You are on page 1of 23

aerospace

Article
Effects of Stroke Amplitude and Wing Planform on the
Aerodynamic Performance of Hovering Flapping Wings †
Hao Li 1 and Mostafa R. A. Nabawy 1,2, *

1 Department of Mechanical, Aerospace and Civil Engineering, The University of Manchester,


Manchester M13 9PL, UK
2 Aerospace Engineering Department, Faculty of Engineering, Cairo University, Giza 12613, Egypt
* Correspondence: mostafa.ahmednabawy@manchester.ac.uk
† This paper is an extended version of our paper published in the Proceedings of the AIAA SciTech 2022 Forum,
San Diego, CA & Virtual, USA, 3–7 January 2022. https://doi.org/10.2514/6.2022-1056.

Abstract: In this paper, the effects of stroke amplitude and wing planform on the aerodynamics of
hovering flapping wings are considered by numerically solving the incompressible Navier–Stokes
equations. The wing planform geometry is represented using a beta-function distribution for an
aspect ratio range of 3–6 and a dimensionless radial centroid location range of 0.4–0.6. Typical normal
hovering kinematics has been employed while allowing both translational and rotational durations to
be equally represented. The combined effects of stroke amplitude with wing aspect ratio and radial
centroid location on the aerodynamic force coefficients and flow structures are studied at a Reynolds
number of 100. It is shown that increasing the stroke amplitude increases the translational lift for
either small aspect ratio or large radial centroid location wings. However, for high aspect ratio or
low radial centroid location wings, increasing the stroke amplitude leads to higher lift coefficients
during the translational phase only up to a stroke amplitude of 160◦ . Further increase in stroke
amplitude results in reduced translational lift due to the increased wingtip stall effect. For all the
Citation: Li, H.; Nabawy, M.R.A.
cases considered, the lift and drag coefficients of the rotational phase decrease with the increase of
Effects of Stroke Amplitude and
stroke amplitude leading to decreased cycle-averaged force coefficients. Furthermore, it is found that
Wing Planform on the Aerodynamic
the significant reduction in the rotational drag as the stroke amplitude increases leads to a consistently
Performance of Hovering Flapping
increasing aerodynamic efficiency against stroke amplitude for all aspect ratio and radial centroid
Wings. Aerospace 2022, 9, 479.
https://doi.org/10.3390/
location cases.
aerospace9090479
Keywords: flapping wings; aerodynamics; wing planform; stroke amplitude; leading-edge vortex
Academic Editors: Lawrence
S. Ukeiley and Bosko Rasuo

Received: 3 July 2022


Accepted: 24 August 2022 1. Introduction
Published: 29 August 2022 Insect flight has received significant research attention in the past few decades due to
Publisher’s Note: MDPI stays neutral the unique flight dynamics involved and has inspired numerous developments of small
with regard to jurisdictional claims in flying robots [1–6]. To exploit the benefits of insect-like flight, it is essential to develop
published maps and institutional affil- adequate understanding of the involved aerodynamics; and thanks to the many and
iations. continuous efforts from the different research communities, our current perception of
the underlying aerodynamic mechanisms of insect wings has improved significantly. To
better classify the different aerodynamic mechanisms, it is instructive to recognise them in
relation to when they take place within an insect-like stroke motion. Such stroke typically
Copyright: © 2022 by the authors. consists of two translational phases (within the forward and backward half-strokes), as
Licensee MDPI, Basel, Switzerland. well as two rotational phases, where the wing flips its motion direction (at the start/end
This article is an open access article of each half-stroke) [7]. In the translational phases, the wing would typically travel at
distributed under the terms and
a relatively constant but high angle of attack (AoA), where a leading-edge vortex (LEV)
conditions of the Creative Commons
resulting from flow separation at the leading-edge is formed on the upper surface of the
Attribution (CC BY) license (https://
wing and persists throughout the translational motion. This persistence of the LEV allows
creativecommons.org/licenses/by/
flow reattachment for insect wings operating at high AoA, where classical wings would
4.0/).

Aerospace 2022, 9, 479. https://doi.org/10.3390/aerospace9090479 https://www.mdpi.com/journal/aerospace


Aerospace 2022, 9, 479 2 of 23

normally stall, thus augmenting lift production [8–12]. In the rotational phases, the wing
rotates and flips the upper and lower surfaces in preparation for the next half-stroke.
Several unsteady aerodynamic mechanisms take place during the wing rotational phases,
including rotational circulation, added mass, and wake capture, which all result in different
forms of force variations that can be significant for certain insect wing motions [10,13,14].
For insect wings, the planform shape has significant influence on the aerodynamic
performance [15]. A widely employed setup to study such influence is the revolving wing
setup. While revolving wing experimental/numerical setups are simple, they are still
valuable in representing the aerodynamic effects within the translational phase of an insect
flapping stroke. Usherwood and Ellington [16] were among the first to experimentally
study the aerodynamic force production of revolving hawkmoth wing models with aspect
ratios ranging between 2.27 and 7.92. They found minor influence of aspect ratio on
force coefficient values, but also observed that higher aspect ratios result in higher lift
curve slopes. Kruyt et al. [17] experimentally visualised the LEV structures on rectangular
revolving wings using Particle Image Velocimetry (PIV). They found that the LEV remains
attached up to a distance along the wingspan that is four times of the mean chord length.
Further outboard, the LEV is detached and lifted away. Hence, they used this observation
to explain the reduced aerodynamic efficiency for wings with aspect ratio greater than
four. Broadley et al. [18,19] experimentally measured the aerodynamic force production
of revolving insect-like wings at Reynolds number of 104 . The effect of wing planform
shape was investigated for a wing aspect ratio range between 2 and 4, and a radial centroid
location range between 40% and 60% of the wingspan. They found that the lift curve
slope decreases with the increase of the area centroid location. Moreover, they showed
that the drag coefficient increases with the decrease of aspect ratio. Harbig et al. [20]
used Computational Fluid Dynamics (CFD) simulations to study aspect ratio effects on a
revolving Drosophila shaped wing planform. They found that the lift coefficient is nearly
constant for aspect ratios within 2.9–5.1; however, when aspect ratio further increases to
7.3, the lift coefficient is decreased.
The planform effect has also been investigated using the complete flapping wing
setup. Clearly, such setup is more representative as it not only simulates the translational
phases of a flapping stroke, but also the rotational phases associated with stroke reversals.
Phillips et al. [21,22] experimentally measured the aerodynamic force production of flapping
rectangular wings with different aspect ratios. They found that for aspect ratios of three
and above, the LEV starts to detach from the wing at approximately 70% of the wingspan
near mid half-stroke. However, they also found that the LEV circulatory lift peaks at an
aspect ratio of six, but further increase of the aspect ratio results in a reduced lift coefficient
due to the increased effect of LEV detachment. Similarly, in the experimental study of
Han et al. [23], it was shown that for aspect ratios greater than three, lifting away of
the LEV results in reduced lift coefficients, hence lift was deemed to peak at an aspect
ratio of three. On the other hand, Luo and Sun [24] used CFD simulations to study the
force production of different wing planforms taken from 10 different insect species, with
aspect ratios ranging between 2.84 and 5.45. Their results showed only minor differences
for the aerodynamic force coefficients of the different wing planforms, hence the effect
of aspect ratio was deemed to be small. Shahzad et al. [25] also used CFD to study the
effect of wing planform on aerodynamic performance. Lift coefficient values were found
to increase with the increase of aspect ratio. Aerodynamic efficiency was also found
to increase with aspect ratio, but only within the aspect ratio range of 1.5–2.96, then it
started to decrease as aspect ratio was further increased. For more detailed discussions
on the effect of aspect ratio on insect wing aerodynamics, the reader is referred to the
review paper [26]. While the previous studies have considered flapping wing motions,
most of the focus was on translational and global aerodynamics. It is worth noting here
that wing rotation/acceleration during stroke reversals can have significant influence on
the aerodynamic performance; however, studies/results focusing on aerodynamic force
contributions during the rotational phases are limited.
Aerospace 2022, 9, 479 3 of 23

