You are on page 1of 15

Aerospace Science and Technology 107 (2020) 106306

Contents lists available at ScienceDirect

Aerospace Science and Technology


www.elsevier.com/locate/aescte

Wind-tunnel and CFD investigations of UAV landing gears and


turrets – Improvements in empirical drag estimation
Falk Götten a,b,∗ , Marc Havermann a , Carsten Braun a , Matthew Marino b , Cees Bil b
a
Department of Aerospace Engineering, FH Aachen University of Applied Sciences, Hohenstaufenallee 6, Aachen, 52064, Germany
b
School of Engineering, RMIT University, 264 Plenty Rd, Bundoora 3083, Australia

a r t i c l e i n f o a b s t r a c t

Article history: This paper analyzes the drag characteristics of several landing gear and turret configurations that
Received 5 August 2020 are representative of unmanned aircraft tricycle landing gears and sensor turrets. A variety of these
Received in revised form 12 October 2020 components were constructed via 3D-printing and analyzed in a wind-tunnel measurement campaign.
Accepted 16 October 2020
Both turrets and landing gears were attached to a modular fuselage that supported both isolated
Available online 22 October 2020
Communicated by Cummings Russell
components and multiple components at a time. Selected cases were numerically investigated with a
Reynolds-averaged Navier-Stokes approach that showed good accuracy when compared to wind-tunnel
Keywords: data. The drag of main gear struts could be significantly reduced via streamlining their cross-sectional
UAV shape and keeping load carrying capabilities similar. The attachment of wheels introduced interference
Landing gear effects that increased strut drag moderately but significantly increased wheel drag compared to isolated
Turret cases. Very similar behavior was identified for front landing gears. The drag of an electro-optical
Drag estimation and infrared sensor turret was found to be much higher than compared to available data of a clean
Wind-tunnel
hemisphere-cylinder combination. This turret drag was merely influenced by geometrical features like
CFD
sensor surfaces and the rotational mechanism. The new data of this study is used to develop simple drag
estimation recommendations for main and front landing gear struts and wheels as well as sensor turrets.
These recommendations take geometrical considerations and interference effects into account.
© 2020 Elsevier Masson SAS. All rights reserved.

1. Introduction system [3]. Studies suggest that streamlining and fairing of UAV
landing gears can offer drag savings and performance gains [3,4].
The relative parasitic drag of small-to-medium-sized fixed-wing Recent research shows an increased interest in aerodynamic
unmanned aircraft (UAVs) is usually higher than larger aircraft analyses of UAVs both in the context of aircraft design and para-
categories like airliners or military transport aircraft [1]. This is sitic drag estimation [5–8]. An accurate parasitic drag estimation in
mainly due to their specific mission scenarios and their compa- UAV design is of utmost importance to provide reliable flight per-
rably small size. Fixed-wing UAVs are often used on surveillance formance predictions. Both sensor turrets and landing gears play
and reconnaissance missions and therefore equipped with gyro- an outstanding role for UAVs, as they represent large external at-
stabilized electro-optical (EO/IR) sensor turrets. These turrets are tachments protruding from the airframe. In general, the empirical
externally attached to the lower airframe and can produce a con- drag estimation of bluff objects like landing gears and turrets re-
siderable amount of drag. Most current fixed-wing surveillance and lies on experience with comparable designs for which drag data is
reconnaissance UAVs are also equipped with non-streamlined and available. This technique does not directly account for the individ-
fixed landing gears. Such landing gears can cause a substantial ual shape and aerodynamic behavior of the specific components,
amount of parasitic drag [2]. Retractable landing gears might not which can result in significant errors in force approximation. The
result in any significant flight performance gains due to their in- level of accuracy is usually accepted in early aircraft design stages.
creased weight and add a considerable level of complexity to the A substantial amount of drag data for large airliner landing gears
is available due to the high interest in landing gear noise reduc-
tion [9–12]. Such data is, however, not transferable to smaller UAVs
*
Corresponding author at: School of Engineering, RMIT University, 264 Plenty Rd, as their landing gear configurations are vastly different from large
Bundoora 3083, Australia.
airliners. Additionally, UAV landing gears operate at significantly
E-mail addresses: goetten@fh-aachen.de (F. Götten), havermann@fh-aachen.de
(M. Havermann), c.braun@fh-aachen.de (C. Braun), matthew.marino@rmit.edu.au lower Reynolds numbers. Recently available literature for aircraft
(M. Marino), cees.bil@rmit.edu.au (C. Bil). and UAV design [13–17] refers directly (or indirectly via Hoerner

https://doi.org/10.1016/j.ast.2020.106306
1270-9638/© 2020 Elsevier Masson SAS. All rights reserved.
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Nomenclature

A streamlined main gear strut x length coordinate


c circular front gear strut/chord x non-attached front wheel
CD drag coefficient X non-attached main wheel
CDf friction drag coefficient y length coordinate
C Dp pressure drag coefficient EARSM Explicit Algebraic Reynolds Stress Model
Cp pressure coefficient EO/IR Electro-optical Infrared
d diameter exp. Experiment
h/d height-to-diameter ratio
FLG Front Landing Gear
L large main gear strut/wheel
IFF Interference Factor
m medium front gear strut/wheel
IWT Industrial Wind Tunnel
M medium main gear strut/wheel
MALE Medium Altitude Long Endurance
r rectangular front gear strut
R rectangular main gear strut MLG Main Landing Gear
Re Reynolds number MTOM Maximum Take-off Mass
s small front gear strut/wheel RANS Reynolds-averaged Navier-Stokes
S small main gear strut/wheel SIMPLE Semi-Implicit Method for Pressure Linked Equations
t thickness SST Shear Stress Transport
S frontal frontal area UAV Unmanned Aerial Vehicle
S ref . reference area θ circumferential coordinate

