Arfmtsv96 N1 P82 95

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Advanced Research in Fluid Mechanics and Thermal Sciences 96, Issue 1 (2022) 82-95

Journal of Advanced Research in Fluid


Mechanics and Thermal Sciences
Journal homepage:
https://semarakilmu.com.my/journals/index.php/fluid_mechanics_thermal_sciences/index
ISSN: 2289-7879

CFD Analysis on Propeller at Varying Propeller Disc Angle and Advance


Ratio
Leong Chee Hang1, Nur Athirah Nadwa Rosli1, Mastura Ab Wahid1,*, Norazila Othman1, Shabudin
Mat1, Mohd Zarhamdy Md Zain2

1
Department of Aeronatical, Automotive and Offshore Engineering, Faculty of Mechanical Engineering, Universiti Teknologi Malaysia, 81310
Skudai, Johor, Malaysia
2 Department of Applied Mechanics and Design Engineering, Faculty of Mechanical Engineering, Universiti Teknologi Malaysia, 81310 Skudai,
Johor, Malaysia

ARTICLE INFO ABSTRACT

Article history: The purpose of this study is to see the feasibility of using Computational Fluid Dynamic
Received 20 February 2022 (CFD) analysis over wind tunnel testing for propeller performance measurement.
Received in revised form 5 May 2022 Computational Fluid Dynamic (CFD) analysis is conducted on APC 6X4E propeller and
Accepted 10 May 2022 9X6E propeller at propeller disc angle, α of 0°, 30°, 60°, and 90° using commercially
Available online 9 June 2022
available software ANSYS FLUENT at advance ratio ranging from 0 to 0.88 and angular
velocity ranging from 1000rpm to 8000rpm. CFD analysis was also performed for different
advance ratio at different rotational speed of 4000rpm and 8000rpm. Multiple Reference
Frame model is used to simulate the rotating propeller in a flowing airstream by
constructing a rotating and static domain around the propeller. The non-zero propeller
disc angle is achieved by changing the inlet direction of the static domain. The SST k-ω
turbulence model is used, and tetrahedral mesh is constructed. The propeller thrust and
torque obtained from the CFD simulation is used to calculate the aerodynamic
characteristics of the propellers. The thrust coefficient, torque coefficient and propeller
efficiency obtained for both propellers follow the trend of the wind tunnel testing. The
results obtained from the CFD simulation matches with the results trend obtained by
another researcher performing wind tunnel analysis, where the thrust coefficient
decreases with increasing advance ratio at propeller disc angle less than 60° and increases
Keywords: with increasing advance ratio at propeller disc angle more than 70°. The error produce
Non-zero propeller disc angle; APC for thrust estimation is lower than 12% for both 6x4E and 9x6E propellers. The error for
propeller; ANSYS FLUENT; SST k-ω torque and efficiency estimation is between 15-30% and 12% respectively. In conclusion,
model; multiple reference frame CFD simulation can predict the aerodynamic characteristics of Low Reynolds Number
(MRF) propeller at different propeller disc angle.

1. Introduction

Before the rise of quadrotor and multirotor UAV’s, past research on Low Reynolds Number
propeller were mainly focused on the aerodynamic characteristics of the propeller when it is parallel
to the free stream or the propeller disc angle of attack, α p = 0°. Due to the unique propeller

*
Corresponding author.
E-mail address: [email protected]