For insect-like flapping motions, the stroke amplitude is a fundamental parameter that
varies between the different species. Until now, limited studies have addressed the effect of
stroke amplitude on the aerodynamic performance of insect wings with sufficient details.
Sane and Dickinson [27] used a model Drosophila wing within a robotic experimental rig to
study the aerodynamic force production for different wing kinematics at a Reynolds number
of 100. In their experiments, the effect of varying the stroke amplitude was investigated
for a range of stroke amplitude between 60◦ and 180◦ . Their results showed that, for
the wing planform considered, lift is highest at a stroke amplitude of 180◦ and angle of
attack of 50◦ . Later, Phillips and Knowles [28] also used a robotic flapping apparatus to
study the effect of Reynolds number and stroke amplitude on the LEV structures of a
model Drosophila wing using PIV. However, they found that for all Reynolds numbers
investigated (Re = 3850–18,210), the LEV breaks down at a constant stroke amplitude of
approximately 132◦ , whereas for smaller stroke amplitudes, the breakdown of the LEV is
suppressed. This observation led them to propose that LEV breakdown depends more on
stroke amplitude, rather than Reynolds number. A couple of previous studies have noted
that stroke amplitude can play an important role in determining the degree of aerodynamic
unsteadiness during the rotational phases. Altshuler et al. [29] experimentally investigated
the effect of stroke amplitude on aerodynamic force production of hovering honeybees. By
employing the measured wing kinematics of honeybees while artificially modulating the
stroke amplitude between 80◦ and 180◦ , they showed that the transient force peaks during
the wing rotational phases decrease with the increase of stroke amplitude up till a stroke
amplitude of 135◦ . They concluded that for low stroke amplitudes, unsteady mechanisms
during the rotational phases become a much larger contributor to the net upward force
required for hovering. Lentink and Dickinson [30] experimentally studied the effect of the
dimensionless stroke amplitude, A∗ = ΦAR, on the LEV structure and force production
for Drosophila wing models. They considered a unidirectional wing motion with infinite A∗
and a reciprocal motion with A∗ = 3.5. They showed that the attachment state of LEV on
the wing was not influenced by A∗ ; however, the smaller dimensionless stroke amplitude
case contributed to lift augmentation by keeping the LEV close to the wing during the wing
rotational phases.
Clearly from the previous studies, for insect-like flapping wings, both the wing plan-
form shape and the stroke amplitude have direct influence on the LEV structure/topology
and aerodynamic force production. However, previous studies have primarily focused on
either the effect of wing planform for a given stroke amplitude or the effect of stroke ampli-
tude for a given wing planform, but their combined effect on aerodynamic performance has
never been investigated. As such, in this study, we aim to provide the most comprehensive
investigation, up to date, of the combined stroke amplitude/wing planform effects on
inset-like wing aerodynamics. Moreover, we provide detailed comparisons between the
relative contributions of both the translational and rotational phases of a flapping stroke
to aerodynamic force production, an aspect that is usually overlooked in the literature.
To enable our study aims, the wing planform shape was defined using the beta-function
distribution, with three aspect ratios (within the range 3–6) and three dimensionless radial
centroid locations (within the range 0.4–0.6) being investigated. The stroke peak-to-peak
amplitude in this study was chosen to vary between 60◦ –180◦ , and its effect was simultane-
ously investigated with the variation of the wing aspect ratio or radial centroid location.
The effects of all these parameters were considered at a Reynolds number of 100, most
relevant to small insects such as fruit flies.
Aerospace 2022, 9, 479 4 of 23

2. Methods
2.1. Wing Planform and Kinematics Definitions
The wing planform shape was defined using the beta-function distribution [31], where
the wing chord length, c(r̂ ), is expressed as:

r̂ m−1 (1 − r̂ )q−1
c(r̂ ) = R 1 .c (1)
Aerospace 2022, 9, x FOR PEER REVIEW
0 r̂ m−1 (1 − r̂ )q−1 dr̂ 4 of 24

where r̂ is the non-dimensional radial position (r̂ = 0–1 from wing root to tip) and the
parameters m and q are related to the non-dimensional radii of first and second moments
𝑟̂ 𝑚−1 (1 − 𝑟̂ )𝑞−1
of wing area, r̂1 and r̂2 , via: 𝑐(𝑟̂ ) = 1 𝑚−1" . 𝑐̅ (1)
∫0 𝑟̂ (1 − 𝑟̂ )𝑞−1 d𝑟̂ #
r̂ (1 − r̂1 )
where 𝑟̂ is the non-dimensional radial = r̂1 1(𝑟̂2 = 0–1
m position − 1 wing root to tip) and the
from (2)
r̂2 − r̂12
parameters 𝑚 and 𝑞 are related to the non-dimensional radii of first and second mo-
ments of wing area, 𝑟̂1 and 𝑟̂2 , via: " #
r̂1 (1 − r̂1 )
q = (1 − r̂1 )− 𝑟̂1 ) 2
𝑟̂1 (1 2
−1 (3)
𝑚 = 𝑟̂1 [ 2 2 −
r̂2 −
1] r̂1 (2)
𝑟̂2 − 𝑟̂1
Ellington [31] demonstrated that insect wing planforms, generally, follow the statistical
𝑟̂1 (1 − 𝑟̂1 )
relation: r̂2 = 0.929[r̂1 ]0.732 , hence
𝑞 = (1we
− 𝑟̂1used
) [ 2 this2 relation
− 1] to correlate r̂1 and r̂(3)
2.
𝑟̂ − 𝑟̂
The wing aspect ratio was defined as:2 1
Ellington [31] demonstrated that insect wing planforms, generally, follow the statistical
relation: 𝑟̂2 = 0.929[𝑟̂1 ]0.732 , hence we used this relation
R to correlate 𝑟̂1 and 𝑟̂2 .
The wing aspect ratio was defined as: AR = (4)
c
𝑅
where R is the length of a single wing 𝐴𝑅and= c is the mean geometric chord.(4)Insect wings
𝑐̅
typically have aspect ratios between two and five, however their aspect ratios are most
where 𝑅 is the length of a single wing and 𝑐̅ is the mean geometric chord. Insect wings
clustered between
typically have aspectthree
ratiosand four two
between [31].and
Asfive,
such, in the their
however current study,
aspect ratioswe
areconsidered
most aspect
ratio values
clustered of 3,three
between 4.5,and
and 6:[31].
four theAs first
such,two aspect
in the current ratios
study,were chosen aspect
we considered because of being
closely representative
ratio values of 3, 4.5, andof insect
6: the firstwings, whereas
two aspect the last
ratios were aspect
chosen ratioofrepresented
because being closely a maximum
value, mainlyoftoinsect
representative demonstrate how aerodynamics
wings, whereas the last aspect ratiochanges as theaaspect
represented maximumratio increases.
value,
The mainly to demonstrate
dimensionless how aerodynamics
radial centroid location (firstchanges
moment as theof
aspect ratio increases.
the wing area), r̂1 , was varied
The dimensionless
between the values radial
0.4,centroid
0.5, andlocation (first moment
0.6, which of the
cover most ofwing area), 𝑟̂shapes
planform 1 , was varied
known for insect
between the values 0.4, 0.5, and 0.6, which cover most of planform shapes known for in-
wings [31]. In all simulation cases, the pitching axis was set at 25% of the local chord and
sect wings [31]. In all simulation cases, the pitching axis was set at 25% of the local chord
the
andwing thickness
the wing thicknesshadhada constant
a constantvalue
value set as 5%
set as 5%ofofthe
themean
mean chord
chord length,
length, a representative
a repre-
value from
sentative thefrom
value reported studies
the reported in theinliterature
studies [14,32].
the literature [14,32].Figure
Figure 11 shows
showsthe the wing planform
wing
shapes
planformemployed in our simulations.
shapes employed in our simulations.

Figure 1. Wing planform shapes employed in the numerical investigation.


