind turbine
Fig. 13, we can observe that AP at low nominal power reaches higher capacity, wind speed and turbulence class are shown for comparison. In
values when the average velocity is at 6 m/s than the case of the ve- fact, despite the relatively large number of LCA studies on the topic,
locity of 7.5 m/s. An opposite behavior is obtained for remaining these are the only reporting data on wind conditions. Our results, as

476
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

To further prove the validity of the model, we compare predictions


from our simplified model with the set of available literature data re-
ported in Figs. 4–7. To this aim, we consider the minimum and the
maximum values among those reported in the literature, for each in-
dicator, which fall within 1 and 3 quartiles for the whole range of
power capacity (1–8000 kW). Thus, we verify if these values are in-
cluded within the range as predicted by our model by considering the
minimum and maximum wind conditions. This comparison is presented
in Fig. 17.
From Fig. 17 we notice that all of the values reported in literature
fall within our predicted ranges, for all of the considered indicators.
This further proves the accuracy of the model.

4.2. Practical implications and limitations of the simplified LCA model

Overall the proposed simplified LCA models are suitable for pre-
liminary identification of environmental performance of wind energy
application, comparison of environmental aspect among the same
technology or benchmarking of specific technology against other
technologies. On the other hand, by using a simplified LCA model is not
possible to analyze a breakdown of the processes within the boundary.
Thus, simplified models are not adequate for a detailed optimization or
hot spot analysis. Models are focusing on life cycle GWP, AP, EP and
Fig. 16. LC-CED of electricity as a function of nominal turbine power, for dif- CED impacts for single WTs. Other impacts should also be included to
ferent wind speed. provide a more general viewpoint. Models are presented for both
electricity produced by WT and entire total life cycle of WTs. The model
Table 8 for electricity produced is generalized on different wind conditions by
Fitting parameters for different IEC wind speed class, onshore WTs case. applying Weibull distribution and manufactures power curves.
However, the simplified LCA model for total impact is also provided;
LCIA Wind speed and C1 C2 R2
therefore it is also possible to assess WTs electricity impact using
turbulence class (IEC)
measured data.
Global Warming I A 133.017 −0.387 0.992 Crucial limitations are related to data collection and their techno-
Potential,,g CO2eq/ II A 124.028 −0.381 0.987 logical, temporal, geographical and methodological coverage. Firstly,
kWh III A 107.313 −0.368 0.983
collected data cover onshore and offshore Horizontal Axis Wind
IV A 109.57 −0.37 0.987
Acidification Potential, g I A 0.867 −0.403 0.993 Turbine (HAWT) technology with power capacities ranged from 1 to
SO2eq/kWh II A 0.801 −0.38 0.993 5000 kW and 500 to 8000 kW, respectively. VAWT technologies in
III A 0.525 −0.325 0.99 general and offshore HAWTs were not included in the data collection,
IV A 0.71 −0.387 0.986
because of the lack of available information. Therefore they are not
Eutrophication Potential, I A 0.125 −0.373 0.993
g PO4eq/kWh II A 0.114 −0.374 0.997
considered in simplified LCA models. The study refers to the WTs
III A 0.099 −0.362 0.992 technologies available in the past and present on the market. It is re-
IV A 0.104 −0.361 0.992 ported high maturity of both offshore and onshore WTs technology with
Cumulative Energy I A 1.904 −0.4 0.984 corresponding Technology Readiness Level (TRL) of 9/10 [111].
Demand, MJ/kWh II A 1.312 −0.348 0.984
However, further innovations are also ongoing and they will refer not
III A 1.264 −0.346 0.971
IV A 1.596 −0.366 0.976 only to turbine size but also to components and foundations design.
After new data and new technologies are available, models could be
updated by including new data sets and learning curves.
Table 9 Results from presented simplified LCA models refer to WTs installed
Comparison of values retrieved from literature with the values calculated in this and manufactured in Europe, by considering secondary datasets for
study. background processes: electricity and material production, which are
Power, kW Wind speed and GWP AP EP CED valid for average EU mix. However, after uncertainty analysis, it
turbulence class (IEC) emerged that simplified models can also be applied for other geo-
g CO2eq/ g SO2eq/ g PO4eq/ MJ/kWh graphical conditions with acceptable uncertainty. Nonetheless, more
kWh kWh kWh research is needed to explore different types of uncertainty (primary
2000 III A model 6.564 0.044 0.006 0.091 data, allocation approach or end – of – life scenarios). Further, har-
III A [66] 7.200 0.045 0.005 0.10 monization of background processes is also recommended in order to
3000 II A model 5.859 0.029 0.003 0.08 increase the reliability and consistency of the whole study, followed by
II A [78] 6.200 0.038 0.006 0.08 methodological harmonization of processes based on the entire cycle
3450 IA model 5.667 0.021 0.002 0.07
sensitive check.
IA [67] 5.300 0.022 0.006 0.06