https://doi.org/10.37934/arfmts.96.1.8295

82
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

configuration for quadrotor and multirotor UAV’s where the propeller is perpendicular to the
freestream and the maneuver by tilting the propeller, the aerodynamic performance of the propeller
at different propeller disc angle of attack is essential for the performance of the UAV’s.
While majority of research on aerodynamic characteristics of propellers at different propeller disc
angle of attack focuses on large propellers, there are a few research on Low Reynolds Number
propeller at different propeller disc angle using wind tunnel testing method. Serrano et al., [1]
conducted an investigation to determine the aerodynamic performance of four propeller with
diameter 12-inch with propeller disc angle of attack ranging from 0o to 90o and advance ratio ranging
from 0 to 0.55 [1]. Based on the researches, the propeller thrust increased with as the propeller disc
angle of attack increased, while the power consumption shows low sensitivity to the changes of the
propeller disc angle of attack [1]. Majority of the research on aerodynamic characteristics of Low
Reynolds Number propellers are using the wind tunnel testing method and there is lack of research
on that topic using CFD analysis. Therefore, this study aims to validate the use of CFD analysis in
obtaining the aerodynamic characteristics of Low Reynolds Number propeller at different propeller
disc angle of attack. Study from Mohamed et al., [15] researched on slotted airfoil. Based on the
study, introducing slotted airfoil for low Reynolds number propeller reduces the efficiency of the
propeller.
Reynolds-averaged approach is used to solve the turbulence flow simulation as it is robust,
economical, and relatively accurate. This approach is used by many commercially available software
such as ANSYS Fluent. In Reynolds-average approach, selection of turbulence model is an important
and it influence the simulation results accuracy. The selection of turbulence model depends on the
type of simulation. For turbomachinery, the common turbulence model used are Spalart-Allmaras (S-
A) model, and k-ε model. For example, Seeni et al., [2] conducted an CFD simulation on APC10x7SF
propeller using Spalart-Allmaras (S-A) turbulence model and validated the result by comparing it to
the experimental result. This study will aim to conduct CFD simulation on propeller at varying
propeller disc angle and obtain the propeller aerodynamic performance.

2. Methodology
2.1 K-ω Turbulence Model

The SST k-ω is proposed by Menter [7] by combining the standard k-ε model and standard k-ω.
This model uses a cross-diffusion term in the transport equation for specific dissipation rate, ω. This
enables the model to predict the flow at near-wall location using the k-ω turbulence model and
predict the flow far away from wall using standard k-ε turbulence model [3]. The transport equation
for turbulent kinetic energy, 𝜅 is represented by

𝑅
𝜕𝜅 𝜕 𝜕 𝜐 𝜕𝜅 𝜏̅𝑖𝑗 ̅𝑖
𝜕𝑢
+ 𝜕𝑥 (𝑢̅𝑖 𝜅) = 𝜕𝑥 (𝜐 + 𝜎𝑇 ) 𝜕𝑥 + −𝜀 (1)
𝜕𝑡 𝑖 𝑖 𝑘 𝑖 𝜌 𝜕𝑥𝑗

𝑅
where 𝜏̅𝑖𝑗 is the Reynolds stress tensor, ρ is the density, and 𝜎𝑘 is the diffusion Prandtl number for
turbulent kinetic energy. The transport equation for turbulence energy dissipation rate, ε is
represented by

𝑅
𝜕𝜀 𝜕 𝜕 𝜐 𝜕𝜀 𝜀 𝜏̅𝑖𝑗 𝜕𝑢
̅𝑖 𝜀2
+ 𝜕𝑥 (𝑢̅𝑖 𝜀) = 𝜕𝑥 (𝜐 + 𝜎𝑇 ) 𝜕𝑥 + 𝐶𝜀1 𝜅 − 𝐶𝜀2 (2)
𝜕𝑡 𝑖 𝑖 𝜀 𝑖 𝜌 𝜕𝑥𝑗 𝜅

83
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

where 𝐶𝜀1 = 1.44, 𝐶𝜀2 = 1.92, 𝜎𝜀 is the diffusion Prandtl number for isotropic turbulence energy
dissipation rate and is equal to 1.3 [4]. Meanwhile the eddy viscosity for k-ε model is expressed as

𝜅2
𝜐𝑇 = 𝐶𝜇 (3)
𝜀

where Cμ = 0.09.

The transport equation for specific dissipation rate, ω is expressed as

𝜕𝜌𝜔 𝜕 𝜕 𝜕𝜔 𝛾 𝜌𝜎𝜔2 𝜕𝜅 𝜕𝜔
+ 𝜕𝑥 (𝜌𝜔𝑢𝑖 ) = 𝜕𝑥 [(𝜇 + 𝜎𝜔 𝜇 𝑇 ) 𝜕𝑥 ] + 𝜐 𝑃 − 𝛽𝜌𝜔2 + 2(1 − 𝐹1) (4)
𝜕𝑡 𝑖 𝑗 𝑗 𝑇 𝜔 𝜕𝑥𝑗 𝜕𝑥𝑗

where the turbulent eddy viscosity is represented by


𝜌𝑎 𝜅
𝜇 𝑇 = max⁡(𝑎 1𝜔,Ω𝐹 ) (5)
1 2

The function F1 and F2 are given by

𝐹1 = tanh(𝑎𝑟𝑔14 ) (6)