Figure 1. Wing planform shapes employed in the numerical investigation.
In the current study, the flapping wing is modelled as a rigid structure, hence the
effects of wing flexibility are not considered. For real flapping insects, wing flexibility as-
pects, e.g., bending and twisting deformations, contribute towards controlling the wing
Aerospace 2022, 9, 479 5 of 23

In the current study, the flapping wing is modelled as a rigid structure, hence the
effects of wing flexibility are not considered. For real flapping insects, wing flexibility
aspects, e.g., bending and twisting deformations, contribute towards controlling the wing
load distribution and are expected to influence the flow structure/topology. That said,
it should be noted that several previous studies in the literature have demonstrated that
rigid wings are capable of representing the aerodynamics of hovering flapping wings in an
acceptable fashion. For example, Zhao et al. [33] experimentally measured the aerodynamic
forces on a flapping wing with flexible trailing edge. They showed that decreasing the
flexural stiffness of the trailing edge leads to decreased aerodynamic force production, but
the lift-to-drag ratio of the wing remains approximately constant. Du and Sun [34] also
studied the effects of camber and twist deformations of a hovering hoverfly wing using
computational fluid dynamics simulations. They showed that the flexible wing version can
produce 10% higher lift compared to a rigid wing version, but the time courses of lift, drag,
and aerodynamic power coefficients for a flexible wing were similar to that of the rigid
wing counterpart.
It is worth noting here that there have been significant research efforts focusing
on the fluid–structure interaction (FSI) problem of flexible flapping aerofoils and wings.
These studies provided different insights into insect wing aerodynamics. Toomey and
Eldredge [35] and Eldredge et al. [36] numerically and experimentally studied the effect of
chordwise flexibility on the aerodynamic performance of 2D aerofoils using a simplified
model of rigid sections connected by a hinge with torsional spring. They showed that a
mildly flexible aerofoil has better aerodynamic performance compared to its rigid counter-
part, when compared over a wide range of kinematic variations [36]. Yin and Luo [37] used
FSI simulations to study 2D flexible flapping aerofoils and found that when the mass is
small (hence, deformation is mainly due to aerodynamic force), high aerodynamic power
efficiency is achieved. Kang and Shyy [38] conducted FSI simulations to study the effect
of passive wing rotation on flexible flapping wings, showing that camber deformation
can mitigate the negative wing–wake interaction effects and hence results in higher lift
production. Nakata and Liu [39] used FSI simulations to investigate the effect of wing
flexibility on the aerodynamics of hovering hawkmoths (Manduca Sexta). It was found that
while wing flexibility can lead to increased lift, the power cost is also increased. Hence,
in total, only a 3.4% increase in aerodynamic efficiency was found for the flexible wing
case when compared to the rigid wing case. Shahzad et al. [40] conducted FSI simulations
to study flexible hovering flapping wings of different aspect ratios and radial centroid
locations. They showed that wing flexibility, for wings with aspect ratios ranging between
4.5 and 6, decreases lift but increases power economy. Notably, Dai et al. [41] also used FSI
simulations to study the aerodynamic performance of hovering rectangular flapping wings
of an aspect ratio of two and different rigidities. They showed that when the frequency
ratio (defined as the ratio of flapping frequency to wing natural frequency) is less than or
equal to 0.3, both lift production and efficiency improve. This result was later supported by
the numerical studies of Wang and Tian [42] and Sridhar and Kang [43], where despite the
difference in adopted approaches, both studies showed that optimum efficiency is evident
at frequency ratios of 0.3 and 0.35, respectively. Finally, Kang et al. [44] studied the effects of
flexibility on the aerodynamic performance of flapping wings in forward flight. They found
that maximum propulsive force occurs while flapping near resonance, whereas optimal
propulsive efficiency occurs when flapping at a frequency around half of the wing natural
frequency.
The employed flapping velocity waveform had a sinusoidal variation during the
acceleration and deceleration phases within a half-stroke, and a constant velocity value in
between. Hence, the flapping velocity during stroke reversal was defined using [45]:
" #
.  . π t̂ − t̂sa
φ t̂ = ±φ M cos , t̂sa ≤ t̂ ≤ t̂sa + t̂ a (5)
t̂ a
FOR PEER REVIEW 6 of 24
Aerospace 2022, 9, 479 6 of 23

.
holding a constantwhere
angle-of-attack during the
φ M is the maximum wingvelocity;
flapping translational
t̂ is the phase. Hence, the
non-dimensional pitch-
time; t̂sa is the
ing angular velocity, during
starting timestroke reversal,
of wing wasand
acceleration; defined asnon-dimensional
t̂ a is the [45]: acceleration duration.
Similarly, the pitching waveform had a sinusoidal variation during stroke reversal while
holding Θ
a constant 2𝜋(𝑡̂ − 𝑡during
angle-of-attack ̂𝑠𝑝 ) the wing translational phase. Hence, the pitching
̇ (𝑡̂) = ±
𝜃angular {1 − cos [ ]} , 𝑡̂𝑠𝑝 ≤as𝑡̂ ≤ 𝑡̂ + 𝑡̂𝑝 (6)
𝑡̂𝑝
velocity, during stroke𝑡̂reversal, was defined [45]:𝑠𝑝
𝑝
( "  #)
where 𝑡̂𝑠𝑝 is the starting time ofθ. wing 
t̂ = ±
Θ
pitching
1 − cos
and2π 𝑡t̂̂𝑝− is
t̂sp the non-dimensional pitching
, t̂sp ≤ t̂ ≤ t̂sp + t̂ p (6)
duration. Here the wing pitch amplitude, t̂ p Θ, was chosen t̂ p to allow a 45° AoA at mid half-
stroke (known to produce a maximum lift coefficient value [46]). The stroke amplitude
where t̂sp is the starting time of wing pitching and t̂ p is the non-dimensional pitching
(peak-to-peak) values chosen
duration. Herefor the investigation were Θ,
wing pitch amplitude, 60°,
was120°, 160°,
chosen and a180°,
to allow 45◦ AoA allowing
at mid half-
representative values observed for many insects [7]. Furthermore, the increment
stroke (known to produce a maximum lift coefficient value [46]). The stroke amplitude in stroke
amplitude was chosen to decrease
(peak-to-peak) values linearly allowing
chosen for for awere
investigation more 60◦refined
, 120◦ , 160 ◦ , and 180◦ , allowing
investigation as
representative values observed for many insects [7]. Furthermore,
the flapping stroke amplitude approaches its maximum value of 180°. Note that the cur- the increment in stroke
amplitude was chosen to decrease linearly allowing for a more refined investigation as the
rent study only employed one wing for investigation, hence wing-wing interaction effects
flapping stroke amplitude approaches its maximum value of 180◦ . Note that the current
were not considered. studyThe
onlynon-dimensional
employed one wing for acceleration and
investigation, pitching
hence wing-wing durations, ̂𝑎 andwere
interaction𝑡effects
𝑡̂𝑝 , were chosen asnot
0.25 to produce
considered. The representative
non-dimensional flapping
accelerationand
andpitching waveform
pitching durations, varia-
t̂ a and t̂ p , were
chosen
tions of insect flight, as 0.25in
as shown toFigure
produce2.representative
This choice wasflapping
basedandon
pitching waveform
a previous variations of
assessment
insect flight, as shown in Figure 2. This choice was based on a previous assessment which
which showed that varying 𝑡̂𝑎 and 𝑡̂𝑝 has negligible effect on the trends of the average
showed that varying t̂ a and t̂ p has negligible effect on the trends of the average aerodynamic
aerodynamic forceforce coefficients
coefficients[47].
[47].

Figure 2. Flapping angle, 𝜙,Flapping


Figure 2. and angle ofφ,attack,
angle, 𝛼, ofvariations
and angle employed
attack, α, variations in theincurrent
employed investiga-
the current investigation.
tion. Regions with white color indicate wing translational phases, whereas regions with grey color
Regions with white color indicate wing translational phases, whereas regions with grey color indicate
indicate wing rotational
wingphases.
rotational phases.

It should be noted here that the current kinematics definition leads to an implicit
It should be noted here that the current kinematics definition leads to an implicit
angular velocity difference between flapping and pitching, i.e., for larger stroke amplitudes,
angular velocity difference between
the wing travels flapping
a larger andfor
stroke angle pitching,
the samei.e., for larger
duration, stroke
hence the ampli-
flapping angular
tudes, the wing travels
velocityaislarger stroke
increased, angle
whereas the for
wingthe sameangular
pitching duration, hence
velocity thethe
remains flapping
same due to
angular velocity istheincreased,
fixed strokewhereas the wingThis
reversal kinematics. pitching
leads to aangular
relativelyvelocity remains
smaller wing theeffect
rotational
with respect to the wing translational effect for the larger stroke
same due to the fixed stroke reversal kinematics. This leads to a relatively smaller wing amplitude cases. However,
collected kinematics data for real insects, such as the data from Ellington [7] for craneflies,
rotational effect with respect to the wing translational effect for the larger stroke ampli-
ladybirds, hoverflies, droneflies, and bumblebees, showed that the pitching duration is
tude cases. However, collected kinematics data for real insects, such as the data from El-
lington [7] for craneflies, ladybirds, hoverflies, droneflies, and bumblebees, showed that
the pitching duration is typically confined within a relatively fixed period near stroke re-
versal. In addition, the method of unifying a pitching waveform variation while changing
Aerospace 2022, 9, 479 7 of 23

typically confined within a relatively fixed period near stroke reversal. In addition, the
method of unifying a pitching waveform variation while changing the flapping amplitude
has been typically employed in previous studies, such as in the work of Altshuler et al. [29].
Finally, the average velocity of the wing taken at r̂2 , i.e., U2 , was used to define the
Reynolds number using the well-known relation:

U2 c 2Φ f r̂2 Rc
Re = = (7)
ν ν
where Φ is the peak-to-peak flapping stroke amplitude; f is the flapping frequency; and
ν is the kinematic viscosity. Note that the frequency was kept constant throughout our
simulations, and the kinematic viscosity was always adjusted to ensure a constant Reynolds
number for a given stroke amplitude and wing planform.

