Você está na página 1de 82

Numerical simulations of turbulent flow around a

distributed propelled wing

António Afonso Moutinho Salreta

Thesis to obtain the Master of Science Degree in

Mechanical Engineering

Supervisors: Prof. Luís Rego da Cunha de Eça

Prof. –Ing. Andreas Henze

Examination Committee

Chairperson: Prof. Viriato Sérgio de Almeida Semião


Supervisor: Prof. Luís Rego da Cunha de Eça
Member of the Committee: Prof. João Eduardo de Barros Teixeira Borges

November 2016
Acknowledgement

First and foremost I would like to thank Prof. Dr. –Ing. Wolfgang Schröder for giving me the opportunity
to work in the Aerodynamics Institute of Aachen of Rheinisch –Westfaelische Technische Hochschule,
and facilitating me the resources to do my thesis. I am honoured to be associated with this institute
and university.

Secondly I want to thank all the guidance and patience given to me during the making of this thesis by
Dr. –Ing. Andreas Henze. His door was always open whenever I had problems or doubts, and gave
me much valuable advices for my work.

I also want to thank Michael Kreimeier for giving me the opportunity to work on this project and for
having the patience to answer to all my questions.

Also to my professor back in Lisbon, Prof. Luís Eça of the Aerodynamic Department of IST-UL, that
help me rewrite and reorganize my thesis and gave me fundamental advice for my work.

Finally, I would like to express my gratitude to my family and friends that supported and encouraged
me throughout all my years of study and through the process of researching and writing of this thesis,
without whom I couldn’t have done it.

I
Resumo
Esta tese tem como objectivo estudar numericamente o escoamento em volta de um avião com
sistema de propulsão eléctrico distribuída intitulado “Silent Air Taxi”. Coeficientes aerodinâmicos de
sustentação e de arraste para diferentes configurações de propulsores foram calculados. Alterações
paramétricas ao diâmetro, à posição do hélice em relação ao bordo de ataque e à potência dos
mesmos foram efectuadas para registar diferenças na sustentação e no arraste. Diferentes
velocidades da aeronave e o número de propulsores activos foram também estudadas. Em cada caso
foi estudado o ponto de sustentação máxima com o objectivo final de fornecer informações iniciais
sobre os propulsores que irão ser utilizados no projecto. Os resultados permitiram concluir que as
alterações à distância da hélice ao bordo de ataque são demasiado pequenas para serem relevantes.
Em relação à diferença de potência e diâmetros, certas configurações mostram serem mais
vantajosas que outras, assim como a diferença de velocidades do escoamento (ou avião). Neste caso
velocidades mais baixas produzem coeficientes de sustentação mais altos. Por fim as simulações
mostram que os propulsores mais perto da ponta da asa influenciam mais a sustentação da asa.
Simulações RANS e o modelo de turbulência SST Menter K-Omega foram usados, usando um
programa de mecânica de fluidos computacional chamado STAR-CCM+. Esta tese inclui a geração
da malha e criação de uma curva de desempenho para os propulsores. Este trabalho foi realizado em
Aachen Alemanha, no instituto de aerodinâmica da Universidade Técnica da Renânia do Norte-
Vestfália RWTH, pertencendo ao Instituto de Aeronáutica e Aeroespacial.

Palavras-chave: propulsão eléctrica distribuida, alterações paramétricas, coeficiente de sustentação,


STAR-CCM+, SST Menter K-Omega.

II
Abstract

This thesis has the objective of studying numerically the flow of a distributed electric propulsion plane
called "Silent Air Taxi" (SAT), calculating the aerodynamic coefficients of lift, and drag for different
propeller configurations concluding on its effects. Parametric changes to diameter and position relative
to the leading edge, as well as power alterations were made to the propellers in order to see the
alterations in lift and drag. Other changes were made, namely to the inflow velocity and the number of
working thrusters so to see the influences of each propeller. In each case the maximum lift point was
studied. Propeller design decisions are the final objective of this work. The results conclude that the
changes in the distance of the propeller to the leading edge are too small to be relevant. Some changes
in power and diameter show some benefits while others do not. Lower velocities show higher lift
coefficients and other results also show that propellers closer to the tip of the wing have higher
influence on the lift. RANS simulations with SST Menter K-Omega turbulence model were done using
the commercial CFD program STAR-CCM+. This master thesis also includes mesh generation, and
the creation of a performance curve for the propellers used in the plane used in the CFD program. The
work was done at the Aerodynamic Institute of RWTH (Aerodynamisches Institut von Aachen, AIA)
but belongs to the Institute of Aeronautics and Astronautics of RWTH (Institut für Luft-und Raumfahrt,
ILR).

Keywords: distributed electric propulsion, parametric changes, lift coefficient, STAR-CCM+, SST
Menter K-Omega.

III
Nomenclature
1prop Reference case without the 1st propeller active
2prop Reference case without the 2nd propeller active
3prop Reference case without the 3rd propeller active
4prop Reference case without the 4th propeller active
5prop Reference case without the 5th propeller active
𝑎 Specific power
𝑎𝑓 Face area vector
𝑎𝑖 Constant of SST Menter K-Omega model
𝐴 Axial force
𝑏 Specific energy
𝐵 Number of blades
BWB Blended wing body
CAD Computer assisted design
𝐶𝐴 Axial Force coefficient
𝐶𝑁 Normal Force coefficient
𝐶𝐷 Drag coefficient
𝐶𝐿 Lift coefficient
𝐶𝐿 𝑀𝑎𝑥 Maximum lift coefficient
𝐷 Drag force
DEP Distributed Electric Propulsion
DLE Distance to the Leading Edge
DP Distributed Propulsion
Default Reference solution

𝐷𝑖𝑛𝑛𝑒𝑟 “high-lift” propeller outer diameter


𝐷𝑜 Outer diameter
𝐷𝑇𝑖𝑝 Tip propeller outer diameter
eBPR Effective bypass ratio
𝑓𝑏𝑥 Volume axial force distribution of the actuator disk
𝑓𝑏𝜃 Volume tangencial force distribution of the actuator disk
𝑝𝑟𝑒𝑠𝑠𝑢𝑟𝑒
𝑓𝑆 Force vector due to pressure on the surface face S
𝑓𝑆𝑠ℎ𝑒𝑎𝑟 Force vector due to shear stress on the surface face S
𝐹1 Blended function of the SST Menter K-Omega
𝐹2 Function of the distance to the wall of the SST Menter K-Omega
I30 Inner propeller at 30 kW
𝐽 Advance ratio
𝑘 Turbulent kinetic energy

IV
𝐾𝑇 Thrust coefficient
𝐾𝑄 Torque coefficient
𝐿 Lift force
𝐿/𝐷 Lift to drag ratio
LEAPTech Leading Edge Asynchronous Propulsion Technology
LES Large Eddy simulations
𝑁 Normal force
NASA National Aeronautics and Space Administration
𝑛 Rotation rate (rpm)
𝑃 Shaft Power of a single propeller
𝑄 Torque of a single propeller
𝑅𝐻 Propeller hub radius
𝑅𝑃 Propeller tip radius
𝑟 Radius

𝑟 Function of radius
𝑟𝑡 ′ Function of radius
𝑟′ Function of radius
𝑆 Wing surface area
𝑆𝑡 and 𝑆𝑖𝑗 Strain tensor
SAT Silent Air Taxi
SCEPTOR Scalable Convergent Electric Propulsion Technology and Operation
Research
SFW Subsonic fixed wing
RANS Reynolds Average Navier-Stokes
𝑇 Total thrust of the propellers
𝑇20 Tip propeller at 20 kW case
𝑇70 Tip propeller at 70 kW case
𝑇𝑓 Stress tensor at face f
𝑇𝑡 Reynolds stress turbulent tensor
TeDP Turboelectric distributed propulsion
𝑢′ Velocity fluctuation in the x direction
v Velocity vector
V∞ Inlet flow velocity
𝑣′ Velocity fluctuation in the y direction
VTOL Vertical take-off and landing
𝑤′ Velocity fluctuation in the z direction
𝑋1 0.1 m difference in 𝑥𝐿𝐸
𝑋2 0.2 m difference in 𝑥𝐿𝐸
𝑥𝐿𝐸 Distance from the propeller and the leading edge
+
𝑦 Non dimensional wall distance

V
Greek Letters
𝛼 Angle of attack
𝜀 Turbulent dissipation rate
𝜂 Propeller efficiency
𝜇𝑡 Dynamic turbulent viscosity
𝜌 Air density
𝜔 Specific Dissipation rate
Ω Absolute value of vorticity

Symbols
° Degrees

VI
Table of contents

Acknowledgement .......................................................................................................................... I

Resumo ......................................................................................................................................... II

Abstract ........................................................................................................................................ III

Nomenclature ............................................................................................................................... IV

List of Figures ............................................................................................................................... IX

List of Tables ................................................................................................................................ XI

1. Introduction ............................................................................................................................... 1

1.1. Motivation and comparable work ................................................................................... 1

1.2. Tools available ............................................................................................................... 6

1.3. Objectives ...................................................................................................................... 6

2. Mathematical Models ............................................................................................................... 8

2.1. Computational flow equations and turbulence model ................................................... 8

2.2. Numerical approximations ........................................................................................... 10

2.3. Propeller model ........................................................................................................... 10

3. Numerical Solution ................................................................................................................. 13

3.1. Geometry and domain ................................................................................................. 13

3.2. Boundary conditions .................................................................................................... 16

3.3. Mesh Refinement ........................................................................................................ 19

3.4. Comparison with results from literature ....................................................................... 27

4. Results ................................................................................................................................... 29

4.1. Reference Solution ...................................................................................................... 29

4.2. X1 and X2 .................................................................................................................... 41

4.3. Diameter change ......................................................................................................... 43

4.4. I30 ................................................................................................................................ 46

4.5. T20 and T70 ................................................................................................................ 48

4.6. Velocity change ........................................................................................................... 51

4.7. Influences of propellers at 31.38 m/s (61knots) .......................................................... 53

5. Conclusions and Future Directions ........................................................................................ 57

5.1. Conclusions ................................................................................................................. 57

5.2. Future Work ................................................................................................................. 58

VII
6. References ............................................................................................................................. 59

7. Appendix ................................................................................................................................ 61

7.1. JavaPROP tables .............................................................................................................. 61

7.2. 2D Simulations on Xfoil ...................................................................................................... 64

7.3. No Propulsion .................................................................................................................... 67

7.4. Optimization design process in JAVAProp ........................................................................ 68

VIII
List of Figures

Figure 1 - SCEPTOR Prototype .................................................................................................... 4


Figure 2 – LEAPTech project with Pressure Coefficient on the surface ....................................... 4
Figure 3 – LEAPTech experimental testbed ................................................................................. 5
Figure 4 - Complete Plane Geometry ......................................................................................... 16
Figure 5 - Propellers locations..................................................................................................... 16
Figure 6 - Entire control volume .................................................................................................. 17
Figure 7 - View from the symmetry plane ................................................................................... 17
Figure 8 - Wing with virtual disks................................................................................................. 18
Figure 9 - Schematic of forces [20] ............................................................................................. 18
Figure 10 – Initial mesh with no refinement (Mesh 1) ................................................................. 21
Figure 11 – Mesh with double the cells in the wake and surface close to the wing(Mesh 2) ..... 21
Figure 12 - Mesh with double the cells per transition (Mesh 3) .................................................. 22
Figure 13 - Mesh double the cells in the wake and surface close to the wing compared to
Mesh2(Mesh 4)............................................................................................................................ 22
Figure 14 - Mesh with double the cells per transition compared to Mesh4 (Mesh 5) ................. 23
Figure 15 – Mesh 5 with local refinement at the disk level (Mesh 6) .......................................... 23
Figure 16 – Mesh 5 with local refinement on a whole box surrounding the wing (Mesh 7) ........ 24
Figure 17 - Prism Layer from Mesh 6 .......................................................................................... 24
Figure 18 - Volume Mesh 6 of the whole wing ............................................................................ 25
Figure 19 – Detailed Mesh 6 on Nacelle ..................................................................................... 25
Figure 20 – Hole in tip nacelle ..................................................................................................... 26
Figure 21 – Hole in tip nacelle fixed ............................................................................................ 26
Figure 22 – Hole in “high-lift” nacelle 4 ....................................................................................... 27
Figure 23 - Hole in “high-lift” nacelle 4 fixed ................................................................................ 27
Figure 24 - Lift coefficient for Reference solution ........................................................................ 31
Figure 25 - Drag coefficient for Reference solution..................................................................... 31
Figure 26 – Residual values for Reference Solution at 8° .......................................................... 32
Figure 27 – Normal force monitor for Reference Solution at 8° .................................................. 32
Figure 28 - Polar graph for Reference solution ........................................................................... 33
Figure 29 - Absolute total pressure field on propeller plane for Reference solution at 8° (inlet
absolute total pressure of 75480 Pa) .......................................................................................... 33
Figure 30 - Absolute total pressure field left to propeller plane for Reference at 8° ................... 34
Figure 31 - Absolute total pressure field right to propeller plane for Reference solution at 8° ... 34
Figure 32 - Velocity field propeller plane for Reference solution at 8°(inlet velocity of -110.3 m/s)
..................................................................................................................................................... 35
Figure 33 - Velocity field right to propeller plane for Reference solution at 8° ............................ 35
Figure 34 - Velocity field left to propeller plane for Reference solution at 8° .............................. 36
Figure 35 - Streamlines for Reference solution at 8° .................................................................. 36

