numerical simulations of turbulent flow around a distributed ......of turboelectric distributed...

10
Numerical simulations of turbulent flow around a distributed propeller wing António Salreta [email protected] Instituto Superior Técnico, Universidade Técnica de Lisboa, Portugal November 2016 Abstract: New advances in electric propulsion have been encouraging new studies in aeronautic propulsion. A new field of study that shows great promise is distributed electric propulsion (DEP). The present work aims to study the effects of changes of the propellers in a DEP aircraft. This study looks at the effects in lift and drag coefficients of the aircraft and is divided in three parts: changes made to the geometry and power of the propellers, changes made to the inflow velocity and changes in the number of propellers working. The CFD program used for this work is STAR-CCM+, where RANS simulations were run using SST K-Omega turbulence model. The mesh generation and mesh refinement are also included in this work, as well as the creation of a performance curve for the propellers from the program JavaProp. Leading edge position changes are not big enough to influence differences in lift and drag, while differences in diameters and power lead to some interesting conclusions. Inflow velocity is a big influential factor for DEP aircrafts. Propellers close to the tip of the wing show more influence on lift. Keywords: Distributed electric propulsion, lift and drag coefficients, STAR-CCM+, RANS, SST K-Omega, JavaProp ----------- Introduction Electric engines are a good alternative to combustion engines. These have many benefits including good efficiency, low noise and high reliability. However the most interesting characteristic of electric engines is the possibility of producing smaller engines and locating them in strategic places of the airplane. This way aircraft designers and engineers have more design freedom, creating vastly different designs without being restrained by the weight and volume of the actual combustion propellers. Some early concepts of turboelectric distributed propulsion vehicles with subsonic fixed wing show promise for better efficiency and noise reduction. N3-X is a NASA’s project that predicts the combination of a blended wing body configuration and distributed propulsion system produces a 70% fuel-burn reduction relative to a B777-200LR reference aircraft [1]. Distributed electric propulsion (DEP) has shown great promise especially in light weight aircrafts. In this design many small propellers are distributed along the span of the wing increasing dynamic pressure, and facilitating of flight at lower velocities excluding the need for multi-element high lift devices. While being a promising design, this brings higher structural complexity. Variable pitch loading can be achieved by controlling individual propellers, making it easier to optimally match the needs of the aircraft. It is also a redundant and robust system that can be advantageous in case of engine failure. This work aims to study the influence of inflow conditions as well as parametric changes made to the electric propellers for a DEP small man-controlled airplane. This plane is entitled of Silent Air Taxi (SAT). Configuration Description SAT’s configuration was inspired by NASA’s SCEPTOR configuration, following the same idea of distributed propellers. In this case the airplane geometry is presented in figure 1 and the configurations in table 1, showing five smaller identical propellers distributed along the span of the wing, and one bigger propeller at the tip (this refers only to one wing). The number of propellers was already chosen and quantity of propellers was not tested. The reason for the location of the propellers in front of the leading edge compared to locating them behind the wing is for favourable pitching moment, acoustic, cruise drag and structural complexity [2].

Upload: others

Post on 03-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Numerical simulations of turbulent flow around a distributed

    propeller wing

    António Salreta

    [email protected]

    Instituto Superior Técnico, Universidade Técnica de Lisboa, Portugal

    November 2016

    Abstract: New advances in electric propulsion have been encouraging new studies in aeronautic propulsion.

    A new field of study that shows great promise is distributed electric propulsion (DEP). The present work aims

    to study the effects of changes of the propellers in a DEP aircraft. This study looks at the effects in lift and

    drag coefficients of the aircraft and is divided in three parts: changes made to the geometry and power of

    the propellers, changes made to the inflow velocity and changes in the number of propellers working. The

    CFD program used for this work is STAR-CCM+, where RANS simulations were run using SST K-Omega

    turbulence model. The mesh generation and mesh refinement are also included in this work, as well as the

    creation of a performance curve for the propellers from the program JavaProp. Leading edge position

    changes are not big enough to influence differences in lift and drag, while differences in diameters and power

    lead to some interesting conclusions. Inflow velocity is a big influential factor for DEP aircrafts. Propellers

    close to the tip of the wing show more influence on lift.

