[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (460)

Search Parameters:
Keywords = LCOE

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 8614 KiB  
Article
Techno-Economic Analysis of Combined Production of Wind Energy and Green Hydrogen on the Northern Coast of Mauritania
by Varha Maaloum, El Moustapha Bououbeid, Mohamed Mahmoud Ali, Kaan Yetilmezsoy, Shafiqur Rehman, Christophe Ménézo, Abdel Kader Mahmoud, Shahab Makoui, Mamadou Lamine Samb and Ahmed Mohamed Yahya
Sustainability 2024, 16(18), 8063; https://doi.org/10.3390/su16188063 - 14 Sep 2024
Viewed by 644
Abstract
Green hydrogen is becoming increasingly popular, with academics, institutions, and governments concentrating on its development, efficiency improvement, and cost reduction. The objective of the Ministry of Petroleum, Mines, and Energy is to achieve a 35% proportion of renewable energy in the overall energy [...] Read more.
Green hydrogen is becoming increasingly popular, with academics, institutions, and governments concentrating on its development, efficiency improvement, and cost reduction. The objective of the Ministry of Petroleum, Mines, and Energy is to achieve a 35% proportion of renewable energy in the overall energy composition by the year 2030, followed by a 50% commitment by 2050. This goal will be achieved through the implementation of feed-in tariffs and the integration of independent power generators. The present study focused on the economic feasibility of green hydrogen and its production process utilizing renewable energy resources on the northern coast of Mauritania. The current investigation also explored the wind potential along the northern coast of Mauritania, spanning over 600 km between Nouakchott and Nouadhibou. Wind data from masts, Lidar stations, and satellites at 10 and 80 m heights from 2022 to 2023 were used to assess wind characteristics and evaluate five turbine types for local conditions. A comprehensive techno-economic analysis was carried out at five specific sites, encompassing the measures of levelized cost of electricity (LCOE) and levelized cost of green hydrogen (LCOGH), as well as sensitivity analysis and economic performance indicators. The results showed an annual average wind speed of 7.6 m/s in Nouakchott to 9.8 m/s in Nouadhibou at 80 m. The GOLDWIND 3.0 MW model showed the highest capacity factor of 50.81% due to its low cut-in speed of 2.5 m/s and its rated wind speed of 10.5 to 11 m/s. The NORDEX 4 MW model forecasted an annual production of 21.97 GWh in Nouadhibou and 19.23 GWh in Boulanoir, with the LCOE ranging from USD 5.69 to 6.51 cents/kWh, below the local electricity tariff, and an LCOGH of USD 1.85 to 2.11 US/kg H2. Multiple economic indicators confirmed the feasibility of wind energy and green hydrogen projects in assessed sites. These results boosted the confidence of the techno-economic model, highlighting the resilience of future investments in these sustainable energy infrastructures. Mauritania’s north coast has potential for wind energy, aiding green hydrogen production for energy goals. Full article
(This article belongs to the Special Issue Renewable Energy, Electric Power Systems and Sustainability)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the average annual wind speed pattern in Mauritania.</p>
Full article ">Figure 2
<p>Locations of meteorological measurement masts in Mauritania.</p>
Full article ">Figure 3
<p>Physical site photos of meteorological measurement masts in Mauritania. (<b>a</b>): Nouadhibou Measurement Mast; (<b>b</b>): Boulanoir Measurement Mast; (<b>c</b>): ZX300 Lidar with Solar Power Supply; (<b>d</b>): Nouakchott Measurement Mast with Its Equipment.</p>
Full article ">Figure 4
<p>Variation in wind speed on a monthly basis for the five different locations (Nouakchott, Nouamghar, Tasiast, Boulanoir, and Nouadhibou) from 2022 to 2023.</p>
Full article ">Figure 5
<p>Wind rose diagrams (N: north; NE: north–east; E: east; SE: south–east; S: south; SW: south–west; W: west; NW: north–west) for the five sites (Nouakchott, Nouamghar, Tasiast, Boulanoir, and Nouadhibou).</p>
Full article ">Figure 6
<p>Frequency distribution of wind along with the Weibull distribution curve for Nouakchott, Nouamghar, and Tasiast.</p>
Full article ">Figure 7
<p>Frequency distribution of wind along with the Weibull distribution curve for Boulanoir and Nouadhibou.</p>
Full article ">Figure 8
<p>Variations in wind speeds throughout different seasons and times of the day at 80 m for Nouakchott, Nouamghar, and Tasiast during the period 2022–2023.</p>
Full article ">Figure 9
<p>Variations in wind speeds throughout different seasons and times of the day at 80 m for Boulanoir and Nouadhibou during the period 2022–2023.</p>
Full article ">
23 pages, 8780 KiB  
Article
Techno-Economic Feasibility Analysis of an Offshore Wave Power Facility in the Aegean Sea, Greece
by Evangelos E. Pompodakis, Georgios I. Orfanoudakis, Yiannis Katsigiannis and Emmanouel Karapidakis
Energies 2024, 17(18), 4588; https://doi.org/10.3390/en17184588 - 12 Sep 2024
Viewed by 254
Abstract
The decarbonization goals of each country necessitate the utilization of renewable resources, with photovoltaic (PV) and wind turbine (WT) generators being the most common forms. However, spatial constraints, especially on islands, can hinder the expansion of PV and WT installations. In this context, [...] Read more.
The decarbonization goals of each country necessitate the utilization of renewable resources, with photovoltaic (PV) and wind turbine (WT) generators being the most common forms. However, spatial constraints, especially on islands, can hinder the expansion of PV and WT installations. In this context, wave energy emerges as a viable supplementary renewable source. Islands are candidate regions to accommodate wave power resources due to their abundant wave potential. While previous studies have explored the wave energy potential of the Aegean Sea, they have not focused on the electricity production and techno-economic aspects of wave power facilities in this area. This paper aims to fill this knowledge gap by conducting a comprehensive techno-economic analysis to evaluate the feasibility of deploying an offshore wave power facility in the Aegean Sea, Greece. The analysis includes a detailed sensitivity assessment of CAPEX and OPEX variability, calculating key indicators like LCOE and NPV to determine the economic viability and profitability of wave energy investments in the region. Additionally, the study identifies hydraulic efficiency and CAPEX thresholds that could make wave power more competitive compared with traditional energy sources. The techno-economic analysis is conducted for a 45 MW offshore floating wave power plant situated between eastern Crete and Kasos—one of the most wave-rich areas in Greece. Despite eastern Crete’s promising wave conditions, the study reveals that with current techno-economic parameters—CAPEX of 7 million EUR/MW, OPEX of 6%, a 20-year lifetime, and 25% efficiency—the wave energy in this area yields a levelized cost of energy (LCOE) of 1417 EUR/MWh. This rate is significantly higher than the prevailing LCOE in Crete, which is between 237 and 300 EUR/MWh. Nonetheless, this study suggests that the LCOE of wave energy in Crete could potentially decrease to as low as 69 EUR/MWh in the future under improved conditions, including a CAPEX of 1 million EUR/MW, an OPEX of 1%, a 30-year lifetime, and 35% hydraulic efficiency for wave converters. It is recommended that manufacturing companies target these specific thresholds to ensure the economic viability of wave power in the waters of the Aegean Sea. Full article
(This article belongs to the Section C: Energy Economics and Policy)
Show Figures

