Next Article in Journal
Two Gracilioethers Containing a [2(5H)-Furanylidene]ethanoate Moiety and 9,10-Dihydroplakortone G: New Polyketides from the Caribbean Marine Sponge Plakortis halichondrioides
Previous Article in Journal
Accident Probability Prediction and Analysis of Bus Drivers Based on Occupational Characteristics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aerodynamic Performance Assessment of Distributed Electric Propulsion after the Wing Trailing Edge

1
School of Mechanical Engineering and Automation, Fuzhou University, Fuzhou 350116, China
2
Key Laboratory of Fluid Power and Intelligent Electro-Hydraulic Control, Fuzhou University, Fuzhou 350116, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(1), 280; https://doi.org/10.3390/app14010280
Submission received: 25 October 2023 / Revised: 21 December 2023 / Accepted: 27 December 2023 / Published: 28 December 2023
(This article belongs to the Special Issue Application of Aerodynamics in Aerospace)

Abstract

:
Distributed electric propulsion (DEP) with four propellers distributed along the rear edge of the wing (pusher DEP configuration) promote aerodynamic interactions to a higher level. To study the aerodynamic performance of DEP with the rear wing through simulations and experiments, the multi-reference frame (MRF) with sliding grid is combined with wind tunnel tests. The obtained results demonstrate that the lift and drag of DEP increase with the angle of attack (AoA) and are related to the relative position of the propellers and wing. The propeller has no significant effect on the lift of the wing, and the lift and the AoA remain linear when the AoA is less than 16°. By contrast, the lift coefficient is much higher than the baseline (isolated wing), and the lift is greatly improved with the increasing drag when the AoA is greater than 16°. This is because the flow around the wing of the pusher configuration remains attached due to the suction of the inflow of the propeller on the trailing edge vortex. In addition, the acceleration effect on the free flow improves the kinetic energy of the airflow, which effectively delays the separation of the airflow in the slipstream region.

1. Introduction

With the advantage of reducing noise and improving energy conversion efficiency, electric propulsion aircraft have captured the public fancy [1]. Research on the notion of electric propulsion for future aircraft is characterized by Leading Edge Asynchronous Propellers Technology (LEAP Tech) [2]. For distributed electric propulsion (DEP), propellers are distributed spanwise along the rear wing, increasing the lift at the stage of taking off or landing [3,4]. By changing the size and spacing of the propeller, the balance of the system increases the dynamic pressure over the wing [5]. Specially, it reduces the effect of gusts due to the increase in relative speed on the wing and increases safety through redundancy [6]. Additionally, the propulsion oriented in the wing wake leads to the drag decreasing with an increasing propulsive efficiency [3].
The United States and Europe have already invested in ground testing and flight tests related to electric propulsion [2,6,7,8,9], including LEAP Tech [1,10], X-57 [11], GL-10 [10], DEP propeller [12], and so on. Veldhuis, L.L.M. [13] and Rakshith, B. et al. [14] performed numerical simulations on the structure design with various parameters concerning the propeller/wing configuration. Patterson, M.D. et al. [15] designed the X-57 Maxwell with high-lift propellers and discussed the advantages and challenges of the high-lift propeller (HLP) at low speeds. Ananda Krishnan, G.K. et al. [16] compared the aerodynamic characteristics of the SR22 civil aircraft with a LEAP concept aircraft, demonstrating that the distributed propulsion system can reduce drag. An unmanned aerial vehicle (UAV) equipped with a DEP system is advantageous in a low-speed flight on a smaller scale. However, its unique flow characteristics pose the complexity of studying the aerodynamic performance of the DEP with numerical simulations and experiments [17].
Current studies involving DEP are mainly focused on simulating the effects of propellers on the wing or fuselage for large aircraft such as transport aircraft and solar aircraft. Patterson, M.D. et al. [15] at the Georgia Institute of Technology and Borer, N.K. et al. [18] at NASA conducted preliminary aerodynamic studies on the distributed propeller propulsion system. The vortex lattice calculation method only considered the influence of the slipstream on the wing. Simulations for mutual aerodynamic interference are limited to the analysis the flow field on the wing surface. Patterson, M.D. et al. [19] studied LEAP tech to analyze the aerodynamic interference between the propeller and the wing. Wang, J. et al. [20] showed the influence of wind field disturbance on the stability of a multirotor UAV with a dynamic model. Currently, playing a revolutionary role in the future of aviation, distributed propulsion technology involves the induced velocity of the propeller at each position of the wing set as the far field condition; the aerodynamic force of the wing is calculated separately [21,22,23,24]. Thus, it is not enough to use only one coupling method to study the aerodynamic interference of wings and propellers due to flow interactions.
Above all, considering that the complex low Reynolds aerodynamic environment imposes difficulties for the pusher configuration of the DEP system, the stronger rotor interferences and vortex movement in the outflow need verification through both simulations and experiments. Also, the strict design requirement for the location of the propeller requires knowing the effect of the horizontal spacing and vertical spacing, which is important for further flight tests. Eventually, it has a much wider reach than results from references that include only numerical simulations or experiments.
In Section 2, the pusher DEP configuration is presented with variables. In Section 3, lift and drag are obtained through wind tunnel testing. In Section 4, numerical simulations on the propulsion system are performed with FLUENT. Finally, Section 5 provides the conclusions.