IX
Figure 36 - Streamlines for Reference solution at 8° .................................................................. 37
Figure 37 - Wall Y+ for Reference solution at 8° ......................................................................... 37
Figure 38 - Vorticity Isosurface for Reference solution at 8° ....................................................... 38
Figure 39 – Normal force coefficient along the span for Reference solution case at 8° ............. 38
Figure 40 - Accumulated Normal force coefficient along the span for Reference solution case at
8° ................................................................................................................................................. 39
Figure 41 - Axial Force Coefficient along the span for Reference solution case at 8°................ 39
Figure 42 - Accumulated Axial Force Coefficient along the span for Reference solution case at 8°
..................................................................................................................................................... 40
Figure 43 - Wall Shear Stress x-direction for Reference solution case at 7° .............................. 40
Figure 44 - Wall Shear Stress x-direction for Reference solution case at 8° .............................. 40
Figure 45 - Wall Shear Stress in the x-direction for Reference solution case at 9° .................... 41
Figure 46 - Wall Shear Stress x-direction for Reference solution case at 10° ............................ 41
Figure 47 – Lift coefficient for the X-direction change cases ...................................................... 42
Figure 48 – 𝐶𝐿/𝐶𝐿𝑟𝑒𝑓 vs angle of attack for the X-direction change cases ................................ 43
Figure 49 - Drag coefficient for the X-direction change cases .................................................... 43
Figure 50 - Lift coefficient for Diameter change case.................................................................. 45
Figure 51 - 𝐶𝐿/𝐶𝐿𝑟𝑒𝑓 vs angle of attack for Diameter change case ............................................ 45
Figure 52 - Drag coefficient for Diameter change case............................................................... 46
Figure 53 - Lift Coefficient for I30 case ....................................................................................... 47
Figure 54 – 𝐶𝐿/𝐶𝐿𝑟𝑒𝑓 vs angle of attack for I30 case.................................................................. 47
Figure 55 - Drag Coefficient for I30 case .................................................................................... 48
Figure 56 - Lift Coefficient for different tip power cases .............................................................. 49
Figure 57 - 𝐶𝐿/𝐶𝐿𝑟𝑒𝑓 vs angle of attack for different tip power cases......................................... 49
Figure 58 - Drag Coefficient for different tip power cases ........................................................... 50
Figure 59 – Wall Shear Stress for T20 at 8° ............................................................................... 50
Figure 60 - Wall Shear Stress for T20 at 9° ................................................................................ 50
Figure 61 - Wall Shear Stress for T20 at 10° .............................................................................. 51
Figure 62 - Lift Coefficient for different flow speed case ............................................................. 52
Figure 63 – CLlCLref vs angle of attack for different flow speed case .......................................... 52
Figure 64 - Drag Coefficient for different flow speed case .......................................................... 53
Figure 65 – Lift coefficient for propellers influences cases ......................................................... 55
Figure 66 - 𝐶𝐿/𝐶𝐿𝑟𝑒𝑓vs angle of attack ....................................................................................... 56
Figure 67 - Drag coefficient for propellers influences cases ....................................................... 56
Figure 68 - 2D Simulations .......................................................................................................... 66
Figure 69 - No propulsion case compared to reference case ..................................................... 67
Figure 70 - Wall Shear Stress in the positive direction of x for No propulsion case at 10° ......... 68

X
List of Tables

Table 1 - Aircraft requirements [1] ................................................................................................. 1


Table 2 - Comparison of the reference configuration of LEAPTech with Cirrus, S2 with Van's RV-
7 [6] ................................................................................................................................................ 3
Table 3 - Flow Properties for 3048 meters [20] ........................................................................... 13
Table 4 - Mesh Parameters ......................................................................................................... 13
Table 5 - SAT Design Specifications[7][8] ................................................................................... 14
Table 6 - Propeller Design Conditions ........................................................................................ 15
Table 7 - Mesh Refinement ......................................................................................................... 20
Table 8 - Comparison table [7][21] .............................................................................................. 28
Table 9 - Table of configurations ................................................................................................. 29
Table 10 - Reference Results...................................................................................................... 30
Table 11 - X1 Results .................................................................................................................. 42
Table 12 - X2 Results .................................................................................................................. 42
Table 13 - Small Diameter Results.............................................................................................. 44
Table 14 - Large Diameter Results ............................................................................................. 44
Table 15 - Results of Inner Propellers at 30 kW ......................................................................... 46
Table 16 - Results of Tip propellers at 20 kW ............................................................................. 48
Table 17 - Results of Tip propellers at 70 kW ............................................................................. 49
Table 18 - Results of flow inlet at 150 knots ............................................................................... 51
Table 19 – Results of flow inlet at 61 knots ................................................................................. 52
Table 20 – Results of flow inlet at 61knots without propeller 1 ................................................... 54
Table 21 - Results of flow inlet at 61knots without propeller 2 .................................................... 54
Table 22 - Results of flow inlet at 61knots without propeller 3 .................................................... 54
Table 23 - Results of flow inlet at 61knots without propeller 4 .................................................... 54
Table 24 - Results of flow inlet at 61knots without propeller 5 .................................................... 55
Table 25 - Inner Propellers Performance table ........................................................................... 62
Table 26 - Tip Propellers Performance table .............................................................................. 63
Table 27 - 2D cruise wing profile Xfoil simulation for 110.606 m/s ............................................. 64
Table 28 - 2D Take-off Xfoil Simulation ...................................................................................... 65
Table 29 - No propeller results .................................................................................................... 67

XI
1. Introduction

1.1. Motivation and comparable work

In the present days modern aviation has seen a higher research for electric propulsion due to higher
fuel costs and improvements made in propellers driven by electric motors. In order to replace normal
combustion engines by electric ones some key characteristics have to be met. Only if the specific
power, weight and reliability required are attained can electric propulsion bring benefits for commercial
and general aircrafts. There are six different electric propulsion architectures:

 All electric
 Hybrid electric
o Parallel hybrid
o Series hybrid
o Series/parallel partial hybrid
 Turboelectric
o Full turboelectric
o Partial turboelectric

All electric systems use batteries as the only source of power while hybrid systems use gas turbine
engines for propulsion and to charge batteries. There are even turboelectric systems that do not
require batteries for propulsion but instead drive electric generators which power inverters and
eventually individual direct current (DC) motors that drive the individual distributed fans. Table 1 lists
and summarizes the requirements of the electrical system performance for Hybrid, All-Electric, and
Turboelectric Propulsion Systems considered by a committee comprised of NASA, the National
Academies of Sciences, Engineering and Medicine.

Aircraft requirements Electric system a Battery b

Power Capability (MW) Specific Power (kw/kg) c Specific Energy (Wh/kg)

General aviation and commuter

Parallel hybrid Motor <1 >3 >250


All-electric Motor <1 >6.5 >400
Turboelectric Motor and generator <1 >6.5 N/a
Regional and single-aisle
Parallel hybrid Motor 1-6 >3 >800
All-electricb Motor 1-11 >6.5 >1,800
Turboelectric Motor 1.5-3; generator 1-11 >6.5 N/a
Twin-aisle
Parallel hybrid Not studied
All-electric Not feasible
Turboelectric Motor 4; generator 30 >10 N/a
Apu for large aircraft Generator 0.5-1 >3 Not studied

Table 1 - Aircraft requirements [1]

a Includes power electronics.

1
b Total battery system and usable energy for discharge durations that are relevant to commercial aviation flight times,
nominally 1-10 hours. Values shown are for rechargeable batteries; primary (nonrechargeable) batteries are not considered
relevant to commercial aviation.
c Conversion factors: 1 kW/kg = 0.61 HP/lb; 1 kg/kW = 2.2 lb/kW = 1.64 lb/HP.

Hybrid-electric and all-electric systems are not recommended as a high priority approach because the
batteries with the power capacity and specific power required for commercial aircraft at least as large
as a regional jet are unlikely to be developed enough to the point that products satisfying FAA
certification requirements can be developed within a 30 year time frame. All-electric battery powered
airplane configurations will be limited to small aircraft. For large commercial aircraft it is likely that
turbo-electric will be used and introduced in the next 10 to 30 years. [1]

Directly replacing normal combustion engines by electric ones can present the most obvious limitations
of electric motors: higher weight due to batteries, power capability and specific power requirements,
cost of batteries, life cycle etc. But the most interesting characteristic about electric propeller is the
advantage of producing smaller sized engines without decreased efficiency. This way design
configurations for aircraft designers becomes much broader. Normal combustion engines restrain the
design possibilities because of their weight and size, limiting the number and position of propellers. By
using small electric engines, taking advantage of their low weight, the position of the propellers
becomes much more flexible. Other configurations can be tested with electric propellers that could not
have been possible with other engines, because of small sized propellers. This opens a wide range of
experimental opportunities, creating possibilities to locate engines in strategic locations, increasing
design flexibility, unlike presently where the location of engines is constrained because of the weight
and size of the propeller. The higher number of propellers also increases the reliability of the aircraft.
[2]

One promising design is the distributed electric propulsion (DEP) where small propellers are placed
along the wing span of a plane. This has been shown to increase dynamic pressure over the wing
resulting in increased maximum lift coefficient at low speeds. This allows for a smaller design area
wing which causes a decrease of aerodynamic sensibility, maintaining take-off and landing
performance, as well as noise reduction. NASA has been investigating this type of configuration for
better synergistic benefits of coupling airframe aerodynamics with propulsion thrust stream. Early
concepts of turboelectric distributed propulsion (TeDP) vehicles with subsonic fixed wing (SFW) show
promise for better efficiency and noise reduction [3]. This concept employs a number of
superconducting electric motors to drive the distributed fans rather than using mechanical shafts and
gears. The power to drive these electric fans is generated by remotely located gas-turbine-driven
superconducting electric generators. This arrangement enables the use of many small distributed fans,
allowing a very high effective bypass ratio (eBPR), while retaining the superior efficiency of large core
engines. A NASA research predicts that the combination of a Blended Wing Body (BWB) configuration
and Distributed Propulsion (DP) system produces a 70% fuel-burn reduction relative to a B777-200LR
reference aircraft 22 [4].

2
Joby Aviation in cooperation with NASA has been sellingthe use of DEP through a project called
LEAPTech (Leading Edge Asynchronous Propulsion Technology). This project is still in development
having already an experimental testbed of a wing mounted on a truck with 18 identical small propellers,
as shown in figure 3. This company has also another project for vertical take-off and landing (VTOL)
called S2 that includes a DEP system. [5][6] Increase maximum lift coefficient as well as noise
reduction are the results in these studies. The increases of dynamic pressure over the wing caused
by the DEP system enables lower stall speeds and/or smaller wing area without the need for
structurally complex traditional multi-element high-lift systems. The LEAPTech study uses STAR-
CCM+ commercial program for computational fluid dynamic (CFD) analyses and for mesh generation.
Navier-Stokes simulations were run using SST (Menter) 𝑘 − 𝜔 turbulence model and the 𝛾 − 𝑅𝑒𝜃
transition model on unstructured meshes. [6] Table 2 shows some project specifications of LEAPTech
and S2 studies, as well as two other existing airplanes that were used as reference in each study.
Also shown in figure 2 is the LEAPTech airplane with pressure coefficient on the surface of the plane.

The program SCEPTOR (Scalable Convergent Electric Propulsion Technology and Operation
Research) from NASA is the continuation of LEAPTech study and focuses on designing and building
a demonstrator aircraft for DEP technology validation [1]. This was the starting point of the SAT (Silent
Air Taxi) plane for this study. SCEPTOR began with the same geometry as Tecnam P2006T twin
engine light aircraft, replacing the wing and propellers with high aspect ratio wing with DEP system.
The latest design had an aspect ratio of 15 and a wingspan of 9.63 m, with 75.6 cm of root chord and
53.04 cm of tip chord. Meanwhile Tecnam P2006T has a wingspan of 11.40 m and an aspect ratio of
8.80. Also as seen in figure 1, SCEPTOR has 6 “high-lift” (because their function is to produce higher
lift along the span of the wing) propellers and one “cruise” (because it’s the only propeller intended
functioning during cruise conditions) propeller in each wing. The smaller propellers have a power of
14.4 kW and rotate at a velocity close to 2250 rpm, having five-blade rotors that fold to reduce drag in
cruise conditions. While the bigger propellers have a power of 48,1 kW turning at 4,548 rpm, located
at the tip where interacting with the tip vortices reducing the drag in 5%.