Figure 1

Figure 1
<p>Illustration of Wave Dragon converter. From top to bottom: (<b>a</b>) side view and (<b>b</b>) top view.</p>
Full article ">Figure 2
<p>Wave Dragon converter models for different sea conditions [<a href="#B35-energies-17-04588" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>Fifteen-year mean wave power (kW/m) of Aegean Sea [<a href="#B38-energies-17-04588" class="html-bibr">38</a>].</p>
Full article ">Figure 4
<p>Wind rose diagram for facility location (eastern Crete).</p>
Full article ">Figure 5
<p>Significant wave height <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">H</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the year 2023 between Crete and Kasos [<a href="#B39-energies-17-04588" class="html-bibr">39</a>].</p>
Full article ">Figure 6
<p>Wave energy period <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">T</mi> </mrow> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the year 2023 between Crete and Kasos [<a href="#B39-energies-17-04588" class="html-bibr">39</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Installation location of wave power facility on the eastern coast of Crete (image from Google Maps), and (<b>b</b>) layout and spacing details for Wave Dragon converters.</p>
Full article ">Figure 8
<p>Wave power flux <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <mi mathvariant="bold-italic">P</mi> <mo stretchy="false">(</mo> <mi mathvariant="bold-italic">t</mi> <mo stretchy="false">)</mo> </mrow> </mfenced> </mrow> </semantics></math> for the year 2023 between Crete and Kasos.</p>
Full article ">Figure 9
<p>Pie chart of the breakdown cost of the examined wave power facility.</p>
Full article ">Figure 10
<p>Estimated evolution of CAPEX (<b>top left</b>), OPEX (<b>top right</b>), and lifetime (<b>bottom</b>) of wave converters until 2050 [<a href="#B4-energies-17-04588" class="html-bibr">4</a>].</p>
Full article ">Figure 11
<p>Scenario 1: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 25%, <span class="html-italic">lifetime</span> = 20.</p>
Full article ">Figure 12
<p>Scenario 2: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 30%, <span class="html-italic">lifetime</span> = 20.</p>
Full article ">Figure 13
<p>Scenario 3: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 30%, <span class="html-italic">lifetime</span> = 25.</p>
Full article ">Figure 14
<p>Scenario 4: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 35%, <span class="html-italic">lifetime</span> = 25.</p>
Full article ">Figure 15
<p>Scenario 5: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 35%, <span class="html-italic">lifetime</span> = 30.</p>
Full article ">Figure 16
<p>NPV value of the investment for different CAPEX and electricity selling prices, assuming <span class="html-italic">lifetime</span> = 20 years, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 25%, and OPEX = 3%.</p>
Full article ">Figure 17
<p>NPV value of the investment for different CAPEX and electricity selling prices, assuming <span class="html-italic">lifetime</span> = 30 years, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">n</mi> </mrow> <mrow> <mi mathvariant="bold-italic">y</mi> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">r</mi> </mrow> </msub> </mrow> </semantics></math> = 35%, and OPEX = 3%.</p>
Full article ">Figure 18
<p>Other examined locations in the Aegean Sea (western Crete, Karpathos, Tinos, Skyros).</p>
Full article ">Figure 19
<p>Wind rose diagram for western Crete.</p>
Full article ">Figure 20
<p>Wind rose diagram for Tinos.</p>
Full article ">Figure 21
<p>Floating offshore power plant combining Wave Dragon and wind turbines (source from [<a href="#B59-energies-17-04588" class="html-bibr">59</a>]).</p>
Full article ">
24 pages, 3750 KiB  
Article
Optimal Hybrid Renewable Energy System to Accelerate a Sustainable Energy Transition in Johor, Malaysia
by Pei Juan Yew, Deepak Chaulagain, Noel Ngando Same, Jaebum Park, Jeong-Ok Lim and Jeung-Soo Huh
Sustainability 2024, 16(17), 7856; https://doi.org/10.3390/su16177856 - 9 Sep 2024
Viewed by 1006
Abstract
As the world’s second-largest palm oil producer, Malaysia heavily depends on its extensive oil palm cultivation, which accounts for nearly 90% of the country’s lignocellulosic biomass waste. Approximately 20–22 tonnes of empty fruit bunches (EFBs) can be derived from an initial yield of [...] Read more.
As the world’s second-largest palm oil producer, Malaysia heavily depends on its extensive oil palm cultivation, which accounts for nearly 90% of the country’s lignocellulosic biomass waste. Approximately 20–22 tonnes of empty fruit bunches (EFBs) can be derived from an initial yield of 100 tonnes of fresh fruit bunches (FFBs) from oil palm trees. The average annual amount of EFBs produced in Johor is 3233 tonnes per day. Recognising that urban areas contribute significantly to anthropogenic greenhouse gas emissions, and to support Malaysia’s transition from fossil fuel-based energy to a low-carbon energy system, this research employed HOMER Pro software 3.18.3 to develop an optimal hybrid renewable energy system integrating solar and biomass (EFB) energy sources in Johor, Malaysia. The most cost-effective system (solar–biomass) consists of 4075 kW solar photovoltaics, a 2100 kW biomass gasifier, 9363 battery units and 1939 kW converters. This configuration results in a total net present cost (NPC) of USD 44,596,990 and a levelised cost of energy (LCOE) of USD 0.2364/kWh. This system satisfies the residential load demand via 6,020,427 kWh (64.7%) of solar-based and 3,286,257 kWh (35.3%) of biomass-based electricity production, with an annual surplus of 2,613,329 kWh (28.1%). The minimal percentages of unmet electric load and capacity shortage, both <0.1%, indicate that all systems can meet the power demand. In conclusion, this research provides valuable insights into the economic viability and technical feasibility of powering the Kulai district with a solar–biomass system. Full article
Show Figures