2. Theoretical Analysis

As shown in Figure 1, DEP can be designed as a pusher configuration according to the relative positions of the propeller and the wing.
There are four small propellers equally distributed along the rear wing. The direction of rotation of the propeller is shown in Figure 1. Therefore, the propeller and the wing in the DEP interact with each other. This section analyzes the aerodynamics of the propeller and wing. In Figure 1, D r p is the dimensionless horizontal distance from the center of rotation of the propeller to the rear edge of the wing (where c is the chord length of the propeller). X r   ( x / R ) is the dimensionless vertical distance from the center of the propeller to the chord plane of the wing. The positive value indicates that the propeller is above the chord plane.
For the isolated rotor without interference, the lift and the drag coefficients are as follows [25]:
C l = 2 L ρ V 2 S
C d = 2 D ρ V 2 S
where L is lift, D is drag, ρ is the density of air, V is the rotational velocity, and S is the reference area.
To characterize the aerodynamic interference between the propeller and the wing in the DEP system, the lift coefficient and the drag coefficient are applied to obtain the variation in the rotor interference [26].

3. Wind Tunnel Testing

3.1. Experimental Setup

A test bench was constructed as shown in Figure 2.
Four propellers were individually controlled by four motors and measured with a tachometer to determine the rotor speed. An ATI Gamma F&T sensor was connected on the wing to perform the force measurement. The DEP configuration is shown in Table 1.

3.2. Error Analysis

The main sources of error in these experiments are the standard deviations of the rotational speed and the mean voltages from the force sensors. Typical values of the standard deviations of force are about 1% of the mean values. Rotational speed error is about 2.5 RPM per second. The statistical (random) error in the coefficient values for each run was estimated using the standard deviation of the repeated samples. This is a valid operation because, even though the dimensional values of the test variables are expected to vary from one run to another, the coefficient values are expected to be the same. To determine accuracy of the test measurements, all the key parameters were derived using the Kline–McClintok method [27]. Based on collected data, the partial derivatives computed can then calculate an estimate of systematic error. This was performed for Cl and Cd. Low uncertainty in all measured variables in all tests increases the confidence in the data and demonstrates a good measurement system. The values of uncertainty that are presented in this study were all calculated with 95% confidence levels.