Cirrus SR22 LEAPTech S2 Van’s RV-7


Seating capacity 4 4 2 2
Gross weight 1 542.1 kg 1 360.8 kg 907.2 kg 816.5 kg
Wing area 13.47 m2 5.12 m2 5.00 m2 11.24 m2
Wingspan 11.67 m 9.45 m 9.00 m 7.62 m
Aspect ratio 10.1 17.4 16.2 5.2
Wing loading 0.99 kg/m2 2.29 kg/m2 1.57 kg/m2 0.63 kg/m2

Cruise speed 94.3 m/s 89.4 m/s 89.4 m/s 89.0 m/s
Cruise 𝑪𝑳 (altitude of 0.30 0.77 0.52 0.19
3657.6 m)

Table 2 - Comparison of the reference configuration of LEAPTech with Cirrus, S2 with Van's RV-7 [6]

3
Figure 1 - SCEPTOR Prototype

Figure 2 – LEAPTech project with Pressure Coefficient on the surface

4
Figure 3 – LEAPTech experimental testbed

In the case of SAT’s propeller configuration and requirements as seen in section 3.4 of this thesis, it is
possible to calculate an approximate value of the maximum electric system weight the airplane can
have. Assuming the base weight of the aircraft SAT is the same as Cirrus SR22 of 1028 kg [7][8], and
eliminating the dry weight of the combustion engine Continental IO-550-N of about 186.8 kg [9], will
result in 841.2 kg of the aircraft without the power plant. Considering SAT generates a lift force at cruise
of 17970 N ( at 4º where L/D is maximum), or 1832 kg in mass terms and subtracting the previous value
leaves only 990.8 kg for the electric systems. To know the required specific power and energy looking
at the following weight equation:

300 103
990.8 ≥ + 300 × ( 1)
𝑎 𝑏

Where 𝑎 is the specific power and 𝑏 is the specific energy, and 300 is the total power of SAT aircraft.
Assuming 𝑎 to be 4.3 kW/kg, and assuming that the plane flies for one hour, then b must be higher than
325.72 Wh/kg. The specific energy can be attained using current by commercial lithium-ion battery cells
that can have 350 kW/h. [1] The specific power can be attained, by using a EMRAX 207 [10] for example,
however the remaining weight is 21,9 kg which is below the weight of a normal person. This means
further advances must be made to the electric system for this project to be feasible.

5
1.2. Tools available

This project was carried out as a simple research for distributed electric propulsion system on small
man-controlled aircraft. In this case only numerical simulations were made, and no experimental data
was obtained. To simulate the flow around the aircraft, a commercial CFD program was used called
STAR-CCM+. The plane geometry was studied without considering the propeller blades. These were
simulated by adding virtual disks in order to make several and faster simulations. In order to do so a
second program was used to obtain the propellers’ performance curve. This program is called
JavaPROP were the propellers specifications and initial flow conditions resulted into an optimized
performance curve. This curve was then used in STAR-CCM+ for modelling the virtual disks. JavaPROP
is a simple program written in java language based on the blade element theory. The blade is divided
into small sections, handled independently from each other. Each segment has a chord and a blade
angle and associated airfoil characteristics. [11]

1.3. Objectives

This project aims to study the influence of flow initial conditions and parametric changes made to the
electric propellers for a DEP small man-controlled airplane. The final objective is to gather information
about which properties affect more the aircraft and by how much. To do this a reference configuration
was chosen as the starting and comparison point for the other simulations, while the whole geometry
remained constant. No flaps of other devices were added. This reference case was used not only for
the parametric changes made to the propellers but also for the changes in flow conditions like velocity
and number of propellers.

The plane geometry is similar to the SCEPTOR prototype, having in this case five smaller propellers
distributed along the span of the wing, and one bigger propeller at the tip (this refers only to one wing).
Two different propellers were chosen for the “inner” or “high-lift” (because the higher dynamic pressure
created increases lift) ones, and for the “tip” or “cruise” (because in SCEPTOR and in this study these
will be the only propellers used during cruise conditions) one because of their different power and
diameter. The parametric changes for the simulations are the outer propeller diameter, the distance
between the disk and the leading-edge (DLE), and the propeller’s power.

To study the influence of the parametric configurations of the propellers on the aircraft, coefficients of
lift, drag and pressure were determined. The “high-lift” propellers diameter ranged from 0,5 to 0,9 m,
and the “cruise” propellers diameter ranged from 0.75 to 1.5 m, distance from the leading edge from 0
to 0.2 m and power overall change from 240 kW to 500 kW. These changes are presented in table 9 in
the results, summarizing the changes in the configurations of the propulsion. For the diameter change,
the tip propeller could not be bigger for risk of touching the “inner” adjacent propeller. The Reference
solution values were chosen to be compared to Cirrus SR22 plane, as well as LEAPTech case.

Another objective of this aircraft is to be able to fly at low speeds, producing less noise, hence the name
Silent Air Taxi given to the aircraft (SAT). Three speeds were studied, 61 knots, 150 knots and 215
knots, simulating different cruising speed. The comparisons of different parametric propeller

6
configurations were only simulated for the aircraft at 215 knots, while the influence of each propeller was
simulated only at 61 knots. In this part of the project, a propeller was tested as disabled by eliminating
the virtual disk, testing all “high-lift” propellers this way one by one. The decision for the lower velocity
was made later in the project because it showed more interesting results.

Reynolds Average Navier-Stokes (RANS) aeropropulsive simulations were performed maintaining the
wing geometry constant throughout the whole process of the project. The flap angle was maintained at
0°, so all results are only considering the wing and propellers. Because of similarity between SAT and
SCEPTOR (and by extension LEAPTech), the results were also compared to a Cirrus SR22 and Tecnam
P2006T.

7
2. Mathematical Models

2.1. Computational flow equations and turbulence model

There are three possible ways to study turbulent flows: using RANS, Large Eddy Simulations (LES) for
large scales of turbulence while the small-scale motions are modelled, or by using a hybrid modelling
approach that combines features of RANS simulations in some parts with LES in others. The chosen
approach was the RANS because of simplicity, less time-consuming, and more economical.

To obtain the RANS equations, the Navier-Stokes equations for the instantaneous velocity and pressure
fields are decomposed into a mean value and a fluctuating component. The resulting equations are
almost identical to the original equations except for an additional term that now appears in the
momentum equations. This additional term is a tensor quantity, known as Reynolds stress tensor, which
has the following definition[12][13] :

̅̅̅̅̅̅
𝑢′ 𝑢′ ̅̅̅̅̅̅
𝑢′ 𝑣 ′ ̅̅̅̅̅̅
𝑢′ 𝑤 ′
𝑇𝑡 = −𝜌 ̅̅̅̅̅̅
𝑢𝑖′ 𝑢𝑗′ = −𝜌 [ ̅̅̅̅̅̅
𝑢′ 𝑣 ′ ̅̅̅̅̅̅
𝑣 ′𝑣′ ̅̅̅̅̅̅
𝑣′ 𝑤′ ] ( 2)
̅̅̅̅̅̅
𝑢′ 𝑤 ′ ̅̅̅̅̅̅
𝑣 ′𝑤′ ̅̅̅̅̅̅̅
𝑤′𝑤′
With the bar denoting the time average (refer to the nomenclature for definition of symbols).

The challenge is to model the Reynolds stress tensor 𝑇𝑡 in terms of the mean flow quantities, and hence
provide closure to the governing equations. In STAR-CCM+ there are two possibilities[18]:

 Eddy Viscosity models


 Reynolds stress transport models (RST)

The Reynolds stress transport models (RST) are also known as second-order (or moment) closure
models. These are the most complex turbulence models in STAR-CCM+ and they account for effects of
anisotropy. This modelling approach originates from the work of Launder (1975) [13] where the eddy
viscosity approach is discarded and the 𝑇𝑡 is directly computed.

2
𝑇𝑡 = 2 𝜇𝑡 𝑆𝑖𝑗 − 3 𝜌𝑘𝛿𝑖𝑗 ( 3)

1
𝑆𝑡 = 2 (𝛻𝑣 + 𝛻𝑣 𝑇 ) , Mean Strain Tensor ( 4)

In the Eddy Viscosity models the concept of a turbulent viscosity 𝜇𝑡 is used to model the 𝑇𝑡 as a function
of mean flow quantities. While some simpler models rely on the concept of mixing length to model the
turbulent viscosity in terms of mean flow quantities, eddy viscosity models in STAR-CCM+ solve
additional transport equations for scalar quantities that enable the turbulent viscosity 𝜇𝑡 to be derived.
[13][14]

Eddy Viscosity models can be separated in linear and non-linear, with linear being of more interest for
this thesis. Linear Eddy viscosity models can be separated in algebraic models (zero-equation models),
one-equation models, and two-equation models. The two equation models are one of the most common
type of turbulence models used, like K-Epsilon or K-Omega.[13][14]

8
In this work the Eddy Viscosity model K-Omega was used. K-Omega model is an alternative of the K-
Epsilon models where the transport equations are solved for the turbulent kinetic energy 𝑘, and the
specific dissipation rate per unit turbulent energy 𝜔. [15] One advantage over the K-Epsilon models is
that the boundary layer is well solved under adverse pressure gradients.

One major disadvantage of this model is that boundary layer computations are sensitive to the values
of 𝜔 in the free stream. This translates in extreme sensitivity to inlet boundary condition which does not
occur for K-Epsilon models.

There are two versions of the K-Omega models in STAR-CCM+:

 Standard K-Omega model


 SST K-Omega model

The first is Wilcox’s K-Omega Model [15] that was corrected in 2006 [16]. In the SST K-Omega Model the
problem of the sensitivity to free-stream is addressed by Menter [17], who recognized that the 𝜀 transport
equation could be transformed into a 𝜔 transport equation by variable substitution. This models uses both
K-Epsilon model in the free stream zones with K-Omega in the inner parts of the boundary layer. This
makes it possible to use this model in the viscous sub-layer. Therefore SST model works as a Low-Re
turbulence model.

Finally simulations were made to the aircraft using SST K-Omega turbulence model. No transition model
was selected. Although the flow may be turbulent and probably unsteady once the maximum attack angle
is reached, the simulations are for steady flow. After maximum lift is reached this model ceases to be as
reliable because separation of the boundary layer occurs. This is the reason why in some cases the value
of lift can be seen to rise again after the maximum lift was achieved, but only momentarily. This is due to
the unstable nature of stall conditions. To better study this, unsteady simulations should be done.

As said before SST Menter K-Omega model employs the k-𝜔 model in the inner region of the boundary
layer and changes to k-𝜀 for the outer region of the boundary layer. The boundary condition of the surface
is no-slip. And 𝐹2 is a function that is one for boundary-layer flows and zero for free shear layers. The
turbulent viscosity is given by:

𝑎 𝑘
𝜇𝑡 = 𝜌 𝑚𝑎𝑥(𝑎1 𝜔,𝛺𝐹 ) ( 5)
1 2

where Ω is the absolute value of the vorticity and 𝑎1 = 0.31. 𝑘 and 𝜔 are calculated by two transport
equations.[17]

The wall treatment used in these simulations is a hybrid wall treatment between viscous sub-layer and
log layer, that attempts to emulate the log layer wall treatment for coarse meshes, and viscous sub-layer
wall treatment for fine meshes. It is also formulated with the desirable characteristic of producing
reasonable answers for meshes of intermediate resolution (that is, when the wall-cell centroid falls within
the buffer region of the boundary layer).

9
Wall laws for this wall treatment are chosen automatically based on the behaviour of the turbulence
model. For this case, blended wall laws are used, which include a buffer region that smoothly blends
the laminar and the turbulent profiles together. Checking the wall y+ determines whether the near-wall
spacing is appropriate for the wall treatment that is selected for the simulation. Only by running an
exploratory simulations and plotting the value of the wall y+ is this possible

2.2. Numerical approximations

Segregated flow model was chosen, with second order convection and diffusion. This solves the flow
equations (one for each component of velocity, and one for pressure) in a segregated or uncoupled
manner. The linkage between the momentum and continuity equations is achieved with a predictor
corrector approach. This model also solves the total energy equation with the temperature as the
solved variable. Enthalpy is then computed from temperature according to the equation of state. In this
case the equation of state is given by the ideal gas theory.

2.3. Propeller model

To analyse force coefficients for an aircraft with a very large number of propellers it was decided to
use actuator disks mimicking the propeller energy input in the flow. In this model the propellers were
substituted by actuator disks that imposed a rotation and a thrust to the flow. The geometry of the
blades and flow are not resolved, instead the rotor is modelled as a thin disk inducing a constant
velocity along the axis of rotation and a pressure difference. This model is more practical when the
propeller’s information is available and its effect to the surrounding area is desirable. The rotor is
simulated through source terms in the momentum equations.[18]

There are currently three methods available in STAR-CCM+ to model fans and propellers: Body Force
Propeller Method, Blade Element Method, and 1-D Momentum Method. The Body Force Propeller
Method was chosen for these simulations. In this model the flow that is induced by the propeller
depends on the flow around the aircraft, and the flow is also affected by the propeller. This can also
be used to model the propeller propulsion for DFBI (Dynamic Fluid Body Interaction) simulation.