2.2. Numerical Simulation


The incompressible Navier–Stokes equations, that assume constant viscosity, were
solved for the fluid flow around the investigated 3D wings:

∂ 
uj = 0 (8)
∂x j
!
∂ ∂  1 ∂p ∂ ∂ui
( ui ) + u j ui = − +ν (9)
∂t ∂x j ρ ∂xi ∂x j ∂x j

where ui represents the flow velocity vector; xi represents the Cartesian position vector; t
denotes time; p denotes pressure; and ρ denotes density. The above equations were handled
using the open-source package OpenFOAM with a cell-centred finite volume method. For
the incompressible flow, a combined PISO (Pressure–Implicit with Splitting of Operators)
and SIMPLE (Semi-Implicit Method for Pressure Linked Equations) algorithm known as
PIMPLE was used to solve the pressure-velocity coupling [48,49]. Second order schemes
were employed for both spatial gradient and time evolution terms. The pressure equation
was solved using the GAMG (Geometric–algebraic Multi Grid) method, and the velocity
equation was solved using the Gauss–Seidel method.
The computational domain consisted of an outer spherical domain surrounding an
inner spherical domain with the wing located at its origin. The outer and inner domains
were linked using a spherical sliding interface. At the far-field boundary of the outer
domain, uniform fixed zero values for gauge pressure and velocity gradient were applied.
At the wing surface, a no-slip boundary condition was applied. The inner sphere had a
radius of 36c and the outer sphere had a radius of 40c. Hexahedra and split-hexahedra
cells were generated using the blockMesh and snappyHexMesh utility of OpenFOAM. The
computational domain and mesh setup are demonstrated in Figure 3. Note that throughout
the simulations, the inner domain rotated about the origin to produce the required wing
motions, whereas the outer domain was kept stationary. Flow variables were interpolated
at the sliding interface using an inverse distance algorithm. This allowed the required
wing rotations without mesh deformation or re-meshing, hence ensured high-quality mesh
throughout the simulations.
FOR PEER REVIEW
Aerospace 2022, 9, 479 8 of 24
8 of 23

Figure 3. (a) Simulation domain;


Figure and (b)domain;
3. (a) Simulation planform
and view of an example
(b) planform view of an mesh distribution.
example mesh distribution.

In this work, the instantaneous lift and drag coefficients were obtained using the
In this work, the instantaneous lift and drag coefficients were obtained using the clas-
classical relations:
sical relations:

 2L t̂
C t̂ =
L (10)
ρU22 S
2𝐿(𝑡̂)
𝐶𝐿 (𝑡̂) = 2  2D t̂
 (10)
𝜌𝑈
CD2 𝑆t̂ = (11)
ρU22 S
2𝐷(𝑡̂) lift and drag forces, and S is the wing area. The
 
where L t̂ and D t̂ are the instantaneous
𝐶 𝐷 (𝑡̂ ) = from t̂0 to t̂1 were defined as: (11)
average lift and drag coefficients 𝜌𝑈a 2period
over 𝑆 2

where 𝐿(𝑡̂) and 𝐷(𝑡̂) are the instantaneous lift and dragt̂t̂01 forces,
CL t̂ dt̂ and 𝑆 is the wing area.
R 
C L_average = (12)
The average lift and drag coefficients over a period from t̂𝑡1̂0−to
t̂0 𝑡1̂ were defined as:

1
∫t̂ 𝐶𝐿 (𝑡̂) d𝑡 t̂1̂
R 
t̂0
CD t̂ dt̂
̅
𝐶𝐿_𝑎𝑣𝑒𝑟𝑎𝑔𝑒 0
=D_average
C = (12)(13)
𝑡1̂ − 𝑡̂0 t̂1 − t̂0
Hence, the cycle averaged coefficients, C L and C D , were obtained by taking t̂0 = 4 and
t̂1 = 5 (Note that, investigations in this t̂1 study were for the fifth stroke cycle to ensure that
∫t̂ 𝐶 𝐷 (𝑡̂ ) d𝑡̂
̅
the effect of initial disturbances,
𝐶𝐷_𝑎𝑣𝑒𝑟𝑎𝑔𝑒 typically
= 0 evident in the first few cycles, are damped); (13) the
̂1 − 𝑡̂0
averaged coefficients during the wing 𝑡translational phase, C LT and C DT , were obtained by
taking t̂0 = 4.125 and t̂1 = 4.375; and the averaged coefficients during the wing rotational
Hence, the cycle averaged
phase, C LRcoefficients,
and C DR , were𝐶𝐿̅ obtained
and 𝐶𝐷̅ by , were
takingobtained
t̂0 = 4.375by
andtaking 𝑡̂0 =Moreover,
t̂1 = 4.625. 4 and 𝑡1̂ the
= 5 (Note that, investigations
aerodynamic in this study
efficiency were forusing
was assessed the fifth strokefactor”
the “power cycle to ensure
metric that the
for general wing
motions, using the definition provided in [50]:
effect of initial disturbances, typically evident in the first few cycles, are damped); the
averaged coefficients during the wing translational phase, ̅ ̅
3/2 𝐶𝐿𝑇 and 𝐶𝐷𝑇 , were obtained
CL
by taking 𝑡̂0 = 4.125 and 𝑡1̂ = 4.375; and the averaged Pf = coefficients during the wing rota-(14)
̅ and 𝐶𝐷𝑅
tional phase, 𝐶𝐿𝑅 ̅ , were obtained by taking 𝑡̂0 =C P4.375 and 𝑡1̂ = 4.625. Moreover,
the aerodynamic efficiency R 5 2P(t̂) wing
where C P iswas assessed
the average using
power the “power
coefficient factor” cycle,
over a flapping metricC Pfor
= general
4 ρU 3 S dt̂, and P is
motions, using thethe
definition provided
aerodynamic power. in [50]: 2

To further investigate the distribution


3/2 of the aerodynamic force along the wing length,
𝐶𝐿̅ Cn , was defined at different spanwise sections
the sectional normal force coefficient,
𝑃 = 𝑓 (14)
𝐶𝑃̅
5 2𝑃(𝑡̂ )
where 𝐶𝑃̅ is the average power coefficient over a flapping cycle, 𝐶𝑃̅ = ∫4 3 d𝑡̂, and 𝑃
𝜌𝑈 𝑆 2
is the aerodynamic power.
To further investigate the distribution of the aerodynamic force along the wing
Aerospace 2022, 9, 479 9 of 23

of the wing. This coefficient at a given spanwise location, r̂, was obtained using the
following equation:
R ĉ(r̂)  
2 0 plower (ŷ) − pupper (ŷ) dŷ
Cn (r̂ ) = (15)
ρU22 ĉ(r̂ )
where plower/upper (ŷ) is the pressure on the lower/upper wing surface along the local
dimensionless chordwise coordinate, ŷ. The vortical structures of the simulated flows were
identified using the well-known Q-criterion [51], which represents the contributions of
the antisymmetric and symmetric components of the velocity gradient tensor such that
a positive value of the Q-criterion indicates regions dominated by rotation. Here, the
Q-criterion, Q, was scaled using the following equation:

c2
Q∗ = Q (16)
U22

Finally, for our simulations, mesh systems with approximately 2.3–3.7 million cells
were chosen, depending on the wing planform geometry simulated. The cells close to the
wing surface were refined by subdivision, resulting in approximately 48,000–97,000 nodes
on the wing surface. The time step size, ∆t, was set as 1e-3 (relative to the flapping cycle).
This simulation setup was decided based on a convergence assessment for the mesh and
time step sizes. Moreover, the accuracy of the numerical solver has been validated against
a number of benchmark experimental and numerical data from literature. The convergence
assessment and validations of our numerical method were comprehensively presented in
our previous work [50], hence are not repeated here for brevity.

