Next Article in Journal
Research on Obstacle Avoidance Planning for UUV Based on A3C Algorithm
Previous Article in Journal
Discovery of Prevalent Ciliophora, Discoba, and Copepoda Protists in Deep Sea Water by In Situ Nucleotide Extraction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Performance of an Oscillating Wing Energy Harvester Using a Leading-Edge Flap

School of Mechanical Engineering, Kyungpook National University, Daegu 41566, Republic of Korea
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(1), 62; https://doi.org/10.3390/jmse12010062
Submission received: 28 November 2023 / Revised: 22 December 2023 / Accepted: 23 December 2023 / Published: 26 December 2023
(This article belongs to the Section Marine Energy)

Abstract

:
In this study, we investigated the power generation capability of an oscillating wing energy harvester featuring an actively controlled flap positioned at the wing’s leading edge. The findings revealed that attaching a leading-edge flap reduces fluid flow separation below the wing’s lower surface at the leading edge, resulting in smoother flow and increased velocity near the hinge region. The leading-edge flap increases the pressure difference across the wing’s surface, thereby enhancing the overall performance. In addition, the introduction of the leading-edge flap effectively elongates the wing’s effective projected length in the heaving direction, leading to increased thrust. We examined flap lengths ranging from 10% to 50% of the chord length, with the maximum pitch angles of the wing and flap varying from 75° to 105° and 30° to 55°, respectively. The optimal power generation was achieved using a flap length of 40% of the chord length, combined with maximum wing and flap pitch angles of 95° and 45°, respectively. These conditions yielded a 29.9% overall power output increase and a 20.2% efficiency improvement compared to the case without the leading-edge flap.

1. Introduction

The global shift toward decreasing reliance on fossil fuels has motivated the investigation of alternative energy sources and methods, with a focus on enhancing their effectiveness. Energy harvesters based on flapping foils, inspired from natural movements, have gained traction as an effective means to extract energy from various natural sources, such as streams, rivers, tidal currents, and wind. This approach has garnered substantial interest in recent years [1,2]. The idea of utilizing a flapping motion for energy generation was initially introduced by Wu in 1972 [3], and later, in 1981, McKinney and Delaurier pioneered the extraction of energy from the heaving and pitching motions of a fluid using a flapping foil [4]. Subsequently, extensive research has been conducted on optimizing the energy-harvesting capabilities of flapping foils via the investigation of motion parameters [1,5,6], geometric and viscous parameters [2,7,8,9,10], and flow control methods [11,12,13,14,15].
Xiao et al. [6] introduced an innovative non-sinusoidal motion for oscillating hydrofoils within energy harvesters. They combined sinusoidal plunging movements with a nearly trapezoidal pitching motion to assess the impact on energy extraction performance. They found that a higher angle of attack results in enhanced power extraction. In comparison to sinusoidal motion, they observed substantial enhancements with the optimal pitching profile, boosting the power coefficient by up to 63% and the total efficiency by up to 50%. Wu et al. [16] investigated the effects of stroke deviation and variables such as the horizontal motion amplitude, phase difference, and frequency on the performance of flapping-foil energy harvesters. Their results emphasized that incorporating horizontal motion enhances the lift force, thereby improving the power generation capacity of the energy harvesters. Their study underscored the enhancement of power output performance owing to flapping-foil energy harvesters. Wang et al. [17] proposed a novel reversed-D trajectory and concluded that this specific trajectory motion can significantly enhance the power output. Swain et al. [18] investigated energy extraction performance by altering the flapping trajectory and the spatial arrangement of foils, identifying favorable wake interaction patterns that contribute to increased efficiency.
Shanmugam and Sohn [19] investigated various deflector designs aimed at improving power generation in flapping-foil energy harvesters. They examined the impact of upstream deflectors (N), the angle of inclination of these deflectors, and the spacing between the hydrofoil and the deflector on the performance. Their simulations highlighted the considerable influence of these parameters on the power generation efficiency of the energy harvesters. Wang et al. [20] analyzed the impact of density-stratified flow on energy extraction efficiency and its interaction with the pitching amplitude. They [21] proposed a tandem-hydrofoil-based tidal array with improved density and reduced costs. Petikidis and Papadakis [22] studied fully passive flapping foils in free surface flow, determining optimal submergence depths and assessing the influence of monochromatic waves. Dahmani and Sohn [23] introduced a novel approach using oscillating tandem wings inside a convergent duct to enhance power extraction in energy harvesting. The downstream hydrofoil was shown to be more influential, contributing 4–80% to the improved power output owing to the high incoming fluid velocity due to the duct design. In addition, they explored the impact of vertical interfoil spacing on tandem/parallel coupled oscillating wings, achieving a 23% increase in power extraction efficiency compared to traditional arrangements [24]. In these systems, maintaining stability is important, as it ensures consistent operation and reduces the risk of performance degradation. Zhao [25,26,27,28] explored the principles of stability as they pertain to the design and analysis of such systems. Building on these insights, recent advancements in piezoelectric harvesters, such as the development of a harvester with a U-shaped geometry [29], have led to significant improvements. This proposed model surpasses traditional models in efficiency, showcasing the evolution of energy-harvesting technology. Further contributing to this field, the creation of a semi-submersible piezoelectric harvester optimized for increased power output [30] underscores the growing potential of these technologies in a range of applications, driven by continuous improvements in design and optimization.
Flow control methods for harnessing energy from flapping-foil systems can be categorized into two types, passive control and active control, based on whether additional energy is required within the flow field [31]. Passive control techniques involve incorporating simple structures on or adjacent to the surface of the airfoil to enhance energy harvesting without introducing extra energy. Examples of passive control methods include passive foil deformation [32], Gurney flaps [33,34], and corrugated foils [35]. Passive control methods are cost-effective and relatively straightforward to install. However, they operate within predefined states and lack adaptability for real-time adjustments to meet changing requirements. In contrast, active control methods such as circulation control [36] and plasma actuators [37,38] can modify the flow field by adjusting excitation parameters as needed, making them more efficient compared to passive control methods.
Flaps are commonly used in lifting mechanisms in the aerospace industry owing to their simple structure, robustness, and effectiveness [39,40]. Flaps can be implemented in both passively and actively controlled forms. Although flaps have been extensively studied for their applications in single blades and vertical-axis wind turbines [41,42,43], the investigation into energy harvesting using flapping wings has mainly focused on Gurney flaps [33,34]. For example, Bing [33] revealed the introduction of a Gurney flap-influenced vortex generation at the trailing edge, which increased the pressure difference on the foil surface and improved the lift force, leading to a 21% improvement in energy-harvesting efficiency. On the other hand, any additional energy consumption resulting from the swing of the Gurney flap itself was not mentioned. Totpal et al. [44] studied the impacts of passive leading-edge flaps on flapping-wing energy harvesters at low cutoff frequencies. Although they observed an improvement in the heave force, this improvement was primarily limited to the initial stages of the flapping cycle when the heave speed was low, and the synchronization between force and speed was poor, making overall flapping-wing control challenging. Alam and Sohn investigated [45] the impact of an actively controlled trailing-edge flap, accompanied by a comprehensive parametric analysis [46]. Their findings indicated that the implementation of the flap yielded advantageous outcomes in terms of enhancing power extraction performance.
In this study, we examined the potential for enhancing energy harvesting in oscillating wing systems via the addition of a controllable leading-edge flap. Our primary goal was to improve the performance of the energy harvester. To achieve this, we conducted a quantitative evaluation of numerous factors, including various flap lengths and maximum pitching angles for the wing and flap, to identify the optimized configurations that maximize power generation and enhance overall efficiency. Flap lengths varying from 10% to 50% of the chord length were systematically investigated. In addition, variations in maximum wing pitch angles spanning from 75° to 105° and maximum flap pitch angles ranging from 30° to 55° were examined. Through this comprehensive investigation, we aimed to reveal the relationships among these diverse parameters and their impacts on energy-harvesting performance.