Figure 1

Figure 1
<p>Overall analysis workflow.</p>
Full article ">Figure 2
<p>Location of the study area.</p>
Full article ">Figure 3
<p>Hourly load profile data for residential homes in Kulai on a daily basis.</p>
Full article ">Figure 4
<p>Monthly average Global Horizontal Irradiance (GHI) and temperature.</p>
Full article ">Figure 5
<p>Schematic diagram of the hybrid renewable energy system.</p>
Full article ">Figure 6
<p>Cost summary of the different configurations.</p>
Full article ">Figure 7
<p>Monthly electric production for all configurations: (<b>a</b>) system 1; (<b>b</b>) system 2; (<b>c</b>) system 3; (<b>d</b>) system 4; (<b>e</b>) system 5.</p>
Full article ">Figure 8
<p>NPC and COE at different biomass prices.</p>
Full article ">Figure 9
<p>NPC and COE at different inflation rates.</p>
Full article ">Figure 10
<p>NPC and COE at different discount rates.</p>
Full article ">Figure 11
<p>NPC and COE at different PV cost.</p>
Full article ">
11 pages, 2405 KiB  
Article
A Simple Procedure to Obtain the Grounding Resistance Measurement of Very Large and Urban Electrodes by a Modified Fall-of-Potential Method
by Jorge Moreno, Pascual Simón, Eduardo Faleiro, Daniel García, Gregorio Denche and Gabriel Asensio
Appl. Sci. 2024, 14(17), 8040; https://doi.org/10.3390/app14178040 - 8 Sep 2024
Viewed by 334
Abstract
The measurement of the grounding resistance of grounding grids in large installations as well as grounding electrodes in urban areas is addressed in this article. The resistance value is obtained using a three-pin array by measuring the fall-of-potential on the ground surface. The [...] Read more.
The measurement of the grounding resistance of grounding grids in large installations as well as grounding electrodes in urban areas is addressed in this article. The resistance value is obtained using a three-pin array by measuring the fall-of-potential on the ground surface. The resistance measured by this method is adjusted to its true value using a correction factor that aligns the measured resistance with the actual value. The proposed measurement method obtains correct values of the grounding resistance even when the auxiliary and potential electrodes of the tree-pin array are close to the electrode to be measured. Thus, it can be applied to large electrodes as well as electrodes in urban areas. Several simulated examples are used to illustrate the method, and some real cases with field measurements are presented for a final validation of the method. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of the three-pin method for measuring the grounding resistance <span class="html-italic">R<sub>G</sub></span>.</p>
Full article ">Figure 2
<p>The <b>upper panel</b> shows the electrode being studied, while the <b>lower panel</b> represents the same electrode partially damaged.</p>
Full article ">Figure 3
<p>The measured grounding resistance profile for the original electrode (red line) and for the damaged electrode (beige line).</p>
Full article ">Figure 4
<p>The Balaidos substation grounding grid.</p>
Full article ">Figure 5
<p>Balaidos grid with the indicative line on the ground where the absolute potential is measured by the three-pin method (blue line) when auxiliar rod is 300 m far away. The absolute potential created by the isolated grid is also shown (red line).</p>
Full article ">Figure 6
<p>Representation of the different measurement procedure parameters as a function of the distance from the edge of the Balaidos grid. An interconnected case with other grounding system is also considered. The dashed line represents the edge of the electrode.</p>
Full article ">Figure 7
<p>In the (<b>a</b>), the Loeches substation is marked in red. The blue dot is the auxiliary current electrode 211 m away from the grid edge. (Courtesy of UFD DISTRIB. ELECTRIC. S.A.) The grounding grid is shown in the (<b>b</b>).</p>
Full article ">Figure 8
<p>The different unitary resistances associated with the studied case, as discussed in the text.</p>
Full article ">
27 pages, 5251 KiB  
Article
Development and Analysis of a Global Floating Wind Levelised Cost of Energy Map
by Sergi Vilajuana Llorente, José Ignacio Rapha and José Luis Domínguez-García
Clean Technol. 2024, 6(3), 1142-1168; https://doi.org/10.3390/cleantechnol6030056 - 5 Sep 2024
Viewed by 410
Abstract
Floating offshore wind (FOW) is rapidly gaining interest due to its large potential. In this regard, it is of special interest to determine the best locations for its installation. One of the main aspects when evaluating the feasibility of a project is the [...] Read more.
Floating offshore wind (FOW) is rapidly gaining interest due to its large potential. In this regard, it is of special interest to determine the best locations for its installation. One of the main aspects when evaluating the feasibility of a project is the levelised cost of energy (LCOE), but there are many variables to consider when calculating it for FOW, and plenty of them are hard to find when the scope is all the suitable areas worldwide. This paper presents the calculation and analysis of the global LCOE with particular focus on the best countries and territories from an economic point of view, considering four types of platforms: semi-submersible, barge, spar, and tension leg platform (TLP). The model takes into account, on the one hand, wind data, average significant wave height, and distance to shore for an accurate calculation of delivered energy to the onshore substation and, on the other hand, bathymetry, distances, and existing data from projects to find appropriate functions for each cost with regression models (e.g., manufacturing, installation, operation and maintenance (O&M), and decommissioning costs). Its results can be used to assess the potential areas around the world and identify the countries and territories with the greatest opportunities regarding FOW. The lowest LCOE values, i.e., the optimal results, correspond to areas where wind resources are more abundant and the main variables of the site affecting the costs (water depth, average significant wave height, distance to shore, and distance to port) are as low as possible. These areas include the border between Venezuela and Colombia, the Canary Islands, Peru, the border between Western Sahara and Mauritania, Egypt, and the southernmost part of Argentina, with LCOEs around 90 €/MWh. Moreover, there are many areas in the range of 100–130 €/MWh. Full article
Show Figures

Figure 1

Figure 1
<p>Regression model summary of (<b>a</b>) mooring system costs, (<b>b</b>) substructure costs, (<b>c</b>) turbine costs, (<b>d</b>) electrical system costs, (<b>e</b>) installation costs, and (<b>f</b>) OPEX.</p>
Full article ">Figure 2
<p>Global floating wind LCOE map.</p>
Full article ">Figure 3
<p>Amount of area with values from 80 €/MWh to 250 €/MWh, represented in intervals of 10.</p>
Full article ">Figure 4
<p>Assessment of the main variables affecting the LCOE mean with the one-factor-at-a-time method.</p>
Full article ">Figure 5
<p>Assessment of the life cycle costs affecting the LCOE mean with the one-factor-at-a-time method.</p>
Full article ">Figure 6
<p>Top 10 territory scores and positions per factor.</p>
Full article ">Figure 7
<p>Argentina LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 8
<p>Peru LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 9
<p>Colombia LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 10
<p>Venezuela LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 11
<p>Canada east coast LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 12
<p>Canada west coast LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 13
<p>New Zealand LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 14
<p>Egypt LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 15
<p>Vietnam LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 16
<p>United Kingdom LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 17
<p>Chile south coast LCOE (&lt;130 €/MWh) map.</p>
Full article ">Figure 18
<p>Chile middle coast LCOE (&lt;130 €/MWh) map.</p>
Full article ">
14 pages, 4642 KiB  
Proceeding Paper
Impact of Climate Change on the Thermoeconomic Performance of Binary-Cycle Geothermal Power Plants
by Paolo Blecich, Igor Wolf, Tomislav Senčić and Igor Bonefačić
Eng. Proc. 2024, 67(1), 29; https://doi.org/10.3390/engproc2024067029 - 3 Sep 2024
Viewed by 105
Abstract
The thermoeconomic performance of geothermal power plants is influenced by a variety of site-specific factors, major economic variables, and the type of the involved technology. In addition to those, ambient conditions also play a role in geothermal power generation by acting on the [...] Read more.
The thermoeconomic performance of geothermal power plants is influenced by a variety of site-specific factors, major economic variables, and the type of the involved technology. In addition to those, ambient conditions also play a role in geothermal power generation by acting on the cooling towers. This study focuses on the performance analysis of a binary cycle with isobutane for geothermal power generation under the impact of climate change. Long-term temperature variations in ambient air are described by temperature anomalies under two shared socioeconomic pathways (SSP). These are the intermediate SSP2-4.5 scenario and the extreme SSP5-8.5 scenario, over the period from 2021 to 2100. Different climate models from the most recent Climate Model Intercomparison Project (CMIP6) are compared against each other and against the observed temperature data. The predictive power of the CMIP6 climate models is evaluated using the root mean square error (RMSE) and the Kullback–Leibler (KL) criteria. The thermoeconomic performance of the geothermal power plant is expressed in terms of net power output, annual electricity generation (AEG), and levelized cost of electricity (LCOE). The geothermal power plant achieves a net power output of 10 MW and an LCOE of 79.2 USD/MWh for an ambient air temperature of 12 °C. This temperature is the average temperature over the reference period of 1991–2020 in Bjelovar, Croatia (45.8988° N, 16.8423° E). Under the impact of climate change, the same geothermal power plant will have the AEG reduced by between 0.5% and 2.9% in the intermediate (SSP2-4.5) scenario and by between 2.0% and 8.7% in the extreme (SSP5-8.5) scenario. The LCOE will increase between 0.4% and 1.8% in the intermediate scenario and from 1.3% to 5.6% in the extreme scenario. Full article
(This article belongs to the Proceedings of The 3rd International Electronic Conference on Processes)
Show Figures

Figure 1

Figure 1
<p>The single-stage configuration of the ORC geothermal power plant.</p>
Full article ">Figure 2
<p>The air-cooled condenser (ACC).</p>
Full article ">Figure 3
<p>Daily average near-surface air temperature in Bjelovar, Croatia (45.8988° N, 16.8423° E) for the reference period of 1991–2020: A comparison between observations and predictions of CMIP6 climate models.</p>
Full article ">Figure 4
<p>Daily average near-surface air temperature in Bjelovar, Croatia (45.8988° N, 16.8423° E) for the reference period of 1991–2020: A comparison between observations and CMIP6 multi-model predictions.</p>
Full article ">Figure 5
<p>Average annual near-surface air temperature in Bjelovar (45.8988° N, 16.8423° E) from 2021 to 2100, under the extreme climate-change scenario (SSP5-8.5).</p>
Full article ">Figure 6
<p><span class="html-italic">T</span>-s chart of the single-stage ORC configuration with isobutane as working fluid. For the meaning of line colors and point states refer to <a href="#engproc-67-00029-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 7
<p>Impact of climate change on the AEG of the single-stage ORC geothermal power plant.</p>
Full article ">Figure 8
<p>Impact of climate change on the LCOE of the single-stage ORC geothermal power plant.</p>
Full article ">
20 pages, 6395 KiB  
Article
A Dispatch Strategy for the Analysis of the Technical, Economic, and Environmental Performance of a Hybrid Renewable Energy System
by Mehmet Ali Köprü, Dursun Öztürk and Burak Yıldırım
Sustainability 2024, 16(17), 7490; https://doi.org/10.3390/su16177490 - 29 Aug 2024
Viewed by 467
Abstract
The use of renewable energy sources (RESs) is increasing every day to meet increasing energy demands and reduce dependence on fossil fuels. When designing hybrid renewable energy systems (HRESs), it is necessary to examine their technical, economic, and environmental feasibility. In this study, [...] Read more.
The use of renewable energy sources (RESs) is increasing every day to meet increasing energy demands and reduce dependence on fossil fuels. When designing hybrid renewable energy systems (HRESs), it is necessary to examine their technical, economic, and environmental feasibility. In this study, a new strategy is proposed using the HOMER Matlab Link (ML) connection for an HRES model consisting of a photovoltaic (PV) system, a wind turbine (WT), a biogas generator (BGG), and a battery storage system (BSS) designed to meet the electrical energy needs of Doğanevler village located in the rural area of Bingöl province. The data obtained as a result of the proposed strategy (PS) are compared with HOMER’s loop charging (CC) and load following (LF) optimization results. According to the PS, the optimum capacity values for the HRES components are 10 kW for WT, 10 kW for PV, 8 kW for BGG, 12 kWh for BSS, and 12 kW for the converter. According to the optimum design, 16,205 kWh of the annual energy produced was generated by PV systems, 22,927 kWh by WTs, and 22,817 kWh by BGGs. This strategy’s NPC and LCOE (Levelized Cost of Energy) values are calculated as USD 130,673.91 and USD 0.207/kWh, respectively. For the CC dispatch strategy, the NPC and LCOE values are calculated as USD 141,892.28 and USD 0.240/kWh, while for the LF dispatch strategy, these values are USD 152,456.89 and USD 0.257/kWh. The CO2 emission value for the system using a BGG was calculated as 480 kg/year, while for the system using a DG, this value increased approximately 57 times and was calculated to be 27,709 kg/year. The results show that the PS is more economical than the other two strategies. The PS provides energy security, reduces costs, and increases environmental sustainability. Finally, a sensitivity analysis was conducted based on the availability of renewable resources, fuel cost, and inflation parameters, and the results were analyzed. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of hybrid system.</p>
Full article ">Figure 2
<p>Average monthly electricity load profile.</p>
Full article ">Figure 3
<p>Hourly load profile.</p>
Full article ">Figure 4
<p>Monthly average solar radiation and clarity index.</p>
Full article ">Figure 5
<p>Average monthly wind speed.</p>
Full article ">Figure 6
<p>Biomass technical potential.</p>
Full article ">Figure 7
<p>Proposed dispatch strategy.</p>
Full article ">Figure 8
<p>Generation amount of HRES sources: (<b>a</b>) LF; (<b>b</b>) CC; (<b>c</b>) PS.</p>
Full article ">Figure 9
<p>BSS input–output energies for dispatch strategies.</p>
Full article ">Figure 10
<p>BGG curves for dispatch strategies: (<b>a</b>) LF; (<b>b</b>) CC; (<b>c</b>) PS.</p>
Full article ">Figure 11
<p>The annual time series of electricity generation and consumption is used for the LF dispatch strategy.</p>
Full article ">Figure 12
<p>Annual time series of electricity generation and consumption for the CC dispatch strategy.</p>
Full article ">Figure 13
<p>Annual time series of electricity generation and consumption for the PS dispatch strategy.</p>
Full article ">Figure 14
<p>Impact of different inflation rates on NPC.</p>
Full article ">
44 pages, 25979 KiB  
Article
Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy
by Zarif Aminov, Khusniddin Alikulov and Tran-Dang Xuan
Energies 2024, 17(17), 4264; https://doi.org/10.3390/en17174264 - 26 Aug 2024
Viewed by 355
Abstract
This study introduces a novel hybrid solar–biomass cogeneration power plant that efficiently produces heat, electricity, carbon dioxide, and hydrogen using concentrated solar power and syngas from cotton stalk biomass. Detailed exergy-based thermodynamic, economic, and environmental analyses demonstrate that the optimized system achieves an [...] Read more.
This study introduces a novel hybrid solar–biomass cogeneration power plant that efficiently produces heat, electricity, carbon dioxide, and hydrogen using concentrated solar power and syngas from cotton stalk biomass. Detailed exergy-based thermodynamic, economic, and environmental analyses demonstrate that the optimized system achieves an exergy efficiency of 48.67% and an exergoeconomic factor of 80.65% and produces 51.5 MW of electricity, 23.3 MW of heat, and 8334.4 kg/h of hydrogen from 87,156.4 kg/h of biomass. The study explores four scenarios for green hydrogen production pathways, including chemical looping reforming and supercritical water gasification, highlighting significant improvements in levelized costs and the environmental impact compared with other solar-based hybrid systems. Systems 2 and 3 exhibit superior performance, with levelized costs of electricity (LCOE) of 49.2 USD/MWh and 55.4 USD/MWh and levelized costs of hydrogen (LCOH) of between 10.7 and 19.5 USD/MWh. The exergoenvironmental impact factor ranges from 66.2% to 73.9%, with an environmental impact rate of 5.4–7.1 Pts/MWh. Despite high irreversibility challenges, the integration of solar energy significantly enhances the system’s exergoeconomic and exergoenvironmental performance, making it a promising alternative as fossil fuel reserves decline. To improve competitiveness, addressing process efficiency and cost reduction in solar concentrators and receivers is crucial. Full article
(This article belongs to the Section A5: Hydrogen Energy)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the hybrid biomass–solar driven CHP with steam- and oxygen-based biomass gasification.</p>
Full article ">Figure 2
<p>Schematic diagram of the hybrid biomass–solar driven CHP with sorption-enhanced reforming.</p>
Full article ">Figure 3
<p>Schematic of the hybrid biomass–solar driven CHP with iron chemical looping reforming.</p>
Full article ">Figure 4
<p>Schematic diagram of the hybrid biomass–solar driven CHP with supercritical water gasification.</p>
Full article ">Figure 5
<p>Hourly variation in the DNI for Tashkent and the efficiency of the CSP plant.</p>
Full article ">Figure 6
<p>Average monthly exergy efficiency of the CSP plants.</p>
Full article ">Figure 7
<p>Total purchased equipment cost of the system.</p>
Full article ">Figure 8
<p>Specific CAPEX (<b>a</b>) and OPEX (<b>b</b>) costs for the examined renewable hydrogen and power production.</p>
Full article ">Figure 9
<p>Exergy efficiency, unit cost of hydrogen, and cost rate of the system.</p>
Full article ">Figure 10
<p>Comparison of CO<sub>2</sub> emissions (<b>a</b>) and the ecological footprint (<b>b</b>) of different hydrogen and power production systems.</p>
Full article ">Figure 11
<p>Exergoenvironmental parameters of the proposed models: impact factor, damage effectiveness index, impact improvement, and sustainability index (<b>a</b>); ecological effect factor and social ecological factor (<b>b</b>).</p>
Full article ">Figure 12
<p>The effects of annual variation in the CSP system on the following environmental indicators: ecological effect factor (<b>a</b>); social ecological factor (<b>b</b>).</p>
Full article ">Figure 13
<p>The effects of variations in operation time on the following costs: exergetic performance and exergy destruction (<b>a</b>); specific CAPEX and investment cost rate (<b>b</b>).</p>
Full article ">Figure 14
<p>The effects of operation time on the LCOE, LCOH, and LCOP of the proposed systems.</p>
Full article ">Figure 15
<p>The effects of operation time on the specific exergy hydrogen cost, relative difference, and exergoeconomic factor.</p>
Full article ">Figure 16
<p>The effects of operation time on the environmental indexes.</p>
Full article ">Figure 17
<p>The effects of different modes on the environmental impacts of the system.</p>
Full article ">Figure A1
<p>Framework of the LCA [<a href="#B22-energies-17-04264" class="html-bibr">22</a>].</p>
Full article ">Figure A2
<p>Boundaries of the hydrogen and power production process.</p>
Full article ">
23 pages, 9244 KiB  
Article
Design and Techno-Economic Analysis of Hybrid Power Systems for Rural Areas: A Case Study of Bingöl
by Ferhat Aydın and Dursun Öztürk
Electricity 2024, 5(3), 562-584; https://doi.org/10.3390/electricity5030028 - 23 Aug 2024
Viewed by 470
Abstract
Today, many factors, especially the increasing world population and developing technology, increase the energy needs of people and societies day by day. The fact that fossil resources are both in danger of depletion and have negative environmental impacts has directed countries to new [...] Read more.
Today, many factors, especially the increasing world population and developing technology, increase the energy needs of people and societies day by day. The fact that fossil resources are both in danger of depletion and have negative environmental impacts has directed countries to new resources. The study focuses on the effective use of renewable energy sources (RES) and the evaluation of waste to meet the electricity and heat energy needs of Yiğit Harman Village, located in the Solhan district of Bingöl Province. For this purpose, a renewable-energy-based combined heat and power system (CHP) was designed using HOMER Pro software (version 3.14.2, Homer Energy LLC, Boulder, CO, USA). Solar, wind, biomass, and hydrogen energy sources were used, considering the resources of the region. Using the designed model, the entire electricity energy requirement and half of the heat energy were completely met by the region’s available RESs. In addition to the technical analysis, economic and environmental analyses were also conducted, and LCOE, NPC, and CO2 emission values were obtained as 0.271 USD/kWh, USD 739,772, and 37,958 kg/yr, respectively. These results indicate that with an investment of approximately USD 7000 per household, the electrical and thermal energy needs for 25 years can be met. Full article
Show Figures

Figure 1

Figure 1
<p>The location of the case study region (Yiğit Harman Village).</p>
Full article ">Figure 2
<p>Configuration of the proposed microgrid.</p>
Full article ">Figure 3
<p>Electrical load data for Load-1.</p>
Full article ">Figure 4
<p>Electrical load data for Load-2.</p>
Full article ">Figure 5
<p>Thermal load data for Yiğit Harman Village.</p>
Full article ">Figure 6
<p>Daily solar radiation and openness index for Yiğit Harman Village.</p>
Full article ">Figure 7
<p>Wind speed graph for Yiğit Harman Village.</p>
Full article ">Figure 8
<p>Biomass graph of Yiğit Harman Village.</p>
Full article ">Figure 9
<p>Power curve of the Eocycle EO10 model WT.</p>
Full article ">Figure 10
<p>Optimization results.</p>
Full article ">Figure 11
<p>Annual generation distribution by sources for the model. (<b>a</b>) Electrical energy. (<b>b</b>) Thermal energy.</p>
Full article ">Figure 12
<p>(<b>a</b>) Generation with diesel generator. (<b>b</b>) Generation with natural gas generator.</p>
Full article ">Figure 13
<p>Annual output power variation of PV panels.</p>
Full article ">Figure 14
<p>Annual output power variation of wind turbines.</p>
Full article ">Figure 15
<p>Annual output power variation of the BG.</p>
Full article ">Figure 16
<p>Annual output power variation of the FC.</p>
Full article ">Figure 17
<p>Annual input power variation of the electrolyzer.</p>
Full article ">Figure 18
<p>Annual charge level change of BSSs.</p>
Full article ">
20 pages, 6721 KiB  
Article
Conceptual Design and Dynamic Analysis of a Wind–Wave Energy Converter with a Mass-Adjustable Buoy
by Yifeng Shi, Jiahuan Lin, Zexin Zhuge, Rongye Zheng and Jun Zhang
J. Mar. Sci. Eng. 2024, 12(8), 1460; https://doi.org/10.3390/jmse12081460 - 22 Aug 2024
Viewed by 403
Abstract
To reduce the levelized cost of energy (LCOE) for offshore wind turbines, a novel wind–wave energy converter (WWEC) with a mass-adjustable buoy is designed. To analyze the impact of buoy mass variations on the system, a coupled comprehensive numerical model is established to [...] Read more.
To reduce the levelized cost of energy (LCOE) for offshore wind turbines, a novel wind–wave energy converter (WWEC) with a mass-adjustable buoy is designed. To analyze the impact of buoy mass variations on the system, a coupled comprehensive numerical model is established to simulate the aerodynamics of the turbine and the hydrodynamics of the platform and buoy. It is found that the occurrence of the buoy out of water significantly reduces the output power. Adjusting the buoy’s mass with suitable strategy can prevent the impact of slamming loads and improve the power output. The mass adjustment strategy is determined based on the output power of the wave energy converter under regular wave conditions. It is found that the mass adjustment strategy can significantly enhance the output power of combined system. The buoy does not move out of the water under the extreme conditions, which avoids the impact of slamming loads on system stability. Moreover, mass-adjustable buoys can reduce the risk of mooring line failure compare to a wind turbine without a buoy. Full article
(This article belongs to the Topic Marine Renewable Energy, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Wave energy converter (<b>a</b>) simplified diagram of the connection section; (<b>b</b>) structure of mass adjustment.</p>
Full article ">Figure 2
<p>The schematic diagram of the WWEC.</p>
Full article ">Figure 3
<p>Conceptual drawings of the platform and buoy.</p>
Full article ">Figure 4
<p>Flow chart of F2A.</p>
Full article ">Figure 5
<p>Radiation damping of the mass-adjustable buoy (<b>a</b>) with hydrodynamic interaction; (<b>b</b>) without hydrodynamic interaction.</p>
Full article ">Figure 6
<p>Added mass of the mass-adjustable buoy (<b>a</b>) with hydrodynamic interaction; (<b>b</b>) without hydrodynamic interaction.</p>
Full article ">Figure 7
<p>Comparison of rotor speed and rotor thrust (<b>a</b>) rotor thrust; (<b>b</b>) rotor speed.</p>
Full article ">Figure 8
<p>Comparison of mean power between numerical results and experimental data (H = 2 m).</p>
Full article ">Figure 9
<p>Power output of buoys with different masses for different PTO damping coefficients (<b>a</b>) draft = 2 m; (<b>b</b>) draft = 3 m; (<b>c</b>) draft = 4 m; (<b>d</b>) mean power.</p>
Full article ">Figure 10
<p>Comparison of F−K force for buoys between 1 m draft and 2 m draft. (<b>a</b>) Amplitude = 1 m; (<b>b</b>) amplitude = 2 m; (<b>c</b>) amplitude = 3 m; (<b>d</b>) amplitude = 4 m.</p>
Full article ">Figure 11
<p>Comparison of power output between buoys with different draft in regular wave (<b>a</b>) amplitude = 1 m, (<b>b</b>) amplitude = 2 m, (<b>c</b>) amplitude = 3 m, (<b>d</b>) amplitude = 4 m.</p>
Full article ">Figure 12
<p>Comparison of mean power between buoys in different amplitudes.</p>
Full article ">Figure 13
<p>Righting moment and wind heeling moment curves (<b>a</b>) intact stability, (<b>b</b>) damaged stability.</p>
Full article ">Figure 14
<p>The power output and time of water exit for different mass buoys in ECs1–8.</p>
Full article ">Figure 15
<p>The impact of buoys on the dynamic response of the wind turbine in EC3 (<b>a</b>) pitch, (<b>b</b>) surge.</p>
Full article ">Figure 16
<p>The output power of the wind turbine in EC3.</p>
Full article ">Figure 17
<p>The mean power and time of water exit for different mass buoys in ECs1–8.</p>
Full article ">Figure 18
<p>The power spectral density (PSD) of the loads on the mooring system; (<b>a</b>) mooring 1 (<b>b</b>) mooring 2.</p>
Full article ">
18 pages, 10341 KiB  
Article
Synergistic Integration of Multiple Wave Energy Converters with Adaptive Resonance and Offshore Floating Wind Turbines through Bayesian Optimization
by Aghamarshana Meduri and HeonYong Kang
J. Mar. Sci. Eng. 2024, 12(8), 1455; https://doi.org/10.3390/jmse12081455 - 22 Aug 2024
Viewed by 493
Abstract
We developed a synergistic ocean renewable system where an array of Wave Energy Converters (WEC) with adaptive resonance was collocated with a Floating Offshore Wind Turbine (FOWT) such that the WECs, capturing wave energy through the resonance adapting to varying irregular waves, consequently [...] Read more.
We developed a synergistic ocean renewable system where an array of Wave Energy Converters (WEC) with adaptive resonance was collocated with a Floating Offshore Wind Turbine (FOWT) such that the WECs, capturing wave energy through the resonance adapting to varying irregular waves, consequently reduced FOWFT loads and turbine motions. Combining Surface-Riding WECs (SR-WEC) individually designed to feasibly relocate their natural frequency at the peak of the wave excitation spectrum for each sea state, and to obtain the highest capture width ratio at one of the frequent sea states for annual average power in a tens of kilowatts scale with a 15 MW FOWT based on a semi-submersible, Bayesian Optimization is implemented to determine the arrangement of WECs that minimize the annual representation of FOWT’s wave excitation spectra. The time-domain simulation of the system in the optimized arrangement is performed, including two sets of interactions: one set is the wind turbine dynamics, mooring lines, and floating body dynamics for FOWT, and the other set is the nonlinear power-take-off dynamics, linear mooring, and individual WECs’ floating body dynamics. Those two sets of interactions are further coupled through the hydrodynamics of diffraction and radiation. For sea states comprising Annual Energy Production, we investigate the capture width ratio of WECs, wave excitation on FOWT, and nacelle acceleration of the turbine compared to their single unit operations. We find that the optimally arranged SR-WECs reduce the wave excitation spectral area of FOWT by up to 60% and lower the turbine’s peak nacelle acceleration by nearly 44% in highly occurring sea states, while multiple WECs often produce more than the single operation, achieving adaptive resonance with a larger wave excitation spectra for those sea states. The synergistic system improves the total Annual Energy Production (AEP) by 1440 MWh, and we address which costs of Levelized Cost Of Energy (LCOE) can be reduced by the collocation. Full article
(This article belongs to the Special Issue The Control, Modeling, and the Development of Wave Energy Convertors)
Show Figures

Figure 1

Figure 1
<p>Main components of (<b>a</b>) intermediate scale SR-WEC and (<b>b</b>) IEA 15-MW FOWT [<a href="#B18-jmse-12-01455" class="html-bibr">18</a>] (drawings not to scale).</p>
Full article ">Figure 2
<p>Wave resource characteristic bin for selected sea states at (<b>a</b>) Test Site A (located near NDBC Buoy 44056, 36.11 N, 75.44 W) and (<b>b</b>) Test Site B (located near NDBC Buoy 41002, 31.88 N, 74.92 W).</p>
Full article ">Figure 3
<p>Flowchart of the synergistic integration of the FOWT with the SR-WEC array.</p>
Full article ">Figure 4
<p>(<b>a</b>) Plan view of array with FOWT, SR-WECs (labeled 1–10) (not to scale) (<b>b</b>) FOWT Mooring Line Parameters [<a href="#B16-jmse-12-01455" class="html-bibr">16</a>].</p>
Full article ">Figure 5
<p>Selected mesh of (<b>a</b>) FOWT (<b>b</b>) SR-WEC used in Bayesian optimization. Comparison of wave excitation loads in surge, heave, and pitch for (<b>c</b>) FOWT (<b>d</b>) SR-WEC.</p>
Full article ">Figure 6
<p>Objective function <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mn>1</mn> </mrow> <mrow> <mn>3</mn> </mrow> </mfrac> </mstyle> <mrow> <msub> <mo stretchy="false">∑</mo> <mrow> <mi>j</mi> <mo>=</mo> <mo>[</mo> <mn>1,3</mn> <mo>,</mo> <mn>5</mn> <mo>]</mo> </mrow> </msub> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>Λ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>j</mi> <mo>,</mo> <mn>1</mn> </mrow> </msub> </mrow> </mrow> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">x</mi> </mrow> <mrow> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">i</mi> <mi mathvariant="bold-italic">s</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">y</mi> </mrow> <mrow> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">i</mi> <mi mathvariant="bold-italic">s</mi> </mrow> </msub> </mrow> </semantics></math> for (<b>a</b>) Site A, (<b>b</b>) Site B, and (<b>c</b>) optimization parameters.</p>
Full article ">Figure 7
<p>Density of input parameters and minimum objective function evaluated in Bayesian optimization for (<b>a</b>) Site A and (<b>b</b>) Site B.</p>
Full article ">Figure 8
<p>Excitation loads in (<b>a</b>) surge, (<b>b</b>) heave, and (<b>c</b>) pitch. Excitation spectra in (<b>d</b>) surge, (<b>e</b>) heave, and (<b>f</b>) pitch at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 6.38 s, 8.7 s of collocated FOWT compared with stand-alone FOWT.</p>
Full article ">Figure 9
<p>Excitation loads in (<b>a</b>) surge, (<b>b</b>) heave, and (<b>c</b>) pitch of collocated FOWT compared between WAMIT and Capytaine.</p>
Full article ">Figure 10
<p>Optimal (<b>a</b>) PTO stiffness coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>K</mi> </mrow> <mrow> <mi>P</mi> <mi>T</mi> <mi>O</mi> </mrow> </msub> </mrow> </semantics></math> and (<b>b</b>) PTO damping coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>P</mi> <mi>T</mi> <mi>O</mi> </mrow> </msub> </mrow> </semantics></math> for SR-WEC at individual sea states.</p>
Full article ">Figure 11
<p>Total hydrodynamic load in surge, heave, and pitch for collocated FOWT compared to stand-alone FOWT at (<b>a</b>–<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>d</b>–<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>g</b>–<b>i</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s.</p>
Full article ">Figure 12
<p>FOWT platform motions in surge, heave, and pitch for collocated FOWT compared to stand-alone FOWT at (<b>a</b>–<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>d</b>–<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>g</b>–<b>i</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s.</p>
Full article ">Figure 13
<p>Nacelle acceleration at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m (<b>a</b>) maximum (peak) RMS values as a function of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>) time history at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s.</p>
Full article ">Figure 14
<p>Spatial distribution of (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>P</mi> </mrow> </msub> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>P</mi> </mrow> </msub> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s, and (<b>c</b>) FOWT excitation spectra, and power of SR-WECs as a % of the stand-alone with different <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> 1.25 m.</p>
Full article ">Figure 15
<p>Performance of SR-WEC 2 in floating body pitch, translator sliding motion, and active power compared to stand-alone SR-WEC at (<b>a</b>–<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>d</b>–<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>g</b>–<b>i</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s.</p>
Full article ">Figure 16
<p>Power spectral density of SR-WEC 2 pitch motion at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s. Power spectral density of SR-WEC 2 translator sliding motion at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s. SR-WEC 2 pitch wave excitation spectra at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> = 1.25 m (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.06 s, (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.22 s, and (<b>i</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 8.7 s.</p>
Full article ">
23 pages, 5970 KiB  
Article
Optimizing Virtual Power Plant Management: A Novel MILP Algorithm to Minimize Levelized Cost of Energy, Technical Losses, and Greenhouse Gas Emissions
by Alain Aoun, Mehdi Adda, Adrian Ilinca, Mazen Ghandour and Hussein Ibrahim
Energies 2024, 17(16), 4075; https://doi.org/10.3390/en17164075 - 16 Aug 2024
Viewed by 364
Abstract
The modern energy landscape is undergoing a significant transformation towards cleaner, decentralized energy sources. This change is driven by environmental and sustainability needs, causing traditional centralized electric grids, which rely heavily on fossil fuels, to be replaced by a diverse range of decentralized [...] Read more.
The modern energy landscape is undergoing a significant transformation towards cleaner, decentralized energy sources. This change is driven by environmental and sustainability needs, causing traditional centralized electric grids, which rely heavily on fossil fuels, to be replaced by a diverse range of decentralized distributed energy resources. Virtual power plants (VPPs) have surfaced as a flexible solution in this transition. A VPP’s primary role is to optimize energy production, storage, and distribution by coordinating output from various connected sources. Relying on advanced communication and control systems, a VPP can balance supply and demand in real time, offer ancillary services, and support grid stability. However, aligning VPPs’ economic and operational practices with broader environmental goals and policies is a challenging yet crucial aspect. This article introduces a new VPP management and optimization algorithm designed for quick and intelligent decision-making, aiming for the lowest levelized cost of energy (LCOE), minimum grid technical losses, and greenhouse gas (GHG) emissions. The algorithm’s effectiveness is confirmed using the IEEE 33-bus grid with 10 different distributed power generators. Simulation results show the algorithm’s responsiveness to complex variables found in practical scenarios, finding the optimal combination of available energy resources. This minimizes the LCOE, technical losses, and GHG emissions in less than 0.08 s, achieving a total LCOE reduction of 16% from the baseline. This work contributes to the development of intelligent energy management systems, aiding the transition towards a more resilient and sustainable energy infrastructure. Full article
(This article belongs to the Section F2: Distributed Energy System)
Show Figures