5. Conclusions and recommendations


predicted from the proposed model, are in line with those from litera-
ture. Overall, we found a small deviation for all of the considered cases. In this study, we performed a thorough review of the LCA studies for
However, we emphasize that the available studies are based on dif- wind energy systems, by taking into account the main available works
ferent assumptions. Therefore a direct comparison with other results in literature. Particular emphasis was given to simplified LCA models
must take into account this aspect. and harmonization procedures. The results for the detailed environ-
mental impact for a wide range of wind turbines capacities were

477
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

Fig. 17. Environmental impacts of WTs of different sizes: comparison of values retrieved from literature with the values calculated using simplified LCA models.

collected and summarized thus leading to the realization of a novel turbines. J Clean Prod 2013;39:191–9. https://doi.org/10.1016/J.JCLEPRO.2012.
simplified correlation model. The main feature of such an approach is 08.022.
[8] Martínez E, Latorre-Biel JI, Jiménez E, Sanz F, Blanco J. Life cycle assessment of a
the possibility to decrease the number of input parameters to the model wind farm repowering process. Renew Sustain Energy Rev 2018;93:260–71.
as compared to detailed LCA while keeping an acceptable level of un- https://doi.org/10.1016/j.rser.2018.05.044.
certainty. Our simplified LCA model starts from extensive LCAs and is [9] WWEA. The world wind energy association: small wind Latest News 2013http://
www.wwindea.org/webimages/WWEA_half_year_report_2014.pdf.
then developed for GWP, AE, EP and CED indicators, applied to onshore [10] Udo de Haes HA, Heijungs R. Life-cycle assessment for energy analysis and man-
and offshore wind power systems. Such a model may provide a useful agement. Appl Energy 2007;84:817–27. https://doi.org/10.1016/j.apenergy.
tool to estimate the impact of a given turbine, described according to its 2007.01.012.
[11] Lombardi L, Mendecka B, Carnevale E. Comparative life cycle assessment of al-
nominal power under different environmental conditions. It can be ternative strategies for energy recovery from used cooking oil. J Environ Manag
concluded that the definition of a simplified model can be useful to 2018;216:235–45. https://doi.org/10.1016/j.jenvman.2017.05.016.
understand the broad range of variability of LCA results for different [12] ISO. ISO 14040: environmental management: life-cycle assessment: principles and
framework. Geneva, Switzerland: International Organization for Standardization;
analyzed scenarios. Moreover, it can also be stated that, in general, WTs
2006 Available at: http://www.iso.org.
systems can be adequately investigated by using simplified LCA models. [13] ISO. ISO 14044: environmental management: life-cycle assessment: requirements
However, it should be underlined that this model presents certain and guidelines Available at Geneva, Switzerland: International Organization for
limitations considering data collection. Standardization; 2006. : http://www.iso.org.
[14] European Commission - Joint Research Centre - Institute for Environment and
Further works should focus on: i) inclusion of other environmental Sustainability. International reference life cycle data system (ILCD) Handbook :
indicators, ii) improving prediction of LCA models by considering in- analysing of existing environmental impact assessment methodologies for use in
novative technologies, iii) development of a simplified LCA model for life cycle assessment. Eur Community 2010;115. https://doi.org/10.2788/38479.
[15] Bala A, Raugei M, Benveniste G, Gazulla C, Fullana-i-Palmer P. Simplified tools for