2. Mathematical Modeling

2.1. Problem Description

We investigated the dynamics of unsteady fluid flow in an oscillating wing energy harvester using a two-dimensional model under turbulent flow conditions by examining the impact of a leading-edge flap on the performance of energy harvesting. Figure 1a presents a two-dimensional (2D) representation of the energy harvester with and without the flap. In addition, Figure 1b shows three-dimensional (3D) visualizations to view this configuration more clearly. This research primarily focuses on a comparative analysis of the wing’s performance with and without the leading-edge flap, shedding light on the role of the leading-edge flap in boosting energy-harvesting efficiency.
Table 1 provides an overview of the physical model and the optimal parameter settings needed to maximize the power generation in an oscillating wing energy harvester without a flap, as reported by Usoh et al. [9]. The oscillating wing model, when equipped with a leading-edge flap, maintains identical physical characteristics (e.g., thickness and total length). However, factors such as the length of the flap and the peak pitch angles of both the wing and flap are varied. The wing and flap are separated by a gap equivalent to half a percent of the wing’s chord length, and the junction between the wing and flap features a radius equal to half the plate thickness.
Figure 2 shows the oscillating wing kinematic in two scenarios: a wing without a flap (2a) and another wing with a leading-edge flap (2b). The incoming fluid velocity is U, and the wing’s chord length is c. The wing’s main body is pivotally attached at one-third of the chord length from the leading edge, enabling both heaving and pitching motions. Meanwhile, the flap, hinged at the wing’s leading edge, executes pitching motions relative to the wing’s main body. Notably, changes in the flap’s pitching profile, dictated by the angle of incidence, significantly affect the energy harvester’s performance. To evaluate the impact of attaching a leading-edge flap to an oscillating wing, its power generation capacity is compared with the maximum power output of a wing without a flap based on the optimum conditions obtained for a flapless wing outlined in Table 1 [9].
The wing’s heaving motion is described by Equation (1):
h = H o s i n ( ω t + ϕ )
In this equation, h denotes the instantaneous amplitude of heaving motion, ω is the angular frequency (expressed as ω = 2πf) associated with the wing’s oscillation, t is the time, and φ is the heave and pitch motion phase difference.
The pitch motion of the wing’s main body is governed by Equation (2):
θ = θ o s i n   ( ω t )
Here, θ represents the pitch angle of the main body. Similarly, the pitch motion of the flap is expressed by Equation (3):
ψ = ψ o s i n   ( ω t )
Here, ψ represents the pitch angle of the flap. The force  F Y ( t )  and the moment  M ( t )  are associated with heaving and pitching relative to the wing’s pivot point, respectively. The corresponding average power outputs derived from these forces and moments are represented by  P ¯ y  and  P ¯ m , respectively, and  P ¯  represents the average total power output [33,34].
P ¯ = P y ¯ + P ¯ m = 0 T [ F Y t V y t + M ( t ) Ω ( t ) ] d t
In this equation,  V y t  denotes the instantaneous heaving motion of the wing, and  Ω t  represents the angular velocity at time  t . The coefficients of heaving and pitching power are represented by Cpy and Cpm, respectively. The total averaged power extraction coefficient in a cycle is expressed as  c ¯ p t  and can be calculated by summing up the averaged power output coefficients obtained from  M ( t )  and  F y ( t )  [47,48]:
C ¯ p t = C ¯ p y + C ¯ p m = P ¯ 0.5 ρ U 3 c T   .  
Considering the combined effects of the main body and flap, the forces, moments, and power generated by both components were summed. The computation of the moment power for the flap incorporated the consideration of relative angular motion between the wing’s main body and flap. The efficiency of power generation in the oscillating wing energy harvester is denoted as η, obtained using Equation (6) [49]:
η = C ¯ p t c d