3. Results and Discussion


3.1. Effect of Aspect Ratio
The investigation provided in this section considers the sole effect of aspect ratio,
while fixing the stroke amplitude value to Φ = 160◦ and the dimensionless radial centroid
location value to r̂1 = 0.5. The normal force coefficient distribution along the wingspan
for the different aspect ratio wings throughout a half-stroke are shown in Figure 4. A
clear distinction is evident regarding the aerodynamic force production at the different
spanwise locations. For all aspect ratios, it is found that the normal force coefficient
increases throughout t̂ = 4.7–4.9 for the sections close to the wing root, whereas for the
section closest to the wingtip, the normal force coefficients decrease during the same period.
On the other hand, at stroke reversal, t̂ = 5.0, the normal force coefficients become negative
for all sections except the ones closest to the wing root, where Cn remains small and positive.
Notably, Figure 4 shows that for all aspect ratios, the sections near the wingtip tend to show
a drop in normal force coefficient after mid half-stroke, i.e., t̂ = 4.8–4.9. This drop in normal
force coefficient is related to wingtip stall effects which lead to loss of aerodynamic force
production [50]. In Figure 4, the sections that show drop in normal force coefficient after
mid half-stroke, i.e., when the normal force coefficient at t̂ = 4.9 becomes lower than that at
t̂ = 4.8, are marked with the grey areas. The grey area region is found to increase from 31%
to 57% of the wingspan as the aspect ratio increases from three to six. This confirms that
higher aspect ratio wings are more susceptible to wingtip stall effects.
Aerospace 2022, 9, x2022,
Aerospace FOR9,PEER
479 REVIEW 10 of 23 10 of 24

Aerospace 2022, 9, x FOR PEER REVIEW 10 of 24

Figure Figure
4. Normal4. Normal force
force coefficient distribution
coefficient distribution along the wingspan
along the for wings
wingspan forwith
wingsdifferent
withaspect
different
Figure
ratios. 4. Normal
Grey areasforce coefficient
represent regionsdistribution along
where the normal the wingspan
force foratwings
coefficient with
𝑡̂ = 4.9 different
becomes aspect aspect
lower
ratios. Grey
ratios. Grey areas represent regions where the normal force coefficient
coefficient at t̂radial at 𝑡̂ = 4.9
lower than lower
becomes
than that at areas represent
𝑡̂ = 4.8. regions where
Stroke amplitude the
is fixed tonormal
160° andforce
the dimensionless = 4.9 centroid
becomeslocation
than that
that atatt̂ =to
is fixed ̂
𝑡 4.8.
= 4.8.
0.5. Stroke
Stroke amplitude
amplitude is fixedistofixed◦
160 andto 160° and the dimensionless
the dimensionless radial centroid radial centroid
location is fixed location
is fixed to 0.5.
to 0.5.
Figure 5 shows the surface pressure coefficient and flow structures at different time
Figure
Figure 5 within
instances 5 shows
shows thesurface
surface
a half-stroke
the pressure
forpressure coefficient
the different andand
aspect ratio
coefficient flowflow
cases. structures at different
For structures
all aspect ratios, thetime time
at different
instances within a half-stroke
LEV, the trailing-edge for the
vortex (TEV), anddifferent aspect(TV)
the tip vortex ratioquickly
cases. For all aspect
roll-up and stay ratios,
close the
instances within a half-stroke for the different aspect ̂ ratio cases. For all aspect ratios, the
to the
LEV, thewing surface following
trailing-edge the startand
vortex (TEV), of the
thehalf-stroke
tip vortex (TV)(𝑡 = 4.6–4.7).
quicklyThe LEVand
roll-up has stay
larger close
LEV,tosize
the trailing-edge
thetowards vortex
the wingtip,
wing surface following (TEV),
resulting inand
the start thehalf-stroke
aofconical
the tip vortex
shaped (t̂(TV)
structure quickly Theroll-up
attached
= 4.6–4.7). to
LEV thehas and
wing stay close
larger
to the wing
leading
size surface
edge.
towards the following
At the the
end ofresulting
wingtip, the start
half-stroke
in aofconical
the half-stroke
𝑡̂ = 5.0, the LEVstructure
shaped ̂ = 4.6–4.7).
is(𝑡found to detach
attached Thefrom
to LEV
the the has larger
wing
wing
size leading
towards surface, and
the At
edge. a
wingtip,region of
the end resultinghigh-pressure coefficient
in a conical
of the half-stroke is
shaped
t̂ = 5.0, the found
LEV is near
structurethe wingtip
found to attached due
detach from to
tothe
the wing
the edge.
wing
leading wake capture
surface,Atandtheaeffect.
regionof
end For 𝐴𝑅 half-stroke
ofthe = 4.5 and 6coefficient
high-pressure cases, twoisthe
𝑡̂ = 5.0, notable
found LEV differences
nearisthe intothe
wingtip
found dueflow
detach to the
from the
structures
wake areeffect.
capture foundFor compared
AR = 4.5 the 𝐴𝑅
to and 6 = 3 case:
cases, two (1) the LEV
notable splits at approximately
differences in the flow 60–
structures
wing70%surface, and a region ̂ =of high-pressure coefficient is found near the at wingtip
𝑡̂ = of
due to
of thecompared
are found wingspan at to 𝑡the 4.6;
ARand = 3 (2) the (1)
case: LEV thedecreases
LEV splits in size near the wingtip60–70%
at approximately
the wake
4.8–4.9,capture
indicatingeffect. For 𝐴𝑅 = 4.5 andnear6 thecases, twofor notable differences in the flow
the wingspan at t̂ =detachment
4.6; and (2)ofthe the
LEVLEV decreases wingtip
in size near the
the higher
wingtipaspect
at t̂ =ratio
4.8–4.9,
structures
cases. are found compared to the 𝐴𝑅 = 3 case: (1) the
indicating detachment of the LEV near the wingtip for the higher aspect ratio cases. LEV splits at approximately 60–
70% of the wingspan at 𝑡̂ = 4.6; and (2) the LEV decreases in size near the wingtip at 𝑡̂ =
4.8–4.9, indicating detachment of the LEV near the wingtip for the higher aspect ratio
cases.

Figure 5. Flow evolution at different time instances for (a) 𝐴𝑅 = 3, (b) 𝐴𝑅 = 4.5, and (c) 𝐴𝑅 = 6. Red
Figure 5. Flow evolution at different time instances for (a) AR = 3, (b) AR = 4.5, and (c) AR = 6. Red
and blue colors indicate positive spanwise (from root to tip) and negative spanwise (from tip to
and blue colors indicate positive spanwise (from root to tip) and negative𝑝 spanwise (from tip to root)
root) rotating vortices, respectively. 𝐶𝑝 is the pressure coefficient: 𝐶𝑝 =p 2.
rotating vortices, respectively. C p is the pressure coefficient: C p = 0.5ρU0.5𝜌𝑈
2.
2
2

Figure 5. Flow evolution at different time instances for (a) 𝐴𝑅 = 3, (b) 𝐴𝑅 = 4.5, and (c) 𝐴𝑅 = 6. Red
Aerospace 2022, 9, x FOR PEER REVIEW 11 of 24

Aerospace 2022, 9, 479 11 of 23

The effect of aspect ratio on the average aerodynamic force production is shown via
Figure
The6.effect
Here,oftoaspect
demonstrate
ratio onthe therelevance of the different
average aerodynamic stroke
force stages, the
production averaging
is shown via
of the force coefficients was conducted over different periods
Figure 6. Here, to demonstrate the relevance of the different stroke stages, the averaging including the whole flap-
of
ping
the stroke,
force i.e., 𝐶𝐿̅ and
coefficients 𝐶𝐷̅ ; the translational
was conducted phaseperiods
over different only, i.e., ̅ and 𝐶𝐷𝑇
𝐶𝐿𝑇
including
̅ ; and the rota-
the whole flapping
tional phase ̅ and 𝐶𝐷𝑅
𝐶𝐿𝑅 ̅ . The results show that the interme-
stroke, i.e., C Lonly
and (stroke
C D ; thereversal), i.e., phase
translational only, i.e., C LT and C DT ; and the rotational
diate aspect ratio wing, 𝐴𝑅
phase only (stroke reversal), i.e., C LR and C DR . The resultsaverage
= 4.5, leads to the highest show that liftthe
coefficient during
intermediate aspectthe
translational
ratio wing, ARphase, = 4.5, 𝐶 ̅
𝐿𝑇 , whereas
leads both the
to the highest 𝐴𝑅 = 4.5
average lift and 6 wingduring
coefficient cases result in the same
the translational
average lift coefficient during the rotational phase, 𝐶 ̅ .
phase, C LT , whereas both the AR = 4.5 and 6 wing cases result in the same average
𝐿𝑅 Hence, the global average liftlift
co-
efficient, 𝐶 ̅ ,
coefficient during
𝐿 has its value at 𝐴𝑅
the rotational phase, C LR . Hence, the global average lift coefficient, C L=,
maximum = 4.5; that said, the difference between 𝐴𝑅
has its maximum value small.
4.5 and 6 is negligibly at AR Similarly,
= 4.5; thatthe average
said, drag coefficient
the difference betweenduring theand
AR = 4.5 transla-
6 is
tional phase, 𝐶̅ , reaches its highest value at 𝐴𝑅
negligibly small.𝐷𝑇Similarly, the average drag coefficient during the translational phase, Ccoef-
= 4.5. However, the average drag DT ,
ficient during
reaches its highestthe rotational
value at AR phase,
= 4.5.̅ However,
𝐶𝐷𝑅 , is more dominant
the average anddrag
decreases with the
coefficient increase
during the
of aspect phase,
rotational ratio, resulting in a consistently
C DR , is more dominant and decreasing
decreasesaverage with thedrag coefficient,
increase 𝐶𝐷̅ ,ratio,
of aspect with
aspect ratio.
resulting in a consistently decreasing average drag coefficient, C D , with aspect ratio.