[18]) to a series of NACA investigations of aircraft landing gears be- with wheels of different sizes to analyze interference effects. Sev-
tween 1934 and 1940 ([19–22]) for determination of UAV or small eral simplified turret geometries with varying height-to-diameter
aircraft landing gear drag. This body of work investigated a vari- ratio are investigated and compared to measurements of an actual
ety of landing gear and wheel configurations that were modern EO/IR turret geometry. The study focuses on the drag characteris-
at that time. Only a few of the analyzed geometries are still used tics of all individual components and their combination. It aims to
today, and the applicability of this data on modern UAV landing provide empirical drag estimation recommendations for UAV land-
gears is questionable. However, a detailed literature study revealed ing gears and turrets and to showcase a series of CFD validation
that this data is still widely used, and no other modern studies cases. The presented data allows an improvement in the accuracy
concerning the drag of UAV landing gears could be identified. An of empirical drag estimation procedures in UAV and small aircraft
initial investigation by the authors [1] found that the landing gear design.
drag data provided by the set of NACA reports underestimated the
drag of UAV landing gears compared to a series of CFD studies by 2. Methodology
up to 80%. Further methods to estimate landing gear drag include
using a variety of drag coefficients for simple geometries like rods, 2.1. Wind-tunnel experiments
elliptical struts and wing-like fairings. These are available in the
The experiments were conducted in the RMIT University Indus-
literature based on wind-tunnel results of simple geometries [23].
trial Wind-tunnel (IWT). The wind-tunnel is a closed-loop return
It is currently unclear if and to what extent these might be appli-
facility that operates under atmospheric conditions with a test sec-
cable on UAV landing gears.
tion size of 9 m × 3 m × 2 m (length × width × height). It is
EO/IR sensor turrets for UAVs attracted much more attention
powered by a 330 kW motor that drives a 6-bladed impeller and
in recent times compared to UAV landing gears. Several studies
can produce flow velocities up to 40 m/s. The tunnel’s clean mean
investigated both aerodynamic and aero-optic effects [24–29]. An
turbulence intensity is about 1.5% at the center of the test section,
excellent overview of recent research is provided by Gordeyev and
where the model is located [32]. Previous analyses showed that the
Jumper [30]. However, most of these studies only investigated sim- turbulence intensity stayed rather constant throughout the length
plified and idealized turret geometries consisting of hemisphere- of the test section. It merely decayed by about 0.3% from the test
cylinder combinations. Actual EO/IR sensor geometries feature sev- section inlet to the outlet [33]. Dynamic pressure was measured
eral protruding surfaces for sensors and a joint rotational mech- with an MKS differential pressure system that was attached to a
anism that allows near 360◦ field-of-view. Such specific features pitot-static tube located in the center of the test section. Further
influence a turret’s aerodynamic behavior and might significantly information on the facility is provided in Refs. [32,34]. Turrets and
increase its drag. Again, no aerodynamic data on actual EO/IR sen- landing gears were attached to a modular dummy fuselage to ap-
sor turret geometries for UAVs could be found in the literature. proximate actual flight conditions. The fuselage was constructed
The lack of reliable aerodynamic data for both UAV landing from three layers of foam material. It had a length of 1.67 m,
gears and turrets poses a serious concern as it greatly increases a height of 0.28 m, and was 0.24 m wide. The middle section
the uncertainty in an empiric parasitic drag estimation that is featured a constant rectangular cross-section with well-rounded
necessary for any UAV design process. This publication, therefore, corners. Both nose and tail were streamlined (see Fig. 1a). No spe-
presents a series of wind-tunnel studies and numerical computa- cific shape optimization was performed, as the fuselage is not the
tions of a variety of UAV landing gears and turrets. The geome- primary geometry of interest in this study. The fuselage has three
tries are derived from an in-house database (see [31]) of over 100 modular attachment points that could support a front landing gear,
fixed-wing UAVs between 20 and 1000 kg maximum take-off mass a turret, and the main landing gear. It could be equipped with ei-
and are representative for a wide variety of UAVs. Several differ- ther one of these parts (Fig. 1b) or with all simultaneously (see
ent front and main landing gear strut configurations are inves- Fig. 1c). The attachment points on the fuselage were smoothly
tigated that include both streamlined and non-streamlined parts. covered when no part was attached to them. The landing gear ar-
The struts are investigated as single components and combined rangement was carefully designed considering overturn angle, tip

2
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 1. Fuselage and turret geometries tested in RMIT’s Industrial Wind-tunnel.

back angle, and front landing gear loading according to guidelines (fuselage length based Reynolds numbers: 2.35 × 106 – 3.54 × 106 ).
from Raymer [35]. This range was chosen as a compromise between measurement ac-
The model was attached to the wind-tunnel wall via a round curacy at low speeds, structural considerations, and the maximum
aluminum sting in a horizontal and inverted position. This en- allowable tunnel velocity. It was used to validate that the measure-
sured that the landing gears and turrets are close to the tunnel’s ments were independent of Reynolds number or dynamic pressure
center section, where the smoothest flow can be expected. The effects. All values presented here were taken from the measure-
maximum blockage ratio was below 4% and regarded as negligible ments at the highest flow velocity, as it is the most representative
[36]. No blocking corrections were applied thereafter. Forces and one for the UAVs included in the in-house database. Landing gear
moments around all axes were measured with a high-precision and turret drag are expected to change with the angle of attack
digital six degree of freedom force-torque sensor from JR3. It al- due to changes in the pressure distribution and separation charac-
lowed measuring the configuration drag with a nominal accuracy teristics. This is neglected in conceptual drag estimation methods
of 0.25% based on the maximum measuring range of 40 N [32]. The [35] and therefore, not considered here. Further investigations are
force-torque sensor was calibrated using dedicated weights before necessary to analyze these effects.
the measurements. This calibration was double-checked after fin- Forces were measured for one minute after the tunnel velocity
ishing the campaign and showed that sensor drift was minimal had stabilized and averaged later on. Several tests were repeated
and did not need correction. Drag forces were the primary in- multiple times on different days and after several installations. The
terest of this study. The force-torque sensor was situated in the maximum drag force deviations between those repeated tests were
center of the fuselage (cover visible in Fig. 1d). This configura- in the order of 3.2% of the total drag force. The percentage uncer-
tion dictates that the load-cell always measured the forces and tainty increases for the individual component drag due to the force
moments of the total installed system, excluding the attachment subtraction procedure described above. The level of uncertainty is
sting. Landing gear or turret drag forces were, therefore, not mea- indicated with error bars for the experimental data given in this
sured directly but were calculated in a two-step approach: First, paper.
measurements of the clean fuselage configurations without any at-
tachments were performed. Afterward, a landing gear or turret was 2.2. CFD analysis
attached to the fuselage, and the measurement repeated. The land-
ing gear or turret drag was then calculated by subtracting the drag Selected cases were numerically analyzed using the Reynolds-
of the clean fuselage configuration from the one with the com- averaged Navier-Stokes (RANS) approach with the commercial
ponent attached. In such, individual component drag in this study computational fluid dynamics software StarCCM+ v15.02. This soft-
is explicitly treated as installed component drag, which might be ware is widely employed in industry and academia and has been
subject to component-fuselage interference effects. No significant extensively validated for a variety of cases [37]. The authors have
fuselage drag increase due to the changed pressure distribution used the same approach for a detailed aerodynamic validation
caused by the attachment of components could be noted in several study of a simplified turret geometry in a previous investigation
CFD studies. Several more cases were investigated that combined [28]. It was found that the numerical method can represent both
different sets of front landing gears with main landing gears and lift, drag, surface pressure, and flow properties of such geometry
turrets. This allowed analyzing interference effects. to a satisfactory degree. All computational grids were created with
All measurements were performed under zero-degrees angle of a hybrid unstructured finite volume algorithm integrated into Star-
attack and five different velocities ranging from 21 m/s to 33 m/s CCM+. The chosen algorithm is based on a cartesian cut-cell ap-

3
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 2. Computational domain, close-up view of mesh, and grid independence study for Turret 3.

proach that predominantly creates hexahedral cells, but also allows suited for secondary flows that can occur in separated regions of
for the creation of arbitrarily shaped cells. A dedicated prismatic bluff bodies. Transition is not modeled, and the flow is assumed
layer mesh was employed to discretize the boundary layer. The nu- to be fully turbulent. This is meaningful, as the 3D-printed land-
merical solution was second-order accurate in space, whereby an ing gear and turret components are not perfectly smooth, and the
upwind scheme was used to discretize convective fluxes. Diffusive wind-tunnel flow quality is unlikely to promote extensive lami-
fluxes were approximated with central difference schemes. nar flow (see also [34]). The complete wind-tunnel test section
The full unsteady RANS equations were solved with an implicit was re-modeled (Fig. 2a - top), placing a velocity inlet boundary
unsteady SIMPLE (Semi-Implicit Method for Pressure Linked Equa- condition at the test section inlet and a pressure outlet bound-
tions) algorithm that was second-order accurate in time (see [38]). ary condition at its outlet. At the inlet condition, the flow velocity
The time step was first approximated based on Strouhal number and its direction were uniformly fixed and corresponded to the
relationships for spheres and cylinders in the given Reynolds num- wind-tunnel investigations at the highest flow velocity. The pres-
ber regime [39]. The final simulations were all run with a time step sure outlet condition was set to the ambient reference pressure,
of 2.5 × 10−4 s. A further reduction of the time-step did not change while the velocity was extrapolated from the adjacent cells. Tur-
the results. After the simulations had stabilized, they were run for bulence variables were fixed at inlet and outlet using a turbulence
at least one more second (4000 time-steps), which corresponds to intensity of 1.5% and a turbulent viscosity ratio of unity [44]. The
about 30-80 vortex oscillation periods of comparable spheres or non-physical decay of turbulent kinetic energy and specific dissi-
cylinders at the given Reynolds numbers. The aerodynamic coeffi- pation rate in the freestream was counteracted by using controlled
cients of this study were later averaged over this period. The Shear decay terms as described in Ref. [44].
Stress Transport (SST) turbulence model [40] was used with an The geometries used in the CFD simulations exactly corre-
Explicit Algebraic Reynolds Stress Model (EARSM) developed from sponded to the ones used in the wind-tunnel experiments, how-
Ref. [41]. The turbulence model was validated multiple times and ever, to reduce discretization effort, small screws that connected
has proven reliable in recent studies involving full-configuration wheels and struts were neglected. The surfaces were discretized
helicopter aerodynamics [42]. Such flows are largely affected by with predominantly quadratic cells of a size that varied depending
unsteadiness and separation effects, which are challenging to cap- on the individual body geometry. The cell sizes on the surfaces are
ture with simulations. small and ensured having at least 300 cells in the circumference
The EARSM non-linear constitutive relation enhances the pre- of cylindrical objects like turrets or wheels (see Fig. 2a-b). Local
diction capabilities of turbulence anisotropy [43]. It is especially surface refinements provided smaller cells, especially at edges or