√𝜅 500𝜐 4𝜌𝜎𝜔2 𝜅
𝑎𝑟𝑔1 = 𝑚𝑖𝑛 [(𝛽∗𝜔𝑑 , 𝑑2 𝜔 ) , 𝐶𝐷 2
⁡] (7)
𝜅𝜔 𝑑

1 𝜕𝜅 𝜕𝜔
𝐶𝐷𝜅𝜔 = max (2𝜌𝜎𝑤2⁡ 𝜔 𝜕𝑥 , 10−20 ) (8)
𝑗 𝜕𝑥𝑗

𝐹2 = tanh(𝑎𝑟𝑔22 ) (9)

√𝜅 500𝜐
𝑎𝑟𝑔2 = max (2 𝛽∗𝜔𝑑 , 𝑑2 𝜔 ) (10)

The constants used is a1 = 0.31, 𝜅 = 0.41, β* = 0.09 and σω2 = 0.856. The SST k-ω turbulence model
is a blend of k-ε model and k-ω model. This eliminate the weakness for k-ε model which is inaccurate
prediction at near wall location and the weakness of k-ω model which is too sensitive to the
freestream turbulence condition. This model gives better result with flow that involves separation of
flow. However, the weakness of this model is it requires higher computation power due to the
additional function F1 and F2.

2.2 Domain Modelling

Two propellers with different diameter and pitch which is APC 6X4E propeller and APC 9X6E
propeller are purchased and to be measured physically. The aerofoil of the propeller blade consists
of low Reynolds number Eppler E63 aerofoil blended with a Clark-Y similar aerofoil at the tip of the
propeller blade. Figure 1 shows the chord and twist distribution for the propellers.

84
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

(a) (b)
Fig. 1. (a) Top and (b) side view sketch of 9x6E and 6X4E propeller

The multiple reference frame (MRF) model approach is used. This approach is widely used for the
analysis for turbomachinery that consist of rotating part and a fixed part [2,5,6]. To imitate the flow
in wind tunnel testing, the domain of the flow is separated into stationary domain and rotating
domain. The rotating domain is a small cylinder that enclose the propeller with the boundary close
to the propeller while the stationary domain is represented by a larger cylinder that enclose the
rotating domain and propeller with the boundary further away from the propeller. For CFD analysis
on the propellers at different propeller disc angle, the size of the static domain is remained constant
while the direction of the inlet is changed to achieve the effect of desired propeller disc angle. Figure
2 below shows the full domain of the model used for simulation at non-zero propeller disc angle.

(a) (b)
Fig. 2. (a) Isometric view and (b) Top View of the CFD domain for non-zero propeller disc angle

2.2 Mesh Generation

The meshing for the model stationary domain, rotating domain and propeller is generated using
the built-in mesh tool in ANSYS FLUENT. Tetrahedral mesh is constructed with capture proximity and
capture curvature built in settings in ANSYS FLUENT. The element size for the mesh at the propeller
and rotating region are finer compared to the element size of the static region mesh as this project
intend to focus on the flow around the propeller to obtain the lift and drag coefficient of the
propeller. To eliminate the mesh discretization error, mesh convergence test is conducted by
recording the thrust of the propeller at decreasing mesh size until convergence of result is obtained.
Then, the element size is where the result converges is chosen. The meshing of the propeller is shown

85
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

in Figure 3 where the element size for the propeller is smaller and the size increase moving further
away from the propeller. Study from Seeni et al., [2] has come up with a fitting method to overcome
grid independent issues and from the result, the range of the error is about 3-70% of error as the
advance ratio increase from 0.192 to 0.799 respectively.