an entire wind farm, which can be performed by investigating on the global warming potential evaluation: when ‘good enough’ is best. Int J Life Cycle
correlation between the farm size and the additional material con- Assess 2010;15:489–98. https://doi.org/10.1007/s11367-010-0153-x.
sumptions, and, finally, iv) further harmonization of the results fo- [16] Price L, Kendall A. Wind power as a case study. J Ind Ecol 2012;16:S22–7. https://
doi.org/10.1111/j.1530-9290.2011.00458.x.
cusing on the background process modeling and methodological [17] Raadal HL, Gagnon L, Modahl IS, Hanssen OJ. Life cycle greenhouse gas (GHG)
choices. emissions from the generation of wind and hydro power. Renew Sustain Energy
Rev 2011;15:3417–22. https://doi.org/10.1016/j.rser.2011.05.001.
[18] Raadal HL, Vold BI, Myhr A, Nygaard TA. GHG emissions and energy performance
Appendix A. Supplementary data
of offshore wind power. Renew Energy 2014;66:314–24. https://doi.org/10.1016/
j.renene.2013.11.075.
Supplementary data to this article can be found online at https:// [19] Lenzen M, Munksgaard J. Energy and CO 2 life-cycle analyses of wind turbines —
review and applications. Renew Energy 2002;26:339–62.
doi.org/10.1016/j.rser.2019.05.019.
[20] Arvesen A, Hertwich EG. Assessing the life cycle environmental impacts of wind
power: a review of present knowledge and research needs. Renew Sustain Energy
References Rev 2012;16:5994–6006. https://doi.org/10.1016/j.rser.2012.06.023.
[21] Dreyer LC, Niemann AL, Hauschild MZ. LCA discussions comparison of three LCIA
methods comparison of three different LCIA methods: EDIP97, CML2001 and eco-
[1] IEA Wind. IEA WIND technology colaboration programme Annual Report 2017 indicator 99 does it matter which one you choose? Int J LCA 2003;8:191–200.
2018 https://doi.org/10.1065/Ica2003.06.115.
[2] EUROSTAT. Infrastructure – electricity – annual data [nrg_113a]. Accessed 26th [22] Heath GA, Mann MK. Background and reflections on the life cycle assessment
May 2016 [n.d]. harmonization project. J Ind Ecol 2012;16:S8–11. https://doi.org/10.1111/j.
[3] GWEC. Global wind energy report: annual market update. 2017Https://GwecNet/ 1530-9290.2012.00478.x.
Policy-Research/Reports/2018:72. [23] Brandão M, Heath G, Cooper J. What can meta-analyses tell us about the reliability
[4] Tabassum-Abbasi, Premalatha M, Abbasi T, Abbasi SA. Wind energy: increasing of life cycle assessment for decision support? J Ind Ecol 2012;16. https://doi.org/
deployment, rising environmental concerns. Renew Sustain Energy Rev 10.1111/j.1530-9290.2012.00477.x. S3–7.
2014;31:270–88. https://doi.org/10.1016/J.RSER.2013.11.019. [24] Asdrubali F, Baldinelli G, D'Alessandro F, Scrucca F. Life cycle assessment of
[5] Gonzalez E, Ortego A, Topham E, Valero A. Is the future development of wind electricity production from renewable energies: review and results harmonization.
energy compromised by the availability of raw materials? J Phys Conf Ser Renew Sustain Energy Rev 2015;42:1113–22. https://doi.org/10.1016/J.RSER.
2018;1102:012028https://doi.org/10.1088/1742-6596/1102/1/012028. 2014.10.082.
[6] IEA Wind. Long-term research and development needs for wind energy for the time [25] Dolan SL, Heath GA. Life cycle greenhouse gas emissions of utility-scale wind
frame 2012 to 2030. Boulder, USA: International Energy Agency Implementing power: systematic review and harmonization. J Ind Ecol 2012;16. https://doi.org/
Agreement for Co-operation in the R&D of Wind Turbine Systems; 2013. [n.d]. 10.1111/j.1530-9290.2012.00464.x.
[7] Ortegon K, Nies LF, Sutherland JW. Preparing for end of service life of wind [26] Carbajales-Dale M. Life cycle assessment: meta-analysis of cumulative energy