2.2. Numerical Method

The power extraction efficiency of the oscillating wing was simulated using the incompressible unsteady Reynolds-averaged Navier–Stokes equations. The kw SST turbulence model was employed, as adopted by other researchers in similar studies [17,48]. The computational model was solved using ANSYS commercial software version 20 (ANSYS, Canonsburg, PA, USA) and the overset method by dividing it into stationary and moving grid sections, as illustrated in Figure 3.
To ensure domain independence, a computational domain spanning from  20 c   t o   50 c  in length to  25 c   t o   25 c  in breadth was selected based on the observation that enlarging the domain size had an insignificant impact on the obtained results. At the inlet, outlet, and wing surface, the boundary conditions, velocity inlet, pressure outlet, and no-slip wall conditions were implemented. The pressure outlet was placed at  50 c  from the oscillating wing pivot point, as depicted in Figure 3a. Separate subgrids (moving grid) were utilized for the main body and the flap, as shown in Figure 3b. The movable grid areas for each part were organized into rectangular grids measuring 6c and 5c. For pressure, momentum, and turbulent viscosity, the spatial discretization was conducted using the second-order upwind approach. The dynamic mesh method facilitated mesh motion, and the airfoil’s movement was input through user-defined functions. The presented analysis shows the data collected after the completion of the eighth stable periodic cycle, which included 2000 time steps.
To investigate the precision and dependability of the numerical simulations, studies on grid and time independence were conducted by altering the grid resolution and time step. Three distinct mesh densities were employed: a coarse grid (D1), a medium grid (D2), and a fine grid (D3). These simulations were performed at a Reynolds number of 5 × 105 using a chord length of 1 and a pivot point located at c/3. The heave amplitude was set at c = Ho = 1, the maximum pitching amplitude, at 75°; the phase angle, at 90°; and the reduced frequency, at 0.14. Table 2 illustrates the influence of grid resolution and time step on the computed value of  C ¯ P t . Based on the outcomes of the grid independence assessment, the D2 grid with a time step of T/2000 was chosen for the simulations to ensure both accuracy and reliability.
Further, we validated the results obtained upon comparison with the simulation results reported by Usoh et al. [9] and Kinsey et al. [48]. The validation process involved examining a flat plate and a NACA0015 airfoil. For the flat plate scenario, the simulation was conducted at Re = 1100, Ho/c = 1.0, f* = 0.14, and θo = 75°. The calculated Cpt values closely matched the results reported by Usoh et al. [9], as depicted in Figure 4a. For the NACA0015 airfoil, shown in Figure 4b, the simulations at Re = 5 × 105, Ho/c = 1.0, θo = 75°, and f* = 0.14 demonstrated a good match between the pushing force coefficient and the x-direction force coefficient in this study and those documented by Kinsey et al. [50]. These comparisons validate the applied numerical model and mesh used in this study.

3. Results and Discussion

Our investigation presents an active flow control methodology aimed at improving the power generation of an oscillating wing. The effects of a rectangular-plate oscillating wing featuring a flap at the leading edge on the energy-harvesting performance are investigated during coupled heaving and pitching movements. The study further evaluates the wider applicability of this flow control technique (over a range of motion parameters) by varying multiple parameters, such as the maximum pitch angles of the wing and the flap, as well as the flap length. Throughout the study, the following parametric conditions were kept constant: the nondimensional pivot point position (xp) = 0.33c, the heave amplitude (Ho/c) = 1.0, the phase angle (φ) = 90°, and the frequency (f*) = 0.14.

3.1. Effect of Varied Maximum Flap Pitch Angles on Power Output

In this section, a comparative analysis of the energy-harvesting performance of a wing with an integrated flap is performed against a baseline foil without a flap at a fixed maximum wing pitch angle (θo) of 75°. The flap length is maintained at 10% of the chord length (c), and the maximum flap pitch angle (ψo) is varied from 20° to 50°.
Table 3 shows the average pushing power coefficient ( C ¯ p y ), average moment power coefficient ( C ¯ p m ), and average total power coefficient ( C ¯ p t ) for the wing with a flap length of 10% of c at θo = 75°. The power coefficients are tabulated against a range of ψo values, along with the percentage change in  C ¯ p t  relative to the wing without a flap (denoted as  Δ C ¯ p t ). In this context, “plate” refers to the baseline wing without a flap, and the varying flap conditions are indicated by their respective maximum pitch angles, such as ψo = 20°, 25°, and so on. The data exhibit a direct correlation between ψo and  C ¯ p y , suggesting that an increase in the flap’s pitch angle enhances the average heaving power coefficient. Conversely, an inverse relationship is observed between ψo and  C ¯ p m , where larger flap pitch angles correlate with more negative values of the average moment power coefficient.  C ¯ p t  displays a nonlinear association with ψo: it shows a modest increase as ψo ascends from 20° to 40°, followed by a decrease in  C ¯ p t  beyond 40°. Notably, the power output increases by a maximum of 6% with a 10% flap length compared to the plate case.
Figure 5a,b show the variations in the pushing force coefficient (Cy) and the pushing power coefficient (Cpy) in time (T) for different ψo cases outlined in Table 3. Notably, between the time intervals from 0.1 t/T to 0.3 t/T, both Cy and Cpy increase as ψo increases, indicating that the oscillating foil experiences enhanced pushing force and an increase in pushing power as ψo increases during this interval. Figure 5c,d present the moment power coefficient (Cpm) and the total power coefficient (Cpt) throughout a cycle considering different ψo cases. The variation in Cpm values is similar for different cases until 0.3 t/T, whereas from 0.3 t/T to 0.5 t/T, the Cpm values decrease with ψo. During the interval from 0.15 t/T to 0.35 t/T, the Cpt values also increase with increasing ψo. The total power output of the oscillating foil shows an upward trend with increasing ψo values during this interval. However, the maximum  C ¯ p t  is achieved at ψo = 40°, as detailed in Table 3.
In Figure 6, streamline and vorticity, velocity magnitude, and pressure contour plots around the foil’s surface are presented for various ψo cases at 0.15 t/T and θo = 75°. For a flat plate, fluid separation occurs below the leading edge on the lower surface of the wing, as illustrated in Figure 6a. However, this separation is significantly reduced at higher ψo values by attaching a leading-edge flap to the wing, resulting in smoother fluid flow and increased velocity near the hinge region on the lower surface, as depicted in Figure 6b.
Figure 6c displays the pressure contour plot, showing that the leading-edge flap reduces the lower surface pressure compared to the wing without a flap. In Figure 7, the pressure coefficient on the wing’s surface along the heaving motion is illustrated for various ψo cases at 0.15 t/T and θ0 = 75°. The presence of a flap decreases the surface pressure owing to the expanded high-velocity region in comparison to the plate, resulting in a greater pressure differential between the upper and lower surfaces and generating a stronger pushing force in the wing, with higher ψo values compared to the plate. In addition, the flap notably increases the projected length in the wing’s heaving motion, positively influencing the pushing force at this specific moment.