Figure 6.6. Average


Figure Average lift
lift and
and drag
drag coefficients
coefficients for
for the
the translational
translational phase,
phase, rotational
rotational phase,
phase, and
and
whole/global wing stroke, for the different aspect ratios.
whole/global wing stroke, for the different aspect ratios.

Thecurrent
The currentresults
resultson onthe
theeffect
effectof
ofaspect
aspectratio
ratiosupports
supportssome someof ofthe
theinsights
insightsfrom
from
previousexperimental
previous experimental and and numerical
numerical studies
studies [17,20–23,25],
[17,20–23,25], i.e., when
i.e., when aspectaspect ratio in-
ratio increases,
creases,
the LEV the LEV
tends totends
breakdownto breakdown and detach
and detach near the near the wingtip,
wingtip, whichwhich
leadsleads to reduced
to reduced lift
lift production,
production, hence
hence intermediate
intermediate aspect
aspect ratios
ratios areare a good
a good choicewhen
choice whenlift
liftcoefficient
coefficientisis
concerned. However,
concerned. However, here,here,wewefurther
furthershow
showthatthatthe
thedetachment
detachmentof ofLEV
LEVforforhigh
highaspect
aspect
ratio wings
ratio wings mainly influences
influences force
forceproduction
productionduring
duringthethetranslational
translationalphase;
phase; forfor
thethe
ro-
rotational phase,increasing
tational phase, increasingthe theaspect
aspectratio
ratioresults
results in
in minor
minor differences in lift lift coefficient,
coefficient,
but
butdecreases
decreasesthe
thedrag
dragcoefficient.
coefficient.This
Thisdecrease
decreasein indrag
dragcoefficient
coefficientduring
duringthe therotational
rotational
phase
phaseisisinfluential
influentialasasititleads
leadsto
toaapronounced
pronounceddecrease
decreaseininthe
theoverall
overalldrag
dragcoefficient
coefficientwithwith
the
theincrease
increaseof
ofaspect
aspectratio,
ratio,for
forthe
thecases
casesconsidered
consideredininthis
thisstudy.
study.

3.2.
3.2.Interaction
InteractionEffect
EffectofofStroke
StrokeAmplitude
Amplitudeand andAspect
AspectRatio
Ratio
In
Inthis
thissection,
section,the
theinteraction
interactioneffect of stroke
effect amplitude
of stroke and and
amplitude aspect ratio ratio
aspect is investigated
is investi-
by varying the stroke amplitude for the different aspect ratio cases. The
gated by varying the stroke amplitude for the different aspect ratio cases. The instantane- instantaneous
lift
ousandlift drag coefficients
and drag for these
coefficients casescases
for these are shown in Figure
are shown 7. For
in Figure 7. all
Foraspect ratios,
all aspect the
ratios,
small stroke amplitude case (Φ = 60 ◦ ) shows obvious peaks in lift/drag coefficients during
the small stroke amplitude case (Φ = 60°) shows obvious peaks in lift/drag coefficients
the rotational
during phase, which
the rotational phase, are lessare
which obvious/negligible for the
less obvious/negligible forlarger stroke
the larger amplitude
stroke ampli-
cases. This indicates a higher wing rotational effect (hence unsteadiness) on aerodynamic
tude cases. This indicates a higher wing rotational effect (hence unsteadiness) on aerody-
force production for small stroke amplitude cases. During the translational phase, it is
namic force production for small stroke amplitude cases. During the translational phase,
found that a higher stroke amplitude leads to faster growth of the lift coefficient, i.e., the
it is found that a higher stroke amplitude leads to faster growth of the lift coefficient, i.e.,
lift coefficient reaches a stable value earlier during the translational phase. Hence, a
the lift coefficient reaches a stable value earlier during the translational phase. Hence, a
higher stroke amplitude leads to higher lift coefficient values, but the difference between
higher stroke amplitude leads to higher lift coefficient values, but the difference between
the various stroke amplitudes, becomes small/negligible at the end of the translational
phase, Figure 7.
Aerospace 2022, 9, x FOR PEER REVIEW 12 of 24

Aerospace 2022, 9, 479 12 of 23


the various stroke amplitudes, becomes small/negligible at the end of the translational
phase, Figure 7.

Figure7.
Figure 7. Variations
Variations of
of the
the instantaneous
instantaneouslift
liftand
anddrag
dragcoefficients forfor
coefficients thethe
different stroke
different amplitude
stroke amplitude
and aspect ratio cases. The dimensionless radial centroid location is fixed to 0.5. Wing translational
and aspect ratio cases. The dimensionless radial centroid location is fixed to 0.5. Wing translational
phases are indicated by white shading; wing rotational phases are indicated by grey shading.
phases are indicated by white shading; wing rotational phases are indicated by grey shading.
The average lift/drag coefficient and power factor values are further shown in Figure
The average lift/drag coefficient and power factor values are further shown in Figure 8.
8. For the small and intermediate aspect ratio cases, 𝐴𝑅 = 3 and 4.5, increasing the stroke
For the small and intermediate aspect ratio cases, AR = 3 and 4.5, increasing the stroke
̅ . For
amplitude increases the average lift coefficient during the translational phase, 𝐶𝐿𝑇 the
amplitude increases the average
aspect ratio case, 𝐴𝑅 = 6, 𝐶𝐿𝑇 ̅ first increases with the stroke amplitude until Φ = C160°,
lift coefficient during the translational phase, LT . For
the aspect ratio case, AR = 6, C LT the
and then decreases slightly when
firststroke
increases with the stroke amplitude until Φ = 160◦ ,
amplitude further increases to Φ = 180°. On ◦ . On
and
the then
otherdecreases
hand, the slightly
average when the stroke
lift coefficient amplitude
during further phase,
the rotational increases , isΦfound
̅ to
𝐶𝐿𝑅 = 180to
the other with
decrease hand, thethe average
increase liftstroke
of the coefficient
amplitudeduring for the rotational
all aspect In fact,Cthe
ratios.phase, LR ,variation
is found to
decrease
of 𝐶𝐿𝑅 is more pronounced; hence, the global average lift coefficient, 𝐶𝐿 , decreasesvariation
̅ with the increase of the stroke amplitude for all aspect ratios. ̅ In fact, the with
C LR
ofthe is more
increase ofpronounced;
stroke amplitude hence, fortheallglobal
aspectaverage lift coefficient,
ratios. This C L , decreases
said, the variation of 𝐶𝐿̅ with
with the
increase of stroke amplitude for all aspect ratios. This said, the
stroke amplitude for Φ = 120–180° is small. The average drag coefficient during Lthe trans- variation of C with stroke
amplitude for Φ = 120–180 ◦ is small. The average drag coefficient during the translational
lational phase, 𝐶𝐷𝑇 ̅ , is found to first decrease with the increase of stroke amplitude until
phase,
Φ = 120°, C DTand
, is found to first
then keeps decrease
relatively with the
constant. On increase
the other of hand,
strokethe amplitude
average drag Φ = 120◦ ,
untilcoeffi-
and
cientthen keeps
during therelatively
rotational constant.
phase, 𝐶On ̅ the other hand, the average drag coefficient during
𝐷𝑅 , is found to decrease with the increase of stroke
the rotational
amplitude forphase, C DRratios,
all aspect , is foundand to decrease
again, the variation ̅ is of
with theofincrease
𝐶𝐷𝑅 stroke
more amplitude
evident. for all
As such,
aspect ratios,
the overall and is
effect again,
that the variation
global averageof C DR is more
drag evident.
coefficient, 𝐶𝐷̅ ,As such, the
decreases overall effect
consistently
iswith
thatthe theincrease
global of average drag coefficient,
stroke amplitude. Notably C Dwith
, decreases consistently
the increase with the increase
of stroke amplitude, the
of stroke amplitude.
decrease in the average Notably
drag coefficient, 𝐶𝐷̅ , is larger
with the increase of stroke
than theamplitude,
decrease inthe thedecrease in the
average lift
average
coefficient,drag𝐶𝐿̅ coefficient,
. This resultsCin D , is larger
increased 𝑃
than
𝑓 the
values, decrease
hence in the
aerodynamicaverage lift
efficiency, coefficient,
with the CL.
This results
increase in increased
in stroke Pf values,
amplitude for allhence
aspectaerodynamic efficiency,
ratios considered. with the increase in stroke
However, 𝑃𝑓
amplitude
with the stroke for all aspect ratios
amplitude becomes considered.
smaller when However, the ratio
the aspect increase in Pf with the stroke
gets higher.
amplitude becomes smaller when the aspect ratio gets higher.
For all the aspect ratios considered, the aerodynamic force coefficients during the wing
rotational phase decrease consistently with the increase of stroke amplitude. The reduced
lift and drag coefficients during the wing rotational phase are due to the reduced wing
pitching angular velocity with respect to the flapping angular velocity, hence a smaller wing
rotational effect with respect to the wing translational effect for the larger stroke amplitude
cases. Notably, the contribution of the wing rotational effect is also a function of the wing
Aerospace 2022, 9, 479 13 of 23