478
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

demand for wind energy technologies. Wind Energy Eng 2017:439–73. https:// Cycle Assess, [online] 21 (9), pp.1218–1230. Available at:: 10.1007/s11367-016-
doi.org/10.1016/B978-0-12-809451-8.00021-7. 1087-8 > 2016.
[27] Padey P, Blanc I, Le Boulch D, Xiusheng Z. A simplified life cycle approach for [60] Al-Behadili SH, El-Osta WB. Life cycle assessment of dernah (Libya) wind farm.
assessing greenhouse gas emissions of wind electricity. J Ind Ecol 2012;16:S28–38. Renew Energy 2015;83:1227–33. https://doi.org/10.1016/j.renene.2015.05.041.
https://doi.org/10.1111/j.1530-9290.2012.00466.x. [61] Martínez E, Blanco J, Jiménez E, Saenz-Díez JC, Sanz F. Comparative evaluation of
[28] Padey P, Girard R, le Boulch D, Blanc I. From LCAs to simplified models: a generic life cycle impact assessment software tools through a wind turbine case study.
methodology applied to wind power electricity. Environ Sci Technol Renew Energy 2015;74:237–46. https://doi.org/10.1016/j.renene.2014.08.004.
2013;47:1231–8. https://doi.org/10.1021/es303435e. [62] Haapala KR, Prempreeda P. Comparative life cycle assessment of 2.0 MW wind
[29] Guezuraga B, Zauner R, Pölz W. Life cycle assessment of two different 2 MW class turbines. Int J Sustain Manuf 2014;3:170–85. https://doi.org/10.1504/IJSM.
wind turbines. Renew Energy 2012;37:37–44. https://doi.org/10.1016/j.renene. 2014.062496.
2011.05.008. [63] Reimers B, Özdirik B, Kaltschmitt M. Greenhouse gas emissions from electricity
[30] Ardente F, Beccali M, Cellura M, Lo Brano V. Energy performances and life cycle generated by offshore wind farms. Renew Energy 2014;72:428–38. https://doi.
assessment of an Italian wind farm. Renew Sustain Energy Rev 2008;12:200–17. org/10.1016/j.renene.2014.07.023.
https://doi.org/10.1016/j.rser.2006.05.013. [64] Uddin MS, Kumar S. Energy, emissions and environmental impact analysis of wind
[31] Mendecka B, Lombardi L, Stanek W. Analysis of life cycle thermo-ecological cost of turbine using life cycle assessment technique. J Clean Prod 2014;69:153–64.
electricity from wind and its application for future incentive mechanism. Energy https://doi.org/10.1016/j.jclepro.2014.01.073.
Convers Manag 2018;170:73–81. https://doi.org/10.1016/j.enconman.2018.05. [65] Vestas. Life cycle assessment of electricity production from an onshore V100-2.0
084. MW wind plant. 2015.
[32] Cadu M, Huijbregts MAJ, Althaus H-J, Koehler A, Hellweg S, Caduff M, et al. Wind [66] Vestas. Life cycle assessment of electricity production from an onshore V110 - 2.0
power Electricity : the bigger the turbine , the greener the. Environ Sci Technol MW wind plant. 2015.
2012;46:4725–33. https://doi.org/10.1021/es204108n. [67] Vestas. Life cycle assessment of electricity production from an onshore V112-3.3
[33] Turconi R, Boldrin A, Astrup T. Life cycle assessment (LCA) of electricity genera- MW wind plant. 2014.
tion technologies: overview, comparability and limitations. Renew Sustain Energy [68] Vestas. Life cycle assessment of electricity production from an onshore V105-3.3
Rev 2013;28:555–65. https://doi.org/10.1016/j.rser.2013.08.013. MW wind plant. 2014.
[34] Chipindula J, Botlaguduru V, Du H, Kommalapati R, Huque Z. Life cycle en- [69] Vestas. Life cycle assessment of electricity production from an onshore V117-3.3
vironmental impact of onshore and offshore wind farms in Texas. Sustainability MW wind plant. 2014.
2018;10:2022. https://doi.org/10.3390/su10062022. [70] Vestas. Life cycle assessment of electricity production from an onshore V126-3.3
[35] Ozoemena M, Cheung WM, Hasan R. Comparative LCA of technology improve- MW wind plant. 2014.