Fig. 3. Meshing of the propeller and its domain

Mesh convergence test is conducted by recording the thrust of the propeller at decreasing mesh
size until convergence of result is obtained. 4 different meshes consisting of tetrahedral meshes is
constructed to test the convergence of the data. The first 3 mesh constructed is the default meshing
using ANSYS Fluent with coarse, mid, and fine meshing while the fourth meshing constructed is fine
mesh with capture curvature and proximity settings on. The information of the mesh generated is
shown in Table 1. The mesh convergence test is tested by running simulation on the 9x6E APC
propeller model at 1000rpm and 0 ms-1 inlet velocity using standard k-ε model. The propeller thrust
results are plotted against the number of mesh element and compared to the manufacturer data.
The mesh convergence test in Figure 4 shows that the result converges as the number of mesh
element increases and the simulation result for the highest number of mesh elements is closest to
the experimental data. This mesh is chosen as the final mesh as the result converges and further
increase in element number will only cost more computation power and low effect on the accuracy
of the analysis.

Table 1
Number of mesh elements for different mesh configuration
Configuration Mesh Element
Coarse 115729
Medium 308082
Fine 339360
Fine mesh with capture curvature and 352496
proximity settings selected

86
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

0.14

0.12

Propeller Thrust (N)


0.1

0.08
Simulation Data
0.06
Experimental Data
0.04

0.02

0
114000 164000 214000 264000 314000
Number of Mesh Element
Fig. 4. Mesh Convergence Test

2.2 Selection of Turbulence Model

Selection of turbulence model is an important step in setting up a CFD simulation. Static thrust
CFD analysis is conducted at zero propeller disc angle with angular velocity ranging from 1000rpm to
8000rpm using different turbulence model. The CFD simulation result is then compared with the
manufacturer data to determine the most suitable turbulence data that match the manufacturer and
UIUC experiment data [8,16] closely as shown in Table 2. For both 6X4E propeller and 9X6E propeller,
SST k-ω model shows more accurate result compared to standard k-ε model, Spalart-Allmaras (S-A)
model, and standard k-ω model as shown in Figure 5 and Figure 6. Therefore, the SST k-ω model is
used for this study.

Thrust against Angular Velocity (6X4E)


1.6
1.4
1.2
Propeller Thrust (N)

1 EXPERIMENTAL

0.8 K-EPSILON

0.6 Spalart-Allmaras

0.4 k-omega

0.2 SST k-omega

0
0 2000 4000 6000 8000 10000

Angular Velocity (rpm)


(a)

87
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

Thrust against Angular Velocity (9X6E)


9
8
7
Propeller Thrust (N)

EXPERIMENTAL
6
5 K-EPSILON
4 Spalart-Allmaras
3
2 k-omega
1 SST k-omega
0
0 2000 4000 6000 8000 10000
Angular Velocity (rpm)
(b)
Fig. 5. Propeller thrust against propeller angular velocity curve for (a) 6X4E and (b) 9x6E
propeller

Torque against Velocity (6x4E)


0.02
0.018
0.016
Propeller Torque (Nm)

0.014 EXPERIMENTAL
0.012
K-EPSILON
0.01
0.008 Spalart-Allmaras
0.006 k-omega
0.004 SST k-omega
0.002
0
0 2000 4000 6000 8000 10000
Angular Velocity (rpm)
(a)

Torque against Angular Velocity (9x6E)


0.16
0.14
Propeller Torque (Nm)

0.12
0.1 EXPERIMENTAL
0.08 K-EPSILON
0.06 Spalart-Allmaras
0.04 k-omega
0.02 SST k-omega
0
0 2000 4000 6000 8000 10000
Angular Velocity (rpm)

(b)
Fig. 6. Propeller torque against propeller angular velocity curve for (a) 6X4E and (b) 9x6E propeller

88
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

Table 2
Percentage error of different turbulence model simulation result compared to
manufacturer experimental data
Turbulence Model Percentage Error (%)
6x4E 9x6E
Propeller Propeller Propeller Propeller
Thrust Torque Thrust Torque
Standard k-ε model 2.41 17.63 2.69 10.73
Spalart-Allmaras 4.34 16.68 2.76 9.68
model
Standard k-ω model 4.62 18.60 3.02 9.78
SST k-ω model 2.54 15.98 2.50 9.03