The method employs a uniform volume force distribution over an open cylinder. The volume force
varies with the radial direction. The radial distribution varies with the Goldstein optimum and is given
by[18]:

𝑓𝑏𝑥 = 𝐴𝑥 𝑟 ∗ √(1 − 𝑟 ∗ ) ( 6)

𝑟 ∗ √1−𝑟 ∗
𝑓𝑏𝜃 = 𝐴𝜃 . 𝑟 ∗ ( 7)
(1−𝑟ℎ′ )+𝑟ℎ′

𝑟 ′ −𝑟ℎ′
𝑟∗ = 1−𝑟ℎ′
( 8)

𝑅𝐻 𝑟
𝑟ℎ′ = 𝑅𝑃
and 𝑟 ′ =
𝑅𝑃
( 9)

10
where 𝑓𝑏𝑥 is the body force component in the axial direction, 𝑓𝑏𝜃 is the body force component in the
tangential direction, r is the radial coordinate, 𝑅𝐻 is the hub radius and 𝑅𝑃 is the propeller tip radius. 𝐴𝑥
and 𝐴𝜃 are constants given by [18]:

105 𝑇
𝐴𝑥 = . 𝜋𝛥(3𝑅 ( 10)
8 𝐻 +4𝑅𝑃 )(𝑅𝑃 −𝑅𝐻 )

105 𝑄
𝐴𝜃 = . ( 11)
8 𝜋𝛥𝑅𝑃 (3𝑅𝐻 +4𝑅𝑃 )(𝑅𝑃 −𝑅𝐻 )

where T is the thrust and Q the torque.

Even though the force distribution inputted in the actuator disk (equations (6) and (7)) depends only
on the radius, the variation of the force distribution with angle 𝜃 is considered in the iteration process.
To be able to model the propeller, a propeller performance curve is needed in which an operation point
was chosen selecting a thrust. JavaPROP[11] was the program used to build this performance curve
for the desired propeller. JavaPROP program is based on the formulas of Adkins and Liebeck’s work
of “Design of Optimum Propellers” [19]. It is based on the theory of optimum propeller as developed
by Betz, Prandtl and Glauert. Only a small number of parameters need to be specified. These are:

 Number of blades, B
 Axial velocity v of the flow (flight speed or boat speed),
 Diameter 𝐷0 of the propeller,
 Selected distribution of airfoil lift and drag coefficients Cl and Cd along the radius. In this
case the distribution is not directly selected but an airfoil is selected at each station (or
radius) with the angle of attack.
 Desired thrust T or the available shaft power P,
 Density 𝜌 of the medium

The disk is applied directly to the original mesh specifying the diameter, thickness and orientation. The
performance table consists of the advance ratio, propeller efficiency, thrust coefficient 𝐾𝑇 , and torque
coefficient 𝐾𝑄 . Two tables were created for the different propellers (inner and tip), and are shown in
the appendix of this thesis (Tables 25 and 26). The characteristics in the tables are given as: [11][18]

𝑉∞
𝐽= ( 12)
𝑛 𝐷𝑂

𝑉∞
𝑃=𝑇∙ ( 13)
𝜂

𝑇
𝐾𝑇 = 𝜌𝑛2 𝐷4 ( 14)
𝑂

𝑄
𝐾𝑄 = 𝜌𝑛2 𝐷5 ( 15)
𝑂

where 𝑉∞ is the flow velocity, 𝐷𝑂 is the propeller diameter, and 𝑛 the rotation rate.

Since the operation point is given by the thrust, the advance ratio is calculated by solving the following
equation numerically:

11
𝑓(𝐽) = 𝐾𝑇 − 𝐾𝑇′ ( 16)

𝐽2 𝑇𝑜𝑝𝑒𝑟𝑎𝑡𝑖𝑛𝑔 𝑝𝑜𝑖𝑛𝑡
𝐾𝑇′ = 𝜌 2 2 ( 17)
𝑖𝑛𝑓𝑙𝑜𝑤 𝑝𝑙𝑎𝑛𝑒 𝑉∞ 𝐷0

After resolving for the advance ratio, 𝐾𝑇 and 𝐾𝑄 are interpolated from the specified propeller
performance curve. And finally the force components 𝑓𝑏𝑥 and 𝑓𝑏𝜃 are calculated using the equations 8
and 9.

12
3. Numerical Solution

3.1. Geometry and domain

The whole aircraft was designed using a CAD design program called CATIA V5. For faster
computational analyses, the simulations were done to only one wing without the fuselage. The wing
incorporated the nacelles of the propellers, and actuator disks were used in replacement of the
propellers blades for simplification reasons. The flow properties shown in the table 3 are for a velocity
of 110.6 m s-1 , and at the altitude of 3048 meters.

PROPERTY VALUE
AIR DENSITY 0.9046 kg m-3
DYNAMIC VISCOSITY 1.7115 x10-5 Pa s
PRESSURE 69680 Pa
SPEED OF SOUND 328.4 m s-1
TEMPERATURE 268.3 K
FLOW VELOCITY 110.6 ms-1 (= 215 knots)

Table 3 - Flow Properties for 3048 meters [20]

PARAMETER VALUE
CELL TYPE Hexahedral
Nº OF CELLS 15.05 Million Cells
BASE SIZE 0.2 m
NUMBER OF PRISM LAYERS 20
PRISM LAYER STRETCHING 1.3
PRISM LAYER TOTAL THICKNESS 1 % the base size = 0.002 m
FIRST PRISM LAYER HEIGHT 3.2032 x 10-5 m
VOLUME GROWTH RATIO 4 equally spaced cells per transition
MAXIMUM CELL SIZE 3.0 m
WAKE REFINEMENT RELATIVE SIZE ON 10 % the base size = 0.02 m
TRAILING EDGE
WAKE REFINEMENT LENGTH 5.0 m
SURFACE CURVATURE 144 Points per circle
SURFACE GROWTH RATE 1.2
SURFACE MINIMUM SIZE 2 % the base size = 0.004 m
SURFACE TARGET SIZE 200 % the base size = 0.4 m
CYLINDRICAL TRIMMER MESH SIZE 10 % the base size = 0.02 m

Table 4 - Mesh Parameters

The chosen mesh for this simulations has the properties presented in table 4. The specifications of the
aircraft are summarized in table 5. To achieve a better understanding of the effects of the different
configurations, a similar contemporary standard aviation aircraft, Cirrus SR22, was used to draw

13
conclusions. Its specifications are also presented to show similarity between planes of the same class.
These specifications were taken from the user manual and the official makers’ website [7] [8]. Flap angle
was maintained at 0° for all configurations so 𝐶𝐿 was not altered in any way.

SAT Cirrus SR22


Wing Area (x2 wings) 5,529 x 2 = 11,058 m2 13,5 m2
Wing Span (x2 wings) 6,497 x 2 = 12.994 m 11,7 m
Aspect Ratio 15,27 10,1
Cruise Air Speed 215 KTAS = 398,18 km h-1 211 KTAS = 391 km h-1
Maximum Take-off Altitude 3048 m 3048 m
Flap Angle 0 degrees -
Horse Power 300 kW 231,2 kW
Chord root 1m -
Chord tip 0,6 m -
Dihedral ν = 3°
Taper ratio λ = 0,6
Sweep
Leading edge 𝜑𝐿𝐸 = 0°
25% chord 𝜑0.25 = 1°
Trailing edge 𝜑 𝑇𝐸 = 3°

Table 5 - SAT Design Specifications[7][8]

To model the virtual disks, a propeller curve was necessary to input into STAR-CCM+. As previously
explained in the propeller model part, a program called JavaPROP was used to make the performance
curves for the design conditions of the propellers. In table 6 the design conditions of the propellers are
presented. These are the input conditions of JavaPROP. Some of these conditions were imposed in the
beginning of the project, other conditions like the rotation rate were chosen to best suit the needs of the
plane. The value chosen for the rotation rate was based on other researches of electric propelled planes.
[11] The airfoil of the propellers’ blade was chosen to be the most suitable [11] and standard airfoil used,
but since there isn’t a lot of research in electric propellers, and much less experimental results, these
conditions are not meant to be the optimal case for this type of application. Furthermore, propeller
optimization escapes the objective of this work and the focus was made more on the effects for the
different parametric configurations. The optimum point shown on table 6 was chosen by JavaPROP
(Hepperle, n.d.) [11] as the best suited for that propeller for higher efficiency, although the real operating
point used during simulations was chosen to meet the power needed. It is presented to show that the
operation point is not far from the optimum. In the end two table were created, one for the inner (“high-
lift”) propellers and one for the tip (“cruise”) propellers. Both tables appear in the annex 7.1. of this thesis
(tables 24 and 25) and contain the information output from JavaPROP, but the tables used for the
simulations are only a fraction of this table. A smaller table only containing the advance ratio, efficiency,
thrust coefficient and torque coefficient is imported to STAR-CCM+. The rest of the configurations
parameters for the propeller needed for STAR-CCM+ are inner and outer diameter, disk’s location, disk’s

14
normal direction and thrust at the operating point. The thrust was chosen to make the power the desired
value.

When the diameter of the disk was altered in the simulations, the propeller curve remained the same,
although it was created, and best suited, for the reference diameter. This decision was made because
a new curve would make it impossible to compare between results since it would not be the same
propeller anymore. In other words, it would alter too many variables.

Inner Tip
Number of Blades 5 3
Design RPM 4000 3000
Outer diameter (m) 0.75 1.5
Inner diameter (m) 0.056 0.292
Airspeed (m/s) 110.6 110.6
Power (kW) 20 50
Airfoil MH 114 13% Re = 500 000 MH 114 13% Re = 500 000
Twist of angle of attack 6º -> 1º 6º -> 1º
Operation point
Thrust (N) 159.05 409.13
RPM 3321 2234
Torque (Nm) 47.75 159.15

Table 6 - Propeller Design Conditions

Only for the Reference solution (solution to which all other solutions were compared to, as described in
table 9) the range of angle of attack was tested from 0 to 12 degrees to better show the curve of the lift
coefficient. 2D simulations were also made. (Refer to the annex 7.2.)

Figure 4 shows the geometry of the whole plane with propellers. It is noticeable thate “inner” propellers
only have three blades instead of five. This is because it was later changed in the project to have five
blades instead of three.. Figure 5 presents the distances between the propellers.

15
Figure 4 - Complete Plane Geometry

Figure 5 - Propellers locations

3.2. Boundary conditions

Figure 6 shows the entire control volume. On the left a semi-spherical part that is the flow inlet boundary.
Here a velocity magnitude and direction are defined. Also viewed in figure 7, there is the symmetry plane
in contact with the wing hub.. The wing and nacelles have a no-slip condition. And finally, the rest of the
surfaces (semi-cylindrical and semi-circle) are considered as pressure outlets where the flow direction
is extrapolated. Meaning that the flow direction is only controlled by the inlet and the propellers. The
outlet has to cover also the cylindrical walls because of the way the angle of attack is recreated in these
simulations. The semi-cylindrical surfaces cannot be considered walls because the inlet flow is rotated
to increase the angle of attack, not the body as it would be the case in an experimental procedure.

16
Figure 7 shows the measurements chosen for the mesh. The measurements were chosen to give
enough space between the boundaries and the body. The distance between the wing and the inlet was
chosen to be 8 times the wing span. In this case the wing span is around 7 meters, so 56 meters is the
radius of the inlet semi-sphere. The distance to the “back” outlet, marked in figure 6 as “Body1: Outlet”
has to have enough space to accommodate the wake. The length between the leading edge and the
“back” outlet is 40 meters, which is enough since the chord of the plane at the hub is around 1 meter.
[18] The geometry of the wing is shown in figure 8 with the propeller disks.

Figure 6 - Entire control volume

Figure 7 - View from the symmetry plane

17
Figure 8 - Wing with virtual disks

Figure 9 - Schematic of forces [20]

𝐶𝐿 and 𝐶𝐷 were calculated from normal and axial forces acting on the wing given by STAR-CCM+ in
the following fashion:

𝐿 = 𝑁𝑐𝑜𝑠(𝛼) − 𝐴 𝑠𝑖𝑛(𝛼) ( 18)

𝐷 = 𝑁𝑠𝑖𝑛(𝛼) + 𝐴 𝑐𝑜𝑠(𝛼) ( 19)


𝐿
𝐶𝐿 = 1 2𝑆
( 20)
𝜌𝑉∞
2

𝐷
𝐶𝐷 = 1 2𝑆
( 21)
𝜌𝑉∞
2

18
The axial force and normal force on a surface given by the Star-CCM+ are computed as:

𝑓 = ∑𝑓(𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 + 𝑓𝑆𝑆ℎ𝑒𝑎𝑟 ). 𝑛𝑓 ( 22)

𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 = (𝑝𝑓 − 𝑝𝑟𝑒𝑓 )𝑎𝑓 ( 23)

𝑓𝑆𝑆ℎ𝑒𝑎𝑟 = −𝑇𝑓 𝑎𝑓 ( 24)

Where 𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 and 𝑓𝑆𝑆ℎ𝑒𝑎𝑟 are the pressure and shear force vectors on the surface face S, 𝑛𝑓 is the
direction defined by the user, 𝑎𝑓 the face area vector, 𝑝𝑓 if the face static pressure, 𝑝𝑟𝑒𝑓 is the reference
pressure, and 𝑇𝑓 is the stress tensor at face 𝑓.[12]

3.3. Mesh Refinement

No systematic refinement was made in this case. Instead a local refinement of the mesh was done. The
meshing process was made in a first instance automatically, but was later refined reducing the base cell
size (value of the cell size form which all other cell sizes are set) and refining areas of most interest, like
the volume around the virtual disks and the wake of the wing. This means the overall mesh is
unstructured. A hexahedral grid type was chosen for all the meshes because of its simplicity in defining
wake refinement and its superior speed to mesh, although polyhedral grids are a better choice for wider
ranges of angle of attack and better body contour. Polyhedral mesh was not used not only for refinement
ease but also because hexahedral meshes require less computing time. The mesh is able to be refined
near the wing surface and coarse in the far field by controlling the inlet and outlet surfaces cell size.