3.2. Impact of Flap Length Variation on Power Output

To examine the role of the leading-edge flap length on the power output of the energy harvester, this section maintains constant θο = 75° and ψo = 40°. The study evaluates changes in flap length relative to the chord length considering lengths of 0% (equivalent to a flat plate), 10%, 15%, 20%, 33% (pivot point), and 40%.
Table 4 summarizes the average coefficients of power ( C ¯ p y C ¯ p m C ¯ p t ) and the percentage change in  C ¯ p t  (denoted as  Δ C ¯ p t ) for the wing with different flap lengths at fixed θo = 75° and ψo = 40°. The  C ¯ p y  values are maximum for a flap span of 20% of the chord length, beyond which the values start decreasing. Conversely, the magnitude of the  C ¯ p m  value steadily increases with the extension of the flap length. For a flap length of 10% of the chord, the  C ¯ p t  value is higher than the flat plate configuration but decreases progressively as the flap span increases. However, the  Δ C ¯ p t  value steadily decreases as the flap span extends up to 20% of the chord length, experiencing a significant drop beyond this threshold. This indicates that beyond a certain flap length, an increase in flap length adversely affects the power output generation of the energy harvester system for the same θo and Ψo.
Figure 8 illustrates the effects of different flap lengths on the pushing force (Cy), pushing power (Cpy), moment power (Cpm), and total power (Cpt) coefficients within an oscillation cycle at fixed pitch angles (θο = 75° and ψo = 40°). Figure 8a displays the variation in Cy, revealing that adding a flap length of 10% of c increases Cy relative to the baseline “plate” case from 0.15 t/T to 0.4 t/T. This increase becomes more pronounced with a flap span of 15% of c, and it continues to grow up to 20% of c. Beyond this flap length, the Cy values plateau, suggesting a threshold for enhancement through flap extension during this interval. Conversely, the interval from 0.4 t/T to 0.5 t/T is characterized by a marked and consistent reduction in Cy as the flap lengths continue to increase. Given that the wing’s vertical velocity (Vy) is higher from 0.1 t/T to 0.4 t/T, even a modest increase in Cy due to changes in flap length has a significant impact on Cpy, as depicted in Figure 8b. Figure 8c indicates that the magnitude of the negative Cpm values intensifies with the increment in the flap length up to 0.25 t/T, suggesting an increase in adverse moment power. Conversely, the positive Cpm values decline from 0.25 t/T to 0.5 t/T, implying a reduction in favorable moment power during this period, particularly for longer flaps. This corresponds to a notable increase in the average negative moment power for larger flap lengths, consistent with the data in Table 4. As for the total power coefficient displayed in Figure 8d, there is an observable increase in Cpt with the lengthening of the flap from 0.1 t/T to 0.3 t/T. However, this trend is reversed from 0.3 t/T to 0.5 t/T, where Cpt begins to decrease. The results suggest that the total power output initially benefits from the increased flap length, which peaks at a certain point before starting to diminish, suggesting that an optimal flap length exists for maximizing the total power output under the specific pitch angles of θο = 75° and ψo = 40°.
Figure 9 presents the streamline and vorticity, along with pressure contour plots, around the foil’s surface for different flap lengths at specific instances of 0.25t/T. In Figure 9a, a flap span of 0% of c (equivalent to a flat plate) leads to fluid flow separation below the lower surface of the wing’s leading edge. This flow separation decreases as the flap length increases to 10% of c, and it decreases further when the flap length reaches 15% of c. In Figure 9b, the associated pressure contour plots reveal that larger flap lengths result in an expanded region of lower pressures on the lower surface.
Figure 10 displays the variation in pressure across the surfaces of the wing for various flap lengths at 0.25t/T. In Figure 10, with a flap span of 10% of c, the wing’s lower surface pressure is observed to be lower than that of the base plate configuration, attributed to fast fluid flow near the hinge position, where the camber becomes the maximum. As the flap length is extended to 15% of c, a further pressure decrease is observed. However, beyond a flap span of 20% of c, the lower surface pressure begins to increase as flow separation becomes more pronounced, suggesting that up to a certain flap length, the incoming flow stays attached to the wing’s surface, reducing pressure on the lower surface. Conversely, when the flap is excessively long, the flow separates earlier, causing an increase in pressure. Therefore, the variations in pressure across the surfaces of the wing decrease beyond a flap span of 20% of c. In addition, with an increase in flap length, there is also an increase in the flap’s projected length along the x-axis, which is beneficial to enhancing the pushing force, offsetting the effects of the reduced pressure differential.

3.3. Effect of Varied Maximum Wing Pitch Angles

This section examines the influence of altering the maximum wing pitch angle from 75° to 95° while maintaining a constant flap length of 20% of c and a maximum flap pitch angle of 40°. The values in Table 5 reflect the average pushing power ( C ¯ py), average moment power ( C ¯ pm), and average total power output ( C ¯ pt) to assess the performance across varied θo C ¯ py increases from θo = 75° to 85°, reaching its peak at θo = 85°, and it decreases thereafter, suggesting that a wing angle of 85° is optimal for generating the pushing power at this operating condition. Conversely, the absolute  C ¯ pm values exhibit a continuous decrease as θo increases. On the other hand, the variations in  C ¯ P t  show an upward trend with higher wing angles, indicating that greater wing angles enhance power generation for a flap span of 20% of c. These findings emphasize the importance of selecting an appropriate wing angle to maximize the power output of the oscillating wing with a fixed flap length. A pitch angle θo of 90° strikes a balance between pushing power and moment power requirements, resulting in the optimum power generation at a flap span of 20% of c and ψo = 40°.
Figure 11 provides the pushing force, pitching moment, and power coefficients in time T for a flap length of 20% of c, ψo = 40°, and varied θo. Figure 11a,b illustrate the Cy and Cpy variations over a cycle for different wing pitch angles spanning from 75° to 95° in 5° increments. Notably, from 0.1 t/T to 0.4 t/T, Cy and Cpy exhibit higher values when the wing’s maximum pitch angle is lower. However, from 0.4 t/T to 0.5 t/T, Cy and Cpy show an increase with higher maximum wing pitch angles. In addition, the absolute Cpm consistently increases up to 0.25 t/T, indicating an increase in adverse moment power for longer flaps. From 0.25 t/T to 0.5 t/T, the trend of increasing Cpm values continues, suggesting an increase in favorable moment power for the wing, as shown in Figure 11c. Notably, the rise in adverse moment power is less pronounced than that in favorable moment power, resulting in a marked decrease in the average negative moment power with higher θo values, which agrees with the values presented in Table 5. Interestingly, the examination of Cpt values reveals a decreasing trend from 0.1 t/T to 0.4 t/T with a maximum increasing wing pitch angle, followed by an increase from 0.4 t/T to 0.5 t/T as the maximum wing pitch angle increases, as shown in Figure 11d.
Figure 12 presents the streamline and velocity magnitude and pressure contour plots for a wing with a leading-edge flap length of 20% of c and ψo = 40° based on varying θo at intervals of 0.25 t/T. In Figure 12a, the streamline and velocity magnitude contours show a notable increase in velocity around the flap’s hinge area on the lower surface as θo increases. Figure 12b illustrates that the wing’s upper surface pressure significantly increases for θo = 90°.
Figure 13 depicts the distribution of pressure coefficients on the foil surface across different θo cases at 0.25 t/T. Figure 13 displays a continuous decrease in the wing’s lower surface pressure, particularly near the flap–wing junction, with increasing θo. As θo increases, the wing’s projected length along the heaving motion direction decreases, accompanied by an increase in the pressure differential across the wing surface.