4
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 3. Main landing gear configurations a-f are to scale.

small radii, which can be seen in Fig. 2b for the EO/IR turret. All 2.3. Landing gear and turret design
near-wall regions were discretized with a boundary layer mesh
consisting of 30 prismatic cell layers. The first cell height ensured 2.3.1. Main landing gear
a y + on the order of 0.5-0.8 for all cases, and at least 10 cells dis- The main landing gear was designed to be representative of a
cretized the viscous sublayer. Care was taken when ensuring that wide range of unmanned aircraft. Its general geometry was derived
the transition between boundary layer mesh and core mesh was from the data provided in an in-house UAV database. This database
smooth, with cell sizes being very similar (see Fig. 2c). All regions contains geometry information and images of over 100 fixed-wing
close to the bodies were highly refined with a specific focus on UAVs between 20 kg and 1000 kg. A great majority of UAVs use
the bodies’ wakes. The mesh was then gradually coarsened fur- solid spring-type main landing gears in which the strut itself pro-
ther away from the bodies. The fuselage was discretized, ensuring vides shock absorption and damping. Prominent UAVs featuring
at least 180 cells in its circumference normal to the flow. The cell this type of main landing gear are AAI’s RQ-7 Shadow, UAV Fac-
density was locally increased in the vicinity of landing gears and tory’s Penguin B, or IAI’s Searcher MK III. The configuration is also
turrets to provide a fine resolution of interference effects. frequently used for general aviation aircraft and captivates through
The final meshes were derived by several grid independence great simplicity and light-weight [35].
studies and have cell counts between 18 and 45 million depending The main landing gear struts designed for this study were sep-
on the configuration. The results of one exemplary grid indepen- arated into two design principles. The first category used a rect-
dency study for a configuration with only the EO/IR turret (descrip- angularly shaped cross-sectional profile, while the second one em-
tion see below) attached to the fuselage are shown in Fig. 2d. The ployed and airfoil type cross-section to streamline the entire strut.
drag coefficients are normalized with the frontal area of the tur- The edges of the rectangular-shaped cross-section profile were
ret normal to the inflow. The drag of this turret does not change rounded to provide some minor flow guidance. A total number
anymore, using more than 20 million cells. Such grid studies were of six different landing gear struts were designed that are shown
repeated for cases with various landing gear and turret configura- in Fig. 3a-f. Three of them featured a rectangular cross-sectional
tions and are not presented here. The forces acting on the bodies shape, while three had a streamlined shape. The general geometry
were determined by a local surface integration of pressure and of all rectangular and all streamlined struts was the same, and only
shear stresses acting on each of the surface cells. The accuracy the struts’ cross-sections and their dimensions were varied. All
of the simulations is judged by comparing them to the respective struts were attached to the fuselage via a connection mechanism
experiments. An error between CFD and experiment below 10% is that provided a smooth fuselage-strut transition (visible in Fig. 3a-
assumed to be reasonable, corresponding to the maximum individ- f). The main landing gear configurations are numbered according
ual component drag accuracy of the wind-tunnel experiments. This to the description in Fig. 3, whereby the first letter indicates which
is affected by measurement uncertainty, repeatability, and force wheels (S-Small, M-Medium, L-Large) are attached to the configu-
subtraction procedure. ration (X - no wheels). The second letter describes the strut size

5
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 4. Statistical evaluation of main landing gear strut size.

(Small – Medium – Large) and the last one if the strut has a rect- 2.3.2. Front landing gear
angular (R) or a streamlined shape (airfoil - A). The sizes can be The front landing gear struts were designed to represent oleo-
compared in Fig. 3a-f, which shows the struts accordingly scaled. pneumatic shock struts that are in widespread use for fixed-wing
The thicknesses and chord lengths of the rectangular-shaped struts UAVs and a variety of general aviation aircraft. Four different front
were chosen according to the geometric results of the in-house struts were designed, whereby two of them had a cylindrical cross-
UAV database. This database connects geometric measures of fixed- sectional shape, and two had a rectangular one with rounded
wing UAVs to their maximum take-off mass (MTOM) and allows edges. The general design of the struts is the same in all cases. The
performing regression analyses. Statistical data available for main struts’ thicknesses were again derived from a statistical evaluation
landing gear struts is shown in Fig. 4. Both thickness (Fig. 4a) and of front landing gears struts shown in Fig. 5a. The same exem-
plary UAV maximum take-off masses were chosen to determine
chord length (Fig. 4b) increase with increasing maximum take-off
the strut thickness as for the main landing gear. The statistical data
mass as would be expected due to higher load carrying require-
scatter is large for the front landing gear strut thickness, and no
ments. Three exemplary take-off masses (50 kg – 250 kg – 800 kg)
clear trend could be evaluated for maximum take-off masses be-
were chosen to evaluate the thickness and chord length of main
low 300 kg. It was, therefore, decided to take a representative strut
landing gear struts. The maximum take-off masses were selected
thickness for a large UAV at 800 kg MTOM and an average one for
to represent a small UAV, a medium-sized tactical UAV, and a large UAVs below 400 kg MTOM. Again, this study uses half-scales. The
MALE UAV. final strut designs are shown in Fig. 5b-e using the same nomen-
The size limitations of the wind-tunnel only allowed half-scales clature as for the main landing gears but with lowercase letters.
of the values found from the statistical evaluation. The smallest The last letter indicates if a strut has a rectangular (r) or cylin-
main strut chord needed to be slightly increased due to structural drical (c) cross-section. The rectangular struts’ thicknesses were
considerations of the half-scale model. Three different wheel sizes adjusted together with the diameters of the cylindrical struts to
that correspond to the strut configurations were determined in the provide both a similar cross-sectional area for normal forces and
same way, while no diagrams are presented here. The wheels were comparable second moments of area for bending due to drag and
connected to the struts via plastic screws. The half-scale strut di- side forces. These requirements could not entirely be met, and the
mensions found in the UAV database were applied to the main final dimensions were chosen as a compromise. However, the de-
landing gear struts with a rectangular cross-section (Fig. 3a-c). signs suggest that it is reasonable to assume that a UAV could be
The streamlined main landing gear struts were designed for fitted with either a rectangularly shaped front strut or its cylindri-
the same load-carrying capabilities as their rectangular counter- cal counterpart and withstand the same landing conditions. This
parts. This allows a direct analysis of the drag reduction achieved guarantees direct comparability of the drag characteristics of both
by streamlining. In a simplified manner, the same load-carrying configurations.
capabilities can be assumed when both the rectangular-shaped Two wheel sizes were determined from a statistical evaluation
cross-section and the streamlined one have the same second mo- of UAV front landing gear wheels and are shown in Fig. 5f. A strut
section combined with a front-wheel is shown in Fig. 5g, whereby
ment of area in the main bending direction. The streamlined cross-
the wheel is attached to the strut via plastic screws. The front
sectional shape was designed using a 30% thick NACA six-series
gears were constructed in the same way as the main landing gears.
airfoil for the complete strut section. This choice allowed providing
a shape with the same second moment of area as the rectangular
2.3.3. Turret
section. While streamlining the main landing gear strut promises Three different turret configurations were investigated in the
significant drag reductions, it can add a considerable amount of wind-tunnel study, see Fig. 6a. Turrets 1 and 2 were simplified and
complexity and cost to the design. These disadvantages have to idealized turret geometries that consisted of a hemisphere and a
be weighed against aerodynamic benefits for the individual UAV. cylindrical part. They had a diameter of 150 mm and 200 mm re-
Such analyses are beyond the scope of this investigation. A com- spectively and were tested with three different height-to-diameter
parison between the streamlined and rectangular cross-section for ratios (h/d1 = 0.5; h/d2 = 0.6; h/d3 = 0.7). The turrets with the
the largest landing gear strut is shown in Fig. 3g. All main landing smallest height-to-diameter ratio just consisted of a hemisphere.
gear parts were 3D printed, sanded, and coated with clear var- The turrets’ sizes were determined by a statistical evaluation of
nish. turret diameters found in the database and presented in Fig. 6b.

6
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 5. Statistical evaluation of front landing gear thickness and final gear designs, b-e are to scale.

Fig. 6. Turret configurations (to scale) and statistical evaluation of turret size.

Their dimensions are appropriate to both represent half-scales of 3. Results and discussion
large UAV turrets and full-scales of smaller ones. Turret 3 was in-
spired as being representative of a variety of actual EO/IR turret The following sections provide the results of the wind-tunnel
geometries with a diameter of 150 mm and a height-to-diameter studies and the numerical computations, including a detailed dis-
ratio of 1.0. It consisted of a spherical part that was designed cussion of the new findings. The wind-tunnel results are used to
to provide pitch motion while a cylindrical part should provide both analyze turret and landing gear drag of UAVs and to provide
yaw motion. These parts were joined together without any flow- validation cases for the CFD studies. The comparisons are restricted
through gap. The turret featured two circular and one rectangular to drag forces, as no pressure probes were used on the model.
flat surfaces on the spherical part. These protuberances should rep- The results of the individual component analyses are presented
resent different sensors of an actual gyro-stabilized EO/IR turret. first, and further studies using multiple components simultane-
All turret geometries were 3D printed, sanded, and coated with ously are presented afterwards. All drag values are shown as co-
varnish. efficients normalized with the dynamic pressure and a reference

7
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 7. Main landing gear strut drag.

area. For highlighting drag force differences due to differently sized main struts’ drag coefficients normalized with this technique are
components, a constant reference area of 1 m2 is chosen accord- shown in Fig. 7b. The struts with a rectangularly shaped cross-
ing to the drag area nomenclature of Hoerner [18]. To develop new section have very similar drag coefficients that are between 0.492
drag coefficients that can be used in empiric drag estimation pro- and 0.524 depending on the strut size. This is an interesting find-
cedures of turrets and landing gears, the drag coefficients have to ing, as the differently sized struts feature different edge radii of
be referred to a specific area of the given component. This ref- their rectangular cross-sections. This should provide some slightly
erence area depends on the type of component and its size. The different flow guidance and affect their drag coefficient. However,
reference areas chosen for this approach are specifically described the study shows that this influence is not significant here. The data
when needed. indicate that a drag coefficient on the order of 0.5 (referred to
the frontal area of the strut) can approximate the drag of non-
3.1. Main landing gear streamlined main landing gear struts of UAVs. The identified drag
coefficients for the rectangularly shaped struts of this study are
3.1.1. Strut drag slightly lower when compared to the values given by Çengel and
The main landing gear strut drag from both the wind-tunnel Cimbala [23] for rectangular rods with rounded edges.
experiments and the CFD computations is shown in Fig. 7 for all The drag of the struts with a streamlined cross-sectional shape
three streamlined and rectangular strut configurations. The drag is normalized to the strut maximum frontal area in the same way
coefficients, which are shown in Fig. 7a are all normalized to a as previously described for the rectangularly shaped struts. In em-
reference area of 1 m2 . The drag of all configurations increases pirical drag estimation procedures, the drag of wing-like surfaces
with increasing overall strut size. However, this drag increase is is usually estimated by applying a skin friction coefficient on the
much more significant for the configurations featuring a rectan- wetted area of the component. This assumes that the flow over the
gular cross-sectional profile than for the ones with a streamlined
component is mostly attached. The streamlined struts in this study
section. The drag of the largest main strut with a rectangular cross-
have a comparably thick airfoil and operate at Reynolds numbers
section is more than three times as high as for the smallest config-
below 300,000. Therefore, significant separation effects can occur.
uration. The drag of each streamlined strut configuration is lower
This prohibits an adequate use of the skin friction drag estima-
than its rectangularly shaped counterpart at approximately similar
tion method. The data shown in Fig. 7b indicates that the small
load-carrying capabilities. For the smallest configuration, the drag
and medium-sized struts have very similar drag coefficients around
savings due to streamlining the struts are on the order of 28%,
0.3. However, the largest streamlined landing gear strut shows a
while these savings significantly increase for the largest configura-
significantly reduced drag coefficient of about 0.19. These wind-
tion. The largest streamlined main gear strut has a drag that is 60%
lower than its rectangularly shaped counterpart. The drag savings tunnel findings are supported by the results of the CFD studies.
are on the order of what is also reported in Refs. [17,18]. Such abrupt drag reduction is a strong indication for a change in
Besides these results, Fig. 7a also shows the drag values found flow separation behavior, which is very sensitive in this Reynolds
in the corresponding CFD studies. All values align to the wind- number regime. Further CFD studies revealed that both the small-
tunnel data with maximum deviations of 9% of strut drag. This (X-S-A) and medium-sized (X-M-A) streamlined strut suffer early
accuracy is within the data accuracy of the wind-tunnel experi- flow separation at about 50% of their chord length. The flow sepa-
ments. ration is clearly visible in the wall shear stress, and surface stream-
In order to derive drag estimation recommendations from this line contours on the outer part of the struts in Fig. 8. In contrast,
data, the strut drag found from the wind-tunnel experiments and the large strut configuration (X-L-A) shows mostly attached flow,
the CFD studies is normalized to the struts’ maximum frontal area. which reduces its drag. Such flow separation behavior is usually
This is a common approach for describing the drag coefficient of challenging to capture in numerical simulations. The good compa-
objects like cylinders, spheres, or other bluff-body shapes [18]. The rability between experimental and numerical drag coefficient, in
frontal area takes only the strut itself into account and not the at- combination with the separation behavior, shows that the chosen
tachment mechanism to the fuselage, as shown in Fig. 3a-f. The numerical approach can handle such cases.

8
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 8. Wall shear stress of main landing gear strut from CFD analysis (sizes are to scale). (For interpretation of the colors in the figure(s), the reader is referred to the web
version of this article.)

As the streamlined struts of this study are half-scaled, Reynolds lated case, this strut has low drag, as the flow stays attached over
numbers found for real-world applications are double as high con- most of its surface. However, the wheels now cause significant flow
sidering the same freestream velocity. Higher Reynolds numbers disturbances that lead to much earlier separation, especially on the
decrease flow separation tendencies and reduce strut drag. This outer parts of the strut. This effect is not as evident for all other
indicates that the largest strut configuration (X-L-A) is the most strut configurations as they feature early flow separation.
representative one for actual UAVs. The aerodynamic behavior of Wheel drag is usually estimated based on a drag coefficient
such a streamlined strut is, of course, influenced by the chosen referred to the wheel maximum frontal area. Typical values for
cross-sectional shape or airfoil. One driving factor of the parasitic isolated wheels found in the literature range from 0.18-0.23 [15,
drag of streamlined sections is its thickness-to-chord ratio. Higher 17,35]. CFD analyses of isolated wheels used in this study con-
thickness-to-chord ratios generally increase airfoil and therefore firm these drag coefficients. However, the drag coefficients of the
strut drag, while lower ratios decrease parasitic drag. In such, the wheels attached to the struts of this study are much higher. They
results of this study might only be seen as exemplary for this reach up to 0.71, which is more than three times as high as their
unique strut. However, it was shown that the chosen drag coeffi- isolated drag values. This is caused by strong interference effects
cient normalization approach is reasonable, and one might assume between wheels and struts that cause flow disturbances, pressure
that the determined coefficients can be extrapolated to other con- changes, and early separation, especially on the inner side of the
figurations. It is, therefore, recommended to empirically estimate wheels. The behavior is shown in the contour plot of the pres-
streamlined main gear strut drag of UAVs with a drag coefficient sure coefficient, and surface streamlines of the wall shear stress in
of about 0.2 for larger struts and 0.3 for smaller ones. These drag Fig. 9c. The flow on the outer part of the wheel is smooth, while
coefficients are also in the range of data given by Çengel and Cim- it is significantly affected by the presence of the strut on its in-
bala [23] for elliptically streamlined rods. ner side. Wheel drag might be significantly reduced by fairing the
wheel with a streamlined cover. This is a common technique of-
3.1.2. Wheel and complete main landing gear drag ten employed by general aviation aircraft. Drag reductions of more
The three differently sized wheel pairs (see Fig. 3h) were at- than 40% of isolated wheels have been reported [18]. Wheel fair-
tached to their corresponding strut configuration (Small – Medium ings were not investigated in the current study, and it is unclear
– Large) and the complete configurations analyzed in the wind- how they perform given the significant strut-wheel interference
tunnel and with the numerical approach. The drag results from effects found in this investigation. Further research is required
the wind-tunnel analysis of these full main gear configurations to analyze the effect of wheel fairings for typical UAV landing
are shown in Fig. 9a-b, together with the results of the iso- gears.
lated strut cases for comparison. The nomenclature to identify the To provide guidelines for empirically estimating the total main
components in Fig. 9 follows the description provided in Fig. 3. landing gear drag of UAVs, Table 1 gives an overview of the drag
Wheel drag significantly influences the total main landing gear coefficients found in the current study. The drag coefficients are
drag with an increasing tendency towards the larger configura- normalized with the frontal area of the specific components. The
tions. This is influenced by choice of the wheels’ sizes based on isolated struts’ drag coefficients should be increased to account
the in-house database. The drag increase is even more severe, con- for wheel interference effects, whereby the majority of interfer-
sidering streamlined strut sections as their isolated struts have ence factors is between 1.07 and 1.15. For cases where mainly
smaller drag compared to their rectangularly shaped counterparts. attached flow is expected for the isolated strut, a higher interfer-
One could argue to analyze wheel drag by subtracting strut drag ence factor is appropriate. Isolated wheel drag can be accurately
from the total main landing gear drag. However, this does not ap- estimated with the guidelines provided in the literature. However,
propriately account for wheel-strut interference effects. The local large interference factors must be used to account for wheel-strut
surface integrations in the CFD simulations, however, allow to ac- interference effects that range from 2.1 up to 3.3. In this study,
curately separate wheel and strut drag. The analysis shows that the wheels are directly attached to the struts without any gap
the attachment of wheels increases the drag of most struts only to that would realistically be necessary to provide the wheels with
a small degree compared to their isolated drag values (Fig. 9a-b). rotational freedom. The CFD analysis showed that wheel-strut in-
This allows determining a strut interference factor. The strut drag terference effects reduce when there is a certain gap between strut
increase is more severe for the largest streamlined strut. In the iso- and wheel. The wheel interference factors presented in Table 1 are,

9
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 9. Main landing gear drag with struts and wheels.

Table 1 lar cross-sectional profile are below 5% for the small and medium
Drag coefficients and interference factors for main landing gear. configuration. The edge rounding of the rectangularly shaped struts
Main gear C D isolated strut CD Strut IFF Wheel IFF reduces their geometric differences to the circular-shaped cross-
size rectangular/ isolated rectangular/ rectangular/ sections, which might explain the similar drag values. Again, the
streamlined wheel streamlined streamlined
CFD simulations are close to the drag values found in the wind-
Small 0.524/0.304 0.213 1.14/1.15 2.58/3.25 tunnel and slightly underpredict strut drag in all cases. Neverthe-
Medium 0.509/0.307 0.234 1.07/1.07 2.57/2.10
less, the differences are below 8%, which is an acceptable accuracy
Large 0.493/0.187 0.215 1.15/1.51 3.30/2.93
for this study.
Front strut drag is normalized to the frontal area of the strut in
therefore, conservative. Interference effects further reduced when the same way as described for the main gear struts. The drag co-
the gap between strut and wheel was increased. Deriving a rela- efficients are presented in Fig. 10b. The normalization now shows
tionship between gap size and interference factor proved difficult that the rectangularly shaped struts have slightly larger drag co-
and dependent on the strut and wheel combination. A significant efficients than compared to their circular counterparts. The frontal
amount of further investigations would be necessary but are be- area normalization also achieves that both small and medium con-
yond the scope of this study. figurations have very comparable drag coefficients, which is nec-
essary to derive empirical drag estimation relationships. All drag
3.2. Front landing gear coefficients determined with this technique are between 0.801 and
0.923. It is, therefore, recommended to apply an averaged drag
3.2.1. Strut drag coefficient of 0.86 for empirically estimating front strut drag of
The front landing gear strut drag from both the wind-tunnel UAVs.
experiments and the CFD simulations is shown in Fig. 10. The
drag coefficients presented in Fig. 10a are normalized with a ref- 3.2.2. Wheel drag and complete front landing gear drag
erence area of 1 m2 , which allows a direct drag force comparison. The front landing gear drag measurements, including wheels,
The differences between the struts with a rectangular and circu- are shown in Fig. 11, together with the values of the isolated struts

10
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 10. Front landing gear strut drag.

Fig. 11. Front landing gear drag with struts and wheels.

as a comparison (nomenclature see Fig. 5). Wheel drag significantly Table 2


affects overall front landing gear drag and increases it by over 30% Drag coefficients and interference factors for front landing gear.
compared to the isolated strut cases. The total front gear drag is Main gear C D isolated strut CD Strut IFF Wheel IFF
divided into the strut’s and wheel’s contribution in the CFD sim- size rectangular/ isolated rectangular/ rectangular/
ulations and presented in Fig. 11. All configurations show only a circular wheel circular circular

slight increase in strut drag due to strut-wheel interference, similar Small 0.923/0.846 0.213 1.06/1.04 3.33/3.44
Medium 0.874/0.835 0.224 1.01/1.04 3.34/3.11
to the case of the main landing gear. Again, wheel-strut interfer-
ence significantly increases wheel drag compared to isolated wheel
cases. The presence of the strut greatly affects the flow around the 3.3. Turret
wheel. The detected effects are very similar to the ones described
for the main wheels and not repeated here. Both Turrets 1 and 2 are simplified turret models and geometri-
The drag coefficients identified for the front landing gear based cally identical except for their different diameters. This affects the
on the wind-tunnel and CFD studies are presented in Table 2. All relationship between turret diameter and fuselage diameter. While
drag coefficients are normalized with the respective body frontal the diameter of Turret 1 is 62.5% of the fuselage width, this value
area. The investigated struts have very similar drag coefficients increases to 83.3% for Turret 2. Both turrets were tested with three
that only vary by about 10% for all cases. The struts’ drag coef- different height-to-diameter ratios, and their drag coefficients are
ficients might be increased with small interference factors due to presented in Fig. 12. These turrets do not represent the shape of
wheel-strut interference effects. The wheel interference factors for actual EO/IR turrets of UAVs, and their significance in terms of tur-
all front strut and wheel combinations are very close and similar ret drag prediction is, therefore, limited. However, the analyses of
to the ones identified for the main wheels. these turrets allow a further validation of the wind-tunnel data of

11
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

clude dynamic rotational effects. The drag of Turret 3 dependent


on its yaw orientation is presented in Fig. 13a. The turret yaw an-
gles are visualized in Fig. 13a, whereby the reader is looking with
the flow direction. The turrets sensors are pointing forward and in
flight direction for turret yaw angle 0◦ . Drag results from Sluder et
al. [26] for a clean turret model with a height-to-diameter ratio of
one are plotted as a comparison.
The results show a significant drag dependency on the turret
yaw orientation. The drag is highest when the camera sensors are
facing upstream (0◦ yaw angle) and reduces up to a yaw angle of
90◦ , where the sensors are facing sideways. This configuration re-
duces the drag of Turret 3 by about 28% compared to the 0◦ yaw
angle. A further increase in the yaw angle then again increases its
drag up to 180◦ yaw angle and the sensors facing downstream.
Fig. 12. Drag of turrets 1 and 2 with data from Sluder et al. [26]. This position, however, still shows a 12% reduction in turret drag
compared to the 0◦ yaw angle. The drag of Turret 3, which repre-
this study, as comparable data is available from Sluder et al. [26]. sents an EO/IR turret, is significantly larger than compared to the
They investigated clean turrets with a variety of height-to-diameter clean turret model from Sluder et al. [26]. Its drag coefficient is
ratios at similar Reynolds numbers. The results of both this wind- more than 60% higher compared to the clean counterpart at the
tunnel study and the one from Sluder et al. are close and indicate same height-to-diameter ratio and yaw angle 0◦ . At the optimal
an increase in turret drag coefficient with respect to their height- yaw orientation of 90◦ the drag of Turret 3 is still about 12% higher
to-diameter ratio. This study finds slightly smaller drag coefficients than the clean turret’s one.
than Sluder et al., which might be a result of the different test set- The findings are supported by numerical analyses that show
tings. Sluder et al. placed their turrets on a large flat plate, while significant flow distortion and separation on the flat sensor sur-
this study used a dummy fuselage which width was only slightly faces in the 0◦ turret yaw position. This early separation leads to a
larger than the turrets’ diameters. However, the wind-tunnel data comparably high turret drag in this case. The analyses also indicate
of this study also indicates an increased uncertainty, as the ab- that a significant part of the high drag in the 0◦ /180◦ position is
solute drag forces of Turret 1 and Turret 2 are comparably small. a reason of the geometrical connection (closed gap) between the
Remarkably, both this wind-tunnel study and the one from Sluder spherical and cylindrical parts of the turret. Strong flow recircu-
et al. revealed a significant increase in drag coefficient based on lation and intense flow separation are observed in this area for
the turret frontal area for a comparably small increase in height- both the 0◦ and 180◦ yaw position. These effects further increase
to-diameter ratio. The turrets with the largest height-to-diameter turret drag. The turret side that faces the inflow in the 90◦ yaw
ratio of this study (0.7) show a more than 40% increased drag co- position (sensors facing sideways) gives a more beneficial behav-
efficient compared to the hemispherical ones (h/d = 0.5). Sluder ior. Here, the spherical part of the turret smoothly joins into the
et al. further show that this drag coefficient increase reaches a cylindrical one. This reduces flow disturbances and delays separa-
plateau for clean turrets with a height-to-diameter ratio of 1.0. A tion compared to the 0◦ /180◦ yaw orientation. Consequently, this
further increase in the height-to-diameter ratio up to 2.0 does not yaw orientation reduces the drag of the turret.
change the drag coefficient to a noteworthy degree. The major aerodynamic effect of the turret yaw orientation is
Turret 3 represents an EO/IR sensor turret, and its drag was further visible in the pressure coefficient distribution on its center-
measured multiple times, while the turret was rotated around its line in Fig. 13b. The pressure coefficients were analyzed with the
yaw axis in nine steps. This is representative of mission scenarios, numerical analysis and present instantaneous values. The pressure
in which a turret performs yaw rotations to adjust its field-of-view. is determined on the centerline of the spherical part, indicated
The measurements were taken consecutively and, therefore, ex- with green lines around the turrets in Fig. 13b. The pressure on

Fig. 13. Drag and surface pressure of Turret 3 with data from Sluder et al. [26].

12
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 14. Landing gear and turret combination tests (wind-tunnel results).

the centerline of the front half (0◦ < θ < 90◦ ) of turret yaw angle ret. Combining all available front- and main landing gear config-
0◦ is greatly affected by the sensor surfaces. However, it generally urations would result in an excessive number of tests. Therefore,
decreases over the half-sphere, which coincides with a flow accel- rectangularly shaped front struts were only tested together with
eration over the front surface. The flow separates early and stays rectangularly shaped main struts. Circular front struts were only
detached. The pressure on the front half centerline (0 < θ < 90◦ ) tested together with streamlined main gear struts. Furthermore,
behaves comparably to a sphere for yaw orientations 90◦ and 180◦ . small front gears were only combined with small main gears,
Similar results were also reported in other investigations with sim- while medium front gears were tested together with medium and
plified turrets (see [27,30]). The 180◦ yaw angle shows a higher large main gears. The drag results of these tests are presented in
flow acceleration and slightly later flow separation compared to Fig. 14a. The solid bars show the drag measured of the complete
the 90◦ case, while the pressure in the separated aft region is configuration (front and main gear) while the dotted and dashed
comparable. This would usually indicate a reduction in drag in bars show the isolated values from previous tests as a compari-
analogy to the flow behavior of a sphere. However, this advantage son. The summation of the individual component drag matches the
is counteracted by increased drag caused by the geometric con- measurement of the complete gear configuration rather accurately.
nection between the spherical and cylindrical parts, facing towards Differences are in the order of 3%, which is within the given mea-
the flow in the 180◦ position (see above). surement accuracy. In such, no interference effects between front
The results indicate that the drag behavior of EO/IR sensor tur- and main gear could be observed.
rets is complex and can depend on the orientation towards the In a second step, Turret 3 was added to the model and tested
flow. The specific geometrical features of such turrets can affect together with the medium-sized, front, and main gears. The tur-
their aerodynamic behavior to a significant degree. Therefore, it ret’s sensors are facing upstream (yaw angle 0◦ ). Such a configura-
might not be easy to derive universally valid empirical drag esti- tion is shown in Fig. 1c. The drag results of the complete config-
mation recommendations for general EO/IR sensor turrets. Never- uration are presented in Fig. 14b as solid bars. Again, dotted and
theless, this study shows that the drag of EO/IR turrets is much dashed bars show the individual components’ drag as a compari-
higher compared to simplified and clean turret geometries. A drag son. The complete configuration consisting of front gear, main gear,
estimation using available data for clean turrets might significantly and turret produces about 20% less drag than the summation of
underestimate the drag of actual EO/IR turrets. The authors, there- their components. This is consistent for both the rectangular and
fore, recommend increasing the drag coefficients that are found in streamlined/circular gear configurations. The turret is situated be-
the literature for clean turrets. This study shows that drag coeffi- hind the front landing gear, which generates a significant wake.
cients on the order of 0.55-0.78 are appropriate depending on the The wake structure is visualized with the magnitude of the vortic-
expected flow disturbances caused by the turret geometry. Further ity as a volume rendering in Fig. 15a. The pressure is significantly
investigations are necessary to confirm this range for other turret reduced inside the wake, as shown in Fig. 15b, on the symmetry
geometries. plane of the model. The reduced pressure in front of the turret re-
duces its drag considerably. The CFD analysis further shows that
3.4. Landing gear and turret interference effects both front, and main landing gear drag are not affected by the
presence of the turret. The drag coefficient of Turret 3 is reduced
The previously presented results investigated front landing gear, to 0.48 compared to 0.78 (normalized to the frontal area) in its
main landing gear, and turret on an isolated basis. However, in real isolated configuration. This is almost a 40% drag reduction of the
applications, these would all be attached to a UAV’s fuselage si- turret and a significant result.
multaneously. This makes these components prone to interference Wake effects are generally complex and highly dependent on
effects. To analyze these effects, several wind-tunnel tests were the upstream geometry and the distance between the two bod-
performed with a variety of simultaneously attached components. ies. The turret drag reduction results found in this study, therefore,
The results are compared to the components’ isolated behaviors may not be universally applicable. The relative positioning between
and further studied with CFD simulations. First, only landing gear the turret and front landing gear depends on multiple factors like
components were tested together, without the presence of a tur- the UAV configuration, the center of gravity, or field-of-view con-

13
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Fig. 15. Contours of pressure coefficient on the symmetry plane of the model and volume rendering of vorticity magnitude (instantaneous values).

siderations. Deriving simple semi-empirical relationships that in- combinations due to flow disturbances caused by protruding sen-
corporate these effects is not possible due to the large variety of sor surfaces and the rotational mechanism. These geometrical fea-
contributing factors. Hoerner [18] recommends reducing the drag tures promote early separation and have the potential of increasing
coefficient of bluff bodies that are located in wakes of upstream turret drag by more than 60% compared to clean hemisphere-
components by a ratio of the local dynamic pressure in front of cylinder combinations of the same height-to-diameter ratio. EO/IR
the body and the freestream dynamic pressure. The local dynamic turret drag further depends on the turret orientation towards the
pressure, however, is usually unknown and might not be easily es- flow and might be significantly reduced when protruding sensors
timated in the design stages that target empirical drag estimation. are not facing towards the flow. Placing a turret in the wake of the
In this sense, it is recommended to estimate EO/IR turret drag as front landing gear significantly reduced the turret drag caused by
an isolated component as this yields conservative estimates. a reduction of dynamic pressure. While drag savings are possible,
optical effects might be worsened due to the increased turbulence
4. Conclusion intensity, possibly increasing vibrations of the turret. Further re-
search is required to investigate the drag of other EO/IR turret
The paper presents the results of a combined wind-tunnel and geometries, as drag can be significantly affected by individual sen-
numerical approach to analyze the drag characteristics of typical sor arrangements and the rotational mechanism.
UAV landing gears and sensor turrets. Both landing gear and turret The findings of this study cumulated in drag estimation rec-
geometries are based on statistical evaluations and are represen- ommendations for both front, and main landing gear struts and
tative of a wide variety of actual configurations used on fixed- wheels as well as EO/IR turrets. An application of these recom-
wing unmanned aircraft. Isolated main landing gear strut drag was mendations in early conceptual UAV design has the potential to
found to be a major contributor to overall landing gear drag. The enhance the accuracy of empirical drag estimations.
analysis showed that great drag savings are possible by streamlin-
ing main gear struts. At similar load carrying capabilities, stream- Funding
lined main gear struts have the potential of reducing strut drag
by over 50%. The drag of the isolated main gear struts behaved Funding for this work was provided by the Department of
proportionally to the struts’ frontal areas, which allowed deriving Aerospace Engineering of FH Aachen University of Applied Sciences
simple empirical drag estimation recommendations. Wheels play a and the College of Science Engineering and Health of the Royal
very important part in overall main landing gear drag, as their at- Melbourne Institute of Technology.
tachment introduces significant interference effects. These interfer-
ence effects lead to a minor increase in strut drag while installed Declaration of competing interest
wheel drag can more than triple compared to isolated cases. This
is a reason of flow disturbances around the wheel caused by the The authors declare that they have no known competing finan-
presence of the struts. These disturbances affect the pressure dis- cial interests or personal relationships that could have appeared to
tribution around the wheels and support early flow separation. influence the work reported in this paper.
Further research might investigate the effect of wheel fairings.
Front landing gear struts were designed as oleo-pneumatic Acknowledgement
damping struts with a circular and rectangular cross-section. The
drag characteristics were largely independent of the cross-sectional The authors would like to express their gratitude to Siemens
shape as the rectangular cross-section had well-rounded edges. PLM Software for providing academic licenses of their software
The front strut drag was found to be proportional to the frontal Star-CCM+.
area of the strut, and simple drag estimation recommendations
were developed. As for the main gear, wheel-front-strut interfer- References
ence effects significantly increased wheel drag compared to iso-
lated cases. Combined tests of front and main landing gears did [1] F. Götten, M. Havermann, C. Braun, F. Gómez, C. Bil, On the applicability of
not reveal any interference effects, and their combined drag was empirical drag estimation methods for unmanned air vehicle design, in: 18th
AIAA Aviation Technology Integration and Operations Conference, AIAA, Reston,
merely equal to the total installed landing gear drag. VA, 2018.
The drag of an isolated EO/IR sensor turret was found to be [2] P.D. Bravo-Mosquera, H.D. Cerón-Muñoz, G. Díaz-Vázquez, F. Martini Catalano,
much larger than the drag of comparable hemisphere-cylinder Conceptual design and CFD analysis of a new prototype of agricultural aircraft,

14
F. Götten, M. Havermann, C. Braun et al. Aerospace Science and Technology 107 (2020) 106306

Aerosp. Sci. Technol. 80 (2018) 156–176, https://doi.org/10.1016/j.ast.2018.07. [24] C.H. Snyder, M.E. Franke, M.L. Masquelier, Wind-tunnel tests of an aircraft tur-
014. ret model, J. Aircr. 37 (3) (2000) 368–376, https://doi.org/10.2514/2.2625.
[3] F. Götten, D.F. Finger, M. Havermann, C. Braun, F. Gómez, C. Bil, On the flight [25] S. Gordeyev, M.L. Post, T. McLaughlin, J. Ceniceros, E.J. Jumper, Aero-optical
performance impact of landing gear drag reduction methods for unmanned air environment around a conformal-window turret, AIAA J. 45 (7) (2007)
vehicles, in: German Aerospace Congress 2018, Bonn, 2018. 1514–1524, https://doi.org/10.2514/1.26380.
[4] B. Pedro, C. Elsa, A. Andrea, Drag clean-up process of the unmanned air- [26] R. Sluder, L. Gris, J. Katz, Aerodynamics of a generic optical turret, J. Aircr.
plane for ecological conservation, Aerotec. Missili Spaz. 85 (2) (2016) 53–62, 45 (5) (2008) 1814–1815, https://doi.org/10.2514/1.36804.
https://doi.org/10.19249/ams.v85i2.233, http://www.aerotecnica.eu/index.php/ [27] R. Jelic, S. Sherer, R. Greendyke, Simulation of various turrets at subsonic and
AMS/article/view/233 [online]. transonic flight conditions using OVERFLOW, J. Aircr. 50 (2) (2013) 398–409,
[5] P. Panagiotou, K. Yakinthos, Aerodynamic efficiency and performance enhance- https://doi.org/10.2514/1.C031844.
ment of fixed-wing UAVs, Aerosp. Sci. Technol. (2019) 105575, https://doi.org/ [28] F. Götten, M. Havermann, C. Braun, F. Gómez, C. Bil, RANS simulation validation
10.1016/j.ast.2019.105575. of a small sensor turret for UAVs, J. Aerosp. Eng. 32 (5) (2019) 4019060, https://
[6] V. Ahuja, R. Hartfield, Optimization of UAV designs for aerodynamic perfor- doi.org/10.1061/(ASCE)AS.1943-5525.0001055.
mance using genetic algorithms, in: 51st AIAA/ASME/ASCE/AHS/ASC Structures, [29] E.R. Mathews, K. Wang, M. Wang, E.J. Jumper, LES of an aero-optical turret flow
Structural Dynamics, and Materials Conference, Orlando, Florida, 2010. at high Reynolds number, in: 54th AIAA Aerospace Sciences Meeting, American
[7] S.G. Kontogiannis, J.A. Ekaterinaris, Design, performance evaluation and op- Institute of Aeronautics and Astronautics, 2016.
timization of a UAV, Aerosp. Sci. Technol. 29 (1) (2013) 339–350, https:// [30] S. Gordeyev, E. Jumper, Fluid dynamics and aero-optics of turrets, Prog. Aerosp.
doi.org/10.1016/j.ast.2013.04.005. Sci. 46 (8) (2010) 388–400, https://doi.org/10.1016/j.paerosci.2010.06.001.
[8] P. Panagiotou, P. Kaparos, K. Yakinthos, Winglet design and optimization for a [31] F. Götten, D.F. Finger, C. Braun, M. Havermann, C. Bil, F. Gómez, Empirical corre-
MALE UAV using CFD, Aerosp. Sci. Technol. 39 (2014) 190–205, https://doi.org/ lations for geometry build-up of fixed wing unmanned air vehicles, in: APISAT
10.1016/j.ast.2014.09.006. 2018 Proceedings, in: Lecture Notes in Electrical Engineering, Springer, Singa-
[9] Z. Xiao, J. Liu, K. Luo, J. Huang, S. Fu, Investigation of flows around a rudimen- pore, 2019, pp. 1365–1381.
tary landing gear with advanced detached-eddy-simulation approaches, AIAA J. [32] M. Marino, Unsteady Pressure Sensing on a MAV Wing for Control Inputs in
51 (1) (2013) 107–125, https://doi.org/10.2514/1.J051598. Turbulence, RMIT University, Melbourne, 2013.
[10] S. Spagnolo, X. Zhang, Z. Hu, D. Angland, Aerodynamic interactions between [33] E.G. Cruz, The Effect of Turbulence on Micro Air Vehicle Airfoils, RMIT, Mel-
landing-gear components, in: 34th AIAA Applied Aerodynamics Conference, bourne, 2012.
American Institute of Aeronautics and Astronautics, 2016. [34] R.M. Pagliarella, On the Aerodynamic Performance of Automotive Vehicle Pla-
[11] P. Spalart, K. Mejia, Analysis of experimental and numerical studies of the rudi- toons Featuring Pre and Post-Critical Leading Forms, RMIT University, Mel-
mentary landing gear, in: 49th AIAA Aerospace Sciences Meeting including the bourne, 2009.
New Horizons Forum and Aerospace Exposition, American Institute of Aero- [35] D.P. Raymer, Aircraft Design. A Conceptual Approach, 6th edn., AIAA Education
nautics and Astronautics, 2011. Series, Reston, VA, ISBN 978-1-62410-490-9, 2019.
[12] Y. Li, M. Smith, X. Zhang, Measurement and control of aircraft landing gear [36] J.B. Barlow, W.H. Rae, A. Pope, Low-Speed Wind Tunnel Testing, 3rd edn., Wiley,
broadband noise, Aerosp. Sci. Technol. 23 (1) (2012) 213–223, https://doi.org/ New York, NY, ISBN 0-471-55774-9, 1999.
10.1016/j.ast.2011.07.009. [37] CD-adapco, Validation of star-CCM+ for external aerodynamics in the aero-
[13] J. Gundlach, Designing Unmanned Aircraft Systems: A Comprehensive Ap- space industry, https://mdx2.plm.automation.siemens.com/sites/default/files/
proach, 2nd edn., AIAA Education Series, Reston, VA, ISBN 978-1-62410-261-5, Presentation/CD-adapco_AeroValidation_v7.pdf-aerospace-industry, retrieved
2014. 22 January 2020.
[14] R. Austin, Unmanned Aircraft Systems. UAVS Design, Development and Deploy- [38] C. Hirsch, Computational Methods for Inviscid and Viscous Flows, Wiley, Chich-
ment, 2nd edn., Wiley, Chichester, Hoboken, ISBN 978-0-470-05819-0, 2010. ester, ISBN 0471924520, 2002.
[15] S. Gudmundsson, General Aviation Aircraft Design: Applied Methods and Pro- [39] J.H. Lienhard, Synopsis of Lift, Drag, and Vortex Frequency Data for Rigid Circu-
cedures, Butterworth-Heinemann, Oxford, ISBN 1-299-84718-8, 2014. lar Cylinders: Bulletin 300, Washington State University, Pullman, Washington,
[16] J. Roskam, Methods for Estimating Drag Polars of Subsonic Airplanes, University 1966.
of Kansas, 1973. [40] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering
[17] E. Torenbeek, Synthesis of Subsonic Airplane Design. An Introduction to applications, AIAA J. 32 (8) (1994) 1598–1605, https://doi.org/10.2514/3.12149.
the Preliminary Design of Subsonic General Aviation and Transport Aircraft, [41] A.K. Hellsten, New advanced k-w turbulence model for high-lift aerodynamics,
with Emphasis on Layout, Aerodynamic Design, Propulsion and Performance, AIAA J. 43 (9) (2005) 1857–1869, https://doi.org/10.2514/1.13754.
Springer Netherlands, Dordrecht, ISBN 978-94-017-3202-4, 2010. [42] A.F. Antoniadis, D. Drikakis, B. Zhong, G. Barakos, R. Steijl, M. Biava, L. Vigevano,
[18] S.F. Hoerner, Fluid-Dynamic Drag. Practical Information on Aerodynamic Drag A. Brocklehurst, O. Boelens, M. Dietz, M. Embacher, W. Khier, Assessment of
and Hydrodynamic Resistance, published by the author, Bakersfield CA, CFD methods against experimental flow measurements for helicopter flows,
ISBN 978-9991194448, 1965. Aerosp. Sci. Technol. 19 (1) (2012) 86–100, https://doi.org/10.1016/j.ast.2011.
[19] W. Herrenstein, D. Biermann, The Drag of Airplane Wheels, Wheel Fairings and 09.003.
Landing Gears I: NACA TR-485, 1934. [43] P.R. Spalart, Strategies for turbulence modeling and simulations, Int. J. Heat
[20] D. Biermann, W. Herrenstein, The Drag of Airplane Wheels, Wheel Fairings and Fluid Flow 21 (3) (2000) 252–263, https://doi.org/10.1016/S0142-727X(00)
Landing Gears II: NACA TR-518, 1936. 00007-2.
[21] W. Herrenstein, D. Biermann, The Drag of Airplane Wheels, Wheel Fairings, and [44] P.R. Spalart, C.L. Rumsey, Effective inflow conditions for turbulence models in
Landing Gears - III: NACA TR-522, NACA, 1936. aerodynamic calculations, AIAA J. 45 (10) (2007) 2544–2553, https://doi.org/10.
[22] H.N. Harmon, Drag Determination of the Forward Component of a Tricycle 2514/1.29373.
Landing Gear: NACA TN-788, NACA, Langley Field, VA, 1940.
[23] Y.A. Çengel, J.M. Cimbala, Fluid Mechanics. Fundamentals and Applications,
McGraw-Hill Education, New York, NY, ISBN 978-1-259-69653-4, 2017.

15

You might also like