kinematics, i.e., a smaller wing rotation duration will lead to larger wing pitching angular
velocity relative to the translational velocity and vice versa. As such, the contribution of
wing rotational effect to the global aerodynamic force is defined by the combined effect
of the stroke amplitude and the wing rotation duration. However, for real insect cases,
the available data within the literature show no obvious relationship between the wing
pitching duration and the stroke amplitude. In fact, a typical insect wing generally has its
pitching rotation confined near stroke reversals, with supination or pronation durations
lasting 10–20% of a full stroke [7] and can reach up to 30% for some insect species [52].
This suggests that insects may rely more on stroke amplitude to modulate the relative
contribution of the rotational and translational effects for aerodynamic force production. In
fact, there are some experimental studies showing that honeybees [29] and bumblebees [53]
Aerospace 2022, 9, x FOR PEER REVIEW 13 of 24
can change their stroke amplitude up to 50% in response to environmental changes, e.g., air
density change, to modulate aerodynamic force production.

Figure8.8.Comparison
Figure Comparisonofofthe
theaverage
averagelift
liftand
anddrag
dragcoefficients
coefficientsasaswell
wellasaspower
powerfactor
factorfor
fordifferent
different
stroke amplitudes and aspect ratios. The dimensionless radial centroid location is fixed to 0.5.
stroke amplitudes and aspect ratios. The dimensionless radial centroid location is fixed to 0.5.

Forflow
The all the aspect ratios
evolutions of theconsidered, the aerodynamic
different stroke amplitude cases force coefficients
near the middleduring
and end the
ofwing rotational (phase
the half-stroke t̂ = 4.7decrease consistently
and 4.9) for the AR = 3, with
4.5,the
andincrease of stroke
6 cases are shownamplitude. The
in Figure 9a–c.
reduced
These timelift and drag
instances coefficients
are during
chosen to best the wing
represent therotational phase areprocess.
LEV development due to the reduced
Generally,
pitching angular ◦
in most cases, the Φ = 60 flow structures show a closely attached LEV near the middle
wing velocity with respect to the flapping angular velocity, hence a
and end of a half-stroke; however, for the larger stroke amplitude cases, Φ = 120–180
smaller wing rotational effect with respect to the wing translational effect for the ◦,
larger
astroke amplitude LEV
more developed cases.atNotably,
the samethe timecontribution
instances isoffound,
the wing rotational
which effectnear
is detached is also
thea
functionnear
wingtip of the
thewing
end of kinematics, i.e., aThis
a half-stroke. smaller wing rotation
is consistent with duration will lead
the variations to force
in the larger
wing pitching angular velocity relative to the translational velocity and vice versa. As
such, the contribution of wing rotational effect to the global aerodynamic force is defined
by the combined effect of the stroke amplitude and the wing rotation duration. However,
for real insect cases, the available data within the literature show no obvious relationship
between the wing pitching duration and the stroke amplitude. In fact, a typical insect wing
c. These time instances are chosen to best represent the LEV development process. Gener-
ally, in most cases, the Φ = 60° flow structures show a closely attached LEV near the mid-
Aerospace 2022, 9, 479 dle and end of a half-stroke; however, for the larger stroke amplitude cases, Φ = 120–180°, 14 of 23
a more developed LEV at the same time instances is found, which is detached near the
wingtip near the end of a half-stroke. This is consistent with the variations in the force
coefficients shown in
coefficients shown inFigure
Figure7,7,where
where thethe larger
larger stroke
stroke amplitude
amplitude casescases
showedshowed
fasterfaster
reach
reach
to a stable value in lift coefficient during the translational phase. It is also notednoted
to a stable value in lift coefficient during the translational phase. It is also that
that when
when the stroke amplitude increases, the LEV detachment region near
the stroke amplitude increases, the LEV detachment region near the wingtip increases the wingtip in-
creases for the 𝐴𝑅 = 6 cases. This was the reason behind the observed
for the AR = 6 cases. This was the reason behind the observed decrease in the average decrease in the
average translational
translational lifthighest
lift for the for the aspect
highestratio
aspect
caseratio case
at the at thestroke
largest largest stroke amplitude,
amplitude, shown in
shown
Figure in
8. Figure 8.

Flow
Figure9.9.Flow
Figure structures
structures forfor
thethe different
different stroke
stroke amplitudes
amplitudes for (a) (a)=AR
for𝐴𝑅 = 3,
3, (b) 𝐴𝑅(b) ARand
= 4.5, = 4.5,
(c) and
𝐴𝑅
6. AR = 6.
=(c)

The results,
The results, shown
shownin in this
this section,
section,indicate
indicatethatthatincreasing
increasingthe
thestroke
strokeamplitude
amplitudein-
in-
creases the
creases theLEV
LEVdevelopment
developmentrate, rate,hence
hencethethelift
liftcoefficient
coefficientduring
duringthe
thetranslational
translationalphase.
phase.
However, this effect is countered by the effect of LEV detachment near the wingtip,
However, this effect is countered by the effect of LEV detachment near the wingtip, which which
leads to reduced lift production for high stroke amplitudes. Nevertheless, since increasing
leads to reduced lift production for high stroke amplitudes. Nevertheless, since increasing
the aspect ratio also leads to faster development of the LEV, the stroke amplitude for maxi-
mum translational lift decreases with the increase of aspect ratio. Therefore, for a given
aspect ratio, the optimum stroke amplitude for translational lift should be large enough for
the LEV development, whilst keeping the wingtip detachment effect small. For example, for
the small aspect ratio case, AR = 3, no obvious detachment of the LEV is evident throughout
the translational phase, hence the average translational lift coefficient increases with the
Aerospace 2022, 9, 479 15 of 23

increase of stroke amplitude; whereas, for the high aspect ratio case, AR = 6, it is found
that the optimum stroke amplitude for translational lift is 160◦ , and further increasing the
stroke amplitude leads to decreased average translational lift coefficient, Figure 8.
Notably, the interaction effect of aspect ratio and stroke amplitude can be correlated to
the mean dimensionless stroke amplitude, Λ = 0.5ΦAR, proposed by Ellington [54] (Note
that Λ is the dimensionless stroke amplitude, A∗ , proposed by Lentink and Dickinson [30],
scaled by a factor of 0.5). Ellington showed that for typical hovering insect wings, Λ
falls within a range between three and six [54]. This supports the results for optimum
translational lift found in the current study: small aspect ratio wings can benefit from larger
stroke amplitudes, whereas large aspect ratio wings are constrained to a maximum stroke
amplitude for optimum lift production. On the other hand, the current results also show
that rotational drag significantly increases at small aspect ratio and stroke amplitude. This
implies that insects with small aspect ratio wings while operating at small stroke amplitudes
may rely more on the ‘drag based’ kinematics, rather than the ‘lift based’ kinematics for
flight weight support.