A prism layer is also present near the wing surface. It’s thickness depends on the base cell size of the
entire mesh. Having 20 layers, all amounting to 1% of the base cell size with a 1.3 stretching ratio. The
angle of the wake refinement in all the figures 10 to 19 is 8 degrees. This is only an example of the angle
of the wake that matches the angle of attack for visualization purposes. The wake refinement had the
same angle as the angle of attack in every simulation, and is extended for 8 meters behind the wing.[18]

It is also to note that problems arose with the creation of the mesh because of errors in the geometry.
Although it seemed okay in CATIA V5, the geometry had holes and corrections had to be made. These
corrections were made in STAR-CCM+ program itself. Figures 20 to 23 show the holes not treated and
corrected. These holes appeared near the nacelles. Other smaller holes were also present but were not
shown.

Table 7 summarizes the mesh refinement process showing how the base cell size and growth were
altered. Additional information about a local refinement is also presented. The cell growth rate for the
wake was maintained constant except from Mesh 5 where it has four equally spaced cells per transition.
For mesh 1, there is only one cell per transition (or layer). This is the initial mesh (figure 10). Every layer
closer to the wing has twice the number of cells. For example, mesh 5, 6 and 7 have four equal sized
cells per transition, while the rest have either two or one equally spaced cells per transition.

To compare between meshes, axial and normal forces coefficients were observed since these are the
values from which the rest of the results are computed from. Refinement convergence criteria was

19
decided to be 0.001 (<10−3 ) for both aerodynamic coefficients (𝐶𝑁 and 𝐶𝐴 ). It is important to note that
these simulations were made many times in order to pin-point the angle of maximum lift, meaning that
a too refined mesh would take too long to simulate, so a compromise was made. Also the value of wall
y+ was analysed and seen to converge to 1. This tells that the cell size of the prism layer is well defined.

From Mesh 1 to Mesh2 the base size is reduced but only a change of the wake and surface cell size is
noticeable, because the far field boundaries surface cells have a fixed size. The same happens from
Mesh 3 to Mesh 4. When template growth changes, Mesh 2 to Mesh 3 and Mesh 4, the wing’s surface
cell size doesn’t change but the number of cells around the wing does. This is due to having more cells
per transition. Cylinders were also used to refine just the volume around the propellers in Mesh 6, and
in Mesh 7 a block surrounding the whole wing was refined (figure 16). Mesh 6 was chosen to be the
best relation between number of cells and accuracy. The refined block around the wing is not necessary.
Additionally in the table, there can also be observed that the value of wall y+ is near the value of 1, which
means the value of 2 mm for the Prism Layer Thickness is well chosen for this problem for Mesh 6. The
difference between Mesh 7 and Mesh 6 is 0.0002 (0.4%) for the normal and axial force coefficients.

The wake refinement is linked with the base size for the whole mesh that was changed according to the
table below. The template growth was also changed and it affected mostly the wake of the aircraft as
seen by comparing figure 11 to 12, and figure 13 to 14. Finally the surface size is in the minimum cases
2% of the base size and the first cell of the boundary layer (first cell on the wing surface) has a size of
3.2032 x 10-5 m as seen in figure 17. The slight fluctuation of wall y+ value between Mesh4, Mesh5 and
Mesh6 is due to the unsteady nature of the flow. These simulations were made for angle of attack of 8º
where the model starts to become accurate.

Cell
Nº of
Base Local Cell growth (number Wall
Mesh cells (in 𝑪𝑨 𝑪𝑵
size refinement of cells per transition) y+
millions)
(m)
Mesh 1 1.0 - 1 0.88 -0.0034 0.5127 4.01
Mesh 2 0.5 - 1 2.38 0.0287 0.6636 2.02
Mesh 3 0.5 - 2 2.57 0.0294 0.6532 2.05
Mesh 4 0.2 - 2 12.16 0.0479 0.7418 0.95
Mesh 5 0.2 Higher wake cell 4 15.00 0.0486 0.7363 0.98
growth
Mesh 6 0.2 Cylinder 4 15.05 0.0489 0.739 0.99
Mesh 7 0.2 Block 4 18.29 0.0487 0.7388 0.99
surrounding the
wing

Table 7 - Mesh Refinement

20
Figure 10 – Initial mesh with no refinement (Mesh 1)

Figure 11 – Mesh with double the cells in the wake and surface close to the wing(Mesh 2)

21
Figure 12 - Mesh with double the cells per transition (Mesh 3)

Figure 13 - Mesh double the cells in the wake and surface close to the wing compared to
Mesh2(Mesh 4)

22
Figure 14 - Mesh with double the cells per transition compared to Mesh4 (Mesh 5)

Figure 15 – Mesh 5 with local refinement at the disk level (Mesh 6)

23
Figure 16 – Mesh 5 with local refinement on a whole box surrounding the wing (Mesh 7)

Figure 17 - Prism Layer from Mesh 6

24
Figure 18 - Volume Mesh 6 of the whole wing

Figure 19 – Detailed Mesh 6 on Nacelle

25
Figure 20 – Hole in tip nacelle

Figure 21 – Hole in tip nacelle fixed

26
Figure 22 – Hole in “high-lift” nacelle 4

Figure 23 - Hole in “high-lift” nacelle 4 fixed

3.4. Comparison with results from literature

Since verification and validation were impossible to do in this project, comparisons with other similar
works was done. Nevertheless experimental data should be done in the future of this project to validate
these results. The aircraft Cirrus SR22 was used as a comparison case because of its comparable
size and power. Cirrus SR22 design configurations were already presented in table 5. The table 8
below completes some missing information that was not needed to be show in table 5. Higher
aerodynamic efficiency is achieved, improving L/D ratio from 11 to 17. The value of lift coefficient is
close for cruise conditions, being in the same order of magnitude. Other airplanes were also added for
comparison in table 8. P2600T was added because SCEPTOR project started by modifying the plane’s
wings with DEP. [8][21]

27
Plane Cirrus SR22 Cessna 172R P2006T SAT
Gross Weight 1542 kg 1111 kg 1180 kg 1361 kg
L/D max 18 - - 26.5
L/D cruise 11 - - 17
Wing Span 11.68 m 10.97 m 11.40 m 13.00 m
Aspect Ratio 10.1 7.42 8.8 15.27
Max Cruise 94.14 m/s 62.76 m/s 69.45 m/s 110.6 m/s
Speed
Total Power 231.17 kW 119.3 kW 147 kW 300 kW
RPM 2700 rpm 2400 rpm 5800 rpm 3321 and 2234
rpm
Lift 0.30 0.40 0.30 0.59
coefficient at
cruise
Stall speed 56.54 m/s 33.34 m/s 33.95 m/s -

Table 8 - Comparison table [7][21]

Other researches like NASA LEAPTech were made with the same commercial program and turbulence
models as this work, except the propeller model that was different although virtual disks were also used.
LEAPTech project shows values of 5.2 for maximum lift coefficient at 7° for a velocity of 33.34 m/s and
40º flap (landing wing configurations), with a wing configuration similar to this project only having higher
number of propellers (18 identical propellers) and a total power of 220 kW (and also no flaps) [6]. For
cruise conditions the 𝐶𝐿 is 0.77, at a velocity of 89.41 m/s. The same order of magnitude is reported in
this project. In the LEAPTech project the 𝐶𝐿 𝑀𝐴𝑋 for 33.34 m/s and Fowler flap at 40° is equal to 5.2, and
L/D ratio of 22.4. These values cannot be compared to because of the existence of a flap, but we gain
an idea of how much more lift would be produced with the presence of a flap. [11] A similar project that
combines DEP with VTOL aircraft is S2 Joby Project also using the same process has a 𝐶𝐿 𝐶𝑅𝑈𝐼𝑆𝐸 of 0.52
for a gross weight of 907.18 kg and cruise speed of 89.4 m/s. [5]

28
4. Results

4.1. Reference Solution

CONFIGURATIONS 𝑷𝒊𝒏𝒏𝒆𝒓 𝑷𝑻𝒊𝒑 𝒙𝑳𝑬 𝑫𝒊𝒏𝒏𝒆𝒓 𝑫𝒕𝒊𝒑


REFERENCE 20 kW 50 kW 0m 0.75 m 1.5 m
X1 20 kW 50 kW 0.1m 0.75 m 1.5 m
X2 20 kW 50 kW 0.2 m 0.75 m 1.5 m
DS 20 kW 50 kW 0m 0.50 m 0.75 m
DL 20 kW 50 kW 0m 0.9 m 1.5 m
POWER I30 30 kW 50 kW 0m 0.75 m 1.5 m
T70 20 kW 70 kW 0m 0.75 m 1.5 m
T20 20 kW 20 kW 0m 0.75 m 1.5 m

Table 9 - Table of configurations

Table 9 summarizes the parametric changes made to the propellers in power and geometry in every
configuration. The results are presented both in tables and in graphical form for better viewing. Table
10 and figure 24 clearly shows for the Reference solution that 𝐶𝐿 and 𝐶𝐷 increases with the angle of
attack as it should. The angle of attack was varied from 0° to 12° for the Reference solution with a 1°
interval. For almost all of the different propellers configurations, 𝐶𝐿 seems to be reaching a maximum
value at the angle of attack of 8°. This is the reason why the results for other configurations are only
presented for angles 7° to 9° . Lift is shown to decrease after the maximum point at which stall is
happening. This model is no longer accurate after stall is reached because it ceases to be a steady flow,
producing results that do not correspond with reality. That is why it shows a slight increase of 𝐶𝐿 between
the angles of 9º and 10º even though the values of both simulations have stabilized. This slight increase
is also noticed in the drag coefficient curve. This model is enough to find the maximum angle of attack
but should not be used for higher angles.

2D simulations for NACA 2412 wing profile without nacelles were made using the program Xfoil [22]
where the maximum angle of attack was found to be at 13.8° with a 𝐶𝐿 𝑀𝑎𝑥 of 1.52 (presented in the
annex). In table 10 it can be seen that the 𝐶𝐿 𝑀𝑎𝑥 for the Reference solution is equal to 0.85 at 8°. Not
only the finite wing has a decrease of lift but the presence of propellers causes it to stall sooner at 8º.
The results of the rest of the parametric changes are presented in phases, depending on the design
modifications made.

29
𝜶 0 2 4 6 7
N 6278.4 12180.8 17974.3 22708.2 24958.6
A 366.7 63.8 -576.9 -1375.6 -1805.6
L 6278.381 12171.209 17970.722 22727.554 24992.662
D 366.661 488.817 678.305 1005.585 1249.569
Cl 0.205 0.398 0.587 0.742 0.816
Cd 0.012 0.016 0.022 0.033 0.041
𝜶 8 9 10 11 12
N 26017.6 25649.8 26117.211 25174.689 24315.525
A -2095.5 -1856.5 -2044.921 -1771.342 -1451.142
L 26056.084 25624.485 26075.529 25050.147 24085.882
D 1545.860 2178.846 2521.352 3064.759 3636.051
Cl 0.851 0.837 0.852 0.818 0.787
Cd 0.050 0.071 0.082 0.100 0.119

Table 10 - Reference Results

Presented in figures 29 to 34 taken from STAR-CCM+ are the absolute total pressure and velocity fields
in the x direction (flow direction at 8° angle of attack) on the sagittal plane that crosses the nacelle
closest to the hub (“high-lift”), and parallel planes to the right and left side respectively of the nacelle.
Figures 30,31, 33 and 34 show the difference caused by the rotation of the propeller on each side of the
propeller. First a higher pressure is noticeable due to the propeller’s existence, creating a box of higher
absolute total pressure around the wing (figures 29, 30 and 31). This pressure value takes into account
the kinetic turbulence energy,that is why it is maximum in the wake. Streamlines are shown to better
visualize the vortices created in figures 35 and 36, and vorticity isosurface in figure 38. In these figures
the tip propellers creates clearly more vortex due to having higher thrust (table 6). Vorticity isosurfaces
are also shown below in figure 38. It can also be seen in figure 37 that the value of wall y+ is almost
always below or near 1, only having a maximum value of 1.03, strengthening that the mesh refinement
was well done. Finally the residuals and normal force monitor are shown in figures 26 and 27 to prove
that convergence was obtained. Convergence of residuals,lift and drag forces is shown in figures 26
and 27. Converge does not equate to accuracy but presents that the simulations were run for the
sufficient amount of time.