3.3. Experimental Results

Figure 3 shows the distribution of Cl and Cd with different rotor locations, where D r p is the horizontal position, X r is the vertical offset, and ‘Baseline’ is the isolated wing without interference from any rotor.
It can be seen that the lift coefficient and drag coefficient increase linearly with the AoA in horizontal positions. The propeller has no significant effect on the lift of the wing when the AoA is less than 16°. It also can be seen that the drag coefficient of DEP with the pusher structure almost coincides with the baseline. However, the lift coefficient curve begins to show a decreasing trend when the AoA is higher than 16°. Specially, the lift coefficient of the DEP configuration is significantly higher than the baseline, which indicates that the lift is greatly improved. Also, the drag coefficient is also significantly increased in this case. Considering the effect of the horizontal position, as shown in Figure 3a,c, it can be seen that the lift and drag coefficients increased with the decreased horizontal distance. Additionally, the vertical offset seems have a greater impact on the aerodynamic performance of the wing. As a result, both the lift coefficient and drag coefficient of the wing gradually increase with the effect of the propeller. For the configuration with X r = 1 ,     D r p = 0.1 c , it is interesting to note that the lift coefficient obtained a maximum at 30° with an increasing tendency.
To determine the final performance of the DEP, Figure 4 presents the distribution of the L/D with different DEP configurations.
The lift is much improved at a higher AoA for the pusher configuration of the DEP, with a consistent tendency for a lower AoA. The effect of the horizontal position is more obvious for the performance of the wing, where the outflow is deformed by the propellers. Although L/D variation is within the experimental uncertainty limits with no significant difference inferred, delay in stall and increase in lift at a higher AoA are observed when the pusher props are placed above the wing chord line. Thus, it is important to study the flow field with a higher AoA.

4. Numerical Simulations

4.1. Simulation Setup

This step was taken to figure out the aerodynamic interference between the wing and propellers located at the rear of the wing as a pusher DEP. The geometry parameters are shown in Table 2. For the simulation setup, the RANS model was applied to analyze the characteristics of the external flow field. Also, mesh refinement was performed on the propeller tip to capture the gradient of the physical field flow. In addition, the finite volume method was used to discrete the differential equations. Considering the low Reynolds number, the turbulence model k-w was selected to obtain the flow field of the DEP. Pressure correction and interpolation were performed using the semi-implicit method (SIMPLE) algorithm with the standard format. Momentum, energy equation, and turbulent viscosity are all in the second-order upwind discrete schemes. Finally, the sliding grid was used to deal with the interaction between the rotating and the stationary regions, and the multiple reference frame (MRF) method was applied as the initial condition for the slip mesh transient calculation. The mesh distribution is shown in Figure 5.
Four different mesh grids were used to verify the grid independence, as shown in Table 3.
It can be found that the error decreased with the increasing grids, and it is stable for the grid around 25 ×   10 5 , which was applied for the following simulations.