3. Results
3.1 CFD Simulation of Propeller at Different Advance Ratio

To further validate the use of CFD simulation in determining the aerodynamic characteristics of
the Low Reynolds Number propeller. The thrust coefficient, torque coefficient and propeller
efficiency of the 9X6E and 6X4E propellers are obtained through CFD simulation and compared with
the experimental data. The advance ratio used ranges from 0 to 0.88 while the angular velocity ranges
from 4000rpm and 8000rpm. The figures show that thrust coefficient and torque coefficient
decreases with the increasing advance ratio for both propellers. For the propeller efficiency, the
propeller efficiency increases with the increasing advance ratio and then decreases with the
increasing advance ratio at advance ratio ≥ 0.6. The simulation is conducted for 4000rpm and
8000rpm to test the consistency of the CFD simulation at different angular velocity.
Based on the result of CFD simulation, the coefficient of thrust, CT and CQ at the same advance
ratio result show consistent value when the simulation is run on different angular velocity. Based on
Figure 7 and Figure 8, the CFD simulation results follows the experimental result closely. For 6X4E
propeller, the Ct value is overestimated at higher advance ratio as shown in Figure 7. Meanwhile, the
CQ value for 6X4E is overestimated at all advance ratios. The average percentage error for C T and CQ
at different angular velocity is 11.50% and 27.49% respectively. For 9X6E propeller, both C T and CQ
simulation result matches the experimental result closely and the CQ value is overestimated at lower
and higher advance ratio. The average percentage error for CT and Cq at different angular velocity is
8.40% and 16.7% respectively. For propeller efficiency, NP, the CFD simulation results follows the
experimental value closely with slight overestimation for both 6X4E and 9X6E propeller. The average
percentage error for NP at different angular velocity is 13.59% and 11.52% for 6X4E and 9X6E
respectively.
There are different reasons that contribute to the error of the CFD simulation. The main source
of error for the CFD simulation is the CAD modelling of the propeller. Firstly, the information on the
aerofoil of the propeller blade is limited. The aerofoil of the propeller blade is only described as low
Reynolds number Eppler E63 aerofoil blended with a Clark-Y similar aerofoil at the tip of the propeller
blade. Besides, the limitation of measuring tools used to measure the dimension of the propellers
also contribute to the error of the simulation. The top and side view of the propeller is sketched and
a digital vernier calliper is used to measure the width at different distance of the propeller blade. The
limitation of the vernier calliper to measure the distance accurately will contribute to the error.
Besides, the blade angle of twist cannot be measured without the specific tools such as 3D scanner.
Information on the propeller blade chord and twist distribution of 9X6E propeller are obtained from
UIUC propeller database and combined with the dimension measured using vernier calliper,
therefore the error of CFD simulation is lower. Meanwhile there is no information on the chord and

89
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

twist distribution of 6X4E propeller blade. Hence, the error of the CFD simulation on 6x4E propeller
is higher than that of 9X6E propeller.

Fig. 7. CT, CQ and Np against Advance Ratio for 6x4E at 4000rpm and 8000rpm

90
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

Fig. 8. CT, CQ and Np against Advance Ratio for 9X6E at 4000rpm and 8000rpm

3.2 Propeller Performance at Different Propeller Disc Angle

The thrust and torque coefficient for both 6X4E and 9X6E APC propeller at varying advance ratio
and propeller disc angle is obtained via CFD simulation. The simulation is conducted at angular
velocity of 4000rpm, propeller disc angle of 30°, 60° and 90° and advance ratio ranging from 0 to
0.88. 4000rpm is chosen as the model and analysis due the less error yielded during 0 angle at
different advance ratio. The results for both 6X4E and 9X6E in figures below show that the thrust
coefficient increases with the increasing propeller disc angle at the same advance ratio. For propeller
disc angle of 0°, 30° and 60°, the thrust coefficient decreases as the advance ratio increases.
Meanwhile, the thrust coefficient at 90° propeller disc angle increases with the increasing advance
ratio. It is also evident that the difference of thrust coefficient at different propeller angle is minimal
at advance ratio < 0.3. The results obtained from the CFD simulation matches with the results
obtained by Serrano et al., [1] where the thrust coefficient decreases with increasing advance ratio
at propeller disc angle ≤ 60° and increases with increasing advance ratio at propeller disc angle ≥ 70°.