30
0,9

0,8

0,7

0,6
Lift Coefficient

0,5

0,4

0,3

0,2

0,1

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 24 - Lift coefficient for Reference solution

0,14

0,12

0,10
Drag Coefficient

0,08

0,06

0,04

0,02

0,00
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 25 - Drag coefficient for Reference solution

31
Figure 26 – Residual values for Reference Solution at 8°

Figure 27 – Normal force monitor for Reference Solution at 8°

32
0,90

0,80

0,70

0,60
Lift coefficient

0,50

0,40

0,30

0,20

0,10

0,00
0,00 0,01 0,02 0,03 0,04 0,05 0,06 0,07 0,08 0,09 0,10

Drag Coefficient

Figure 28 - Polar graph for Reference solution

Figure 29 - Absolute total pressure field on propeller plane for Reference solution at 8° (inlet absolute total
pressure of 75480 Pa)

33
Figure 30 - Absolute total pressure field left to propeller plane for Reference at 8°

Figure 31 - Absolute total pressure field right to propeller plane for Reference solution at 8°

34
Figure 32 - Velocity field propeller plane for Reference solution at 8°(inlet velocity of -110.3 m/s)

Figure 33 - Velocity field right to propeller plane for Reference solution at 8°

35
Figure 34 - Velocity field left to propeller plane for Reference solution at 8°

Figure 35 - Streamlines for Reference solution at 8°

36
Figure 36 - Streamlines for Reference solution at 8°

Figure 37 - Wall Y+ for Reference solution at 8°

37
Figure 38 - Vorticity Isosurface for Reference solution at 8°

0,004

0,0035
Normal Force Coefficient

0,003

0,0025

0,002

0,0015

0,001

0,0005

0
0 1 2 3 4 5 6 7 8
Spanwise distance (m)

Figure 39 – Normal force coefficient along the span for Reference solution case at 8°

38
0,9

Accumulated Normal Force Coefficient


0,8

0,7

0,6

0,5

0,4

0,3

0,2

0,1

0
0 1 2 3 4 5 6 7 8
Spanwise distance (m)

Figure 40 - Accumulated Normal force coefficient along the span for Reference solution case at 8°

0,0006

0,0004
Axial Force Coefficient

0,0002

0
0 1 2 3 4 5 6 7 8
-0,0002

-0,0004

-0,0006

-0,0008
Spanwise distance (m)

Figure 41 - Axial Force Coefficient along the span for Reference solution case at 8°

39
0,08

Accumulated Axial Force Coefficient 0,07

0,06

0,05

0,04

0,03

0,02

0,01

0
0 1 2 3 4 5 6 7 8
Spanwise distance (m)

Figure 42 - Accumulated Axial Force Coefficient along the span for Reference solution case at 8°

Figure 43 - Wall Shear Stress x-direction for Reference solution case at 7°

Figure 44 - Wall Shear Stress x-direction for Reference solution case at 8°

40
Figure 45 - Wall Shear Stress in the x-direction for Reference solution case at 9°

Figure 46 - Wall Shear Stress x-direction for Reference solution case at 10°

The last figures 39, 40, 41 and 42 show the axial and normal force along the span of the wing. Small
increases are noticeable for either the drag and the lift force because of the propellers’ presence,
although that change is not big enough to be seen in the accumulated normal force case, but enough
for the accumulated axial force. This is because of a low angle of attack. Rising lift from the hub to the
tip is seen as it was expected, reaching the same maximum value in table 10.

Four more pictures of the wall shear stress and the velocity profile are shown. Figure 43, 44, 45 and 46
show the wall shear stress on the wing that is in the positive direction of the x-axis (towards the leading
edge). From 7° to 10°, the blues areas on the wing are becoming larger meaning the separation of the
boundary layer grows with the increase of angle attack. In figure 44 (8°) the separation of the boundary
layer is in each side of every propeller. Only in figure 45 (9°) the two outer most “high-lift” propellers
develop big separation on the left hand side (also seen in figures 31 and 34). Figure 46 (10°) just
confirms that separation keeps growing.

4.2. X1 and X2

For X1 configuration 𝐶𝐿 𝑀𝑎𝑥 value remains at 0.85 at an angle of 8 degrees. The variations of the lift
coefficient are seen to vary 0.004 (0.48%) (figure 47 and 48) from the Referencesolution, while the drag
coefficient varies 0.001 (2%) with 𝐶𝐷 being 0.0508 (figure 49). The case of X2 has no decrease of 𝐶𝐿 𝑀𝑎𝑥 ,
remaining at a value of 0.85 to an angle of also 8 degrees, and the drag coefficient is unaltered. Figure
48 shows 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 ratio of both cases.

41
The changes for both cases are not big enough to conclude a significant difference, which leads to
believe that these changes in the configuration should not be considered important for the design of the
propellers.

7 8 9 10
N 25046.7 25896.4 25350.9 23897.4
A -1823.7 -2070.5 -1838.1 -1422.8
L 25082.2 25932.5 25326.3 23781.4
D 1242.3 1553.7 2150.3 2748.6
Cl 0.8194 0.8472 0.8274 0.7769
Cd 0.0406 0.0508 0.0702 0.0898

Table 11 - X1 Results

7 8 9
N 25099.6 26018.6 25966.0
A -1839.4 -2098.3 -1875.4
L 25136.70 26057.4 25939.7
D 1233.1 1543.2 2209.6
Cl 0.8212 0.8513 0.8475
Cd 0.0403 0.0504 0.0722

Table 12 - X2 Results

0,9

0,8

0,7

0,6
Lift Coefficient

0,5
DEFAULT
0,4 X1
X2
0,3

0,2

0,1

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 47 – Lift coefficient for the X-direction change cases

42
1,02

1,00
X1 X2

0,98
CL / CLref

0,96

0,94

0,92
6 7 8 9 10 11
Angle of attack (°)

Figure 48 – 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 vs angle of attack for the X-direction change cases

0,14

0,12

0,10
Drag Coefficient

0,08

0,06 DEFAULT
X1
0,04
X2

0,02

0,00
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 49 - Drag coefficient for the X-direction change cases

4.3. Diameter change

While for the small diameter configuration, the value of 𝐶𝐿 𝑀𝑎𝑥 seems to increase to 0.88 (3.74 %)(figure
50), for the larger diameter configuration, this value drops slightly to 0.84 (0.58 %)(figure 51). The drag
coefficient doesn’t change much for both cases in relation to the Reference solution (figure 52), which
causes an increase in aerodynamic efficiency of L/D equal to 17.32 for the smaller diameters
configuration and a reduction of L/D to 16.80 for the larger diameter configuration. In this case, and any
other cases, the propeller curve was not changed, as well as the thrust. This means that the rotation
rate had to change if the diameter changes and power maintains constant. For larger diameters, the

43
rotation rate is around 2400 rpm, while in the case of the smaller diameters, propellers rotation is around
the speed of 5500 rpm. Both the area affected by the propeller and the rotation rate change the drag
caused by the propeller. However these two values change differently and in opposite directions. This
explains why the drag for the Big Diameter case is the same as the Reference solution even though the
affected area is bigger. In the case of the Small Diameter case the higher rotation rate increases the
drag slightly even though less area is affected by the propellers. Looking at the lift the Small Diameter
case has higher values, since the flow over the wing is less disturbed and higher pressure caused by
the propellers increases the pressure difference between sides of the airfoil.

7 8 9
N 25341.1 26986.2 25441.7
A -1834.1 -2216.5 -1897.1
L 25375.7 27032.0 25425.2
D 1267.9 1560.8 2106.2
Cl 0.8290 0.8831 0.8306
Cd 0.0414 0.0510 0.0688

Table 13 - Small Diameter Results

7 8 9
N 24986.3 25868.0 24982.5
A -1832.4 -2078.4 -1775.2
L 25023.4 25905.5 24952.6
D 1226.4 1542.0 2154.8
Cl 0.8175 0.8463 0.8152
Cd 0.0401 0.05038 0.0704

Table 14 - Large Diameter Results

44
1,0

0,9

0,8

0,7
Lift Coefficient

0,6

0,5 DEFAULT
DS
0,4 DL

0,3

0,2

0,1

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 50 - Lift coefficient for Diameter change case

1,06

1,04

1,02
CL / CLref

1,00

DS DL
0,98

0,96
6 7 8 9 10
Angle of attack (°)

Figure 51 - 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 vs angle of attack for Diameter change case

45
0,14

0,12

0,10
Drag Coefficient

0,08

0,06
DEFAULT
0,04 DS
DL
0,02

0,00
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 52 - Drag coefficient for Diameter change case

4.4. I30

In this configuration, only the inner propellers have more power, while the tip propeller remain the same
as the Reference solution. As can be observed by the table 15, the lift coefficient 𝐶𝐿 increases to 0.87
at 8° of angle of attack. It’s an increase of 0.022 (2.6 %) (figure 53) for 50 kW of total power increase.
This corresponds to an increase of 640.9 Newtons in the lift force. The reason for this increase is due
to the higher induced velocity over the wing that ultimately generates higher lift. Obviously that also
means higher drag which in this case is a 0.0013 (2.6 %) of difference of 𝐶𝐷 for the same angles, in this
case 8°. There is an increase of L/D to 17 but is a small difference, leading to conclude this increase in
power is not worth it. The overall change in 𝐶𝐿 is not big enough (figure 54).

7 8 9
N 25357.3 26655.6 25630.7
A -1848.4 -2161.3 -1848.1
L 25393.6 26697.0 25604.3
D 1255.7 1569.5 2184.1
Cl 0.8296 0.8722 0.8365
Cd 0.0410 0.0513 0.0714
Table 15 - Results of Inner Propellers at 30 kW

46
1,0

0,9

0,8

0,7
Lift Coefficient

0,6

0,5 DEFAULT

0,4 I30

0,3

0,2

0,1

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack (°)

Figure 53 - Lift Coefficient for I30 case

1,05

1,03
CL/CLref

1,01

0,99

0,97
6 6,5 7 7,5 8 8,5 9 9,5
Angle of attack (°)

Figure 54 – 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 vs angle of attack for I30 case

47
0,10

0,09

0,08
Drag Coefficient

0,07

0,06

0,05 DEFAULT
I30
0,04

0,03

0,02
6 7 8 9 10
Angle of attack (°)

Figure 55 - Drag Coefficient for I30 case

4.5. T20 and T70

Other effects studied were the ones created by the increase and decrease of power of the tip propellers.
Both higher and lower power changes were tested. For the case of T20, the lift not only becomes bigger
to a value of 0.87, a change of 0.02 (2.4 %) (figure 56), but also happens in a higher angle of attack of
9°. However for the T70 case the change in lift is not noticeable, being 0.86 for an angle of 8°. The drag
difference is also not very noticeable for the same angle of attack with changes of -0.016 (31.9%) for
the T20 case, and 0.0003 (0.6%) for the T70 case (figure 57). Neither one of these parametric changes
bring benefits in terms of lift, but only T20 presents a delay of maximum angle of attack, concluding that
the tip (or “cruise”) influences the flow next to the “inner” propellers closer to the tip. This can be seen
by looking at figures 59 to 61 where the shear stress tensor is shown to alter in comparison to the
reference configuration for the same angles of attack. Figure 57 corroborates this conclusion showing
the ratio of 𝐶𝐿 between cases.

7 8 9 10
N 24371.8 25935.8 26672.6 26159.5
A -1845.0 -2087.6 -2160.4 -1973.6
L 24415.0 25973.9 26682.2 26104.8
D 1138.9 1542.3 2038.7 2599.0
Cl 0.7976 0.8486 0.8717 0.8528
Cd 0.0372 0.0504 0.0666 0.0849

Table 16 - Results of Tip propellers at 20 kW

48
7 8 9
N 25093.3 26160.1 25885.2
A -1836.7 -2105.3 -1865.3
L 25130.1 26198.5 25858.3
D 1235.1 1556.0 2207.0
Cl 0.8210 0.8559 0.8448
Cd 0.0404 0.0508 0.0721

Table 17 - Results of Tip propellers at 70 kW

0,90

0,85
Lift Coefficient

0,80 DEFAULT
T20
T70

0,75

0,70
6 7 8 9 10 11
Angle of attack (°)

Figure 56 - Lift Coefficient for different tip power cases

1,05

1,03
CL/CLref

1,01
T20
T70
0,99

0,97
6,5 7 7,5 8 8,5 9 9,5 10 10,5
Angle of attack (°)

Figure 57 - 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 vs angle of attack for different tip power cases

49
0,11

0,10

0,09

0,08
Drag Coefficient

DEFAULT
0,07 T20
T70
0,06

0,05

0,04

0,03

0,02
6 7 8 9 10 11
Angle of attack (°)

Figure 58 - Drag Coefficient for different tip power cases

Figure 59 – Wall Shear Stress for T20 at 8°

Figure 60 - Wall Shear Stress for T20 at 9°

50
Figure 61 - Wall Shear Stress for T20 at 10°

4.6. Velocity change

Another case to study was the effect of speed. The changes are obviously much more noticeable than
the previous ones. Two different cases of 77.2 m/s (150 knots) and 31.4 m/s (61 knots) were tested.
The maximum angle of attack still seems to be at 8° for the 77.2 m/s. Resulting in an increase of 0.042
(4.95 %) for the lift coefficient and decrease of 0.0027 (5.45 %) for the drag coefficient. This shows that
lower cruise velocity helps the plane to produce higher L/D ratios, 16.86 for the Reference solution case
and 18.71 for the 150 knots case, both at 8°.

Also shown in the same figure 62, there is the case of 31.4 m/s. This is the lowest speed tested in
LEAPTech project in order to understand if the “high-lift” propellers lower the stall speed and it shows
much better results for the 𝐶𝐿 and 𝐶𝐷 . First the maximum angle of attack is 12°, and 𝐶𝐿 and 𝐶𝐷 have
respectively, an increase of 0.409 (48.12%) and an increase of 0.098 (194.10%). Clearly this type of
“high-lift” devices behave better and bring more efficiency at lower speeds. This is because these
propeller have a higher influence at lower speeds. Figures 62 and 64 show 𝐶𝐿 and 𝐶𝐷 alphas a function
of alpha respectively. Figure 63 show also the large increase in 𝐶𝐿 already discussed.

7 8 9
N 12344.6 13280.0 12416.0
A -916.9 -1148.0 -902.0
L 12364.3 13310.5 12404.2
D 594.4 711.4 1051.4
Cl 0.8299 0.8934 0.8326
Cd 0.0399 0.0478 0.0706

Table 18 - Results of flow inlet at 150 knots

51
8 9 10 12 14
N 2764.6 2870.0 2975.0 3115.0 3055.0
A -242.8 -280.0 -283.0 -288.0 -221.0
L 2771.4 2878.5 2978.9 3106.8 3017.7
D 144.3 172.4 237.9 365.9 524.6
Cl 1.125 1.168 1.209 1.261 1.225
Cd 0.0586 0.0700 0.0966 0.1485 0.2129

Table 19 – Results of flow inlet at 61 knots

1,4

1,2

1,0
Lift Coefficient

0,8

0,6
DEFAULT
0,4
150 knts
0,2 61knots

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Angle of attack (°)

Figure 62 - Lift Coefficient for different flow speed case

1,50

1,40

1,30
CL/CLref

1,20 150 knts


61knots
1,10

1,00

0,90
6,5 7,5 8,5 9,5 10,5 11,5 12,5 13,5 14,5
Angle of attack (°)

Figure 63 – CLlCLref vs angle of attack for different flow speed case

52
0,25

0,20
DEFAULT
150 knts
Drag Coefficient

0,15
61knots

0,10

0,05

0,00
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Angle of attack (°)

Figure 64 - Drag Coefficient for different flow speed case

4.7. Influences of propellers at 31.38 m/s (61knots)

To study the influences of propellers at 31.38 m/s (61 knots), the “high-lift” propellers actuator disks
were removed one by one. In these cases the power was maintained but with one less propeller so it is
expected that 𝐶𝐿 and 𝐶𝐷 decrease. What we are looking for here is the amount of change of 𝐶𝐿 and 𝐶𝐷
and what propellers are causing more “damage”(difference) when absent. Tables 19 to 23 show the
results of each simulation. The range of the attack angle is from 8° to 14° as it was for the Reference
solution at 61 knots. The numbering of the propellers from the hub to the tip goes from 1 to 6, 6 th being
the tip propeller.

It is clear that for every following configuration the 𝐶𝐿 is lower than for the Reference solution case at 61
knots. However there is a 0.01 of difference in 𝐶𝐿 for Prop1 and Prop2, but a much higher difference of
0.05 and 0.04 for Prop4 and Prop5 respectively. There is no big and noticeable change in the drag
coefficients for any case, around 0.012 (8%) for the 1prop and 2prop cases and around 0.006 (4%) for
the other cases.

In conclusion, the propellers 4 and 5 are the most important “high-lift” propellers on the wing. This seems
to have a correlation with the distance of the propeller to body of the aircraft. The more close to the tip
of the aircraft the more lift the propeller seems to produce. However this doesn’t happen with the fifth
propeller. It seems that propeller 4 causes more damage than propeller 5. In 4prop there is a drop of
132.8 N in lift. Likewise in 5prop there is a drop of 98.6 N in lift. This may be explained with the presence
of bigger and stronger propeller on the tip of the wing. Without the 5 th propeller the tip vortex can be
better used to generate lift.

53
8 10 12 14
N 2656.0 2845.0 3087.0 2885.0
A -240.0 -270.0 -312.0 -235.0
L 2663.6 2848.7 3084.4 2856.2
D 132.0 228.1 336.6 469.9
Cl 1.081 1.156 1.252 1.159
Cd 0.0536 0.0926 0.1366 0.1907

Table 20 – Results of flow inlet at 61knots without propeller 1

8 10 12 14
N 2642.0 2863.0 3082.0 2915.0
A -232.0 -283.0 -315.0 -240.0
L 2648.6 2868.6 3080.1 2886.5
D 138.0 218.4 332.7 472.3
Cl 1.0749 1.1643 1.2501 1.1715
Cd 0.0560 0.0887 0.1350 0.1917

Table 21 - Results of flow inlet at 61knots without propeller 2

8 10 12 14
N 2658.8 2927.0 3066.0 2985.0
A -237.8 -294.0 -295.0 -266.0
L 2666.1 2933.6 3060.3 2960.7
D 134.6 218.7 348.9 464.0
Cl 1.082 1.191 1.242 1.202
Cd 0.0546 0.0888 0.1416 0.1883

Table 22 - Results of flow inlet at 61knots without propeller 3

8 10 12 14
N 2629.0 2881.0 2982.0 2861.0
A -232.0 -277.0 -275.0 -161.0
L 2635.7 2885.3 2974.0 2815.0
D 136.1 227.5 351.0 535.9
Cl 1.070 1.171 1.207 1.143
Cd 0.0552 0.0923 0.1425 0.2175

Table 23 - Results of flow inlet at 61knots without propeller 4

54
8 10 12 14
N 2653.0 2907.0 3015.0 2905.0
A -235.0 -280.0 -284.0 -205.0
L 2659.9 2911.5 3008.2 2868.3
D 136.5 229.0 349.1 503.9
Cl 1.080 1.182 1.221 1.164
Cd 0.0554 0.0930 0.1417 0.2045

Table 24 - Results of flow inlet at 61knots without propeller 5

1,30

61knots

1,25 1prop
2prop
3prop
1,20 4prop
Lift ceofficient

5prop

1,15

1,10

1,05
6 7 8 9 10 11 12 13 14 15
Angle of attack (°)

Figure 65 – Lift coefficient for propellers influences cases

55
1,00

0,99

0,98

0,97 1prop

2prop
C L/C Lref

0,96
3prop

4prop
0,95
5prop
0,94

0,93

0,92
7 8 9 10 11 12 13 14 15 16
Angle of attack(°)

Figure 66 - 𝐶𝐿 /𝐶𝐿𝑟𝑒𝑓 vs angle of attack

0,23

0,21 61knots
0,19 1prop
2prop
Drag coefficient

0,17
3prop
0,15
4prop
0,13
5prop
0,11

0,09

0,07

0,05
7 8 9 10 11 12 13 14 15
Angle of attack (°)

Figure 67 - Drag coefficient for propellers influences cases

56
5. Conclusions and Future Directions

5.1. Conclusions

Simulations for various propeller configurations were successfully made being always compared with a
reference solution. The mesh was optimized for this case until convergence of 𝐶𝐿 and 𝐶𝐷 was achieved.
Aerodynamic coefficients of lift and drag were compared respectively between configurations. Flow was
studied for all the configurations. The decision on the relevance of propeller parametric configurations
was ultimately the objective of this work, so assumptions were made to study these propellers. Since
there is not verification or validation of this work, the results obtained could not be confirmed. This means
these results should only be used to show variation between cases and not conclude on the magnitude
of the coefficients.

As can be seen in the comparison part, the values do not differ that much compared with Cirrus SR22
for the lift coefficient from the results presented having the same order of magnitude.

Results show that DEP is best suited for low velocities. Looking at the 2D results of the wing airfoil
without the propeller geometry, it can be concluded that 2D assumptions cannot be compared to 3D
cases. DEP increases 𝐶𝐿 for lower velocities. It can also be seen that the presence of propeller increases
the 𝐶𝐿 compared to a case without propeller, although stall happens sooner. As already discussed, the
left side of the propeller stalls first than the right side. This concludes that an airfoil with gentle stall
characteristics should be used to achieve higher 𝐶𝐿 instead of a airfoil with higher 𝐶𝐿𝑚𝑎𝑥 but which stall
more abruptly.

Propellers’ diameter and distance to the leading edge have not showed significant change in 𝐶𝐿 to be
conclusive about their effect. While power change in the propellers show that the increasing power of
the “high-lift” propellers is not advantageous, as it increases more drag than lift. Higher powered tip
propeller also doesn’t show benefits. However reducing the power level of the tip propeller to the same
as the “high-lift” ones show benefits because of increased stall angle of attack. This is seen to occur
because the tip propeller has a big influence in the flow near the fourth and fifth propeller. Meaning the
power difference shouldn’t be very big between propellers.

The simulations without one alternating propeller not working demonstrates again that the propellers
farther away from the hub of the wing influence more the 𝐶𝐿 . In this case, without the fourth propeller
active a lower overall 𝐶𝐿 is verified.

The advantages of STAR-CCM+ is the highly automated CAD healing software and mesh generation
abilities that allow for faster analysis. The range of turbulence models is extensive. However the mesh
refinement can greatly increase the computation time, and a lot of time and experience is needed to
optimize a mesh. The importation of the initial geometry files had errors which causes problems in the
meshing process.

This plane could have higher efficiency with a smaller propeller and potentially with higher number of
propellers. The distance from the leading edge is not important characteristic since the lift coefficient

57
difference was negligible. Optimal propellers was not the objective of this thesis a better designs may
be needed to reach better results.

5.2. Future Work

The results presented in this work need to be compared with experimental data for validation. The
propellers have to be designed for take-off operation point including flaps. This was not demanded in
this work. A take-off criteria should be made to analyse if take off without flaps is possible. Propeller
power don’t have to match and should be tested to have different thrust to maximize performance.
Counter rotating propellers should be tested so to reduce vortices created in the wake, as well as
minimizing influences of different vortices. Unsteady simulations should also be made if higher angles
of attack are of importance, and to better study the flow around the wing.

58
6. References

[1] The National Academy of Sciences (2016) Commercial Aircraft Propulsion and Energy Systems
Research: Reducing Global Carbon Emissions, pp. 53-64, The National Academies Press

[2] Arthur Dubois A. . Martin van der Geest M.V.D., JoeBen Bevirt J.B.,Clarke S., Christie R.J., Borer
N.K., (2016) Design of an Electric Propulsion System for SCEPTOR. Joby Aviation, NASA. USA

[3] Kim H.D.. (2011). Distributed Propulsion Vehicles. 27th International Congress of Aeronautical
Sciences, NASA Glenn Research Center. Cleveland USA, ICAS 2010

[4] Felder J.L.. Brown G.V.. Kim H.D.. and Chu J.. (2011) Turboelectric Distributed Propulsion in a
Hybrid Wing Body Aircraft. NASA Langley Research Center and NASA Glenn Research Center.
Hampton and Cleveland USA, ISABE-2011-1340, E-18064

[5] Stoll A. M. . Bevirt J.B.. Pei P.P. . and Stilson E.V. . (2014). Conceptual Design of the Joby S2
Electric VTOL PAV. Joby Aviation. Santa Cruz. California, Aviation Technology, Integration, and
Operations Conference

[6] Stoll A. M. . Bevirt J.B.. Pei P.P., Moore M.D., Fredericks W.J., and Borer N.K., (2014). Drag
Reduction Through Distributed Electric Propulsion. Joby Aviation. Santa Cruz. California. Aviation
Technology, Integration, and Operations Conference

[7] Cirrus Aircraft Official Website, Accessed on 16-06-2016 at: http://cirrusaircraft.com/aircraft/sr22/

[8] Cirrus Aircraft Corporation, (2011), Airplane information manual for Cirrus Design SR22, Duluth,
Minnesota

[9] Continental Motors 550 Series Specifications Sheet, Accessed on 02-12-2016 at:
http://www.continentalmotors.aero/uploadedFiles/Content/xImages/550Series-SpecSheet-WEB.pdf

[10] ENSTROJ EMRAX 207 Series Manual Specification Sheet, Accessed on 02-12-2016 at
http://www.enstroj.si/Electric-products/emrax-200.html

[11] Hepperle M.. (2013). Java PROP – Design and Analysis of Propellers. Accessed in 15-03-2016
at: http://www.mh-aerotools.de/airfoils/javaprop.htm

[12] Tennekes H. and Lumley J.L., (1972) A First Course in Turbulence, 1st edition, MIT Press.

[13] Ferziger J.H.. Peric M.. (2002) Computational Methods for Fluid Dynamics. 3rd Edition. Springer.
Berlin. pag. 265-307.