3.4. Combined Impact of Varying Flap Length and Maximum Pitch Angles of the Wing and Flap

Figure 14 illustrates the combined influence of the span of the flap (ranging from 10% to 50% of c), a maximum wing pitch angle (θo) ranging from 75° to 105°, and a maximum flap pitch angle (ψo) varying from 30° to 55° on the resulting power output ( C ¯ P t ). When the flap’s extent is at 10% of c C ¯ P t  is maximized at θo = 75° across various ψo, as shown in Figure 14a. The maximum  C ¯ P t  for varied θo occurs at ψo = 40°, although the variation in power output across different θo is small. As the span of the flap increases to 20% of c, the optimum θo significantly rises to 90°, and the maximum  C ¯ P t  values are notably higher compared to the flap span of 10% of c, as depicted in Figure 14b. The optimal θo is found to be the same for flap lengths of 25% and 33% of c, as shown in Figure 14c,d, respectively. However, the maximum power generation is attained for a flap span of 40% of c, beyond which it starts to decrease for various θo and ψo values depicted in Figure 14e,f, respectively.
The optimum power output was achieved for the flap spans ranging from 40% to 50% of c, θo = 95°–100°, and ψo = 45° and 50°. As the length of the flap increases, the optimum θo gradually increases from 75° to 100°, whereas ψo varies between 45° and 50°. The optimal power generation was achieved with a flap span of 40% of c at θo = 95° and ψo = 45°. This configuration resulted in a 29.9% enhancement in power output and a 20.2% boost in efficiency relative to the wing configuration without a leading-edge flap.

4. Conclusions

We investigated a flat plate oscillating wing with an integrated leading-edge flap by focusing on its impact on energy-harvesting performance. Numerical simulations were conducted using the overset grid method under transient conditions, and the k–ω SST turbulence model was employed. Various parameters such as the flap length (from 10% to 50% of the chord length) and maximum pitch angles for the wing and flap (from 75° to 105° and from 30° to 55°, respectively) were quantitatively assessed to identify the configurations that maximize power generation and efficiency.
The results showed that the addition of the leading-edge flap leads to smoother fluid flow along the wing’s lower surface and a high-velocity region compared to the wing without a flap. In addition, an increase in the flap’s maximum pitch angle or the flap length enhances the wing’s effective projected length in the heaving direction. By analyzing the combined effects of varying flap length, wing angle, and flap angle, we identified the optimal configurations that yielded substantial power output enhancements. In particular, flap lengths within the range of 40%–50% of the chord length, combined with θo values from 95° to 100° and ψo of 45°–50°, deliver peak power generation. The results demonstrate a remarkable 29.9% increase in overall power output and a commendable 23% efficiency enhancement compared to the baseline wing configuration without the flap.

Author Contributions

Conceptualization, M.A. and C.H.S.; Methodology, M.A. and C.H.S.; Simulations, M.A.; Investigation, M.A. and C.H.S.; Writing and editing, M.A. and C.H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from the Korean government’s Ministry of Science and ICT (MSIT), provided through a grant from the National Research Foundation of Korea (NRF), under grant number 2022R1F1A1061903.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting the findings of this study are available upon personal request to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

SymbolDescription
twThe thickness of the wing
dThe total vertical movement
xWing’s projected length along heaving motion
xpThe pitching center of the wing
cAirfoil chord length
hWing’s heave amplitude (instantaneous)
HoWing’s heave amplitude (maximum)
θ(t)Wing’s pitch angle (instantaneous)
θoWing’s pitch angle (maximum)
Ψ(t)Flap’s pitch angle (instantaneous)
ΨoFlap’s pitch angle (maximum)
fOscillation frequency
f*Nondimensional reduced frequency (fc/U)
CyCoefficient of pushing force
CmCoefficient of pitching moment
CpyCoefficient of pushing power
CpmCoefficient of pitching moment power
CptCoefficient of total power
C ¯ p y Coefficient of average pushing power
C ¯ p m Coefficient of average moment power
C ¯ p t Coefficient of average total power
Δ C ¯ p t   Change in the average total power coefficient compared to the wing without the flap
FYPushing force
FXDrag force