4.2. Simulation Result

To figure out the increased lift for a higher AoA, as presented in the experiments, as shown in Figure 6, we selected the velocity distribution at the 43% span section of pusher structures for an AoA ranging from 20° to 30°.
For the horizontal position at D r p = 0.1 c , by changing the vertical offset of the propeller, it can be found that there is a significant difference in the inflow of the propeller where the upper inflow is mixed with the outflow of the wing. For the flow field near the trailing edge of the propeller, a deflection was found in the inflow direction of the propeller at X r = 0.5 with the increasing velocity. It was more obvious for the propeller mounted above the paddle. On the contrary, the interaction with the lower wing was weak. For the two pusher structures at a 30° AoA with a distance of 0.1 c from the trailing edge and a vertical offset of X r = 0.5 and X r = 1 , the airflow through the wing interacted with the rotor flow field, creating a stabilized leading edge vortex [28]. The vortex at X r = 0.5 interacted with the inflow of the propeller, and the vortex at X r = 1 began to affect the propeller’s tip vortex and down wash flow, forming another smaller vortex behind the paddle. Therefore, when the trailing edge of the wing approached the tip of the blade, the eddy current action was more intense, and the X r = 1 with a large vertical offset weakened the eddy current intensity above the wing, thereby increasing the lift.
Figure 7 shows a pressure contour for DEP structures ranging from 20° to 30°.
The pressure difference in the leading edge of the wing increased slightly with the positions for a lower AoA. Thus, the lift of the DEP system did not increase significantly. For AoA = 30°, the negative pressure area increased significantly, which directly increased wing lift and drag. In Figure 7c, it can be seen that the increment for the negative pressure on the wing reached the maximum when the distance from the propeller center to the rear edge was 0.1 c to 0.15 c . Within a certain range, the small change in the flow direction of the propeller in the DEP has no significant effect on the aerodynamic performance. Similarly, the vertical installation position of the propeller significantly changes the DEP flow field distribution. When the central axis of the propeller is above the wind chord, the negative pressure region of the wing decreases. In short, the aerodynamic performance of the pusher DEP structure is affected by the vertical installation position of the propeller. Specifically, the vertical offset from the bottom improves the increment in the lift of the wing.
Figure 8 shows the vorticity distribution of the pusher configuration at 20° and 30°.
It can be seen that when the propeller is installed above the chord, the tip vortex of the wing acts on the propeller and merges with the propeller downwash flow. Also, the vortex of the wing interacts with the propeller vortex, where the trailing edge vortex is more obvious with the propeller vortex system at X r = 1 . This indicates that the interference between the propeller and the wing may be advantageous to improve the performance of the DEP.
Figure 9 shows the pressure distribution at 43% spanwise.
In Figure 9a, it can be seen that the streamline changed with the pusher structure, which also affected the pressure distribution. The negative pressure of the trailing edge of the wing at 0.1 c was slightly higher than that at 0.15 c , indicating that the change along the flow direction is advantageous to improve the aerodynamic interference and flight duration for a hover state. Figure 9b shows the pressure of the DEP with different vertical offsets at 30°. The pressure for X r = 0.5 is affected by the upper airfoil vortex with a sudden change. The pressure at the middle section is convexly distributed, with a slightly higher positive pressure than X r = 1 . For a larger vertical offset X r = 1 , the pressure distribution is relatively stable, with a higher maximum than the X r = 0.5 , which indicated that this structure has better lift performance.
Figure 10 shows the lift distribution of propellers for the pusher DEP configuration.
It can be seen that propellers 1 and 3 obtained less lift, and propeller 2 was less affected by the wing tip vortex or the adjacent propellers with a larger lift. Propeller 4 was totally emerged in the wing tip vortex, with an increasing force variation. In addition, lift variation in X r = 0 ,   D r p = 0.1 c , and X r = 0 ,   D r p = 0.15 c was completely consistent, and X r = 0.5 ,   D r p = 0.1 c , and X r = 1 ,   D r p = 0.1 c also showed a similar tendency. However, X r = 0 ,   D r p = 0.1 c and X r = 0.5 ,   D r p = 0.1 c showed better performance. This further verifies that the closer the propeller is, the more interference there is. From the vertical offset, the propeller is mounted above the trailing edge string for better lift performance.

5. Conclusions

This paper proposed a pusher DEP configuration to study the aerodynamic performance of the mutual interference between the propeller and wing with wind tunnel tests and numerical simulations. Conclusions are as follows:
(1)
The propeller has no significant effect on the lift of the wing when the AoA is less than 16°. However, the lift coefficient is much higher than the baseline (isolated wing), and the lift of the pusher DEP system is significantly increased up to 73–87% without any decrease in L/D. In contrast, the drag coefficient is also higher than that of the isolated wing. With a higher AoA, it is interesting to note that the lift for the pusher configuration was also improved.
(2)
The wing of the pusher structure stalled at 16° AoA. This may be caused by the acceleration of the axial airflow affected by the propeller and the interference from the inflow and downwash caused by the slipstream. Accordingly, it increased the drag coefficient, especially for the DEP with a smaller horizontal distance between the propeller and the trailing edge of the wing. Also, a delay in stall and an increase in lift at a high AoA were observed when the pusher props were placed above the wing chord line.
(3)
The trailing edge vortex of the wing for the pusher DEP structure interacted with the flow field of the propeller, and it showed an increase in lift and drag increase. L/D variation was within the experimental uncertainty limits of the baseline case (isolated wing), and no significant difference was found. The propeller in the pusher DEP was subjected to the blocking effect of the wing or the eddy current of the trailing edge of the wing, and the lift was significantly increased.