[14] Rumsey C.. (2016). Turbulence Modeling Resource. Langley Research Center. NASA. Accessed
on 23-07-2016 at : https://turbmodels.larc.nasa.gov

[15] Wilcox. D.C. (1998). Turbulence Modeling for CFD. 3rd edition. DCW Industries. Inc.

59
[16] Wilcox. D.C. (2008). Formulation of the k-𝜔 turbulence model revised. AIAA Journal. Vol. 46 No.
11 pp. 2823-2838.

[17] Menter. F.R. (1994). Two-equation eddy-viscosity turbulence modelling for engineering
applications. AIAA Journal. Vol. 32 No. 8. pp 1598-1605.

[18] STAR-CCM+ User Guide. Star-CD version 4.02 (2006)

[19] Adkins C. N.. Liebeck R.H.. (1994) Design of Optimum Propellers. Douglas Aircraft Company.
Long Beach. California. Journal of Propulsion and Power. Vol 10. No. 5.

[20] Aerospaceweb Accessed on 08-04-2013 at: http://www.aerospaceweb.org/

[21] Tecnam P2006T Light Twin General Aviation Aircraft Flight Manual (2010) . 2nd Edition.

[22] XFoil Subsonic Development System Website, Drela M.. Youngren H., Accessed on 06-04-2016
at: http://web.mit.edu/drela/Public/web/xfoil/

60
7. Appendix
7.1. JavaPROP tables
Tables 24 e 25 are the performance tables generated by JavaProp for both cases of inner and tip
propellers for the conditions described in table 6.

61
Table 25 - Inner Propellers Performance table

62
Table 26 - Tip Propellers Performance table

63
7.2. 2D Simulations on Xfoil

Tables 26 e 27 are the generated tables of Xfoil simulations for Naca2412 profile at different flow
speeds of 110.6 and 31.38 m/s.

2D NACA 2412 MACH = 0.337 RE = 1.8026E+6


ALPHA CL CD CM
0 0,2565 0.0054 -0.0562
0,5 0,3153 0.00535 -0.0563
1 0,3742 0.00531 -0.0563
1,5 0,4329 0.00521 -0.0563
2 0,4902 0.00524 -0.0562
2,5 0,5461 0.00533 -0.056
3 0,6033 0.00547 -0.0557
3,5 0,671 0.00568 -0.0551
4 0,7444 0.00595 -0.0543
4,5 0,8223 0.00628 -0.053
5 0,8696 0.0067 -0.0511
5,5 0,9168 0.00719 -0.0505
6 0,9661 0.00823 -0.0515
6,5 1,0173 0.00873 -0.0497
7 1,0694 0.00925 -0.048
7,5 1,122 0.00978 -0.0464
8 1,174 0.01034 -0.0446
8,5 1,225 0.01096 -0.0426
9 1,2744 0.01165 -0.0402
9,5 1,3212 0.01234 -0.0377
10 1,3604 0.01318 -0.0345
10,5 1,3976 0.01404 -0.031
11 1,427 0.01499 -0.0261
11,5 1,457 0.01623 -0.0207
12 1,4807 0.01771 -0.0153
12,5 1,4947 0.01942 -0.0103
13 1,5086 0.02162 -0.0054
13,5 1,5147 0.02446 -0.0009
14 1,5127 0.02827 0.0029
14,5 1,4979 0.03364 0.0057
15 1,4699 0.0398 0.0072
15,5 1,4429 0.04746 0.0071
16 1,4092 0.05677 0.0055
16,5 1,3719 0.06739 0.0026

Table 27 - 2D cruise wing profile Xfoil simulation for 110.606 m/s

64
2D NACA 2412 MACH = 0.092 RE = 5.1141E+5
ALPHA CL CD CM
0 0,2307 0.00549 -0.0528
0,5 0,2821 0.00549 -0.0528
1 0,3452 0.00535 -0.0526
1,5 0,4193 0.00524 -0.0525
2 0,5023 0.00519 -0.0522
2,5 0,5664 0.00513 -0.0517
3 0,6144 0.00517 -0.0507
3,5 0,6623 0.0053 -0.0495
4 0,7104 0.00559 -0.0498
4,5 0,7584 0.006 -0.0517
5 0,8065 0.00651 -0.0552
5,5 0,8543 0.00707 -0.056
6 0,901 0.00766 -0.0539
6,5 0,9465 0.00827 -0.0521
7 0,9904 0.00891 -0.0505
7,5 1,0331 0.00952 -0.0493
8 1,0755 0.01011 -0.0482
8,5 1,1175 0.0107 -0.0472
9 1,1564 0.0113 -0.0461
9,5 1,1939 0.01192 -0.045
10 1,2306 0.0126 -0.0437
10,5 1,259 0.01329 -0.0424
11 1,2792 0.01399 -0.0409
11,5 1,306 0.01481 -0.039
12 1,322 0.01562 -0.0371
12,5 1,3386 0.01649 -0.0347
13 1,3554 0.01767 -0.0306
13,5 1,3608 0.01865 -0.0274
14 1,367 0.01987 -0.0239
14,5 1,3732 0.0214 -0.0205
15 1,376 0.02367 -0.0165
15,5 1,3748 0.02563 -0.0138
16 1,3644 0.0281 -0.0113
16,5 1,3581 0.0311 -0.0091
17 1,3474 0.03478 -0.0075
17,5 1,3333 0.03937 -0.0064
18 1,3172 0.04532 -0.0061
18,5 1,2997 0.05334 -0.0071
19 1,2814 0.06183 -0.0089
19,5 1,2625 0.07068 -0.0114
20 1,2395 0.08093 -0.0147
Table 28 - 2D Take-off Xfoil Simulation

65
To see clearly the change between the 2D and the 3D cases, a study case of 2D simulations was done
for the wing profile. The figure for the 2D scenarios follows showing a 𝐶𝐿 𝑀𝑎𝑥 of 1.52 at 13.5° for 110.6
m/s, and 𝐶𝐿 𝑀𝑎𝑥 of 1.38 at 15° for 31.38 m/s. As it was to be expected none of the 3D cases present
higher values of 𝐶𝐿 𝑀𝑎𝑥 , this is because of induced drag and the presence of nacelles. The 2D simulations
were done with Xfoil while STAR-CCM+ was used for the 3D case. A big reduction in the drag is seen
to occur as well as a reduction of the angle of attack. This was done to have an idea of how the wing
would react but as seen from the results, the lift coefficient is completely different. While the increase in
maximum angle of attack is seen to occur for lower velocities, in 3D the 𝐶𝐿𝑚𝑎𝑥 increases value, while in
2D the opposite happens. This serves to show that 3D simulations for this work differ a lot from 2D
simulations, and other effects have to be accounted for.

CL
1,6

1,4

1,2

110,606 m/s
0,8
31.38 m/s

0,6

0,4

0,2

0 α
0 5 10 15 20 25

Figure 68 - 2D Simulations

66
7.3. No Propulsion

1,0

0,9

0,8

0,7
Lift Coefficient

0,6

0,5

0,4
DEFAULT NoProp
0,3

0,2

0,1

0,0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Angle of attack

Figure 69 - No propulsion case compared to reference case

The case of the geometry with no propulsion was also studied and compared with the reference case
solution. The geometry was maintained with the nacelles and inflow velocity of 110,6 m/s, only not
considering virtual disks. In this case the maximum angle of attack is near 10° but has a lower slope.
This means the overall lift generated is higher for the case with DEP. However stall occurs sooner.
This can also be explained by the different local angle of attack of the wing cause by the propellers.
Further studies of DEP with flaps should be made to acquire more data.

𝜶 0 2 4 6 8
NF 5424 10868 16022 21003 25450
AF 310 66 -457 -1256 -2005
L 5424,0 10859,1 16014,8 21019,2 25481,4
D 310,0 445,2 661,8 946,3 1556,5
CL 0,177 0,355 0,523 0,687 0,832
CD 0,010 0,015 0,022 0,031 0,051
𝜶 9 10 11 12
NF 27401 28604 26606 24709
AF -2650 -2630 -2180 -1830
L 27478,2 28626,1 26533,1 24549,5
D 1669,1 2377,0 2936,7 3347,3
CL 0,898 0,935 0,867 0,802
CD 0,055 0,078 0,096 0,109
Table 29 - No propeller results

67
Figure 70 - Wall Shear Stress in the positive direction of x for No propulsion case at 10 °

7.4. Optimization design process in JAVAProp

To be able to model the propeller, a propeller performance curve is needed in which an operation point
was chosen selecting a thrust. JavaPROP [11] was the program used to build this performance curve
for the desired propeller. JavaPROP program is based on the formulas of Adkins and Liebeck’s work
of “Design of Optimum Propellers” [19]. It is based on the theory of optimum propeller as developed
by Betz, Prandtl and Glauert. The design proceeds in the following summarized steps:

1) Select an initial estimate for 𝜁 = 𝑣 ′ /𝑉 . the displacement velocity ratio (𝜁 = 0 will work).
2) Determine the values for F (Prandtl momentum loss factor) and 𝜙 (flow angle) at each blade
station by:

2
𝐹 = 𝑎𝑟𝑐𝑐𝑜𝑠(𝑒 −𝑓 ) ( 25)
𝜋

𝐵
𝑓 = (1 − 𝜉)𝑠𝑖𝑛𝜙𝑡 ( 26)
2

𝑡𝑎𝑛 𝜙𝑡 = 𝜆(1 + 𝜁/2) ( 27)

𝑡𝑎𝑛 𝜙 = (𝑡𝑎𝑛𝜙𝑡 )/𝜉 ( 28)

where 𝜙𝑡 is the flow angle at the tip. 𝜆 is a constant and 𝜉 is the nondimensional radius r/R.
3) Determine the product 𝑊𝑐. and Reynolds number from:

𝑊𝑐 = 4𝜋𝜆𝐺𝑉𝑅𝜁/𝐶𝑙 𝐵 ( 29)

Where G is circulation function given by 𝐺 = 𝐹 × 𝑐𝑜𝑠𝜙𝑠𝑖𝑛𝜙.


4) Determine 𝜀 (drag-to-lift ratio)and 𝛼 (angle of attack) for the airfoil section data. This is selected
by the user.
5) If 𝜀 is to be minimized. change 𝐶𝑙 and repeat steps 3 and 4 until this is accomplished at each
station.
6) Determine a (axial interference factor) and a’ (rotational interference factor). and W (local total
velocity) from:

𝑎 = (𝜁/2) 𝑐𝑜𝑠 2 𝜙 (1 − 𝜀𝑡𝑎𝑛𝜙) ( 30)

68
𝑎′ = (𝜁/2𝑥) 𝑐𝑜𝑠𝜙𝑠𝑖𝑛𝜙(1 + 𝜀/𝑡𝑎𝑛𝜙) ( 31)

𝑊 = 𝑉(1 + 𝑎)/𝑠𝑖𝑛𝜙 ( 32)

where 𝑥 = Ωr/V
7) Compute the chord from step 3 and the blade twist 𝛽 = 𝛼 + 𝜙
8) Determine the four derivatives in I and J (advance ratio) and numerically integrate these from
𝜉 = 𝜉0 to 𝜉 = 1 from:

𝐼1′ = 4𝜉𝐺(1 − 𝜀𝑡𝑎𝑛𝜙) ( 33)

𝐼2′ = 𝜆(𝐼1′ /2𝜉)(1 + 𝜀/𝑡𝑎𝑛𝜙)𝑠𝑖𝑛 𝜙𝑐𝑜𝑠𝜙 ( 34)

𝐽1′ = 4𝜉𝐺(1 + 𝜀/𝑡𝑎𝑛𝜙) ( 35)

𝐽2′ = (𝐽1′ /2)(1 − 𝜀𝑡𝑎𝑛𝜙) 𝑐𝑜𝑠 2 𝜙 ( 36)

where the prime denotes derivatives with respect to 𝜉.


9) Determine 𝜁 and 𝑃𝐶 (power coefficient. 2𝑃/𝜌𝑉 3 𝜋𝑅2 ) from equations:

1
𝐼1 𝐼1 2 𝑇𝑐 2
𝜁=( ) − [( ) − ] ( 37)
2𝐼2 2𝐼2 𝐼2

𝑃𝐶 = 𝐽1 𝜁 + 𝐽2 𝜁 2 ( 38)

where 𝑇𝐶 is the thrust coefficient given by 2𝑇/𝜌𝑉 3 𝜋𝑅2 .


10) If the value of 𝜁 is not sufficiently close to the old one. within 0.1%. start over from step 2 using
new 𝜁.
11) Determine propeller efficiency as 𝑇𝐶 /𝑃𝐶 . and other features such as solidity.

The above steps take around three or four cycles to converge. This is an extension of the propeller
theory of Glauert to improve design of optimal propellers and refine the calculation of the performance
of arbitrary propellers. This extension includes the elimination of the small angle assumptions in the
optimal design theory, accurate calculation of the vortex displacement velocity which properly accounts
for the blade section drag, and elimination of the small angle assumptions in the Prandtl momentum
loss function for both design and analysis. This provides a good method even though contraction of
the propeller wake is neglected.

69

Você também pode gostar