References

  1. Jones, K.D.; Davids, S.; Platzer, M.F. Oscillating-Wing Power Generator. In Proceedings of the 3rd ASME/JSME Joint Fluids Engineering Conference; American Society of Mechanical Engineers: New York, NY, USA, 1999. [Google Scholar]
  2. Shimizu, E.; Isogai, K.; Obayashi, S. Multiobjective design study of a flapping wing power generator. J. Fluids Eng. Trans. ASME 2008, 130, 0211041–0211048. [Google Scholar] [CrossRef]
  3. Wu, T.Y. Extraction of flow energy by a wing oscillating in waves. J. Ship Res. 1972, 16, 66–78. [Google Scholar] [CrossRef]
  4. McKinney, W.; DeLaurier, J. Wingmill: An Oscillating-Wing Windmill. AIAA Pap. 1981, 5, 109–115. [Google Scholar] [CrossRef]
  5. Lindsey, K. A Feasibility Study of Oscillating-Wing Power Generators. Doctoral Thesis, Naval Postgraduate School, Monterey, CA, USA, 2002. Available online: https://core.ac.uk/download/pdf/36698644.pdf (accessed on 8 August 2023).
  6. Xiao, Q.; Liao, W.; Yang, S.; Peng, Y. How motion trajectory affects energy extraction performance of a biomimic energy generator with an oscillating foil? Renew. Energy 2012, 37, 61–75. [Google Scholar] [CrossRef]
  7. Li, W.; Wang, W.Q.; Yan, Y.; Tian, F.B. Effects of pitching motion profile on energy harvesting performance of a semi-active flapping foil using immersed boundary method. Ocean Eng. 2018, 163, 94–106. [Google Scholar] [CrossRef]
  8. Xie, Y.; Lu, K.; Zhang, D. Investigation on energy extraction performance of an oscillating foil with modified flapping motion. Renew. Energy 2014, 63, 550–557. [Google Scholar] [CrossRef]
  9. Usoh, C.O.; Young, J.; Lai, J.C.S.; Ashraf, M.A. Numerical analysis of a non-profiled plate for flapping wing turbines. In Proceedings of the 18th Australasian Fluid Mechanics Conference, Launceston, Australia, 3–7 December 2012; pp. 3–6. [Google Scholar]
  10. Zhu, Q. Optimal frequency for flow energy harvesting of a flapping foil. J. Fluid Mech. 2011, 675, 495–517. [Google Scholar] [CrossRef]
  11. Zhu, B.; Han, W.; Sun, X.; Wang, Y.; Cao, Y.; Wu, G.; Huang, D.; Zheng, Z.C. Research on energy extraction characteristics of an adaptive deformation oscillating-wing. J. Renew. Sustain. Energy 2015, 7, 023101. [Google Scholar] [CrossRef]
  12. Rushen, J.; Siala, F.; Totpal, A.D.; Planck, C.J.; Liburdy, J.A. The effects of a passively actuated trailing edge on the aerodynamics of an oscillating wing. Am. Soc. Mech. Eng. Fluids Eng. Div. FEDSM 2014, 1B, V01BT12A012. [Google Scholar] [CrossRef]
  13. Zhu, B.; Xia, P.; Huang, Y.; Zhang, W. Energy extraction properties of a flapping wing with an arc-deformable airfoil. J. Renew. Sustain. Energy 2019, 11, 023302. [Google Scholar] [CrossRef]
  14. Sun, G.; Wang, Y.; Xie, Y.; Ma, P.; Zhang, Y. Hydrodynamic and energy extraction properties of oscillating hydrofoils with a trailing edge flap. Appl. Ocean Res. 2021, 110, 102530. [Google Scholar] [CrossRef]
  15. Drost, K.J.; Johnson, H.; Aptey, S.V.; Liburdyz, J.A. Low reynolds number flow dynamics of a thin airfoil with an actuated leading edge. In Proceedings of the 41st AIAA Fluid Dynamics Conference Exhibition, Honolulu, HI, USA, 27–30 June 2011. [Google Scholar] [CrossRef]
  16. Wu, J.; Chen, Y.; Zhao, N.; Wang, T. Influence of stroke deviation on the power extraction performance of a fully-active flapping foil. Renew. Energy 2016, 94, 440–451. [Google Scholar] [CrossRef]
  17. Wang, B.; Zhu, B.; Zhang, W. New type of motion trajectory for increasing the power extraction efficiency of flapping wing devices. Energy 2019, 189, 116072. [Google Scholar] [CrossRef]
  18. Swain, P.K.; Dora, S.P.; Barik, A.K. Energy extraction performance of tandem flapping foil undergoing elliptical motion trajectory. Ocean Eng. 2023, 268, 113390. [Google Scholar] [CrossRef]
  19. Shanmugam, A.R.; Park, K.S.; Sohn, C.H. Comparison of the Power Extraction Performance of an Oscillating Hydrofoil Turbine with Different Deflector Designs. Energies 2023, 16, 3420. [Google Scholar] [CrossRef]
  20. Wang, J.; Deng, J.; Kandel, P.; Sun, L. Numerical study on the energy extraction performance by flapping foils in a density stratified flow. J. Fluids Struct. 2023, 118, 103865. [Google Scholar] [CrossRef]
  21. Wang, G.; Ng, B.F. Energy harvesting performance of a tandem-hydrofoil based closely-interconnected tidal array. Energy Convers. Manag. 2023, 280, 116796. [Google Scholar] [CrossRef]
  22. Petikidis, N.; Papadakis, G. Investigation of Submergence Depth and Wave-Induced Effects on the Performance of a Fully Passive Energy Harvesting Flapping Foil Operating Beneath the Free Surface. 2023. Available online: http://arxiv.org/abs/2304.09611 (accessed on 25 February 2023).
  23. Dahmani, F.; Sohn, C.H. Effect of convergent duct geometry on the energy extraction performance of tandem oscillating hydrofoils system. J. Fluids Struct. 2020, 95, 102949. [Google Scholar] [CrossRef]
  24. Dahmani, F.; Sohn, C.H. Effects of the downstream spatial configuration on the energy extraction performance of tandem/parallel combined oscillating hydrofoils. J. Mech. Sci. Technol. 2020, 34, 2035–2046. [Google Scholar] [CrossRef]
  25. Zhao, K. Stability of a Nonlinear Fractional Langevin System with Nonsingular Exponential Kernel and Delay Control. Discret. Dyn. Nat. Soc. 2022, 2022, 9169185. [Google Scholar] [CrossRef]
  26. Zhao, K. Stability of a Nonlinear Langevin System of ML-Type Fractional Derivative Affected by Time-Varying Delays and Differential Feedback Control. Fractal Fract. 2022, 6, 725. [Google Scholar] [CrossRef]
  27. Zhao, K. Existence and UH-stability of integral boundary problem for a class of nonlinear higher-order Hadamard fractional Langevin equation via Mittag-Leffler functions. Filomat 2023, 37, 1053–1063. [Google Scholar] [CrossRef]
  28. Zhao, K. Generalized UH-stability of a nonlinear fractional coupling (p1, p2)-Laplacian system concerned with nonsingular Atangana–Baleanu fractional calculus. J. Inequalities Appl. 2023, 96. [Google Scholar] [CrossRef]
  29. Aramendia, I.; Saenz-Aguirre, A.; Boyano, A.; Fernandez-Gamiz, U.; Zulueta, E. Oscillating U-shaped body for underwater piezoelectric energy harvester power optimization. Micromachines 2019, 10, 737. [Google Scholar] [CrossRef] [PubMed]