Author Contributions

Y.L. carried out experiments; X.Z. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (Grant No. 52275095) and Fujian Provincial Industrial Robot Basic Components Technology Research and Development Center (2014H2004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are already included in the manuscript.

Acknowledgments

The authors thank the Key Laboratory of Fluid Power and Intelligent Electro-Hydraulic Control (Fuzhou University), Fujian Province University, and Fuzhou University Jinjiang Science and Education Park for applying the experimental field.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Borer, N.K.; Patterson, M.D.; Viken, J.K.; Moore, M.D.; Bevirt, J.; Stoll, A.M.; Gibson, A.R. Design and performance of the NASA SCEPTOR distributed electric propulsion flight demonstrator. In Proceedings of the 16th AIAA Aviation Technology, Integration, and Operations Conference, Washington, DC, USA, 13–17 July 2016; p. 3920. [Google Scholar]
  2. Stoll, A.M.; Bevirt, J.; Moore, M.D.; Fredericks, W.J.; Borer, N.K. Drag reduction through distributed electric propulsion. In Proceedings of the 14th AIAA Aviation Technology, Integration, and Operations Conference, Atlanta, GA, USA, 16–20 June 2014; p. 2851. [Google Scholar]
  3. Moore, K.R.; Ning, A. Distributed electric propulsion effects on existing aircraft through multidisciplinary optimization. In Proceedings of the 2018 AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Kissimmee, FL, USA, 8–12 January 2018; p. 1652. [Google Scholar]
  4. Kong, X.; Zhang, Z.; Lu, J.; Li, J.; Yu, L. Review of electric power system of distributed electric propulsion aircraft. Acta Aeronaut. Et Astronaut. Sin. 2018, 39, 46–62. [Google Scholar]
  5. Moore, M.D. Misconceptions of electric aircraft and their emerging aviation markets. In Proceedings of the 52nd Aerospace Sciences Meeting, National Harbor, MD, USA, 13–17 January 2014; p. 0535. [Google Scholar]
  6. Leifsson, L.; Ko, A.; Mason, W.H.; Schetz, J.A.; Grossman, B.; Haftka, R.T. Multidisciplinary design optimization of blended-wing-body transport aircraft with distributed propulsion. Aerosp. Sci. Technol. 2013, 25, 16–28. [Google Scholar] [CrossRef]
  7. Juvé, L.; Fosse, J.; Joubert, E.; Fouquet, N. Airbus Group electrical aircraft program, the E-FAN project. In Proceedings of the 52nd AIAA/SAE/ASEE Joint Propulsion Conference, Salt Lake City, UT, USA, 25–27 July 2016; p. 4613. [Google Scholar]
  8. Christie, R.J.; Dubois, A.; Derlaga, J.M. Cooling of Electric Motors Used for Propulsion on SCEPTOR; NASA: Washington, DC, USA, 2017. [Google Scholar]
  9. Rothhaar, P.M.; Murphy, P.C.; Bacon, B.J.; Gregory, I.M.; Grauer, J.A.; Busan, R.C.; Croom, M.A. NASA Langley distributed propulsion VTOL tiltwing aircraft testing, modeling, simulation, control, and flight test development. In Proceedings of the 14th AIAA Aviation Technology, Integration, and Operations Conference, Atlanta, GA, USA, 16–20 June 2014; p. 2999. [Google Scholar]
  10. Fredericks, W.J.; McSwain, R.G.; Beaton, B.F.; Klassman, D.W.; Theodore, C.R. Greased Lightning (GL-10) Flight Testing Campaign; NASA: Washington, DC, USA, 2017. [Google Scholar]