ment opportunities for a 1.5-MW wind turbine in the context of an onshore wind [71] Utomo Adi, Tim Daly, Thibaut Chanvrier IH. H2OCEAN Developement of a wind-
farm. Clean Technol Environ Policy 2018;20:173–90. https://doi.org/10.1007/ wave power open-sea platform equipped for hydrogen generation with support for
s10098-017-1466-2. multiple users of energy. D9.9 Life cycle assessment report vol. 288145. 2014.
[36] Stanek W, Mendecka B, Lombardi L, Simla T. Environmental assessment of the [72] Demir N, Taşkın A. Life cycle assessment of wind turbines in Pınarbaşı-Kayseri. J
wind turbine systems based on thermo-ecological cost. Energy 2018. https://doi. Clean Prod 2013;54:253–63. https://doi.org/10.1016/j.jclepro.2013.04.016.
org/10.1016/j.energy.2018.07.032. [73] Greening B, Azapagic A. Environmental impacts of micro-wind turbines and their
[37] Yang J, Chang Y, Zhang L, Hao Y, Yan Q, Wang C. The life-cycle energy and en- potential tocontribute to UK climate change targets. Energy 2013;59:454–66.
vironmental emissions of a typical offshore wind farm in China. J Clean Prod https://doi.org/10.1016/j.energy.2013.06.037.
2018;180:316–24. https://doi.org/10.1016/j.jclepro.2018.01.082. [74] Oebels KB, Pacca S. Life cycle assessment of an onshore wind farm located at the
[38] Vestas. Life cycle assessment of electricity production from an onshore V116-2.0 northeastern coast of Brazil. Renew Energy 2013;53:60–70. https://doi.org/10.
MW wind plant. 2018. 1016/J.RENENE.2012.10.026.
[39] Gamesa. EPD: electricity from a European GAMESA G128 - 5.0 MW onshore wind [75] Rashedi A, Sridhar I, Tseng KJ. Life cycle assessment of 50 MW wind firms and
farm. 2018. strategies for impact reduction. Renew Sustain Energy Rev 2013;21:89–101.
[40] Gamesa. EPD: electricity from a European GAMESA G132 - 5.0 MW onshore wind https://doi.org/10.1016/J.RSER.2012.12.045.
farm. 2018. [76] Kabir MR, Rooke B, Dassanayake GDM, Fleck BA. Comparative life cycle energy,
[41] Siemens. A clean energy solution – from cradle to grave. Offshore wind power emission, and economic analysis of 100 kW nameplate wind power generation.
plant employing SWT-3.2-113. 2018. Renew Energy 2012;37:133–41. https://doi.org/10.1016/j.renene.2011.06.003.
[42] Siemens. A clean energy solution – from cradle to grave. Offshore wind power [77] Wang Y, Sun T. Life cycle assessment of CO 2 emissions from wind power plants:
plant employing SWT-6.0-154. 2018. methodology and case studies. Renew Energy 2012;43:30–6. https://doi.org/10.
[43] Siemens. A clean energy solution – from cradle to grave. Offshore wind power 1016/j.renene.2011.12.017.
plant employing SWT-7.0-154. 2018. [78] Vestas. Life cycle assessment of electricity production from an onshore V90 - 3.0
[44] Huang Y-F, Gan X-J, Chiueh P-T. Life cycle assessment and net energy analysis of MW wind plant. 2013.
offshore wind power systems. Renew Energy 2017:98–106. https://doi.org/10. [79] Vestas. Life cycle assessment of electricity production from an onshore V100-2.6
1016/j.renene.2016.10.050. MW wind plant. 2013.
[45] Lombardi L, Mendecka B, Carnevale E, Stanek W. Environmental impacts of [80] Gamesa. EPD: electricity from a European GAMESA G90 - 2.0 MW onshore wind
electricity production of micro wind turbines with vertical axis. Renew Energy farm. 2013.
2018;128:553–64. https://doi.org/10.1016/j.renene.2017.07.010. [81] Arvesen A, Hertwich EG. Corrigendum: environmental implications of large-scale
[46] Siddiqui O, Dincer I. Comparative assessment of the environmental impacts of adoption of wind power : a scenario-based life cycle assessment. 2011. https://doi.