    Keywords: Distributed electric propulsion, lift and drag coefficients, STAR-CCM+, RANS, SST K-Omega,

    JavaProp

    -----------

    Introduction

    Electric engines are a good

    alternative to combustion engines. These

    have many benefits including good

    efficiency, low noise and high reliability.

    However the most interesting characteristic

    of electric engines is the possibility of

    producing smaller engines and locating

    them in strategic places of the airplane. This

    way aircraft designers and engineers have

    more design freedom, creating vastly

    different designs without being restrained by

    the weight and volume of the actual

    combustion propellers. Some early concepts

    of turboelectric distributed propulsion

    vehicles with subsonic fixed wing show

    promise for better efficiency and noise

    reduction. N3-X is a NASA’s project that predicts the combination of a blended wing

    body configuration and distributed

    propulsion system produces a 70% fuel-burn

    reduction relative to a B777-200LR

    reference aircraft [1].

    Distributed electric propulsion (DEP)

    has shown great promise especially in light

    weight aircrafts. In this design many small

    propellers are distributed along the span of

    the wing increasing dynamic pressure, and

    facilitating of flight at lower velocities

    excluding the need for multi-element high lift

    devices. While being a promising design,

    this brings higher structural complexity.

    Variable pitch loading can be achieved by

    controlling individual propellers, making it

    easier to optimally match the needs of the

    aircraft. It is also a redundant and robust

    system that can be advantageous in case of

    engine failure.

    This work aims to study the

    influence of inflow conditions as well as

    parametric changes made to the electric

    propellers for a DEP small man-controlled

    airplane. This plane is entitled of Silent Air

    Taxi (SAT).

    Configuration Description

    SAT’s configuration was inspired by

    NASA’s SCEPTOR configuration, following

    the same idea of distributed propellers. In

    this case the airplane geometry is presented

    in figure 1 and the configurations in table 1,

    showing five smaller identical propellers

    distributed along the span of the wing, and

    one bigger propeller at the tip (this refers

    only to one wing). The number of propellers

    was already chosen and quantity of

    propellers was not tested. The reason for the

    location of the propellers in front of the

    leading edge compared to locating them

    behind the wing is for favourable pitching

    moment, acoustic, cruise drag and structural

    complexity [2].

  • Comparisons were made to Cirrus SR22, Cessna 172R and Tecnam P2006T airplanes because these represent the current state of the art for this class of aircraft and share a similar wing span (table 2). Propeller configurations are displayed in table 1. These propeller were chosen to have the same characteristics as pre-existing propellers. Simulations were done only to one wing taking advantage of symmetry and excluding the fuselage to be less computationally demanding. The geometry was tested without flap.

    Figure 1 - SAT's geometry

    Objectives

    To study changes in lift a reference

    configuration was chosen as the starting and

    comparison point for the other simulations,

    while the whole geometry remained

    constant. Only numerical simulations were

    done. Different speeds were tested, for

    cruise at 110.6 m/s and take-off/landing

    31.38 m/s. Parametric changes to the

    propellers were made to the diameter,

    distance to the leading edge and power of

    the engine. These were tested for higher

    velocity (110.61 m/s) to produce higher

    differences. While to study the influences of

    each inner propeller, the simulations were

    made for an inflow velocity of 31.38 m/s. To

    do these last simulations these propellers

    were switched off one by one.

    Mathematical models

    RANS simulations were performed using the

    commercial program STAR-CCM+. Both

    mesh generation and simulations were

    accomplished using this program. SST

    (Menter) K-Omega was the turbulence

    model because it behaves well in adverse

    pressure gradient, and solves the flow inside

    and outside the boundary layer with

    relatively accuracy.

    SST (Menter) K-Omega model is a hybrid of

    two other models using K-Omega in the

    inner boundary layer and K-Epsilon in the far

    field. Using the following equation to

    calculate the turbulent viscosity:

    𝜇𝑡 = 𝜌𝑎1𝑘

    𝑚𝑎𝑥(𝑎1𝜔,𝛺𝐹2) ( 1)

    where 𝑎1 is a constant, Ω is the absolute value

    of the vorticity, and 𝐹2 is a function that is one

    for boundary-layer flows and zero for free

    shear layers. 𝑘 and 𝜔 are calculated by two

    transport equations. [3]

    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.

    The axial force and normal force on a surface

    are computed by the Star-CCM+ as:

    𝑓 = ∑ (𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 + 𝑓𝑆

    𝑆ℎ𝑒𝑎𝑟). 𝑛𝑓𝑓 ( 2)

    𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 = (𝑝𝑓 − 𝑝𝑟𝑒𝑓)𝑎𝑓 ( 3)

    𝑓𝑆𝑆ℎ𝑒𝑎𝑟 = −𝑇𝑓𝑎𝑓 ( 4)

    where 𝑓𝑆𝑃𝑟𝑒𝑠𝑠𝑢𝑟𝑒 and 𝑓𝑆

    𝑆ℎ𝑒𝑎𝑟 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,

    Inner Tip

    Number of Blades 5 3 Design RPM 4000 3000 Outer diameter 0.75 m 1.5 m Inner diameter 0.056 m 0.292 m Airspeed (m/s) 110.6 Power (kW) 20 50 Airfoil MH 114 13% Re =

    500 000

    Table 1 - Propellers' configurations

  • 𝑝𝑟𝑒𝑓 is the reference pressure, and 𝑇𝑓 is the

    stress tensor at face 𝑓.

    Table 2 - Comparison of reference configuration with similar plane

    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.

    Propeller model

    For simulating the propellers, a

    virtual disk was used to induce force

    distributions instead of resolving the

    propellers blade geometry. This is a faster way

    of simulating a large number of propellers. The

    rotor is simulated through source terms in the

    momentum equations. The method employs a

    uniform volume force distribution over the

    cylindrical virtual disk. The volume force

    varies with the radial direction. The radial

    distribution varies with the Goldstein optimum

    and is given by: [4]

    𝑓𝑏𝑥 = 𝐴𝑥𝑟∗√(1 − 𝑟∗) ( 5)

    𝑓𝑏𝜃 = 𝐴𝜃 .𝑟∗√1−𝑟∗

    𝑟∗(1−𝑟ℎ′)+𝑟ℎ

    ′ ( 6)

    𝑟∗ =𝑟′−𝑟ℎ

    1−𝑟ℎ′ ( 7)

    𝑟ℎ′ =

    𝑅𝐻

    𝑅𝑃, and 𝑟′ =

    𝑟

    𝑅𝑃 ( 8)

    where 𝐴𝑥 and 𝐴𝜃 are functions of the propeller

    hub radius 𝑅𝐻 and tip radius 𝑅𝑃.

    Outside STAR-CCM+, JavaPROP

    [5] is used to obtain the propeller curve for

    both types of propellers. In this program

    equations based on the formulas of Adkins

    and Liebeck’s work of “Design of Optimum

    Propellers” are used to arrive to an optimum

    solution.[6] 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 D 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 table from the

    resulting performance curve consists of the

    advance ratio, propeller efficiency, thrust

    coefficient 𝐾𝑇, and torque coefficient 𝐾𝑄 given

    as: [4][5]

    𝐽 =𝑉∞

    𝑛 𝐷𝑂 ( 9)

    𝑃 = 𝑇 ∙𝑉∞

    𝜂 ( 10)

    𝐾𝑇 =𝑇

    𝜌𝑛2𝐷𝑂4 ( 11)

    𝐾𝑄 =𝑄

    𝜌𝑛2𝐷𝑂5 ( 12)

    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:

    𝑓(𝐽) = 𝐾𝑇 − 𝐾𝑇′ ( 13)

    𝐾𝑇′ =

    𝐽2𝑇𝑜𝑝𝑒𝑟𝑎𝑡𝑖𝑛𝑔 𝑝𝑜𝑖𝑛𝑡

    𝜌𝑖𝑛𝑓𝑙𝑜𝑤 𝑝𝑙𝑎𝑛𝑒𝑉∞2 𝐷0

    2 ( 14)

    Boundary Conditions

    The flow inlet is made from a semi-

    spherical surface where a velocity magnitude

    and direction are defined. (figure 2) This way

    the wing is maintained static, and the flow

    direction is only controlled by the inlet and the

    propellers. In contact with the wing hub a

    symmetry plane is placed. The wing and

    nacelles have a no-slip condition. And finally,

    the rest of the surfaces (semi-cylindrical and

    semi-circle) are considered as a pressure

    outlet where the flow is extrapolated. 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 determines the

    angle of attack, not the rotation of the body as

    it would be the case in an experimental

    procedure.

    Mesh Refinement

    Mesh refinement was not made

    systematically but was instead local. At first

    the meshing process was made automatically

    but was then refined locally. A hexahedral grid

    type was chosen for all the mesh because of

    its simplicity in defining wake refinement and

    its superior speed to mesh. A prism layer was

    used near the body with 20 layers, all

    amounting to 1% of the base cell size (cell size

    that every other parameter is related to). 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 decided to be 0.001

    (

  • size of the first cell closer to the surface is

    correct for Mesh 6. The difference between

    Mesh 7 and Mesh 6 is 0.0002 (0.4%) for the

    normal and axial force coefficients so further

    refinement was not made. 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. Mesh 6 is the final

    mesh used for this project.

    Mesh Nº of cells (in millions)

    𝑪𝑨𝑭 𝑪𝑵𝑭 Wall y+

    Mesh 1 0.88 -0.0034 0.5127 4.01 Mesh 2 2.38 0.0287 0.6636 2.02 Mesh 3 2.57 0.0294 0.6532 2.05 Mesh 4 12.16 0.0479 0.7418 0.95 Mesh 5 15.00 0.0486 0.7363 0.98 Mesh 6 15.05 0.0489 0.739 0.99 Mesh 7 18.29 0.0487 0.7388 0.99

    Table 3 – Mesh refinement process

    Results

    Configs 𝑷𝒊𝒏𝒏𝒆𝒓(kW)

    𝑷𝑻𝒊𝒑

    (kW)

    𝒙𝑳𝑬(m)

    𝑫𝒊𝒏𝒏𝒆𝒓(m)

    𝑫𝒕𝒊𝒑(m)

    Ref 20 kw 50 kw 0 0.75 1.5 X1 20 kw 50 kw 0.1 0.75 1.5 X2 20 kw 50 kw 0.2 0.75 1.5 DS 20 kw 50 kw 0 0.50 0.75 DL 20 kw 50 kw 0 m 0.9 1.5

    Po

    we

    r I30 30 kw 50 kw 0 m 0.75 1.5 T70 20 kw 70 kw 0 m 0.75 1.5 T20 20 kw 20 kw 0 m 0.75 1.5

    Table 4 - Table of configurations

    Table 4 summarizes the parametric

    changes made to the propellers in power and

    geometry in every configuration. Only for the

    reference solution (solution to which every

    other result is compared) the variation of angle

    of attack has been from 0° to 12°. For the rest

    of the cases only the point of maximum lift was

    interesting studying. For almost all of the

    different propellers configurations, 𝐶𝐿 reaches

    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°. Unfortunately this model is no longer

    accurate after separation is reached because

    it ceases to be a steady flow and it does not

    take into account possible vortex shedding,

    producing results that do not correspond with

    reality. That is why it shows a slight increase

    of 𝐶𝐿 between the angles of 9° and 10° even

    though the values of both simulations have

    stabilized.

    2D simulations for NACA 2412 wing

    profile without nacelles were made using the

    program Xfoil [7] 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 figure 4

    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.

    Figure 3 shows the wall shear stress

    in the x-direction for the reference solution for

    9°, after the maximum angle of attack. There is clearly a separation happening more in one

    side on the left of the propeller. This is due to

    the rotation of the propeller that increases the

    local angle of attack on the left side of the

    propeller forcing it to stall first. This means the

    chosen wing airfoil cannot stall abruptly for

    DEP systems.

    In figure 4 every case for geometry

    and power change are presented and the first

    thing noticeable is that all cases have very

    close values, stalling near the same angle of

    attack. For X1 configurations 𝐶𝐿 𝑀𝑎𝑥 value

    decreases to 0.84 at an angle of 8 degrees.

    The variations of the lift coefficient are seen to

    vary 0.004 (0.47%) from the Reference

    configuration, while the drag coefficient varies

    0.00026 (0.5%) with 𝐶𝐷 being 0.051 (figure 4).

    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. This is a very low and not relevant change to conclude such differences.