  30. Teso-Fz-betoño, D.; Aramendia, I.; Martinez-Rico, J.; Fernandez-Gamiz, U.; Zulueta, E. Piezoelectric energy harvesting controlled with an IGBT H-bridge and bidirectional buck–boost for low-cost 4G devices. Sensors 2020, 20, 7039. [Google Scholar] [CrossRef] [PubMed]
  31. Kral, L.D.; The, M.I.; Classification, C. active flow control technology. Br. Med. J. (Clin. Res. Ed.) 1985, 291, 219. [Google Scholar] [CrossRef]
  32. Yin, B.; Luo, H. Effect of wing inertia on hovering performance of flexible flapping wings. Phys. Fluids 2010, 22, 111902. [Google Scholar] [CrossRef]
  33. Zhu, B.; Huang, Y.; Zhang, Y. Energy harvesting properties of a flapping wing with an adaptive Gurney flap. Energy 2018, 152, 119–128. [Google Scholar] [CrossRef]
  34. Xie, Y.H.; Jiang, W.; Lu, K.; Zhang, D. Numerical investigation into energy extraction of flapping airfoil with Gurney flaps. Energy 2016, 109, 694–702. [Google Scholar] [CrossRef]
  35. Le, T.Q.; Ko, J.H.; Byun, D. Morphological effect of a scallop shell on a flapping-type tidal stream generator. Bioinspiration Biomim. 2013, 8, 036009. [Google Scholar] [CrossRef]
  36. Shi, F.; Sun, X. Study on performance enhancement of a flapping foil energy harvester using circulation control. J. Fluids Eng. Trans. ASME 2021, 143, 071207. [Google Scholar] [CrossRef]
  37. Feng, L.H.; Jukes, T.N.; Choi, K.S.; Wang, J.J. Flow control over a NACA 0012 airfoil using dielectric-barrier-discharge plasma actuator with a Gurney flap. Exp. Fluids 2012, 52, 1533–1546. [Google Scholar] [CrossRef]
  38. Mahboubidoust, A.; Ramiar, A.; Dardel, M. Numerical investigation of plasma actuated and non-actuated Gurney flaps on aerodynamic characteristics of a plunging airfoil. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 2016, 230, 1423–1437. [Google Scholar] [CrossRef]
  39. Lee, T.; Su, Y.Y. Unsteady airfoil with a harmonically deflected trailing-edge flap. J. Fluids Struct. 2011, 27, 1411–1424. [Google Scholar] [CrossRef]
  40. Visconti, U.; Eun, W.J.; Sim, J.S.; Lee, S.W.; Shin, S.J. Design improvements and flap deflection evaluations with considering centrifugal load on active trailing edge flap. Aircr. Eng. Aerosp. Technol. 2019, 91, 10–19. [Google Scholar] [CrossRef]
  41. Bofeng, X.; Junheng, F.; Qing, L.; Chang, X.; Zhenzhou, Z.; Yue, Y. Aerodynamic performance analysis of a trailing-edge flap for wind turbines. J. Phys. Conf. Ser. 2018, 1037, 022020. [Google Scholar] [CrossRef]
  42. Liu, Q.; Miao, W.; Li, C.; Hao, W.; Zhu, H.; Deng, Y. Effects of trailing-edge movable flap on aerodynamic performance and noise characteristics of VAWT. Energy 2019, 189, 116271. [Google Scholar] [CrossRef]
  43. Bianchini, A.; Balduzzi, F.; Di Rosa, D.; Ferrara, G. On the use of Gurney Flaps for the aerodynamic performance augmentation of Darrieus wind turbines. Energy Convers. Manag. 2019, 184, 402–415. [Google Scholar] [CrossRef]
  44. Totpal, A.D.; Siala, F.F.; Liburdy, J.A. Energy harvesting of an oscillating foil at low reduced frequencies with rigid and passively deforming leading edge. J. Fluids Struct. 2018, 82, 329–342. [Google Scholar] [CrossRef]
  45. Alam, M.; Sohn, C.H. Enhanced performance of oscillating wing energy harvester using active controlled flap. J. Mech. Sci. Technol. 2023, 37, 2405–2415. [Google Scholar] [CrossRef]
  46. Alam, M.; Sohn, C.H. Parametric analysis of an oscillating wing energy harvester with a trailing edge flap. J. Mech. Sci. Technol. 2023, 37, 3563–3573. [Google Scholar] [CrossRef]
  47. Kinsey, T.; Dumas, G. Parametric study of an oscillating airfoil in a power-extraction regime. AIAA J. 2008, 46, 1318–1330. [Google Scholar] [CrossRef]
  48. Kinsey, T.; Dumas, G. Computational fluid dynamics analysis of a hydrokinetic turbine based on oscillating hydrofoils. J. Fluids Eng. Trans. ASME 2012, 134, 021104. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Wang, Y.; Han, J.; Sun, G.; Xie, Y. Effects of hydrofoil motion parameters and swing arm parameters on power extraction of a flexible hydrofoil in swing arm mode. Ocean Eng. 2022, 245, 110543. [Google Scholar] [CrossRef]
  50. Kinsey, T.; Dumas, G. Optimal operating parameters for an oscillating foil turbine at reynolds number 500,000. AIAA J. 2014, 52, 1885–1895. [Google Scholar] [CrossRef]
Figure 1. Configurations of an oscillating wing with and without a flap: (a) 2D view; (b) 3D view.
Figure 1. Configurations of an oscillating wing with and without a flap: (a) 2D view; (b) 3D view.
Jmse 12 00062 g001aJmse 12 00062 g001b
Figure 2. Kinematics of (a) oscillating wing without a flap; (b) oscillating wing with a leading-edge flap.
Figure 2. Kinematics of (a) oscillating wing without a flap; (b) oscillating wing with a leading-edge flap.
Jmse 12 00062 g002
Figure 3. Illustrations of the computational domain and boundary conditions: (a) computational domain; (b) sub-region grid.
Figure 3. Illustrations of the computational domain and boundary conditions: (a) computational domain; (b) sub-region grid.
Jmse 12 00062 g003
Figure 4. Comparative analysis of the numerical results obtained in this study against the validation data: (a) comparison of the instantaneous Cpt values for a rectangular-plate wing at Re = 1100 in comparison to the findings reported by Usoh et al. [9]; (b) comparison of the instantaneous CL and CD values for NACA0015 at Re = 5 × 105, benchmarked against the findings reported by Kinsey et al. [48].
Figure 4. Comparative analysis of the numerical results obtained in this study against the validation data: (a) comparison of the instantaneous Cpt values for a rectangular-plate wing at Re = 1100 in comparison to the findings reported by Usoh et al. [9]; (b) comparison of the instantaneous CL and CD values for NACA0015 at Re = 5 × 105, benchmarked against the findings reported by Kinsey et al. [48].
Jmse 12 00062 g004
Figure 5. Comparative analyses of Cy, Cpy, Cpm, and Cpt values within a cycle for the wing with a flap length of 10% of c at θo = 75° and varied ψo (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Figure 5. Comparative analyses of Cy, Cpy, Cpm, and Cpt values within a cycle for the wing with a flap length of 10% of c at θo = 75° and varied ψo (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Jmse 12 00062 g005