Figure 1

Figure 1
<p>VPP MILP-based optimization algorithm flow chart.</p>
Full article ">Figure 2
<p>Scenario #1 with a total load of 3926 kW and no spinning reserve.</p>
Full article ">Figure 3
<p>Scenario #2 with a total load of 1963 kW and no spinning reserve.</p>
Full article ">Figure 4
<p>Scenario #3 with a total load of 3926 kW and no spinning reserve, with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>L</mi> <mi>C</mi> <mi>O</mi> <mi>E</mi> </mrow> <mrow> <mi>g</mi> <mi>r</mi> <mi>i</mi> <mi>d</mi> <mo> </mo> </mrow> </msub> </mrow> </semantics></math> = 0.24 <span>$</span>/kWh.</p>
Full article ">Figure 5
<p>Scenario #4 with a total load of 3926 kW and 25% spinning reserve.</p>
Full article ">Figure 6
<p>Sensitivity analysis taking into consideration the variation in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>L</mi> <mi>C</mi> <mi>O</mi> <mi>E</mi> </mrow> <mrow> <mi>P</mi> <mi>V</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Sensitivity analysis taking into consideration the variation in the carbon price (<math display="inline"><semantics> <mrow> <mo>∂</mo> </mrow> </semantics></math> in <span>$</span>/kg of CO<sub>2</sub>e).</p>
Full article ">Figure A1
<p>IEEE 33 bus VPP Grid Model.</p>
Full article ">
27 pages, 56161 KiB  
Article
Locating Insulation Defects in HV Substations Using HFCT Sensors and AI Diagnostic Tools
by Javier Ortego, Fernando Garnacho, Fernando Álvarez, Eduardo Arcones and Abderrahim Khamlichi
Sensors 2024, 24(16), 5312; https://doi.org/10.3390/s24165312 - 16 Aug 2024
Viewed by 506
Abstract
In general, a high voltage (HV) substation can be made up of multiple insulation subsystems: an air insulation subsystem (AIS), gas insulation subsystem (GIS), liquid insulation subsystem (power transformers), and solid insulation subsystem (power cables), all of them with their grounding structures interconnected [...] Read more.
In general, a high voltage (HV) substation can be made up of multiple insulation subsystems: an air insulation subsystem (AIS), gas insulation subsystem (GIS), liquid insulation subsystem (power transformers), and solid insulation subsystem (power cables), all of them with their grounding structures interconnected and linked to the substation earth. Partial discharge (PD) pulses, which are generated in a HV apparatus belonging to a subsystem, travel through the grounding structures of the others. PD analyzers using high-frequency current transformer (HFCT) sensors, which are installed at the connections between the grounding structures, are sensitive to these traveling pulses. In a substation made up of an AIS, several non-critical PD sources can be detected, such as possible corona, air surface, or floating discharges. To perform the correct diagnosis, non-critical PD sources must be separated from critical PD sources related to insulation defects, such as a cavity in a solid dielectric material, mobile particles in SF6, or surface discharges in oil. Powerful diagnostic tools using PD clustering and phase-resolved PD (PRPD) pattern recognition have been developed to check the insulation condition of HV substations. However, a common issue is how to determine the subsystem in which a critical PD source is located when there are several PD sources, and a critical one is near the boundary between two HV subsystems, e.g., a cavity defect located between a cable end and a GIS. The traveling direction of the detected PD is valuable information to determine the subsystem in which the insulation defect is located. However, incorrect diagnostics are usually due to the constraints of PD measuring systems and inadequate PD diagnostic procedures. This paper presents a diagnostic procedure using an appropriate PD analyzer with multiple HFCT sensors to carry out efficient insulation condition diagnoses. This PD procedure has been developed on the basis of laboratory tests, transient signal modeling, and validation tests. The validation tests were carried out in a special test bench developed for the characterization of PD analyzers. To demonstrate the effectiveness of the procedure, a real case is also presented, where satisfactory results are shown. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic three-phase circuit of a MV/LV substation: a. general view of the three MV cabinets: MV cabinet of the input line L1, MV cabinet of the output line L2, and MV cabinet of the PTP, b. HFCT sensors hugging the braids of each end of the cables, and c. HFCT sensors hugging each input and output power cable.</p>
Full article ">Figure 2
<p>(<b>a</b>) Three MV cabins of the emulated MV/LV substation. (<b>b</b>) Equivalent circuit of the testing setup.</p>
Full article ">Figure 3
<p>(<b>a</b>) Model emulating a set of five MV/LV substations (TC1, TC2,…, TC5) interconnected by 250 m length 12/20 kV cable systems with a 240 mm<sup>2</sup> aluminium section; (<b>b</b>) electrical parameters of each network element used for the transient simulation.</p>
Full article ">Figure 4
<p>Schematic circuit of ATP for the transient analysis corresponding to the equivalent circuit model shown in <a href="#sensors-24-05312-f003" class="html-fig">Figure 3</a> when the ideal current pulses are injected: (a) in the cable end and (b) in the MV cabinet in which the cable end is connected.</p>
Full article ">Figure 5
<p>Pulse front polarity of raw signal in green and filtered signal in red, and pattern polarity depending on the traveling direction (<b>a</b>) when the PD source is at the cable system side for the HFCT installed at the grounding of the cable terminal and (<b>b</b>) when the PD source is outside of the cable at the installation side.</p>
Full article ">Figure 6
<p>Flowchart of the procedure to determine PRPD pattern polarity.</p>
Full article ">Figure 7
<p>Examples (<b>a</b>,<b>b</b>) of arrival time identification based on energy criterion and pulse front polarity analysis based on the slope at 50% of the peak.</p>
Full article ">Figure 8
<p>Schematic circuit of the scale model representing a GIS and cable system to check the PD location capability of any PD analyzer using HFCT sensors.</p>
Full article ">Figure 9
<p>On the left, schematic circuit of the MV cable monitored; on the right, detailed view of HFCT sensors installed hugging the braids of cable terminations. The pulse polarity described in <a href="#sec6dot2-sensors-24-05312" class="html-sec">Section 6.2</a> leads to the red up and down arrow symbol corresponding to the PD pulse currents.</p>
Full article ">Figure 10
<p>(<b>a</b>) Affected phase analysis, (<b>b</b>) pattern polarity analysis, and (<b>c</b>) pulse polarity analysis in the positive half-period.</p>
Full article ">Figure 11
<p>Schematic circuit of the adjacent MV cable system supervised in measurement #2.</p>
Full article ">
19 pages, 8100 KiB  
Article
Techno-Economic Assessment of Bifacial Photovoltaic Systems under Desert Climatic Conditions
by Osama Ayadi, Bilal Rinchi, Sameer Al-Dahidi, Mohammed E. B. Abdalla and Mohammed Al-Mahmodi
Sustainability 2024, 16(16), 6982; https://doi.org/10.3390/su16166982 - 15 Aug 2024
Viewed by 715
Abstract
The decaying prices and improving efficiency of bifacial solar photovoltaic (PV) technologies make them most promising for harnessing solar radiation. Deserts have a high solar potential, but harsh conditions like high temperatures and dust negatively affect the performance of any proposed solar system. [...] Read more.
The decaying prices and improving efficiency of bifacial solar photovoltaic (PV) technologies make them most promising for harnessing solar radiation. Deserts have a high solar potential, but harsh conditions like high temperatures and dust negatively affect the performance of any proposed solar system. The most attractive aspect of deserts is their long-term sustainability, as they are free from urban and agricultural expansion. In this work, the System Advisor Model (SAM) software version 2023.12.17 was used to model a 100 MW PV plant and evaluate the techno-economic performance of fixed, 1-axis, and 2-axis bifacial systems under the climatic conditions of six deserts from around the world. This study explores technical parameters such as the performance ratio, specific yield, and capacity factor. Additionally, the levelized cost of energy (LCOE) indicator was used to compare the economic performance of the different systems. Results showed high specific yield: the averages for the three systems in six deserts were 2040, 2372, and 2555 kWh/kWp, respectively. Economic analysis found that an LCOE below 4 ¢/kWh is achievable in all deserts, reaching a minimum of 2.45 ¢/kWh under favorable conditions. These results emphasize the high potential of utility-scale PV projects in deserts to advance a green, sustainable energy future. Full article
(This article belongs to the Section Energy Sustainability)
Show Figures

Figure 1

Figure 1
<p>Map of the selected locations around the world.</p>
Full article ">Figure 2
<p>Research methodology.</p>
Full article ">Figure 3
<p>Sahara Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 4
<p>Arabian Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 5
<p>Gobi Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 6
<p>Atacama Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 7
<p>Sturt Stony Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 8
<p>Mojave Desert results: (<b>a</b>) hourly production of the fixed and 1-axis systems; (<b>b</b>) hourly production of the fixed and 2-axis systems; (<b>c</b>) hourly production of the three systems.</p>
Full article ">Figure 9
<p>Direct comparison of daily power production: (<b>a</b>) Fixed-tilt systems; (<b>b</b>) 1-axis tracking systems; (<b>c</b>) 2-axis tracking systems.</p>
Full article ">Figure 10
<p>The annual specific yield of the fixed, 1-axis, and 2-axis system under the six selected deserts.</p>
Full article ">Figure 11
<p>The performance ratio of the fixed, 1-axis, and 2-axis system under the six selected deserts.</p>
Full article ">Figure 12
<p>The capacity factor of the fixed, 1-axis, and 2-axis system under the six selected deserts.</p>
Full article ">Figure 13
<p>The LCOE of the fixed, 1-axis, and 2-axis system under the six selected deserts.</p>
Full article ">
19 pages, 6875 KiB  
Article
Implementing Tri-Brid Energy Systems for Renewable Integration in Southern Alberta, Canada
by Mohammad Adnan Aftab, James Byrne, Paul Hazendonk, Dan Johnson, Locke Spencer and Tim Weis
Clean Technol. 2024, 6(3), 1038-1056; https://doi.org/10.3390/cleantechnol6030052 - 13 Aug 2024
Viewed by 632
Abstract
The steep decline in the price of wind turbines and solar photovoltaics provides a possibility to decarbonize electricity deeply and affordably. This study uses the HOMER Pro energy modeling tool to model an optimized grid-connected renewable energy system for a community in southern [...] Read more.
The steep decline in the price of wind turbines and solar photovoltaics provides a possibility to decarbonize electricity deeply and affordably. This study uses the HOMER Pro energy modeling tool to model an optimized grid-connected renewable energy system for a community in southern Alberta, Canada. The study’s goal is to identify the best renewable energy technology combinations that can provide electricity at the lowest levelized cost of energy (LCOE) and has lower greenhouse gas emissions as compared to the electricity produced by traditional fossil fuel. Gleichen is a small town in southern Alberta that is close to numerous commercial wind and solar projects given the region’s high quality renewable resources. “Tri-brid” systems consisting of wind turbines, solar photovoltaics, and battery energy storage systems (BESS) are considered and compared based on electricity prices, net present cost, and greenhouse gas emissions savings. This tri-brid system is connected to the grid to sell excess generated electricity or buy electricity when there is less or no availability of solar and wind energy. The tri-brid energy system has an estimated LCOE of 0.0705 CAD/kWh, which is competitive with the price of electricity generated by natural gas and coal, which is 0.127 CAD/kWh. Full article
Show Figures

Figure 1

Figure 1
<p>Potential site for PV and wind turbine Installation.</p>
Full article ">Figure 2
<p>Homer Pro simulation flowchart.</p>
Full article ">Figure 3
<p>Solar potential map of southern Alberta.</p>
Full article ">Figure 4
<p>Wind potential map of southern Alberta.</p>
Full article ">Figure 5
<p>Monthly average wind speed in Gleichen, Alberta.</p>
Full article ">Figure 6
<p>Monthly average solar global horizontal irradiance (GHI) in Gleichen, Alberta.</p>
Full article ">Figure 7
<p>Schematic presentation of grid and AC load (base case).</p>
Full article ">Figure 8
<p>Schematic presentation of grid-connected PV system.</p>
Full article ">Figure 9
<p>Schematic presentation of grid-connected wind turbine system.</p>
Full article ">Figure 10
<p>Schematic presentation of grid-connected PV and wind turbine system.</p>
Full article ">Figure 11
<p>Cost analysis of different power system configurations.</p>
Full article ">Figure 12
<p>Cash flow of national grid-only system configuration.</p>
Full article ">Figure 13
<p>Cash flow of PV–grid system configuration.</p>
Full article ">Figure 14
<p>Cash flow of WT–grid system configuration.</p>
Full article ">Figure 15
<p>Cash flow of grid-tied PV-WT system configuration.</p>
Full article ">Figure 16
<p>ESS quantity for different power system configurations.</p>
Full article ">Figure 17
<p>GHG emissions for different power system configurations.</p>
Full article ">
Back to TopTop