    Smaller propellers’ diameters are also tested (DS table 4). The value of 𝐶𝐿 𝑀𝑎𝑥 seems

    to increase to 0.88 (3.7 %), for the larger

    diameter configuration, this value drops

    slightly to 0.84 (0.59 %). The drag coefficient

    doesn’t change much for both cases in relation

    to the reference solution, which leads to

    conclude that this difference in lift is due to

    different rotation rates of the propellers. For

    simulations where larger propeller diameters

    were tested (DL), the rotation rate is around

    2400 rpm, while in the case of the smaller

    diameters, propellers rotation is around the

  • speed of 5500 rpm. This also explains why the

    drag for the larger diameter case is the same

    as the reference solution even though the

    affected area is bigger. In the case of the

    smaller diameters, the higher rotation rate

    increases the drag 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.

    For I30 configuration, 𝐶𝐿 increases to

    0.87 at 8° of angle of attack. It’s an increase of

    0.021 (2.5 %) in 𝐶𝐿 for 50 kW of total power

    increase. 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 0.001 (1.5 %) of difference of 𝐶𝐷 for the

    same angles, in this case 8°. There is a

    significant change in drag but the lift is not

    changing enough, leading to conclude this

    increase in power is not worth it. The overall

    change in 𝐶𝐿 is not big enough.

    Others studied effects were the ones

    created by the increase and decrease of

    power in the tip propellers. Both higher and

    lower power changes were tested. For the

    case of T20, the lift not only gets bigger to a

    value of 0.872, a change of 0.02 (2.4 %), but

    also happens in a higher angle of attack of 9°.

    However for the T70 case the change in lift is

    not noticeable, being 0.85 for the angle of 8°.

    The drag difference is clearly more noticeable

    with changes of 0.016 (32%) for the T20 case,

    and 0.0003 (0.6%) for the T70 case. Either

    one of these parametric changes bring

    benefits in terms of lift, but only T20 presents

    also a decrease in drag, concluding that the tip

    (or “cruise”) propeller does not bring benefits

    when at a higher power than the “high-lift”

    propellers.

    Finally a change in inflow velocity is studied

    and a clear change in results happens. Two

    different cases of 77.2 m/s (150 knots) and

    31.38 m/s (61 knots) were tested. The

    maximum angle still seems to be at 8° for the

    77.2 m/s. There is 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 produces

    higher L/D ratios, 16.86 for 110.6 m/s and

    18.71 for 31.38 m/s, both at 8°.

    Shown in figure 9, there are the lift

    coefficients for different speeds. First the

    maximum angle of attack is at 12° for 31.38

    m/s, and 𝐶𝐿 and 𝐶𝐷 have respectively, an

    increase of 0.409 (48.12%) and an increase of

    0.098 (194.10%). Not only stall happens later,

    as the overall lift coefficient of the wing is

    higher. Clearly this type of “high-lift” devices

    behave better and bring more efficiency at

    lower speeds.

    Figure 3 - Wall Shear Stress in the positive x-direction for Reference solution case at 9°

  • Figure 4 - Lift coefficient for all of the propeller change cases

    Figure 5 - Cl/Clref vs angle of attack for all of the propeller change cases

    Figure 6 – Drag coefficient for all of the propeller change case

    0,2

    0,3

    0,4

    0,5

    0,6

    0,7

    0,8

    0,9

    0 1 2 3 4 5 6 7 8 9 10 11 12 13

    Lift

    Co

    effi

    cien

    t

    Angle of attack (°)

    DEFAULTX1X2DSDLT20I30T70

    0,92

    0,94

    0,96

    0,98

    1,00

    1,02

    1,04

    1,06

    6 7 8 9 10 11

    CL

    / C

    Lref

    Angle of attack (°)

    X1X2DSDLT20I30

    0,03

    0,04

    0,05

    0,06

    0,07

    0,08

    0,09

    0,10

    6 7 8 9 10 11

    Dra

    g C

    oef

    fici

    ent

    Angle of attack (°)

    DEFAULT

    X1

    X2

    DS

    DL

    T20

    I30

    T70

  • Figure 7 – Lift coefficient for all the cases with one propeller turned off

    Figure 8 - Cl/Clref for all the cases with one propeller turned off

    1,05

    1,10

    1,15

    1,20

    1,25

    1,30

    6 7 8 9 10 11 12 13 14 15

    Lift

    ceo

    ffic

    ien

    t

    Angle of attack (°)

    61knots

    1prop

    2prop

    3prop

    4prop

    5prop

    0,92

    0,93

    0,94

    0,95

    0,96

    0,97

    0,98

    0,99

    1,00

    6 7 8 9 10 11 12 13 14 15

    Cl/

    Clr

    ef

    Angle of attack (°)

    1prop

    2prop

    3prop

    4prop

    5prop

  • Figure 9 - Lift coefficient for all the cases with one propeller turned off

    Figure 10 – Lift coefficient change for velocity change, and without propellers

    ----------------

    Conclusion

    A series of different cases have

    converged successfully showing both

    expected and unexpected results.