Figure 6. (a) Streamline and vorticity patterns, (b) velocity magnitude, and (c) pressure distribution of the wing with varied flap lengths at θo = 75° with varied ψo at 0.15 t/T.
Figure 6. (a) Streamline and vorticity patterns, (b) velocity magnitude, and (c) pressure distribution of the wing with varied flap lengths at θo = 75° with varied ψo at 0.15 t/T.
Jmse 12 00062 g006
Figure 7. Pressure coefficient across the wing surface during heaving motion at θο = 75° with varied ψo at 0.15 t/T.
Figure 7. Pressure coefficient across the wing surface during heaving motion at θο = 75° with varied ψo at 0.15 t/T.
Jmse 12 00062 g007
Figure 8. Comparative analyses of Cy, Cpy, Cpm, and Cpt values within a cycle for the wing with varied flap lengths at θo = 75° and ψo = 40° (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Figure 8. Comparative analyses of Cy, Cpy, Cpm, and Cpt values within a cycle for the wing with varied flap lengths at θo = 75° and ψo = 40° (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Jmse 12 00062 g008
Figure 9. (a) Streamline and vorticity and (b) pressure contours of a wing with varying flap lengths at θο = 75° and ψo = 40° at 0.25 t/T.
Figure 9. (a) Streamline and vorticity and (b) pressure contours of a wing with varying flap lengths at θο = 75° and ψo = 40° at 0.25 t/T.
Jmse 12 00062 g009
Figure 10. Pressure coefficient on wing’s surfaces during heaving motion for varied leading-edge flap lengths at θo = 75° and ψo = 40° at 0.25 t/T.
Figure 10. Pressure coefficient on wing’s surfaces during heaving motion for varied leading-edge flap lengths at θo = 75° and ψo = 40° at 0.25 t/T.
Jmse 12 00062 g010
Figure 11. Comparison of Cy, Cpy, Cpm, and Cpt within a cycle for the wing with a leading-edge flap of 20% of c, varied θo and ψo = 40° (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Figure 11. Comparison of Cy, Cpy, Cpm, and Cpt within a cycle for the wing with a leading-edge flap of 20% of c, varied θo and ψo = 40° (a) Cy (b) Cpy (c) Cpm (d) Cpt.
Jmse 12 00062 g011
Figure 12. (a) Streamline and velocity magnitude and (b) pressure contour plots of the wing with a leading-edge flap of 20% of c, varied θo, and ψo = 40° at 0.25 t/T.
Figure 12. (a) Streamline and velocity magnitude and (b) pressure contour plots of the wing with a leading-edge flap of 20% of c, varied θo, and ψo = 40° at 0.25 t/T.
Jmse 12 00062 g012
Figure 13. Pressure coefficient on the wing surface during heaving motion for a flap length of 20% of c, ψo = 40°, and varied θo at 0.25 t/T.
Figure 13. Pressure coefficient on the wing surface during heaving motion for a flap length of 20% of c, ψo = 40°, and varied θo at 0.25 t/T.
Jmse 12 00062 g013
Figure 14. Average total power coefficient ( C ¯ P t )  for an oscillating wing with variations in flap length, θo, and Ψo.
Figure 14. Average total power coefficient ( C ¯ P t )  for an oscillating wing with variations in flap length, θo, and Ψo.
Jmse 12 00062 g014
Table 1. Physical model parameters of the oscillating wing energy harvester.
Table 1. Physical model parameters of the oscillating wing energy harvester.
DescriptionSymbolValue
Profile Rectangular plate
Widtht0.04c
Length of chordc1.0
Pivot pointxpc/3
Heaving amplitudeHo/c1
Pitching amplitudeθo75°
Frequencyf*0.14
Reynolds numberRe5 × 105
Phase angleϕ90o
Table 2. Analysis of mesh independence in the flat plate case.
Table 2. Analysis of mesh independence in the flat plate case.
Mesh in Stationary DomainMesh in Moving DomainTime Steps in a Cycle C ¯ P t %   Variation   in   C ¯ P t
D158.7 × 10324.5 × 103 T/20000.985
D288.2 × 10337.3 × 103T/10000.976
T/20000.963D1 to D2 = 2.28
T/40000.959
D3102.4 × 103 43.9 × 103T/20000.956D2 to D3 = 0.73
Table 3. Total power extraction coefficients for the oscillating wing at θo = 75°, associated with different ψo values, and the percentage increase in total power output compared to the wing without a flap.
Table 3. Total power extraction coefficients for the oscillating wing at θo = 75°, associated with different ψo values, and the percentage increase in total power output compared to the wing without a flap.
ψoPlate (0°)20°25°30°35°40°45°50°
C ¯ p y 0.9290.9530.970.9851.011.0561.0781.073
C ¯ p m 0.0340.0180.009−0.015−0.026−0.034−0.073−0.098
C ¯ p t 0.9630.9720.9790.970.9841.0221.010.974
Δ C ¯ p t (%)-0.91.60.72.26.14.81.1
η 0.3730.3770.3790.3760.3810.3960.3910.377
Table 4. Overall power coefficients for a wing with different flap lengths at θο = 75° and ψo = 40°.
Table 4. Overall power coefficients for a wing with different flap lengths at θο = 75° and ψo = 40°.
Plate10% of c15% of c20% of c33% of c40% of c
C ¯ p y 0.9291.0561.1121.1591.1241.110
C ¯ p m 0.034−0.034−0.102−0.166−0.242−0.259
C ¯ p t 0.9631.0221.010.9940.8800.852
Δ C ¯ p t   (%)-6.14.93.2−8.6−11.5
Table 5. Values of  C ¯ p y C ¯ pm, and  C ¯ P t  for oscillating wing equipped with leading-edge flap length 20% of c, ψo = 40°, and varied θo.
Table 5. Values of  C ¯ p y C ¯ pm, and  C ¯ P t  for oscillating wing equipped with leading-edge flap length 20% of c, ψo = 40°, and varied θo.
θo75°80°85°90°95°
C ¯ p y 1.1601.2071.2121.1631.071
C ¯ p m −0.166−0.139−0.115−0.056−0.017
C ¯ p t 0.9941.0681.0971.1071.054
Δ C ¯ p t (%)3.210.913.914.99.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alam, M.; Sohn, C.H. Enhancing the Performance of an Oscillating Wing Energy Harvester Using a Leading-Edge Flap. J. Mar. Sci. Eng. 2024, 12, 62. https://doi.org/10.3390/jmse12010062

AMA Style

Alam M, Sohn CH. Enhancing the Performance of an Oscillating Wing Energy Harvester Using a Leading-Edge Flap. Journal of Marine Science and Engineering. 2024; 12(1):62. https://doi.org/10.3390/jmse12010062

Chicago/Turabian Style

Alam, Maqusud, and Chang Hyun Sohn. 2024. "Enhancing the Performance of an Oscillating Wing Energy Harvester Using a Leading-Edge Flap" Journal of Marine Science and Engineering 12, no. 1: 62. https://doi.org/10.3390/jmse12010062

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop