Conference PaperPDF Available

Analysis and Design of a Small-Scale Wingtip-Mounted Pusher Propeller

Authors:

Abstract and Figures

The wingtip-mounted pusher propeller, which experiences a performance benefit from the interaction with the wingtip flowfield, is an interesting concept for distributed propulsion. This paper examines a propeller design framework and provides verification with RANS CFD simulations by analysing the wing of a 9-passenger commuter airplane with a wingtip-mounted propeller in pusher configuration. In the taken approach, a wingtip flowfield is extracted from a CFD simulation, circumferentially averaged and provided to a lower order propeller analysis and optimisation routine. Possible propulsive efficiency gains for the propeller due to installation are significant, up to 16% increase at low thrust levels, decreasing to approximately 7.5% at the highest thrust level, for a range of thrust from 5% up to 100% of the wing drag. These gains are found to be independent of propeller radius for thrust levels larger than 30% of the wing drag. Effectively, the propeller geometry is optimized for the required thrust and to a lesser degree for the non-uniformity in the flowfield. Propeller blade optimization and installation result in higher profile efficiency in the blade root sections and a more inboard thrust distribution.
Content may be subject to copyright.
Analysis and Design of a Small-Scale Wingtip-Mounted
Pusher Propeller
Tom C. A. Stokkermans, Sebastiaan Nootebosand Leo L. M. Veldhuis
Delft University of Technology, 2629 HS Delft, The Netherlands
The wingtip-mounted pusher propeller, which experiences a performance benefit from the
interaction with the wingtip flowfield, is an interesting concept for distributed propulsion. This
paper examines a propeller design framework and provides verification with RANS CFD sim-
ulations by analysing the wing of a 9-passenger commuter airplane with a wingtip-mounted
propeller in pusher configuration. In the taken approach, a wingtip flowfield is extracted from
a CFD simulation, circumferentially averaged and provided to a lower order propeller analysis
and optimisation routine. Possible propulsive eciency gains for the propeller due to installa-
tion are significant, up to 16
%
increase at low thrust levels, decreasing to approximately 7.5
%
at the highest thrust level, for a range of thrust from 5
%
up to 100
%
of the wing drag. These
gains are found to be independent of propeller radius for thrust levels larger than 30
%
of the
wing drag. Eectively, the propeller geometry is optimized for the required thrust and to a
lesser degree for the non-uniformity in the flowfield. Propeller blade optimization and instal-
lation result in higher profile eciency in the blade root sections and a more inboard thrust
distribution.
Nomenclature
b=Wing span, m
CD=Drag coecient
CL=Lift coecient
Cp=(pp)/qpressure coecient
CQ=Q/ρn2D5
ptorque coecient
CT=T/ρn2D4
pthrust coecient
c=Chord, m
Dp=Propeller diameter, m
D=Wing drag, N
hi=Average cell size of grid i, m
J=V/(nDp)advance ratio
L=Wing lift, N
n=Propeller rotational speed, s1
P=Shaft power, W
p=Static pressure, Pa
Q=Torque, Nm
Q0=Torque distribution, Nm ·m1
q=Dynamic pressure, Pa
Rp=Propeller radius, m
Rp,ref=Reference XPROP propeller radius, m
r=
Radial coordinate in propeller reference frame, m
T=Thrust, N
T0=Thrust distribution, N ·m1
U
s=
Standard deviation of a fit based on theoretic order
of convergence
Uφ=Estimated discretization uncertainty
V=Velocity, m ·s1
x=Axial coordinate, m
y=Vertical coordinate, m
y+=Dimensionless wall distance
z=Lateral coordinate, m
α=Angle of attack, deg
β=Blade pitch angle, deg
ηp=TV/Ppropeller propulsive eciency
ρ=Density, kg ·m3
ϕ=Propeller blade phase angle, deg
φi=Numerical solution obtained using grid i
ω=Vorticity, s1
Subscripts
0=Extrapolated
0.7Rp=Radial coordinate, in which r/Rpis equal to 0.7
a=Axial
c=Chord based
iso =In isolated condition
h=Hub
k=Kink
p=Propeller
r=Root
t=Tangential, tip
=Freestream
Ph.D. Candidate, Flight Performance and Propulsion Section, Faculty of Aerospace Engineering, AIAA Member, t.c.a.stokkermans@tudelft.nl.
M.Sc. Graduate, Flight Performance and Propulsion Section, Faculty of Aerospace Engineering
Full Professor, Head of Flight Performance and Propulsion Section, Faculty of Aerospace Engineering, AIAA Member.
1
I. Introduction
Today’s research on propellers is driven by their potential for reduced fuel consumption compared to turbofan
propulsion. Not only the high propulsive eciency of the propeller itself, but also its location on the airframe can
enhance the overall eciency of the aircraft. Wingtip-mounted propellers have been envisaged for their favorable
interaction eects. For the tractor propeller variant, the interaction of the wing with the slipstream results in a reduction
of the wing induced drag if the rotation direction of the propeller is opposite to that of the wingtip vortex [
1
3
]. The
pusher propeller variant experiences a reduction in shaft power due to the swirling vortex inflow from the wingtip in
case the propeller rotates against the direction of the wingtip vortex [
2
,
4
6
]. Moreover, the modification of the wingtip
vortex may reduce wing induced drag as well. Yet, the ingestion of the non-uniform inflow field may result in a noise
penalty for the pusher variant. Adverse aeroelastic eects due to the large weight of an engine at the tip of a wing and
the large yawing moment arm of the thrust vector in case of a one-engine-inoperative condition have prevented the
application of wingtip-mounted propellers up to now. However, the emergence of electric propulsion in aircraft allows
almost penalty-free downscaling of the propeller due to the scale independence of the electric motor [
7
]. This makes it
possible to distribute the propulsion for instance through a hybrid-electric architecture, and enables to scale down the
wingtip-mounted propeller like proposed in the SCEPTOR concept [8].
a) Modified Learfan 2100 b) Modified Tecnam P2012 Traveller
Fig. 1 Impression of airplane concepts with small-scale wingtip-mounted pusher propellers.
In this research the pusher variant is investigated in the context of distributed propulsion. An impression of such a
concept, combining i.e. fuselage boundary layer ingestion and wingtip-mounted pusher propellers through a hybrid
electric architecture, is shown in Fig. 1 a. A more conventional airplane layout with small-scale wingtip-mounted
pusher propellers is sketched in Fig. 1 b. The design freedom to scale propellers and to distribute propulsion, results in
a non-unique thrust requirement. For example, the propeller can be designed to balance just the induced drag of the
wing or balance the entire cruise drag of the aircraft. The resulting performance benefit of propeller installation at the
wingtip may depend on the thrust level. Hence the following research question: How does propeller scale influence the
propulsive eciency benefit and the upstream aerodynamic loading on the wing? Up to now, only research on a full
scale wingtip-mounted pusher propeller has been performed, with a propeller not specifically designed for this task
[
4
]. The inflow to the propeller is non-uniform, especially when the propeller becomes smaller relative to the wingtip
flowfield. For a boundary layer ingestion propeller, Ref. [
9
] has shown that maximizing the propulsive eciency gain
requires a dierent design to cope with the non-uniform inflow experienced on the aft fuselage. Analogue to that, the
wingtip-mounted pusher propeller may also benefit from design optimization, resulting in the second research question:
To what extent can the propulsive eciency benefit be increased by designing the propeller for the non-uniform inflow
experienced at the wingtip?
This research, which is regarded as an extension of Ref. [
10
], gives insight in these questions by analysis of a
specific case through the following steps:
1)
CFD analyses of the wing are performed in order to quantify the wing performance and extract the wingtip
flowfield.
2)
A lower order propeller analysis and optimisation routine PROPR is established and validated for uniform inflow.
3) The wingtip flowfield is fed to PROPR for analysis and design optimization.
4)
The upstream eect of the propeller designs on the wing performance is analyzed through CFD analyses of the
wing with an actuator disk representation of the propeller.
5)
The accuracy of PROPR for the non-uniform wingtip flowfield is checked through a fully resolved propeller–wing
CFD simulation.
2
II. Computational Methods
A. RANS CFD Simulations
Four dierent types of RANS CFD simulations were performed in order to establish the wing performance and wingtip
flowfield, to estimate the upstream eect of the propeller on the wing and to verify the accuracy of PROPR:
Isolated wing simulations
Isolated propeller simulations
Wing simulations with actuator disk propeller representation
A propeller–wing simulation
The wing used for these simulations was derived from the Tecnam P2012 Traveller [
11
], a twin-prop 9-passenger
commuter airplane with a maximum take-omass of 3600
kg
. Only the wing was taken into account, without the
original propeller and nacelles. A sketch of the wing is shown in Fig. 2 a. A minimum radius nacelle was added at
the tip of the wing, extending aft of the trailing edge to accommodate a propeller in pusher configuration. The nacelle
radius was kept equal to the propeller hub radius
Rh
. Hence, so far there is no provision to accommodate an electric
motor. The wing parameters are given in Table 1 and are partly based on Ref. [12].
a) Wing model with nacelle and spinner. b) XPROP propeller model.
y
x
z
ypxp
zp
xp
Rp
Rh
n
yp
0.21b
0.29b
cr = ck
ct
Fig. 2 Isometric view of the wing and TU Delft research propeller XPROP.
The propeller in the isolated propeller and propeller–wing simulations is the 6-bladed XPROP propeller, shown in
Fig. 2 b, a research propeller from Delft University of Technology typical for turboprop airplanes. An extensive grid
study and experimental validation for this propeller was performed in Ref. [
13
], and the same propeller grid density
was used in this research. The spinner was modified to convert it into a pusher propeller. The propeller was used in its
original size with a radius of
Rp,ref =
0
.
2032 m and hub radius of
Rh=
0
.
23
Rp,ref
. Simulations were also performed
with proportionally scaled versions of smaller and larger size.
Table 1 Wing model parameters.
Parameter Value
Span b13.55 m
Root and kink chord cr,ck2.06 m
Tip chord ct1.38 m
c/4 sweep, dihedral, twist 0
Root and kink airfoil NACA23015
Tip airfoil NACA23012
Cruise speed 80 m/s
Cruise altitude 3048 m
Cruise angle of attack 3
Cruise lift coecient 0.35
3
Symmetry BC Pressure
outlet BC
Pressure
farfield BC
OD
ID
WD
PD SD
OD:
ID:
WD:
PD:
SD:
Outer Domain
Inner Domain
Wing Domain
Propeller Domain
Slipstream Domain
15cr
cr
y
x
5cr
20cr
Fig. 3 Computational domain and boundary conditions for the wing simulations.
The RANS equations for compressible flow were used with a 2
nd
order accurate scheme in ANSYS
r
Fluent 18.1
[
14
], a commercial, unstructured, finite volume, cell-centered solver. For the propeller–wing simulations time-dependent
solutions were found by a 2
nd
order backward Euler scheme with a time step equivalent to 2
deg
of propeller rotation.
Discretization of the advection term was done with an upwind scheme using the Barth–Jesperson boundedness principle
[
15
]. For the equation of state, an ideal gas was assumed and Sutherland’s law was used to predict the corresponding
dynamic viscosity. Standard atmospheric conditions at the cruise altitude were assumed. The turbulence model was
selected based on the findings of Kim and Rhee [
16
], who tested several turbulence models to simulate the wingtip
vortex of an isolated wing. The eddy viscosity model in their research that best agreed with experimental data in terms
of static pressure and axial velocity in the wingtip-vortex core was the Spalart–Allmaras (SA) one-equation model [
17
]
with modification proposed by Dacles-Mariani et al. [
18
] to prevent build-up of turbulence viscosity in vortex cores.
Therefore, this model was selected for the current research. Values for the inlet turbulence quantities were based on the
recommendations by Spalart and Rumsey [
19
], which resulted in an eddy viscosity ratio of 0
.
21044 for the SA model.
In order to fully resolve the boundary layer, the y+value on the no-slip walls of the model was less than one.
The computational domain and boundary conditions for the wing simulations are shown in Fig. 3. The outer
dimensions of the domain were chosen to be suciently large with respect to the wing chord, in order to minimize the
influence of the boundary conditions on the flow properties near the wing. At the domain inlet, a total-pressure jump
with respect to the undisturbed static pressure was set to reach the cruise speed. Furthermore, the undisturbed total
temperature was specified. At the domain outlet, the static pressure was prescribed to be on average equal to undisturbed
static pressure. On the outboard side of the domain, a Riemann-invariant pressure farfield condition was specified
with a Mach number, static pressure and static temperature complying with the inlet conditions. On the inboard side
side, a symmetry boundary condition was imposed. The wing, nacelle, propeller and spinner were modeled as no-slip
walls. The computational domain was divided in several domains for refinement of the grid. The propeller domain (PD)
was connected to the other domains through sliding mesh interfaces to allow grid rotation for simulation of propeller
motion. This domain could be replaced by a domain without propeller blades for the isolated wing simulations and
wing simulations with actuator disk representation of the propeller. The actuator disk model described in Ref. [
20
] was
used, requiring the propeller blade radial distribution of thrust and torque as input. Grids were constructed by means
of ANSYS
r
Meshing. For regions adjacent to no-slip walls, the unstructured grid was made up of a triangular wall
mesh, followed by layers of semi-structured prismatic elements of the inflation layer. For the remainder of the domain
tetrahedral elements were used. Grid density in the whole domain was controlled by wall refinement of all no-slip walls,
volume refinement of the domains, a 1
st
layer thickness of the inflation layers, and growth rates of the inflation layers
and the remainder of the grid. Grid study results for the isolated wing are shown in Section III.
The computational domain and boundary conditions for the isolated propeller simulations are described in Ref. [
13
].
Since the wake of a propeller with axisymmetric nacelle is cyclic with the number of the blades, only a single blade was
modeled in a wedge shaped domain with appropriate boundary conditions. The movement of the propeller and spinner
was simulated with a rotating reference frame.
4
B. Propeller Analysis and Optimisation Routine: PROPR
To perform quick propeller analyses and to aerodynamically design the propeller for optimized performance, a
PRopeller analysis and OPtimisation Routine named PROPR was set up based on XROTOR [
21
]. The software program
uses discrete line vortices forming a semi-rigid wake to iteratively determine the induced velocities and has been used
before by Refs. [
9
,
22
25
] for both uniform and non-uniform inflow. XFOIL [
26
] was selected for airfoil analyses. All
details of PROPR are described in Ref. [
10
]. Twenty radial sections were used to define the propeller geometry and
provide XROTOR with airfoil data. Each radial section was supplied with the correct non-uniform flowfield. Fully
turbulent flow for the airfoil data was assumed to allow comparison with the fully turbulent CFD simulations and to
increase stability of running XFOIL in an optimization framework. The non-linear part of the lift curve was modified
using an empirical model by Snel et al. [
27
] to correct two dimensional data for three dimensional rotational eects.
Comparison of PROPR results with validation data is provided in Section IV.
III. Isolated Wing Analysis
CFD analyses of the wing were performed at 3
deg
angle of attack to represent a cruise condition. A grid dependency
study was performed to estimate the discretization uncertainty and to select an appropriate grid density. All refinements
were varied systematically, except for the inflation layer, which was kept constant in line with Roache [
28
]. To estimate
discretization uncertainty, the least-squares version of the grid convergence index (GCI) proposed by Eça and Hoekstra
[
29
] was applied, with the alteration of using only the theoretical order of the solver of 2 to estimate the extrapolated
grid results. Table 2 gives an overview of the grid sizes and wing lift and drag found for each grid. Table 3 presents the
extrapolated lift and drag
φ0
, the standard deviation of the fit based on the theoretical order of convergence
U
s
and the
estimated discretization uncertainty
Uφ
for grid 2. The uncertainty of 0
.
49% and 3
.
10% for the lift and drag coecient
respectively was deemed acceptable for this study.
Table 2 Overview of grids and resulting isolated wing performance.
Grid No. of cells hi/h1CL[-] CD[-]
4 8,561,478 1.82 0.3434 0.01638
3 16,315,794 1.47 0.3440 0.01622
2 32,756,863 1.16 0.3447 0.01600
1 51,424,220 1.00 0.3452 0.01585
Table 3 Grid extrapolation results and estimated discretization uncertainties for grid 2.
CLCD
φ00.3458 0.01568
U
s(%) 0.08 0.57
Uφ(%) 0.49 3.10
From the solution on grid 2 the flowfield that was fed to PROPR was extracted from the propeller plane, the plane
where the propeller will be installed. The flowfield at the wingtip is visualised in Fig. 4 by means of streamtraces and
the wingtip vortex is shown by an axial vorticity isosurface. On the isosurface and the propeller plane the velocity
magnitude is plotted and on the wing surface the pressure coecient distribution is shown. In Fig. 5 this flowfield
at the propeller plane is given for the left wingtip as seen from behind, by means of contour plots of the axial and
tangential velocity components. These velocity components are most relevant for the propeller aerodynamic loading
and, together with the propeller rotational speed and propeller induced velocities, determine the local blade section
angle of attack and dynamic pressure. The tangential velocity component is defined positive in the rotation direction of
the propeller, which is running counterclockwise for the left propeller when seen from behind. In PROPR, the inflow
flowfield is radially varying but assumed to be circumferentially constant. Therefore a circumferential average of this
flowfield was taken, which will impact the resulting propeller response as follows: In the axial velocity contour plot
the reduced velocity in the wake of the wing is clearly visible and its eect on the propeller will be averaged. In the
tangential velocity plot a region of strong negative velocity and a region of positive velocity can be observed which
have an opposite eect on the propeller. These regions will also be circumferentially averaged in a net negative velocity.
5
(V V) / V
−0.40
−0.35
−0.30
−0.25
−0.20
−0.15
−0.10
−0.05
0.00
0.05
Cp
−1.25
−1.00
−0.75
−0.50
−0.25
0.00
0.25
0.50
0.75
1.00
axial vorticity isosurface at ωx = −900 s-1
propeller plane
Fig. 4 Visualisation of wingtip flowfield and pressure coecient distribution on the wing surface.
0.4 0.2 0.0 0.20.4 0.3 0.2 0.1 0.0
(Va V) / VVt / V
zp
yp
Rp,ref
zp
yp
Rp,ref
Fig. 5 Contour plots of axial and tangential velocity components at the propeller plane behind the left wingtip.
The dashed line shows the propeller disc area for the reference XPROP propeller.
The resulting propeller designs for this flowfield are presented in Section V and the consequence of circumferential
averaging will be further discussed in Section VII.
IV. Isolated Propeller Analysis
To establish confidence in PROPR, two comparisons were made. The first was with in-house windtunnel data of
the XPROP propeller in uniform flow. In Fig. 6 a comparison is presented of the thrust and torque coecient versus
advance ratio for
V=
30 m/s. Although at high advance ratios a significant deviation starts to appear for both thrust
and torque coecient, in the region of interest where
CT
is higher, the match with the experimental data is satisfactory.
The second comparison was made with a CFD model of the isolated propeller from Ref. [
13
]. In Fig. 7 the thrust
and torque distribution over a propeller blade are plotted for both PROPR and the CFD model for
V=
30 m/s and an
advance ratio of
J=
0
.
74. This corresponds to a thrust of 16% of the isolated wing drag. The integrated loading is
overestimated by PROPR with 4
.
5% and 7
.
6% in thrust and torque respectively. The shape of the thrust and torque
distribution from PROPR and the CFD model are very similar except for at the tip of the blade. The local rise in thrust
and torque in the CFD model due to the tip vortex is not captured in the PROPR results because of the single lifting line
representation of the blade. Considering that in these comparisons acceptable agreement was found, a design study was
performed with PROPR. This is presented in the next Section V.
6
0.35
0.00
0.30
0.25
0.20
0.15
0.10
0.05
−0.05
Thrust coefficient CT
0.50 0.75 1.00 1.25 1.50
PROPR
Experiment
0.06
0.00
0.05
0.04
0.03
0.02
0.01
Torque coefficient CQ
Advance ratio J
0.50 0.75 1.00 1.25 1.50
Fig. 6 Propeller performance from PROPR and inhouse windtunnel test for V=30 m/s and β0.7Rp=30.
180
Thrust distribution T´ (N / m)
Radial coordinate r / Rp
0.2 0.3
PROPR
CFD
160
140
120
100
80
60
40
20
00.4 0.5 0.6 0.7 0.8 0.9 1.0
14
Torque distribution Q´ (Nm / m)
0.2 0.3
12
00.4 0.5 0.6 0.7 0.8 0.9 1.0
10
8
6
4
2
Fig. 7 Blade loading distributions from PROPR and CFD model for V=30 m/s, β0.7Rp=30and J =0.74.
7
V. Propeller Design Study
In this section the design optimization results from PROPR for a propeller placed in the wingtip flowfield are
discussed. The dierent propeller designs are given in Table 4. The original XPROP propeller was tested in isolated and
installed condition. The XPROP propeller was also optimized for minimum power in terms of blade pitch distribution
and chord distribution, keeping the airfoil sections the same. This was done for isolated and installed conditions,
denoted isoOpt and insOpt respectively. The former was tested both in isolated and installed conditions, while the latter
was only tested in installed conditions. Note that these optimizations were performed for each thrust level separately
and that for each thrust level the performance is given for optimum operating conditions in terms of β0.7Rpand J.
Table 4 Overview of dierent propeller geometries used in design study.
Propeller name Description
XPROP-iso Original XPROP propeller operating in isolated condition
isoOpt-iso Optimized for isolated condition, operating in isolated condition
XPROP-ins Original XPROP propeller operating in installed condition
isoOpt-ins Optimized for isolated condition, operating in installed condition
insOpt-ins Optimized for installed condition, operating in installed condition
0.10 0.14 0.17 0.21 0.24 0.27 0.31 0.34 0.38 0.41
Propulsive efficiency ηp (%)
100
90
80
70
100
95
90
85
Power P / Piso (%)
XPROP-iso
isoOpt-iso
XPROP-ins
isoOpt-ins
insOpt-ins
80
0.44 0.48
Thrust over drag ratio T / Diso
Fig. 8 Propeller optimization results showing the propulsive eciency and power versus the thrust to wing
drag ratio at constant radius Rp/Rp,ref =1.00 for isolated and installed conditions.
A. Propeller Optimization for Constant Radius
First, a design sweep was performed for a range of thrust levels, keeping the propeller radius equal to that of the
original XPROP propeller
Rp,ref
. In Fig. 8 the propeller performance is shown for a range of design thrust levels with
respect to the isolated wing drag
T/Diso
. The propulsive eciency
ηp=TV/P
is plotted, as well as the relative power
required with respect to the XPROP-iso propeller
P/Piso
. Dashed lines refer to performance in isolated conditions, solid
lines indicate installed propeller performance. Put in perspective, the wing induced drag is about 40% of the total wing
drag in this condition, so a range of thrust levels is plotted from 25% up to 115% of the wing induced drag.
8
A number of trends can clearly be observed: The propulsive eciency decreases with increasing design thrust in both
the isolated and installed cases. This is expected since the propeller radius was kept constant. The possible eciency
gains due to installation and optimisation of the propeller are significant. Comparing insOpt-ins with XPROP-iso results,
up to 20% increase in eciency is achieved at lower design thrust levels by the combined eect of optimisation and
installation, decreasing to approximately 10% and remaining constant at higher thrust levels. Although a more fair
comparison is that of the insOpt-ins with the isoOpt-iso propeller, both optimised for their respective flowfield. Then,
an eciency increase up to 15% remains at the lowest thrust level, decreasing to 9% at the highest thrust level.
Looking at the power plot, the eectiveness of geometrical optimization in both isolated and installed conditions
at low and high thrust levels is clearly visible. This eect diminishes at more average thrust levels, as the XPROP
propeller is apparently designed for those thrust levels. The reduction in required power is significant when installing
the propeller, even for the XPROP with non-optimized geometry. Reductions up to 20% are achieved when comparing
insOpt-ins with XPROP-iso results, although looking at the most fair comparison with the isoOpt-iso propeller, a bit
lower maximum power reduction of up to 15% is found. Note that the installed propeller with optimized geometry for
isolated conditions (isoOpt) yields almost identical power reductions compared to the installed propeller with optimized
geometry for installed conditions (insOpt). Thus, eectively the propeller geometry is optimized for the required
thrust level and to a lesser degree for the non-uniformity in the flowfield. Only at the very high design thrust levels of
T/Diso 0.42 a noticeable dierence in performance between the two dierent optimized propellers is observed.
With increasing thrust, the power reduction due to installation converges quickly to a nearly constant value, meaning
that the eective power reduction that can be achieved by installation of the propeller almost does not change with thrust
requirement. One would expect that the eective power reduction would decrease with increasing propeller design
thrust, as there is only a finite amount of energy to be ‘extracted’ from the wingtip flowfield present in the installed case.
Because of the limited maximum thrust that can be delivered by the XPROP propeller it is not possible to investigate the
eective power decrease at even higher thrust levels for the current propeller radius. The experimental work done by
Patterson et al. [
4
] gives already an indication of the power reduction found at higher thrust levels, considering that the
size of that propeller and its thrust relative to the wing was much larger. At similar lift coecient the power reduction
was found to be 14%, although this was for a non-optimized propeller design. Considering that this is of similar order
to what was found in this study thus far, it may be that over a larger thrust range this relative power reduction stays
more or less constant. To confirm this, in Section V.C the higher thrust regime up to a thrust equal to the cruise drag of
the wing will be investigated with larger radius propellers.
B. Blade Loading and Geometry Changes Due to Optimization
The resulting changes in the propeller blade loading and geometry due to optimization and installation are discussed
for two dierent design thrust levels:
T=
0
.
21
Diso
and
T=
0
.
39
Diso
. First, the results from optimization of the XPROP
propeller with
T=
0
.
21
Diso
are discussed. In Fig. 9 the blade pitch angles, chord fractions, the profile eciency and
thrust distribution over the entire blade radius are shown. It is seen that the chord distribution of the insOpt and isoOpt
propellers are nearly identical, even though the insOpt propeller was subjected to the non-uniform wingtip flowfield
during optimization. In both cases, the chord lengths were reduced by as much as 40% compared to the XPROP
propeller. A slight increase in blade pitch angle in the root sections is observed. Both the optimized isoOpt and insOpt
propellers show significantly higher eciency in the root sections. Presence of the non-uniform inflow enables further
eciency gains. Combined, this leads to a higher local profile eciency for the insOpt propeller, especially for the
root sections (
r/Rp<
0
.
4). A clear trend is visible in the thrust distribution over the blade. Geometry optimization
of the XPROP propeller causes the thrust distribution to shift inboard towards the root. Furthermore, the maximum
value decreases. The same behavior to an even greater extent is observed when investigating the insOpt propeller. The
combination of higher profile eciency and the production of thrust there where the profile eciency is higher makes
this the most ecient propeller of the three.
Second, the optimization results with
T=
0
.
39
Diso
are discussed. In Fig. 10 it is seen that the local blade chord
lengths are increased in both the isoOpt and insOpt propeller designs compared to the original XPROP propeller. This
geometry change due to optimization shows a reverse trend than what was seen for the lower design thrust in Fig. 9,
where the chord fractions were decreased due to optimization. The local blade pitch angle is higher for both the isoOpt
and insOpt propeller design, as was the case for the propellers optimized for
T=
0
.
21
Diso
. Also the improvements in
local eciency are similar, except for at the tip. Note that the eciency of the XPROP propeller is relatively low in the
tip region (
r/Rp>
0
.
85) due to tip stall. Optimization of the propeller geometry reduces this tip stall and the overall
eciency distribution is again an almost ideal constant distribution. Finally, again an inboard shift in thrust distribution
is observed for the optimized propeller geometries.
9
Radial coordinate r / Rp
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
Profile efficiency (%)
100
80
60
40
20
200
150
100
50
0
T´ (N / m)
Blade pitch angle β (deg)
20
30
40
50
60
70
Chord c / Rp
0.20
0.15
0.10
0.05
XPROP-iso
isoOpt-iso
insOpt-ins
Fig. 9 Propeller blade loading and geometry changes due to optimization for T =0.21Diso.
Radial coordinate r / Rp
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
Profile efficiency (%)
100
80
60
40
20
200
100
0
T´ (N / m)
Blade pitch angle β (deg)
20
30
40
50
60
70
Chord c / Rp
0.20
0.15
0.10
0.05
XPROP-iso
isoOpt-iso
insOpt-ins
300
400
Fig. 10 Propeller blade loading and geometry changes due to optimization for T =0.39Diso.
10
C. Propeller Optimization with Varying Radius
The performance of the propeller placed in the wingtip flowfield was also investigated for propeller radii dierent
from the XPROP propeller. The radius was varied between 0
.
75
1
.
50
Rp,ref
, where
Rp,ref
is the radius of the XPROP
propeller. A design thrust range up to
T=Diso
was analysed for the largest radius propeller. The hub dimension was
kept constant and was not scaled with propeller radius. In this analysis the performance of the propeller optimized
for installed condition and operating in installed condition, insOpt-ins, was compared to a propeller of equal radius
optimized for isolated condition operating in isolated condition, isoOpt-iso. This is the most fair comparison since
both propellers are optimised for their respective flowfield. The propeller performance results are given in Fig. 11 as
a function of thrust level. For any given thrust level, the propulsive eciency
ηp
increases with increasing propeller
radius for all propeller designs. This is a straight forward result since, with increasing radius and equal thrust, more
mass is accelerated by the propeller but with a lower velocity increase. The corresponding advance ratio is also higher
for larger radii propellers due to the large reduction in required rotational speed n.
80
85
90
95
100
105
1.0
1.2
1.4
1.6
1.8
2.0
0.75Rp,ref
1.00Rp,ref
1.25Rp,ref
1.50Rp,ref
0.0 0.2 0.4 0.6 0.8 1.0
6
8
10
12
14
16
18
0.0 0.2 0.4 0.6 0.8 1.0
82
84
86
88
90
92
Propulsive efficiency ηp (%)ηpηp, iso (%)
Power P / Piso (%) Advance ratio J
Thrust over drag ratio T / Diso
Fig. 11 Optimization results with varying propeller radius for installed condition with insOpt-ins design com-
pared to isolated condition with isoOpt-iso design.
The gain in eciency by installation of the propeller at the wingtip,
ηpηp,iso
, varies less with propeller radius.
Only for low thrust levels of less than 30% of the wing drag, significant dierences occur in the eciency gain. While
the eciency gain decreases with thrust level, it does so at a reducing rate, decreasing to a still significant gain of 7
.
5%
when the thrust equals the wing drag
T=
1
.
00
Diso
. As discussed earlier for the optimisation results at equal radius, the
power ratio
P/Piso
seems to level oto an approximately constant value at larger thrust levels. This is also happening
for the other propeller radii, however the relative power reduction reduces for larger propeller radii, reducing to a 9%
reduction due to the wingtip flowfield for 1.50Rp.
11
VI. Actuator Disk–Wing Analysis
The upstream eect of the propeller designs on the wing performance was analyzed through CFD analyses of the
wing with an actuator disk representation of the propeller. These simulations are of similar computational cost as the
isolated wing simulations but do provide the required time-averaged upstream eect of the propeller at much reduced
computational cost compared to the fully resolved propeller–wing simulation in Section VII. In Fig. 12 the wing lift
coecient
CL
and lift-over-drag ratio
L/D
are plotted as a function of propeller thrust for a number of cases with
varying propeller radius from Fig. 11. Results are also shown for the isolated wing.
0.0 0.2 0.4 0.6 0.8 1.0
0.3446
0.3447
0.3448
0.3449
0.3450
0.3451
0.3452
0.0 0.2 0.4 0.6 0.8 1.0
21.30
21.35
21.40
21.45
21.50
21.55
21.60 Prop off
0.75Rp,ref
1.00Rp,ref
1.25Rp,ref
1.50Rp,ref
Thrust over drag ratio T / Diso
Wing lift coefficient CL
Wing lift-over-drag ratio L / D
1.50Rp,ref no swirl
Fig. 12 Wing performance as function of propeller thrust investigated with an actuator disk.
The changes in wing lift coecient due to propeller thrust are quite small and are only just larger than the estimated
discretization uncertainty given in Table 3. Despite this, an increasing trend in
CL
is visible with increasing propeller
thrust level. This could be explained by a slightly stronger suction over the wing due to the presence of the propeller.
Contrary to the lift coecient, the lift-over-drag ratio decreases with increasing propeller thrust. Apparently the
propeller slightly reduces the wing eciency. Another observation is that with increasing propeller radius, the upstream
eects of the propeller on the wing reduce. For the 1
.
50
Rp
case an additional simulation was performed where no
swirl component was introduced by the actuator disk, denoted no swirl. It shows that the swirl component has only a
minor impact on the wing loading. Thus, the reduction of swirl in the wingtip flowfield due to the propeller does not
cause a significant change in wing loading and the observed changes are mainly an eect of the axial velocity increase.
However in general, it is concluded that the upstream eect of the propeller on the wing loading is very limited for the
investigated thrust range. A closer look at the downstream interaction of the wing on the propeller is presented in the
next section.
VII. Propeller–Wing Analysis
To check the accuracy of PROPR for the non-uniform wingtip flowfield, and to investigate the unsteady behaviour
of a wingtip-mounted pusher propeller, a fully resolved propeller–wing CFD simulation was performed. The original
XPROP propeller was simulated installed on the wingtip at optimum operating condition in terms of
β0.7Rp
and
J
at a
thrust of T=0.30Diso.
In Fig. 13 a the normalized blade thrust and torque as function of blade phase angle are shown, as defined in Fig. 14
b. In addition, the mean of the thrust and torque obtained from the transient CFD simulation are shown, as well as the
thrust and torque values calculated using PROPR. As expected, the blade thrust and torque vary considerably over a
rotation due to the circumferential non-uniformities in the flowfield that were shown in Fig. 5. Following the trajectory
of the blade, clear trends are observed:
ϕ=0: Initial position, blade pointing up and perpendicular to the wing surface.
ϕ=90: Blade is fully outboard and subjected to the largest negative tangential velocities, thus highest thrust.
12
36031527022518013590450 36031527022518013590450
1.2
1.1
1.0
0.9
0.8
0.7
0.6
1.2
1.1
1.0
0.9
0.8
0.7
0.6
Propeller blade phase angle φ (deg)
Blade thrust T / TPROPR
Blade torque Q / QPROPR
CFD
PROPR
CFD mean
Fig. 13 Installed XPROP propeller blade thrust and torque evolution over a complete rotation for
T=0.30Diso.
ϕ=
180
: Blade is again perpendicular to the wing but now pointing down. Thrust delivered with inflow from
under the wing is slightly lower.
ϕ=
270
: At this angle the blade is fully immersed in the wake region of the wing, showing a clear local peak in
thrust, due to the lower axial velocity. This causes an increased angle of attack of the blade sections.
Identical trends are observed when investigating the torque evolution. The dierence in mean thrust and torque
obtained between CFD and PROPR are very similar to those obtained for the isolated propeller in Section IV. The blade
thrust and torque are overestimated 3
.
3% and 4
.
0% by PROPR respectively. This provides confidence in the validity of
PROPR given these non-uniform flowfields.
To get some more insight in the blade loading evolution, in Fig. 14 a the propeller blade thrust distribution is plotted
at various blade positions as defined in Fig. 14 b. The blade distribution from PROPR is also shown. It is seen that the
thrust distribution over the blade from PROPR is nearly identical to that found in the transient CFD simulation over
blade 1. Blade 2 and 3 are subjected to the largest tangential velocity field and indeed show the highest thrust. Most
notably, these blades experience a significantly higher thrust at the root sections of the blade. Blade 5 experiences the
lowest blade loading because, as was shown in Fig. 5, the tangential velocity field at this location is near zero. A drop in
thrust near the root occurring over blade 6 is seen, because it is immersed in a flowfield with locally positive tangential
velocity. Blade 6 experiences this positive tangential velocity near the root, because it crosses the tip vortex of the wing.
This becomes clear from Fig 15, where a tangential and an axial vorticity isosurface is shown at such levels that they
identify the propeller blade tip vortices and wingtip vortex respectively. The propeller blades are at the same position as
in Fig. 14 b. It is clear that blade 6 is starting to cross the wingtip vortex.
13
1
2
3
4
5
6
Thrust distribution T´ (N / m)
Radial coordinate r / Rp
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
100
0
50
−50
150
200
250
300
350
CFD - blade 1
CFD - blade 2
CFD - blade 3
CFD - blade 4
CFD - blade 5
CFD - blade 6
PROPR
a) Propeller blade thrust distribution comparison. b) Definition of blade positions.
yp
zp
φ
5
6
4
3
2
1
Fig. 14 Comparison of XPROP propeller blade thrust distributions at dierent blade positions for
T=0.30Diso.
a) Side view of vorticity isosurfaces.
b) Rear view of vorticity isosurfaces.
ωx = −900 s-1
ωt = 800 s-1
wingtip vortex
blade tip vortices
Fig. 15 Vorticity isosurfaces identifying the XPROP propeller blade tip vortices and wingtip vortex.
14
VIII. Conclusions
A design framework for wingtip-mounted pusher propellers was established and verified with RANS CFD sim-
ulations. It was found that the following approach is valid: First a wingtip flowfield was extracted from an isolated
wing simulation. Then, this flowfield was circumferentially averaged and used in a validated lifting-line based propeller
analysis and optimisation routine named PROPR for aerodynamic design optimization. The upstream eect of the
propeller on the wing performance was investigated with an actuator disk representation of the propeller in multiple
wing simulations and was found to be very limited for the tested thrust levels. This was a requirement for the taken
approach. However, a trend of increasing wing lift coecient and decreasing lift-over-drag ratio was observed with
increasing propeller thrust, which may become significant for higher thrust levels than currently investigated. At last,
the resulting propeller performance from PROPR was in line with time-averaged propeller loading of fully resolved
propeller-wing RANS CFD simulations, providing confidence in the validity of the approach. This was despite the large
fluctuation found in the transient propeller blade loading due to circumferential non-uniformities in the inflow field.
With this design framework for wingtip-mounted pusher propellers, various conclusions on propeller design for
wingtip-mounted pusher propellers are drawn:
The possible propulsive eciency gains for the propeller due to installation are significant: Up to 16% increase in
eciency was achieved at the lowest investigated thrust level of 5% of the wing drag, decreasing with a reducing
slope to approximately 7.5% for the highest investigated thrust level of 100% of the wing drag.
The propulsive eciency gain was only significantly dependent on the propeller radius for low thrust levels of
less than 30% of the wing drag.
It was found that eectively the propeller geometry is optimized for the required thrust level and to a lesser degree
for the non-uniformity in the flowfield.
Propeller blade optimization and installation result in higher profile eciency in the blade root sections and a
more inboard thrust distribution over the blade.
Considering the propulsive eciency benefits found in this study, wingtip-mounted pusher-propellers should be
considered for (hybrid)-electric concepts. The propeller thrust should then be a design variable, determining together
with the wing lift the resulting eciency benefit from this aerodynamic interaction.
References
[1]
Loth, J. L., and Loth, F., “Induced Drag Reduction with Wing Tip Mounted Propellers,” 2nd Applied Aerodynamics Conference,
AIAA Paper 1984-2149, Aug. 1984. doi:10.2514/6.1984-2149.
[2]
Miranda, L. R., and Brennan, J. E., “Aerodynamic Eects of Wingtip-Mounted Propellers and Turbines,” 4th Applied
Aerodynamics Conference, AIAA Paper 1986-1802, June 1986. doi:10.2514/6.1986-1802.
[3]
Sinnige, T., van Arnhem, N., Stokkermans, T. C. A., Eitelberg, G., and Veldhuis, L. L. M., “Wingtip-Mounted Propellers:
Aerodynamic Analysis of Interaction Eects and Comparison with Conventional Layout,Journal of Aircraft, Vol. 56, No. 1,
2019, pp. 295–312. doi:10.2514/1.C034978.
[4]
Patterson Jr, J. C., and Bartlett, G. R., “Eect of a Wing Tip-Mounted Pusher Turboprop on the Aerodynamic Characteristics
of a Semi-Span Wing,AIAA/SAE/ASME/ASEE 21st Joint Propulsion Conference, AIAA Paper 1985-1286, July 1985. doi:
10.2514/6.1985-1286.
[5]
Janus, J. M., Chatterjee, A., and Cave, C., “Computational Analysis of a Wingtip-Mounted Pusher Turboprop,” Journal of
Aircraft, Vol. 33, No. 2, 1996, pp. 441–444. doi:10.2514/3.46959.
[6]
Stokkermans, T., Voskuijl, M., Veldhuis, L., Soemarwoto, B., Fukari, R., and Eglin, P., “Aerodynamic Installation Eects of
Lateral Rotors on a Novel Compound Helicopter Configuration,AHS International 74th Annual Forum &Technology Display,
May 2018.
[7]
Moore, M. D., and Fredericks, B., “Misconceptions of Electric Propulsion Aircraft and their Emergent Aviation Markets,52nd
Aerospace Sciences Meeting, AIAA Paper 2014-0535, Jan. 2014. doi:10.2514/6.2014-0535.
[8]
Borer, N. K., Patterson, M. D., Viken, J. K., Moore, M. D., Clarke, S., Redifer, M. E., Christie, R. J., Stoll, A. M., Dubois, A.,
Bevirt, J., Gibson, A. R., Foster, T. J., and Osterkamp, P. G., “Design and Performance of the NASA SCEPTOR Distributed
Electric Propulsion Flight Demonstrator,” 16th AIAA Aviation Technology, Integration, and Operations Conference, AIAA
Paper 2016-3920, June 2016. doi:10.2514/6.2016-3920.
[9]
Stokkermans, T. C. A., van Arnhem, N., and Veldhuis, L. L. M., “Mitigation of Propeller Kinetic Energy Losses with Boundary
Layer Ingestion and Swirl Recovery Vanes,” Proceedings of the 2016 Applied Aerodynamics Research Conference, Royal
Aeronautical Soc., London, 2016, pp. 56–69.
[10]
Nootebos, S., “Aerodynamic Analysis and Optimisation of Wingtip-Mounted Pusher Propellers,” Master’s thesis, Delft
University of Technology, 2018.
15
[11]
Costruzioni Aeronautiche TECNAM S.p.A., “P2012 Traveller,”
https://www.tecnam.com/aircraft/p2012-
traveller/, Accessed: 2018-10-30.
[12]
Alba, C., “A Surrogate-Based Multi-Disciplinary Design Optimization Framework Exploiting Wing-Propeller Interaction,
Master’s thesis, Delft University of Technology, 2017.
[13]
Stokkermans, T. C. A., and Veldhuis, L. L. M., “Propeller Performance at Large Incidence Angle for Compound Helicopters
and eVTOL,” (unpublished).
[14] Anon., “ANSYSrAcademic Research Release 18.1 Help System, Fluent,” ANSYS, inc., Canonsburg, PA, 2017.
[15]
Barth, T. J., and Jesperson, D. C., “The Design and Application of Upwind Schemes on Unstructured Meshes,27th Aerospace
Sciences Meeting, AIAA Paper 1989-366, Jan. 1989. doi:10.2514/6.1989-366.
[16]
Kim, S.-E., and Rhee, S. H., “Ecient Engineering Prediction of Turbulent Wing Tip Vortex Flows,Computer Modeling in
Engineering &Sciences (CMES), Vol. 62, No. 3, 2010, pp. 291–309. doi:10.3970/cmes.2010.062.291.
[17]
Spalart, P. R., and Allmaras, S. R., “A One-Equation Turbulence Model for Aerodynamic Flows,” 30th Aerospace Sciences
Meeting and Exhibit, AIAA Paper 1992-439, June 1992. doi:10.2514/6.1992-439.
[18]
Dacles-Mariani, J., Zilliac, G. G., Chow, J. S., and Bradshaw, P., “Numerical/Experimental Study of a Wingtip Vortex in the
Near Field,” AIAA Journal, Vol. 33, No. 9, 1995, pp. 1561–1568. doi:10.2514/3.12826.
[19]
Spalart, P. R., and Rumsey, C. L., “Eective Inflow Conditions for Turbulence Models in Aerodynamic Calculations,” AIAA
Journal, Vol. 45, No. 10, 2007, pp. 2544–2553. doi:10.2514/1.29373.
[20] Stokkermans, T. C. A., van Arnhem, N., Sinnige, T., and Veldhuis, L. L. M., “Validation and Comparison of RANS Propeller
Modeling Methods for Tip-Mounted Applications,AIAA Journal, Vol. 57, No. 2, 2019, pp. 566–580. doi:10.2514/1.J057398.
[21]
Drela, M., and Youngren, H., “XROTOR 7.55 (Computer software),” , Retrieved from:
http://web.mit.edu/drela/
Public/web/xrotor/.
[22]
Hirner, A., Dorn, F., Lutz, T., and Krämer, E., “Improvement of Propulsive Eciency by Dedicated Stern Thruster Design,7th
ATIO Conference, AIAA Paper 2007-7702, Sept. 2007. doi:10.2514/6.2007-7702.
[23]
Sinnige, T., “The Eects of Pylon Blowing on Pusher Propeller Performance and Noise Emissions,” Master’s thesis, Delft
University of Technology, 2013.
[24]
Borer, N., and Moore, M., “Integrated Propeller - Wing Design Exploration for Distributed Propulsion Concepts,53rd AIAA
Aerospace Sciences Meeting, AIAA Paper 2015-1672, Jan. 2015. doi:10.2514/6.2015-1672.
[25]
Humpert, B. J., Gaeta, R. J., and Jacob, J. D., “Optimal Propeller Design for Quiet Aircraft Using Numerical Analysis,” 21st
AIAA/CEAS Aeroacoustics Conference, AIAA Paper 2015-2360, June 2015. doi:10.2514/6.2015-2360.
[26]
Drela, M., “XFOIL: An Analysis and Design System for Low Reynolds Number Airfoils,Low Reynolds Number Aerodynamics,
Springer Berlin Heidelberg, 1989.
[27]
Snel, H., Houwink, R., and Bosschers, J., “Sectional Prediction of Lift Coecients on Rotating Wind Turbine Blades in Stall,”
Tech. Rep. ECN-C–93-052, Netherlands Energy Research Foundation, 1994.
[28]
Roache, P. J., “Quantification of Uncertainty in Computational Fluid Dynamics,” Annual Review of Fluid Mechanics, Vol. 29,
No. 1, 1997, pp. 123–160. doi:10.1146/annurev.fluid.29.1.123.
[29]
Eça, L., and Hoekstra, M., “Discretization Uncertainty Estimation Based on a Least Squares Version of the Grid Convergence
Index,Proceedings of the Second Workshop on CFD Uncertainty Analysis, Oct. 2006.
16
... This is a typical example where as a concept, often a relatively simple inflow is envisioned consisting of only a swirling inflow, while in practice the inflow is highly distorted and nonuniform. The mesh and inflow distribution from Ref. [190] were used to obtain the required validation data. The propeller is again the XPROP, but with a blade pitch of β 0. Although the wing-tip and propeller axis are aligned, the wing-tip vortex rolls over the nacelle causing a large variation of J t . ...
... References [48,63,184,201,202] demonstrate that by careful installation, significant propeller efficiency gains can be achieved relative to a propeller that operates in freestream conditions. For example, axisymmetric inflow that covers a large part of the propeller disk, such as a swirling inflow to a wing-tip mounted pusher-propeller or a boundary layer inflow, have shown to enhance the propeller efficiency typically by 5% to 10% [56,57,190,203]. Local disturbances, such as the propeller encountering a wake or vortex, have a small effect on the mean propeller forces, while they do cause significant unsteady loads and noise [89,187,[204][205][206][207][208]. ...
... A pusher propeller that is mounted to the tip of a wing operates in the flowfield induced by the wing tip-vortex and the bound vortex, and in the wing wake, with the objective of high propeller efficiency caused by the swirling inflow [36]. Several authors have confirmed the propeller efficiency benefit [57,190]. A comprehensive analysis of the unsteady blade loads, the shift in the propeller efficiency curve, and the dependency of diameter and vertical position on these performance indicators is not known from literature. ...
Thesis
Full-text available
It is envisioned that the future generations of regional and short to medium-range aircraft employ a high level of propeller integration to achieve low-emission flight. The objectives of unconventional propeller installations include the enhancement of the airframe aerodynamic efficiency, increasing the propeller efficiency, improving cabin comfort, and improving the overall aircraft design by lower operating empty weight. Furthermore, by employing the aerodynamic interaction in specific phases of the flight, beneficial propulsion integration can also enable the use of alternative energy sources and increase the electrification level of the propulsion system. The closer proximity of the propeller and airframe requires a more dedicated integral design (approach) of both the airframe and propulsion unit. The objective of this dissertation is: to characterize the role of the aerodynamic interaction between the propeller and the airframe on the performance and static stability characteristics for selected aircraft configurations which aim for a beneficial propeller-airframe interaction. To this end, three different types of analyses are performed. First, fundamental phenomena are investigated which provide insight for related configurations and derivatives thereof. Second, a configuration study indicates the expected trends on various performance indicators. Finally, two detailed studies on aircraft level demonstrate the relative importance and the coupling between aerodynamic interactions. The first configuration features propellers that are mounted to the horizontal tailplane. This is an example where there is a strong interaction between the propeller and airframe that affects performance, stability, and control, and contains various interaction mechanisms that are of interest for other configurations as well. A second specific case is the a distributed propulsion configuration with propellers mounted to the inboard part of the wing (in front of the high lift devices), together with a propeller mounted to the tip of the wing. One of the focal points of the current study is extending the understanding of nonuniform inflow effects on propeller performance and its role in aircraft stability and trim. Compared to the conventional configuration, for various unconventional propeller installations, the nonuniform inflow to the propeller differs both in type and magnitude, and varies with flight condition. The slipstream shape and consequently its interaction with lifting surfaces are affected as well. These factors directly affect the gradients and offsets of the propeller force curves and therefore the aircraft stability and trim, respectively. By employing CFD results, a study has been performed on the {sensitivity} of the radial load distribution to a change in inflow condition that is expressed as a change in local advance ratio. The constructed distributions provide insight into what region of the disk is responsible for the largest changes of the propeller forces. This has been demonstrated to be the region of highest loading. It is also shown that for a given propeller design, nonuniform inflow can be represented by an `installation coefficient' kappa such that the efficiency curve of this uninstalled propeller is scaled along the advance ratio and efficiency directions by kappa to obtain the installed propeller efficiency. Using the data of the isolated propeller for an arbitrary blade angle, the advance ratio at which the installed case has highest efficiency, as well as the value of the maximum efficiency, can be quantified immediately. The computational intensive analyses to find the optimum blade angle for the installed cases are therefore redundant and the formulation of the installation coefficient is therefore highly valuable to the aircraft designer. The installation coefficient also gives insight in what regime of the efficiency-advance-ratio curve the largest changes occur due to nonuniform inflow. [...]
... Different concepts with interacting propulsion systems such as DP and WTP are currently being investigated. DP configurations are subject of research by the DLR [4,5], at the TU-Delft [6], the NASA with the projects X57 [7] and Pegasus [8], as well as the Onera with the Ampere project [9]. The positive effects on WTP were already demonstrated in wind tunnel tests in the 1960-80s [10][11][12][13][14]. Design studies on WTP configurations have recently been carried out by Stokkermans [6] and by Blaesser [8]. ...
... DP configurations are subject of research by the DLR [4,5], at the TU-Delft [6], the NASA with the projects X57 [7] and Pegasus [8], as well as the Onera with the Ampere project [9]. The positive effects on WTP were already demonstrated in wind tunnel tests in the 1960-80s [10][11][12][13][14]. Design studies on WTP configurations have recently been carried out by Stokkermans [6] and by Blaesser [8]. However, a general aerodynamic comparison of the concepts under the same boundary conditions with regard to their energy efficiency has not yet been carried out. ...
Article
Full-text available
Beneficial interactions that occur between propellers and the wing can be used to increase the overall efficiency of an aircraft in cruise flight. Different concepts with such interacting propellers are distributed propulsion (DP) and wingtip mounted propellers (WTP). For DP, a full distribution over the entire span can be distinguished from a partial distribution, concentrating the propellers at the wing tip area. The paper focuses on the energy efficiency in cruise flight as a result of the interactions and provides a general comparison of the concepts (WTP, full and partial DP) with a Beechcraft 1900D commuter aircraft as a reference. Parametric CFD studies varying the number and the position of the propellers are performed with a half-wing model. The simulations are performed with the second-order finite-volume flow solver TAU, developed by the German Aerospace Center (DLR), employing Reynolds-averaged Navier–Stokes (RANS) equations. The propellers are modeled using an Actuator Disk (ACD). An algorithm is used to reach cruise condition by iteratively adjusting the propeller rotational speed and the wing angle of attack. The CFD results are analyzed and evaluated with respect to the overall efficiency including the aerodynamic efficiency of the wing as well as the propulsive efficiency of the propellers. The parameter study shows that in cruise flight partial DP is more efficient than a full DP. The pure WTP configuration was found as the optimum of the propeller distribution along the wing, resulting in a saving of required power of 5.6%5.6\% 5.6 % , relative to the reference configuration.
... With the subsequent advance in CFD methods, the interaction between propellers and wings on fixed-wing aircraft was numerically analysed with increasing fidelity by conducting Euler [51], RANS [52], URANS [53,54], or delayed detached-eddy simulations (DDES) [55]. Janus et al. [51] confirmed the experimental results of Patterson and Bartlett with their analysis of the same generic geometry and an analysis of the lift breakdown by components for different setups omitting propeller and nacelle. ...
... Janus et al. [51] confirmed the experimental results of Patterson and Bartlett with their analysis of the same generic geometry and an analysis of the lift breakdown by components for different setups omitting propeller and nacelle. The studies of Veldhuis [52], Stokkermans et al. [53], and Schollenberger et al. [54] focussed on the influence of propeller thrust level and position, respectively, with the propellers represented by actuator lines (AL) in the latter analysis, whereas Jayasundara et al. [55] investigated the unsteady aerodynamic phenomena of the National Aeronautics and Space Administration (NASA) X-57's wing-tip-mounted (tractor) propeller and the resulting interaction-related noise characteristics. ...
Thesis
Full-text available
In this thesis, the mutual effects of main rotor, wings, and lateral rotors of Airbus Helicopters' compound helicopter RACER are addressed under different flight conditions. In order to isolate the interactional phenomena, trimmed high-fidelity computational fluid dynamics (CFD) simulations are not only conducted on a detailed representation of the full compound helicopter but also on reduced configurations which are created by removing the respective other components of interest. By comparing the aerodynamic performance of main rotor, wings, and lateral rotors for the different computations, mutual influences can consequently be determined, which are further divided into first- and second-order effects. With the help of a loose coupling between the comprehensive analysis (CA) tool HOST and the CFD solver FLOWer, realistically trimmed free-flight conditions are determined for the relevant flight states. The first of these is RACER's cruise flight at 220kts, which represents the operating condition the compound helicopter was specifically designed for. Due to the elevated advance ratio of over 0.5 and the resulting azimuthal variation of main rotor inflow, a strong asymmetry is witnessed in unsteady Reynolds-averaged Navier-Stokes (URANS) simulations, where most of the interactional phenomena originating from or affecting the advancing side of the rotor disk are significantly stronger than on the retreating side. In order to account for the large regions of separated flow below the wings, RACER's hover as the second flight condition of interest is analysed by means of delayed detached-eddy simulations (DDES). It is not only that second-order effects generally play a more important role for this flight state, but that asymmetries in aerodynamic interactions are not linked to the main rotor and its thrust distribution but rather to the different operating conditions of the lateral rotors where the right-hand rotor generates reverse thrust to provide anti-torque.
... As shown in Fig. 16, the positive direction of spanwise flow is positive y direction, and the positive spanwise velocity component above the wing in inboard-up is significantly smaller than that in inboard-down in both amplitude and duration, indicating that the wing-tip vorticity is effectively offset at this position. Peter Sharpe [12] and Stokkermans [18] et al. pointed out that rotor completely installed on the wing tips can effectively counteract the influence of the wing tip vortex. However, in this paper, because the installation position of the outer rotor is still some distance away from the wing tip, even if its rotation direction is opposite to the direction of the wing tip vortex, the spanwise velocity of the inboard-up at section 8 is still not much different from that of the inboard-down (see Fig. 17). ...
Chapter
Full-text available
n this paper, a specific configuration of tilt-rotor eVTOL with the propeller installed in front of the wing was analyzed. We do the coupling design of pylon and wing and then unsteady RANS CFD method with sliding mesh strategy was applied for the rotation domain, and attention was paid to the propeller-wing coupling effect. For the propeller, compared with the isolated condition, the inner propeller thrust decreased slightly (about −2.1% at α = 0°) and the efficiency reduced a little (about 1% at α = 0°) in the case of coupling with the wing, but the outer propeller thrust increased by 2.3%, presenting a symmetric situation, and the cruise efficiency remains unchanged. For the main wing, the direction of rotation of the outer propeller affects its aerodynamic characteristics: When the rotation direction is inboard-up, the lift coefficient of the whole aircraft was increased by 5.6%, and the lift-drag ratio is reduced by 2.9%. The aerodynamic efficiency of inboard-up is higher than that of the inboard-down, because the propeller slipstream introduced by the propeller whose rotation direction opposite to the wingtip vortex cancels out part of the downwash effect of the wingtip vortex, resulting in lower induced drag. This counteracting effect is also transmitted to the upstream propellers in reverse. The propeller slipstream also changes the angle of attack and velocity of the local air flow on the wing, so the surface pressure distribution and circulation distribution are changed. The installation position and rotation direction of propeller have a profound effect on slipstream, wingtip vortex, etc.
... The maximum thrust is obtained at 83% of the propeller radius, which is consistent with the general propeller characteristics presented in. 30 The symmetric NACA0012 airfoil can be observed to have the highest angle of attack α a (r) compared to the cambered airfoils, which could have been anticipated since the cambered airfoils are intended to produce higher lift at lower α a (r) when compared to the symmetric airfoil. ...
Article
Full-text available
This study presents a predominantly numerical and theoretical investigation into the balance of tonal and broadband noise due to an isolated propeller in uniform motion. The predicted trends in the balance between tonal and broadband noise radiation with varying blade number and speed of rotation is supported by preliminary experimental measurements. Here, we assume that the dominant noise generation mechanisms are the tones due to steady loading and blade thickness, while the broadband noise is due to boundary layer scattering at the trailing edge. The study also provides a detailed comparison between the tonal and broadband formulations to highlight their similarities and differences. In this paper, we show that the main differences in the behaviour and character of the tonal and broadband spectra and directivities are due to the number of acoustic modes that can be excited. This paper presents a parametric study in which the variation in tonal and broadband noise is investigated as a function of blade tip Mach number ( M t ) and blade number ( B) whilst maintaining constant solidity and thrust. This study is repeated for three NACA airfoil profiles. It is found that tonal noise dominates at low blade number and low frequency and/or higher tip speeds, while broadband noise is the major contributor at high-frequencies and at high blade number and low tip speeds. The results show a clear distinction between the combinations of M t and B that are dominated by tonal and by broadband noise. These results are interpreted from fundamental principles relating to modal radiation efficiencies. We confirm this trend of balance between tonal and broadband noise with measured noise at different B and M t . The results of this paper will serve as useful guidelines for preliminary propeller design.
... Among the available methods, Large-eddy simulation (LES) is a high-fidelity alternative that enables a more fundamental analysis of the noise generation and propagation mechanisms. On the other hand, the use of hybrid approaches (Afari et al., 2019;Avallone et al., 2018) as well as unsteady Reynolds Average Navier-Stokes (RANS) simulations (Chirico et al., 2017;Hambrey, 2017;Ben Nasr et al., 2017;Arnhem et al., 2018;Stokkermans et al., 2019;Bento et al., 2020) can deliver reasonable results at reduced computational cost. Analytical, semi-empirical and reduced order methods are desirable in the design process largely because of the ability to quickly estimate the impact of design parameters on noise levels and performance characteristics within an optimization framework. ...
... Large-eddy simulation (LES) is a high-fidelity alternative that makes possible the deeper analyses of the phenomena. On the other hand, the use of hybrid approaches [2,3] as well as unsteady Reynolds Averaged Navier-Stokes (RANS) simulations [4][5][6][7][8][9] can lead to reasonable results with reduced computational cost. Analytical, semi-empirical and reduced order methods are desired for the design process as their use make possible the analyses of multiple configurations and optimization processes. ...
Conference Paper
Full-text available
Propeller-driven aircraft noise has been a problem to passengers and community for decades. With the recent advances and focus on the development of electric propulsion and unmanned aerial vehicles, the noise emitted by propellers is still a concern. The design of low-noise propellers requires a balance between noise emission and aerodynamic performance, which commonly walk in different ways during design. Numerical approaches such computational fluid dynamics (CFD) coupled with noise prediction methods can aid in the design of propellers and in their installation effects. Analytical, semi-empirical or reduced order methods are desired for the design process as their use make possible the analyses of multiple configurations and optimization processes. We propose an analysis and validation of a simulation model for predicting the flow field and aerodynamic characteristics of propellers using the opensource code OpenFOAM. A 4-blade tractor propeller geometry of diameter D = 0.237 m was chosen from the literature. Unsteady Reynolds Averaged Navier-Stokes simulations were conducted using the SST turbulence model. A dynamic mesh approach was adopted to simulate the rotation of the propeller, in which the domain is divided in a static region and a rotating region. Sensitivity to grid refinement and boundary conditions are assessed and results are validated through comparisons with reference data. The acoustic field is obtained by different approaches: acoustic analogies and semi-analytical methods, and the results are compared to other numerical data in the literature. In the first approach the tonal noise field is obtained from the simulated flow field using the Ffowcs-Williams and Hawkings analogy. In the second approach, the distributed thrust and torque coefficients, obtained via CFD, are used as input to semi-analytical methods for the prediction of the thickness and loading components of propeller noise. Results of both approaches are compared in terms of accuracy and computational cost.
... In 2018, [13] determined in wind tunnel tests, that wing tip propellers in tractor configuration lead to a drag reduction from 5 % to 15 % compared to a conventional wing. In 2019, CFD simulations determined that pusher wing tip propellers increase the propulsive efficiency of the propeller by 7.5 % to 16 % depending on the investigated thrust level [14]. ...
Thesis
Full-text available
Hybrid-electric aircraft possess the potential to reduce the CO2 emissions of general aviation aircraft. However, optimized propulsion systems are needed to leverage the advantages and lower the fuel consumption. In order to identify optimized designs, scaling characteristics of the individual propulsion elements are necessary. Furthermore, appropriate power and energy reserves in case of a propulsion component failure are required and the impact of an increased propulsion system mass on the aircraft mass and the aerodynamic drag needs to be considered. The derived trends are implemented into a sizing program and optimized hybrid-electric propulsion systems are determined for a 4-seat hybrid-electric aircraft with a cruise speed of 220 km/h as well as a 9-seat hybrid-electric aircraft with a cruise speed of 400 km/h.
Article
Full-text available
An improved multi-objective design optimization framework is proposed for the efficient design of proprotor blades tailored to specific high-altitude mission requirements. This framework builds upon existing methods by leveraging a reformulated Vortex Particle Method (rVPM) and incorporates three key stages: (1) rapid determination of overall proprotor parameters using a semi-empirical model, (2) optimized blade chord and twist distribution bounds based on minimum energy loss theory, and (3) global optimization with a high-fidelity rVPM-based aerodynamic solver coupled with a multi-objective hybrid optimization algorithm. Applied to a small high-altitude tiltrotor, the framework produced Pareto-optimal proprotor designs with a figure of merit of 0.814 and cruise efficiency of 0.896, exceeding mission targets by over 15%. Key findings indicate that large taper ratios and low twist improve hover performance, while elliptical blade planforms with high twist enhance cruise efficiency, and a tip anhedral further boosts overall performance. This framework streamlines the industrial customization of proprotor blades, significantly reducing the design space for advanced optimization while improving performance in demanding high-altitude environments.
Conference Paper
View Video Presentation: https://doi.org/10.2514/6.2022-3300.vid Wingtip-mounted propellers (WTP) can increase the overall efficiency of an aircraft via positive aerodynamic interactions between the propellers and the wing vortex system. The effect of a WTP thereby depends to a large extent on the propeller design. CFD-based parametric studies on the influence of different propeller parameters (number of blades, radius, blade pitch angle) and aircraft parameters (wingspan and flight velocity) on the interactions are presented for a WTP tractor configuration. Steady-state RANS simulations are performed and the propellers are modelled with an Actuator Disk (ACD) method. It is shown that the consideration of the viscous drag, a realistic representation of the propeller wake by non-uniform loaded propellers and the nacelle geometry have an essential influence on the WTP effects. The rotational speed was identified as the most important parameter for WTP design. Slow-rotating propellers are therefore recommended for WTP applications.
Article
Full-text available
An experimental dataset is presented of propeller performance in static condition and at low subsonic airspeeds for various angles of attack up to 90deg. Numerical investigation through a RANS CFD model revealed the mechanisms behind performance changes with advance ratio, angle of attack and configuration changes. The experimental dataset was found to be free of major errors and is very suitable for validation of propeller models. Furthermore, aerodynamic interaction with an upstream wing was tested with the propeller and wing normal to the flow, to represent the interaction occurring with a time-averaged main rotor slipstream on a compound helicopter. From numerical investigation it was concluded that the results are qualitatively representative of this interaction. The experimental data showed that addition of the wing results in a net reduction of all propeller performance quantities, with thrust reducing up to 20%. A thrust decreasing and increasing mechanism were found numerically. For most tested operating conditions, the wing resulted in a small decrease of propeller thrust-over-power ratio. Decreasing propeller advance ratio, increasing wing distance, and increasing flap deflection generally decreased the effect of the wing on thrust and power, however the influence of flap deflection was found to be small.
Article
Full-text available
The potentially favorable interaction of a propeller slipstream with a wingtip is a complex problem, because of the importance of vortex interaction and viscous effects. This paper examines the capability of different propeller modeling methods in a Reynolds-averaged Navier–Stokes (RANS) solver for the simulation of wingtip-mounted propellers. The applicability of actuator-disk and actuator-line models to reduce the cost of propeller modeling is investigated in its most accurate form, by applying propeller-blade-loading results extracted from simulations, in which the blades are fully resolved. The numerical results are validated by comparison with measurement data from in-house wind-tunnel experiments. It is concluded that the aerodynamic interactions for the wingtip-mounted propeller in tractor configuration can be predicted by RANS simulations with a simple one-equation turbulence model (Spalart–Allmaras), provided that the uncertainty due to numerical diffusion is accounted for by a grid-dependency study, or reduced by local grid refinement. The actuator-line model reduced the computational time by 17% without introducing errors into the time-accurate and time-averaged wing loading. The actuator-disk model reduces the computational cost by 85% by removing time dependency, with a small penalty in the accuracy of the time-averaged flowfield and a 3.9% overprediction of the wing lift.
Article
Full-text available
Wingtip-mounted propellers installed in a tractor configuration can decrease the wing induced drag by attenuating the wingtip vortex by the propeller slipstream. This paper presents an aerodynamic analysis of the propeller–wing interaction effects for the wingtip-mounted propeller configuration, including a comparison with a conventional configuration with the propeller mounted on the inboard part of the wing. Measurements were taken in a low-speed wind tunnel at Delft University of Technology, with two wing models and a low-speed propeller. Particle-image-velocimetry measurements downstream of a symmetric wing with integrated flap highlighted the swirl reductions characteristic of the wingtip-mounted propeller due to wingtip-vortex attenuation and swirl recovery. External-balance and surface-pressure measurements confirmed that this led to an induced-drag reduction with inboard-up propeller rotation. In a direct comparison with a conventional propeller–wing layout, the wingtip-mounted configuration showed a drag reduction of around 15% at a lift coefficient of 0.5 and a thrust coefficient of 0.12. This aerodynamic benefit increased upon increasing the wing lift coefficient and propeller thrust setting. An analysis of the wing performance showed that the aerodynamic benefit of the wingtip-mounted propeller was due to an increase of the wing’s effective span-efficiency parameter.
Conference Paper
Installation effects of the lateral rotors for a compound helicopter were investigated by means of unsteady CFD simulations. The helicopter featured a box-wing design for additional lift in cruise and wingtip-mounted lateral rotors in pusher configuration for additional thrust in cruise and counter-torque in hover. It was found that propeller performance installation effects for a compound helicopter are highly beneficial in cruise, while some penalties exist in extreme and specific cases in hover and autorotation. In cruise the main interaction was between the wing and lateral rotors, resulting in a propulsive efficiency increase up to 10.6% due to wingtip vortex energy recovery. In hover the main rotor slipstream resulted in a near perpendicular inflow to the lateral rotors, with a disturbance from the wings due to the deflection of the main rotor slipstream. For higher than nominal lateral rotor thrust settings, this resulted in a performance penalty. In autorotation a small power installation penalty was present due to inflow disturbances.
Conference Paper
This paper presents multi-disciplinary design optimization (MDO) framework for design of general aviation aircraft wing including the effect of wing mounted propeller on thewing aerodynamic characteristics. In pursuit of achieving such objective, a wing-propeller full-interaction routine was developed, and integrated in a multidisciplinary design analysis including structural and performance models. In this context, innovative solutions to assess wing effects on slipstream development were introduced. Several optimizations were run starting from an existing aircraft design, testing different local and global surrogate-based frameworks, and allowing to assess found optima and recorded computational performances. The achieved optima proved the hypothesis that the inclusion of wing-propeller interaction effects can lead to better designs. Total function calls and running times of the tested frameworks proved that the adoption of surrogate models allows to improve overall optimization performances, provided that suitable surrogate modeling techniques are chosen.