    2D simulations show that 𝐶𝐿 increases

    with higher velocity (110.61 m/s) but stall first

    than for lower velocities (31.38 m/s). The

    same can’t be said for the 3D cases. With

    propellers at lower velocities stall seems to

    also be occurring later as it would be

    expected, but because of the propellers 𝐶𝐿 is

    much higher than for lower velocities. There is

    also a slight increase comparing with the

    geometry without propellers. Here the

    comparison is difficult since the virtual disk are

    not present anymore in NoProp case (figure 9)

    producing much less drag. This means DEP is

    producing higher 𝐶𝐿 overall but stalls first. Also

    DEP systems seem to work much better at

    lower velocities, and should not be used at

    higher velocities.

    Parametric changes to diameter and

    distance to the leading edge seem to have

    produce no conclusive data, even for higher

    velocities. However power changes hints that

    0,05

    0,07

    0,09

    0,11

    0,13

    0,15

    0,17

    0,19

    0,21

    0,23

    7 8 9 10 11 12 13 14 15

    Dra

    g co

    effi

    cien

    t

    Angle of attack (°)

    61knots

    1prop

    2prop

    3prop

    4prop

    5prop

    0,0

    0,2

    0,4

    0,6

    0,8

    1,0

    1,2

    1,4

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

    Lift

    Co

    effi

    cien

    t

    Angle of attack (°)

    DEFAULT

    150 knts

    61knots

  • the tip propeller should not have relatively

    more power than the smaller ones. Vortices

    produced by the propellers seem to be

    interfering with one another, producing more

    harm than good.

    Finally the propellers that seemed to

    influence more the lift are the ones positioned

    more to the tip of the wing. However without

    the fifth propeller (counting from the hub to the

    tip of the wing) 𝐶𝐿 is slightly higher than without

    the fourth propeller. As explained before, one

    possible explanation is the influence of

    different vortices produced. One possible

    explanation is the influence of different

    vortices produced from the fifth and the tip

    propeller on one another. This may cause an

    increase of dynamic pressure over the wing

    due to not contra-rotating propellers.

    Acknowledgement

    The author would like to thank first

    and foremost Prof. Dr. -Ing. Andreas Henze

    and Prof. Dr. Luís Eça for all the guidance and

    support in the making of this work. This project

    was possible thanks to the opportunity given

    by Prof. Dr. –Ing. Wolfgang Schröder who

    made it possible to work at the Aerodynamic

    Institute of RWTH Aachen, and also to

    Michael Kreimeier for the opportunity to work

    on this project. This project belongs to the

    Aeronautics and Aerospace Department of

    RWTH Aachen that was done in collaboration

    with the Aerodynamic Institute of the same

    university.

    References

    [1] Felder J.L., Brown G.V., DaeKim H., and

    Chu J., (2011) Turboelectric Distributed

    Propulsion in a Hybrid Wing Body Aircraft.

    NASA Glenn Research Center, USA.

    [2] Moore, M. D. and Fredericks, W. J., (2014) Misconceptions of Electric Aircraft and their Emerging Aviation Markets," AIAA SciTech, American Institute of Aeronautics

    and Astronautics.

    [3] Menter F.R., (1994) Two-Equation Eddy-

    Viscosity Turbulence Models for Engineering

    Applications. AIAA Journal, 32(8):1598-1605

    [4] STAR-CCM+ User Guide. Star-CD version

    4.02 (2006)

    [5] Hepperle M.. (2013). Java PROP – Design

    and Analysis of Propellers. Accessed in 15-

    03-2016 at: http://www.mh-

    aerotools.de/airfoils/javaprop.htm

    [6] 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.

    [7] XFoil Subsonic Development System

    Website, Drela M.. Youngren H., Accessed on

    06-04-2016 at:

    http://web.mit.edu/drela/Public/web/xfoil/

    http://www.mh-aerotools.de/airfoils/javaprop.htmhttp://www.mh-aerotools.de/airfoils/javaprop.htm