  11. Schnulo, S.L.; Chin, J.; Smith, A.; Paul-Dubois-Taine, A. Steady state thermal analyses of SCEPTOR X-57 wingtip propulsion. In Proceedings of the 17th AIAA Aviation Technology, Integration, and Operations Conference, Denver, CO, USA, 5–9 June 2017; p. 3783. [Google Scholar]
  12. Traub, L.W. Propeller characterization for distributed propulsion. J. Aerosp. Eng. 2021, 34, 04021020. [Google Scholar] [CrossRef]
  13. Veldhuis, L.L.M. Review of propeller-wing aerodynamic interference. In Proceedings of the 24th lnternational Congress of Aeronautical Sciences, Yokohama, Japan, 29 August–3 September 2004. [Google Scholar]
  14. Rakshith, B.; Deshpande, S.; Narasimha, R.; Praveen, C. Optimal low-drag wing planforms for tractor-configuration propeller-driven aircraft. J. Aircr. 2015, 52, 1791–1801. [Google Scholar] [CrossRef]
  15. Patterson, M.D.; Borer, N.K. Approach considerations in aircraft with high-lift propeller systems. In Proceedings of the 17th AIAA Aviation Technology, Integration, and Operations Conference, Denver, CO, USA, 5–9 June 2017; p. 3782. [Google Scholar]
  16. Ananda Krishnan, G.K.; Deters, R.W.; Selig, M.S. Propeller induced flow effects on wings at low Reynolds numbers. In Proceedings of the 31st AIAA Applied Aerodynamics Conference, San Diego, CA, USA, 24–27 June 2013; p. 3193. [Google Scholar]
  17. Zhou, Z.; Gan, W. Studying aerodynamic performances of the low-Reynolds-number airfoil of solar energy UAV. J. Northwestern Polytech. Univ. 2014, 32, 163–168. [Google Scholar]
  18. Borer, N.K.; Moore, M.D. Integrated propeller-wing design exploration for distributed propulsion concepts. In Proceedings of the 53rd AIAA Aerospace Sciences Meeting, Kissimmee, FL, USA, 5–9 January 2015; p. 1672. [Google Scholar]
  19. Patterson, M.D.; German, B. Wing aerodynamic analysis incorporating one-way interaction with distributed propellers. In Proceedings of the 14th AIAA Aviation Technology, Integration, and Operations Conference, Atlanta, GA, USA, 16–20 June 2014; p. 2852. [Google Scholar]
  20. Wang, J.; Yang, J.; Yang, Z. Dynamics Modeling and Simulation of Multi-rotor UAV based on the Composite Wind Field Model. In Proceedings of the 2022 13th International Conference on Reliability, Maintainability, and Safety (ICRMS), Kowloon, Hong Kong, 21–24 August 2022; pp. 127–134. [Google Scholar]
  21. Gohardani, A.S. A synergistic glance at the prospects of distributed propulsion technology and the electric aircraft concept for future unmanned air vehicles and commercial/military aviation. Prog. Aerosp. Sci. 2013, 57, 25–70. [Google Scholar] [CrossRef]
  22. Seddon, J.M.; Newman, S. Basic Helicopter Aerodynamics; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  23. Luo, J.; Zhu, L.; Yan, G. Novel quadrotor forward-flight model based on wake interference. AIAA J. 2015, 53, 3522–3533. [Google Scholar] [CrossRef]
  24. Leishman, G.J. Principles of Helicopter Aerodynamics with CD Extra; Cambridge University Press: Cambridge, UK, 2006. [Google Scholar]