91
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

As for torque coefficient, the CFD simulation result shows that the torque coefficient increases
with increasing propeller disc angle at the same advance ratio except for thrust coefficient at α = 60°
smaller than that of 30°. For propeller disc angle of 0°, 30° and 60°, the torque coefficient decreases
as the advance ratio increases. Meanwhile, the torque coefficient at 90° propeller disc angle increases
with the increasing advance ratio. The torque coefficient at advance ratio ≤ 0.3 also shows overlap
and minimal difference indicating that the propeller disc angle only affects the torque coefficient
slightly at low advance ratio.
According to Figure 9, Figure 10 and Figure 11, the propeller efficiency increases with increasing
propeller disc angle of attack at the constant advance ratio. For 0° propeller disc angle, the propeller
efficiency increases with increasing advance ratio until the maximum efficiency around advance ratio
of 0.6 and then decreases with the increasing advance ratio. For propeller disc angle > 0°, the
propeller efficiency increases with increasing advance ratio. At advance ratio ≤ 0.3, the propeller
efficiency at different propeller disc angle overlaps and there ARE only small differences of the
efficiency. These CFD simulation results matches the wind tunnel results obtained by Serrano et al.,
[1] closely.

(a)

(b)
Fig. 9. Coefficient of thrust, CT against advance ratio at different angle for
(a) 6X4E propeller (b) 9x6E

92
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

(a)

(b)
Fig. 10. Coefficient of torque, CQ against advance ratio for (a) 6X4E propeller (b)
9x6E

93
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

1.8
1.6
1.4
1.2
1 0 degree
Np

0.8 30 degree
0.6 60 degree
0.4 90 degree
0.2
0
0 0.2 0.4 0.6 0.8 1

Advance Ratio
(a)
1.8
1.6
1.4
1.2
1 0 degree
Np

0.8 30 degree
0.6 60 degree
0.4 90 degree
0.2
0
0 0.2 0.4 0.6 0.8 1

Advance Ratio
(b)
Fig. 11. Efficiency, NP against advance ratio for (a) 6X4E propeller (b) 9x6E

4. Conclusions

Two CFD analysis was conducted, one is the CFD simulation on 9x6E and 6x4E APC propellers
conducted using ANSYS FLUENT at different advance ratio from 0 to 0.88 and different rpm at
4000rpm and 8000rpm. The percentage error of the CFD simulation on 6X4E propeller is higher than
that of 9X6E propeller. Second, the CFD simulation on 9X6E and 6X4E APC propellers conducted at
advance ratio ranging from 0 to 0.8 and propeller disc angle of 0°, 30°, 60°, and 90° at 4000rpm. The
CT, CQ and NP against advance ratio curves generated from CFD simulation followed the trend of the
wind tunnel testing data conducted by Serrano et al., [1].
Multiple reference frame method is used to simulate the rotating of the propeller in the flowing
airstream. Tetrahedral mesh is constructed, and mesh convergence test is conducted to minimize the
discretization error. By comparing the simulation results from different turbulence model, SST k-ω
model is found to provide more result with higher accuracy compared to standard k-ε model, Spalart-
Allmaras (S-A) model, and standard k-ω model.
The CFD analysis shows that the turbulence model and the mesh used are acceptable in
comparison to the wind tunnel test for thrust performance estimation but improvement on meshing
configuration can be done for torque performance estimation. The error produce for thrust

94
Journal of Advanced Research in Fluid Mechanics and Thermal Sciences
Volume 96, Issue 1 (2022) 82-95

estimation is lower than 12% for both 6x4E and 9x6E propellers. The error for torque and efficiency
estimation is between 15-30% and 12% respectively.

Acknowledgement
This research is funded by Universiti Teknologi Malaysia under UTM FR vote Q.J130000.2551.21H61.
This research would also like to acknowledge the facilities provided by AEROLAB UTM to aid the
research.