nuclear, wind and hydro-electric power plants in Ontario: a life cycle assessment. J org/10.1088/1748-9326/7/3/039501.
Clean Prod 2017;164:848–60. https://doi.org/10.1016/j.jclepro.2017.06.237. [82] Wagner H-JJ, Baack C, Eickelkamp T, Epe A, Lohmann J, Troy S. Life cycle as-
[47] Wang W-C, Teah H-Y. Life cycle assessment of small-scale horizontal axis wind sessment of the offshore wind farm alpha ventus. Energy 2011;36:2459–64.
turbines in Taiwan. J Clean Prod 2017;141:492–501. https://doi.org/10.1016/J. https://doi.org/10.1016/j.energy.2011.01.036.
JCLEPRO.2016.09.128. [83] Zhong ZW, Song B, Loh PE. LCAs of a polycrystalline photovoltaic module and a
[48] Bonou A, Laurent A, Olsen SI. Life cycle assessment of onshore and offshore wind wind turbine. Renew Energy 2011;36:2227–37. https://doi.org/10.1016/j.renene.
energy-from theory to application. Appl Energy 2016;180:327–37. https://doi. 2011.01.021.
org/10.1016/j.apenergy.2016.07.058. [84] Amor M Ben, Lesage P, Pineau PO, Samson R. Can distributed generation offer
[49] Vestas. Life cycle assessment of electricity production from an onshore V112-3.45 substantial benefits in a Northeastern American context? A case study of small-
MW wind plant. 2017. scale renewable technologies using a life cycle methodology. Renew Sustain
[50] Vestas. Life cycle assessment of electricity production from an onshore V105-3.45 Energy Rev 2010;14:2885–95. https://doi.org/10.1016/j.rser.2010.08.001.
MW wind plant. 2017. [85] Arvesen A, Hertwich EG. Environmental implications of large-scale adoption of
[51] Vestas. Life cycle assessment of electricity production from an onshore V117-3.45 wind power: a scenario-based life cycle assessment. Environ Res Lett
MW wind plant. 2017. 2011;6:045102https://doi.org/10.1088/1748-9326/6/4/045102.
[52] Vestas. Life cycle assessment of electricity production from an onshore V126-3.45 [86] Vestas. Life cycle assessment of electricity production from an onshore V80-2.0
MW wind plant. 2017. MW GridStreamerTM. 2011.
[53] Vestas. Life cycle assessment of electricity production from an onshore V136-3.45 [87] Vestas. Life cycle assesment of electricity production from an onshore V90-2.0 MW
MW wind plant. 2017. GridStreamerTM wind plant. 2011.
[54] Vestas. Life cycle assessment of electricity production from an onshore V120-2.0 [88] Vestas. Life cycle assessment of electricity production from a V100-1.8 MW grid-
MW wind plant. 2018. streamer wind plant. 2011.
[55] Gamesa. EPD: electricity from a European GAMESA G126 - 2.625 MW onshore [89] Vestas. Life cycle assessment of electricity production from an onshore V112-3.0
wind farm. 2017. MW wind plant. 2011.
[56] Gamesa. EPD: electricity from a European GAMESA G132 - 3.465 MW onshore [90] Parsons DJ, Chatterton JC, Brennan FP, Kolios AJ. Carbon Brainprint Case Study:
wind farm. 2017. novel offshore vertical axis wind turbines. 2011.
[57] Gamesa. EPD: electricity from a European GAMESA G114 - 2.5 MW onshore wind [91] Crawford RH. Life cycle energy and greenhouse emissions analysis of wind tur-
farm. 2017. bines and the effect of size on energy yield. Renew Sustain Energy Rev
[58] Gamesa. EPD: electricity from a European GAMESA G114 - 2.0 MW onshore wind 2009;13:2653–60. https://doi.org/10.1016/j.rser.2009.07.008.
farm. 2017. [92] Martínez E, Jiménez E, Blanco J, Sanz F. LCA sensitivity analysis of a multi-
[59] Wernet, G., Bauer, C., Steubing, B., Reinhard, J., Moreno-Ruiz, E., and Weidema B. megawatt wind turbine. Appl Energy 2010;87:2293–303. https://doi.org/10.
The ecoinvent database version 3 (part I): overview and methodology. Int J Life 1016/j.apenergy.2009.11.025.

479
B. Mendecka and L. Lombardi Renewable and Sustainable Energy Reviews 111 (2019) 462–480

[93] Martínez E, Sanz F, Pellegrini S, Jiménez E, Blanco J. Life-cycle assessment of a 2- [102] Vestas. Life cycle assessment of electricity produced from onshore sited wind
MW rated power wind turbine: CML method. Int J Life Cycle Assess power plants based on Vestas V82-1.65 MW turbines. 2006.
2009;14:52–63. https://doi.org/10.1007/s11367-008-0033-9. [103] Turconi R, O'Dwyer C, Flynn D, Astrup T. Emissions from cycling of thermal power
[94] Tremeac B, Meunier F. Life cycle analysis of 4.5 MW and 250 W wind turbines. plants in electricity systems with high penetration of wind power: life cycle as-
Renew Sustain Energy Rev 2009;13:2104–10. https://doi.org/10.1016/j.rser. sessment for Ireland. Appl Energy 2014;131:1–8. https://doi.org/10.1016/j.
2009.01.001. apenergy.2014.06.006.
[95] Weinzettel J, Reenaas M, Solli C, Hertwich EG. Life cycle assessment of a floating [104] Huijbregts Mark AJ, Rombouts Linda JA, Hellweg Stefanie, Frischknecht Rolf, Jan
offshore wind turbine. Renew Energy 2009;34:742–7. https://doi.org/10.1016/j. Hendriks A, Meent Dik van de, et al. Is cumulative fossil energy demand a useful
renene.2008.04.004. indicator for the environmental performance of products? 2005. https://doi.org/
[96] Lenzen M, Wachsmann U. Wind turbines in Brazil and Germany: an example of 10.1021/ES051689G.
geographical variability in life-cycle assessment. Appl Energy 2004;77:119–30. [105] Szargut J. Exergy method: technical and ecological applications. WIT Press; 2005.
https://doi.org/10.1016/S0306-2619(03)00105-3. [106] Stanek W, editor. Thermodynamics for sustainable management of natural re-
[97] Jungbluth N, Bauer C, Dones R, Frischknecht R. The ecoinvent database energy sources Cham: Springer International Publishing; 2017. https://doi.org/10.1007/
supply 1 life cycle assessment for emerging technologies: case studies for photo- 978-3-319-48649-9.
voltaic and wind power. Int J LCA 2004;2004:1–11. https://doi.org/10.1065/ [107] Bare JC, Hofstetter P, Pennington DW, de Haes HAU. Midpoints versus endpoints:
lca2004.11.181.3. the sacrifices and benefits. Int J Life Cycle Assess 2000;5:319–26. https://doi.org/
[98] Schleisner L. Life cycle assessment of a wind farm and related externalities. Renew 10.1007/BF02978665.
Energy 2000;20:279–88. https://doi.org/10.1016/S0960-1481(99)00123-8. [108] Download page for WindPower Program, UK wind speed database program and
[99] Martínez E, Sanz F, Pellegrini S, Jiménez E, Blanco J. Life cycle assessment of a turbine power curve database [n.d].
multi-megawatt wind turbine. Renew Energy 2009;34:667–73. https://doi.org/10. [109] Wind turbines e Part 1: design requirements (IEC 61400-1:2005 þ A1:2010) [n.d].
1016/j.renene.2008.05.020. [110] Justus CG, Hargraves WR, Yalcin A. Nationwide assessment of potential output
[100] Vestas. Life cycle assessment of electricity production from an onshore V80-2.0 from wind-powered generators. J Appl Meteorol 1976;15:673–8. https://doi.org/
MW wind plant. 2004. 10.1175/1520-0450(1976)015<0673:NAOPOF>2.0.CO;2.
[101] Vestas. V80-2MW Life Cycle Assessment of offshore and onshore sited wind farms. [111] Krasteva K. Renewable energy technologies. Sustain Innov Impact 2018:237–50.
2004. https://doi.org/10.4324/9781351174824-31.

480

You might also like