  25. Patterson, M.D.; Daskilewicz, M.J.; German, B. Simplified aerodynamics models to predict the effects of upstream propellers on wing lift. In Proceedings of the 53rd AIAA Aerospace Sciences Meeting, Kissimmee, FL, USA, 5–9 January 2015; p. 1673. [Google Scholar]
  26. Lei, Y.; Wang, J. Aerodynamic performance of quadrotor UAV with non-planar rotors. Appl. Sci. 2019, 9, 2779. [Google Scholar] [CrossRef]
  27. Lei, Y.; Wang, H. Aerodynamic optimization of a micro quadrotor aircraft with different rotor spacings in hover. Appl. Sci. 2020, 10, 1272. [Google Scholar] [CrossRef]
  28. Nguyen, D.H.; Liu, Y.; Mori, K. Experimental study for aerodynamic performance of quadrotor helicopter. Trans. Jpn. Soc. Aeronaut. Space Sci. 2018, 61, 29–39. [Google Scholar] [CrossRef]
Figure 1. The pusher configuration of the DEP system: (a) 3D view; (b) side view.
Figure 1. The pusher configuration of the DEP system: (a) 3D view; (b) side view.
Applsci 14 00280 g001
Figure 2. Experimental setup: (a) sketch; (b) models.
Figure 2. Experimental setup: (a) sketch; (b) models.
Applsci 14 00280 g002
Figure 3. Distribution of lift coefficient (Cl) and drag coefficient (Cd). (a) Cl in various horizontal positions; (b) Cl in various vertical offsets; (c) Cd in various horizontal positions; (d) Cd in various vertical offsets.
Figure 3. Distribution of lift coefficient (Cl) and drag coefficient (Cd). (a) Cl in various horizontal positions; (b) Cl in various vertical offsets; (c) Cd in various horizontal positions; (d) Cd in various vertical offsets.
Applsci 14 00280 g003
Figure 4. Distribution of L/D: (a) horizontal position; (b) vertical position.
Figure 4. Distribution of L/D: (a) horizontal position; (b) vertical position.
Applsci 14 00280 g004
Figure 5. The mesh distribution: (a) face mesh; (b) side view of cell.
Figure 5. The mesh distribution: (a) face mesh; (b) side view of cell.
Applsci 14 00280 g005
Figure 6. Velocity distribution at 43% spanwise. (a) X r = 0 , D r p = 0.1 c , AoA = 20°; (b) X r = 0 , D r p = 0.1 c , AoA = 24°; (c) X r = 0 , D r p = 0.1 c , AoA = 30°; (d) X r = 0 ,   D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 , D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 ,   D r p = 0.1 c , AoA = 30°.
Figure 6. Velocity distribution at 43% spanwise. (a) X r = 0 , D r p = 0.1 c , AoA = 20°; (b) X r = 0 , D r p = 0.1 c , AoA = 24°; (c) X r = 0 , D r p = 0.1 c , AoA = 30°; (d) X r = 0 ,   D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 , D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 ,   D r p = 0.1 c , AoA = 30°.
Applsci 14 00280 g006aApplsci 14 00280 g006b
Figure 7. Pressure distribution. (a) X r = 0 , D r p = 0.1 c , AoA = 20°; (b) X r = 0 ,   D r p = 0.1 c , AoA = 24°; (c) X r = 0 , D r p = 0.1 c , AoA = 30°; (d) X r = 0 ,   D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 ,   D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 , D r p = 0.1 c , AoA = 30°.
Figure 7. Pressure distribution. (a) X r = 0 , D r p = 0.1 c , AoA = 20°; (b) X r = 0 ,   D r p = 0.1 c , AoA = 24°; (c) X r = 0 , D r p = 0.1 c , AoA = 30°; (d) X r = 0 ,   D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 , D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 ,   D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 , D r p = 0.1 c , AoA = 30°.