References
[1] Serrano, David, Max Ren, Ahmed Jawad Qureshi, and Sina Ghaemi. "Effect of disk angle-of-attack on aerodynamic
performance of small propellers." Aerospace Science and Technology 92 (2019): 901-914.
https://doi.org/10.1016/j.ast.2019.07.022
[2] Seeni, Aravind, Parvathy Rajendran, and Hussin Mamat. "A CFD Mesh Independent Solution Technique for Low
Reynolds Number Propeller." CFD Letters 11, no. 10 (2019): 15-30.
[3] Cable, Matthew. "An evaluation of turbulence models for the numerical study of forced and natural convective
flow in Atria." PhD diss., Queen's University, 2009.
[4] Zienkiewicz, O. C., R. L. Taylor, and P. Nithiarasu. "Chapter 8-Turbulent flows." The Finite Element Method for Fluid
Dynamics, 7th edn. Butterworth-Heinemann (2014): 283-308. https://doi.org/10.1016/B978-1-85617-635-
4.00008-X
[5] Kutty, Hairuniza Ahmed, and Parvathy Rajendran. "3D CFD simulation and experimental validation of small APC
slow flyer propeller blade." Aerospace 4, no. 1 (2017): 10. https://doi.org/10.3390/aerospace4010010
[6] Duan, Xiaoxia, Xin Feng, Chao Yang, and Zaisha Mao. "CFD modeling of turbulent reacting flow in a semi-batch
stirred-tank reactor." Chinese Journal of Chemical Engineering 26, no. 4 (2018): 675-683.
https://doi.org/10.1016/j.cjche.2017.05.014
[7] Menter, Florian R. "Two-equation eddy-viscosity turbulence models for engineering applications." AIAA Journal 32,
no. 8 (1994): 1598-1605. https://doi.org/10.2514/3.12149
[8] Brandt, John, and Michael Selig. "Propeller performance data at low reynolds numbers." In 49th AIAA Aerospace
Sciences Meeting including the New Horizons Forum and Aerospace Exposition, p. 1255. 2011.
https://doi.org/10.2514/6.2011-1255
[9] Silvestre, Miguel, João Morgado, Pedro Alves, Pedro Santos, Pedro Gamboa, and José Páscoa. "Low Reynolds
Number Propeller Performance Testing." Recent Advances in Mechanical Engineering (2014).
[10] Wendt, John F., ed. Computational fluid dynamics: an introduction. Springer Science & Business Media, 2008.
[11] Wei, Xinli, Jie Ren, and Xiangrui Meng. "Simulation and Experiment Study of Flow Field and Dynamic Performance
in Stirred Reactor." In Challenges of Power Engineering and Environment, pp. 1408-1413. Springer, Berlin,
Heidelberg, 2007. https://doi.org/10.1007/978-3-540-76694-0_265
[12] Yilmaz, Serdar, Duygu Erdem, and Mehmet S. Kavsaoglu. "Performance of a ducted propeller designed for UAV
applications at zero angle of attack flight: An experimental study." Aerospace Science and Technology 45 (2015):
376-386. https://doi.org/10.1016/j.ast.2015.06.005
[13] Ab Wahid, M., A. N. Asmi, N. Othman, M. I. Ardani, MZ Md Zain, and S. Mansor. "The effect of different motor
constants to an commercial propeller." In IOP Conference Series: Materials Science and Engineering, vol. 884, no.
1, p. 012096. IOP Publishing, 2020. https://doi.org/10.1088/1757-899X/884/1/012096
[14] Loureiro, Eric Vargas, Nicolas Lima Oliveira, Patricia Habib Hallak, Flávia de Souza Bastos, Lucas Machado Rocha,
Rafael Grande Pancini Delmonte, and Afonso Celso de Castro Lemonge. "Evaluation of low fidelity and CFD methods
for the aerodynamic performance of a small propeller." Aerospace Science and Technology 108 (2021): 106402.
https://doi.org/10.1016/j.ast.2020.106402
[15] Mohamed, Wan Mazlina Wan, Nirresh Prabu Ravindran, and Parvathy Rajendran. "A CFD Simulation on the
Performance of Slotted Propeller Design for Various Airfoil Configurations." CFD Letters 13, no. 3 (2021): 43-57.
https://doi.org/10.37934/cfdl.13.3.4357
[16] Brandt, John B, Robert W Deters, and Gavin K Ananda. "UIUC Propeller Database - Volume 2." UIUC Propeller Data
Site. UIUC Applied Aerodynamics Group, May 27, 2015.

95

You might also like