Applsci 14 00280 g007
Figure 8. Vorticity distribution. (a) X r = 0 ,   D r p = 0.1 c , AoA = 20°; (b) X r = 0 , D r p = 0.1 c , AoA = 24°; (c) X r = 0 ,   D r p = 0.1 c , AoA = 30°; (d) X r = 0 , D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 , D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 , D r p = 0.1 c , AoA = 30°.
Figure 8. Vorticity distribution. (a) X r = 0 ,   D r p = 0.1 c , AoA = 20°; (b) X r = 0 , D r p = 0.1 c , AoA = 24°; (c) X r = 0 ,   D r p = 0.1 c , AoA = 30°; (d) X r = 0 , D r p = 0.15 c , AoA = 20°; (e) X r = 0.5 ,   D r p = 0.1 c , AoA = 20°; (f) X r = 0.5 , D r p = 0.1 c , AoA = 20°; (g) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (h) X r = 0.5 ,   D r p = 0.1 c , AoA = 24°; (i) X r = 0.5 , D r p = 0.1 c , AoA = 30°; (j) X r = 1 ,   D r p = 0.1 c , AoA = 20°; (k) X r = 1 , D r p = 0.1 c , AoA = 30°.
Applsci 14 00280 g008
Figure 9. Pressure distribution along chordwise direction at a 43% span section. (a) Xr = 0, 0.1c and Xr0, 0.15c; (b) Xr0.5, 0.1c and Xr1, 0.1c, AoA = 30°.
Figure 9. Pressure distribution along chordwise direction at a 43% span section. (a) Xr = 0, 0.1c and Xr0, 0.15c; (b) Xr0.5, 0.1c and Xr1, 0.1c, AoA = 30°.
Applsci 14 00280 g009
Figure 10. Lift distribution.
Figure 10. Lift distribution.
Applsci 14 00280 g010
Table 1. Experimental parameters.
Table 1. Experimental parameters.
PartParameterNumerical Value
Wind tunnel cross-section-600 mm × 600 mm
NACA4415 airfoil used for the wingSpan300 mm
Chord150 mm
Aspect Ratio2
Four propellersDiameter76.2 mm
Pitch76.2 mm
No. of blades4
Operating conditionsFree-stream speed18 m/s (typical)
Prop RPM18,000 rpm
Wing Reynolds number180,000
Table 2. Simulation parameters.
Table 2. Simulation parameters.
PartParameterNumerical Value
Simulation calculation domain-600 mm × 600 mm
NACA4415 airfoil used for the wingSpan300 mm
Chord150 mm
Aspect Ratio2
Four propellersDiameter76.2 mm
Pitch76.2 mm
No. of blades4
Boundary conditionsSpeed inlet18 m/s (typical)
Rotation domain RPM18,000 rpm
Table 3. Grid independence for X r = 0 ,   D r p = 0.15 c , and AoA = 20°.
Table 3. Grid independence for X r = 0 ,   D r p = 0.15 c , and AoA = 20°.
Rotated   Region   Grid   ( × 10 5 ) Total   Grid   ( × 10 5 ) ClCl Error (%)CdCd Error (%)
1.8985.470.95−10.20.44822+5.8
2.14107.510.987−7.90.43678+3.1
2.55133.641.013−5.50.43466+2.6
2.83167.291.014−5.40.43424+2.5
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

Lei, Y.; Zhao, X. Aerodynamic Performance Assessment of Distributed Electric Propulsion after the Wing Trailing Edge. Appl. Sci. 2024, 14, 280. https://doi.org/10.3390/app14010280

AMA Style

Lei Y, Zhao X. Aerodynamic Performance Assessment of Distributed Electric Propulsion after the Wing Trailing Edge. Applied Sciences. 2024; 14(1):280. https://doi.org/10.3390/app14010280

Chicago/Turabian Style

Lei, Yao, and Xiangzheng Zhao. 2024. "Aerodynamic Performance Assessment of Distributed Electric Propulsion after the Wing Trailing Edge" Applied Sciences 14, no. 1: 280. https://doi.org/10.3390/app14010280

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