3.3. Effect of Radial Centroid Location


In this section, the effect of the radial centroid location on aerodynamic force produc-
tion is investigated, while fixing the stroke amplitude value to Φ = 160◦ and the aspect ratio
to its intermediate value of AR = 4.5. The normal force coefficient distributions along the
wingspan throughout a half-stroke are shown in Figure 10. Again, for all radial centroid
location cases, throughout t̂ = 4.7–4.9, the sections close to the wing root show continuously
increasing normal force coefficient, whereas for the section closest to the wingtip, the
normal force coefficients decrease. In Figure 10, the wing sections that have a decreasing
normal force coefficient after mid half-stroke, i.e., when the normal force coefficient at
t̂ = 4.9 becomes lower than that at t̂ = 4.8, are marked with the grey areas. It is evident that
with the increase of the dimensionless radial centroid location from 0.4 to 0.6, the grey area
Aerospace 2022, 9, x FOR PEER REVIEW region is decreased from 53% to 30% of the wingspan. This confirms that smaller radial 16 of 24
centroid location wings are more susceptible to wingtip stall effects due to the smaller wing
area near the wingtip.

Figure 10. Normal


Figure 10. Normal force coefficient
force coefficient distribution along
distribution along thethe wingspan
wingspan for wings
for wings with different
with different radial radial
centroid locations.
centroid locations.Grey
Greyareas represent
areas represent regions
regions wherewhere the normal
the normal force coefficient
force coefficient at 𝑡̂ = 4.9 be-
at t̂ = 4.9 becomes
comes lower than that at 𝑡̂ = 4.8. Stroke amplitude is fixed
◦ to 160° and the aspect
lower than that at t̂ = 4.8. Stroke amplitude is fixed to 160 and the aspect ratio is fixed to 4.5. ratio is fixed to
4.5.
The LEV structures at the different time instances for the dimensionless radial centroid
locations
The LEV r̂1 structures
= 0.4 and 0.6at arethe
shown in Figure
different time11. instances
Note that the
forsame
the demonstration
dimensionless forradial
the cen-
troid locations 𝑟̂1 = 0.4 and 0.6 are shown in Figure 11. Note that the same demonstration
for the dimensionless radial centroid location of 𝑟̂1 = 0.5 was previously shown in Figure
5b. The flow structures in Figure 11 show that the smaller radial centroid location results
in earlier detachment of the LEV near the wingtip and that increasing the radial centroid
Figure 10. Normal force coefficient distribution along the wingspan for wings with different radial
centroid locations. Grey areas represent regions where the normal force coefficient at 𝑡̂ = 4.9 be-
comes lower than that at 𝑡̂ = 4.8. Stroke amplitude is fixed to 160° and the aspect ratio is fixed to
4.5.
Aerospace 2022, 9, 479 16 of 23
The LEV structures at the different time instances for the dimensionless radial cen-
troid locations 𝑟̂1 = 0.4 and 0.6 are shown in Figure 11. Note that the same demonstration
for the dimensionless
dimensionless radial centroid
radial centroid locationlocation of 𝑟̂was
of r̂1 = 0.5 1 = 0.5 was previously
previously shown in shown
Figurein5b.
Figure
The
5b.
flow The flow structures
structures in Figurein11 Figure
show11 show
that that theradial
the smaller smaller radial centroid
centroid location
location results inresults
earlier
in earlier detachment
detachment of the LEV of near
the LEV near the and
the wingtip wingtip
that and that increasing
increasing the radialthe radial centroid
centroid location
location
results inresults in less
less LEV LEV detachment,
detachment, which iswhich is consistent
consistent with thewith the reduced
reduced wingtip wingtip stall
stall effects
effects
for the for ther̂larger
larger 1 cases 𝑟̂ cases
shown
1 inshown
Figure in Figure
10. 10. Interestingly,
Interestingly, the currentthe current
results areresults are
consistent
consistent with the numerical
with the numerical study of Bhatstudyetof
al.Bhat
[55]et al.revolving
for [55] for revolving wings,anwhere
wings, where an evo-
evolutionary
lutionary
optimization optimization
frameworkframework
was employed wastoemployed to findplanform
find the optimal the optimal planform
shapes shapes
with respect to
lift coefficient
with respect tovalues. Their results
lift coefficient showed
values. that theshowed
Their results optimum wing
that the shapes
optimum have larger
wing area
shapes
near the
have largerwingtip,
area nearwiththe dimensionless
wingtip, with radial centroid locations
dimensionless varying between
radial centroid locations0.68–0.71
varying
for the different
between 0.68–0.71 aspect
for theratios considered.
different aspect ratios considered.

Figure 11. Flow evolution at different time instances for (a) 𝑟̂1 = 0.4 and (b) 𝑟̂1 = 0.6.
Figure 11. Flow evolution at different time instances for (a) r̂1 = 0.4 and (b) r̂1 = 0.6.

The
The average
average liftlift and
and drag drag coefficients
coefficients for for the
the different
different radial
radial centroid
centroid location
location cases
cases
are
are presented
presented in in Figure
Figure 12. 12. TheThe results
results show
show thatthat the
the average
average lift lift coefficient
coefficient during
during the the
translational
translational phase,
phase, ̅ , is
𝐶C𝐿𝑇 , highest
is highest for the
for intermediate
the intermediate dimensionless
dimensionless radial centroid
radial centroidlo-
Aerospace 2022, 9, x FOR PEER REVIEW LT 17 of 24
location,𝑟̂1r̂1= =0.5.
cation, 0.5.Note
Note that,
that, although
althoughthe thewingtip
wingtipstallstalleffect
effectisisreduced
reducedfor higher𝑟̂1r̂1cases,
forhigher cases,
the
the formation
formation of of LEV
LEV also also gets
gets delayed
delayed with with the
the increase
increase in in 𝑟̂r̂11 [50]. Hence,
Hence, thethe average
average
translational lift for the highest
translational lift for the highest r̂11 𝑟̂ case of 0.6 is in fact lower than the intermediate𝑟̂r̂11 case
0.6 is in fact lower than the intermediate case
ofof0.5.
0.5.Furthermore, increasing r̂1𝑟̂1leads
Furthermore,increasing leadstotoa adecreased
decreasedaverageaveragelift liftcoefficient
coefficientduring
duringthe the
rotational phase,
rotational phase, C LR 𝐶 ̅ . As such, the global average lift coefficient values,
𝐿𝑅. As such, the global average lift coefficient values, C L𝐿, for both the 𝐶̅ , for both the
r̂1𝑟̂1==0.4 andr̂1𝑟̂=
0.4and = 0.5
1 0.5 casescases areare similar,
similar, whereas
whereas thethe
casecaser̂1 =𝑟̂0.6
1 =results
0.6 results inlowest
in the the lowest 𝐶𝐿̅ .
C L . On
theOnother
the other
hand,hand,
all averageall average drag coefficient
drag coefficient values (translational,
values (translational, rotational,rotational,
and global)and
global) decrease
decrease consistentlyconsistently
with the with increasethe increase
of radial of radial centroid
centroid location. location.

Figure12.
Figure 12.Average
Averagelift
liftand
anddrag
dragcoefficients
coefficientsfor
forthe
thetranslational
translationalphase,
phase,rotational
rotationalphase,
phase,and
andthe
the
whole/global wing stroke, for the different dimensionless radial centroid locations.
whole/global wing stroke, for the different dimensionless radial centroid locations.

3.4. Interaction Effect of Stroke Amplitude and Radial Centroid Location


The instantaneous lift and drag coefficients for the different stroke amplitudes and
the three radial centroid locations considered in this study are shown in Figure 13. Here,
for all results shown in this section, the aspect ratio is fixed at 4.5. The results show that
for the smallest stroke amplitude value of Φ = 60°, clear peaks of lift and drag coefficients
Aerospace 2022, 9, 479 17 of 23

3.4. Interaction Effect of Stroke Amplitude and Radial Centroid Location


The instantaneous lift and drag coefficients for the different stroke amplitudes and the
three radial centroid locations considered in this study are shown in Figure 13. Here, for
all results shown in this section, the aspect ratio is fixed at 4.5. The results show that for
the smallest stroke amplitude value of Φ = 60◦ , clear peaks of lift and drag coefficients are
found during the wing rotational phase. It is also noted that increasing the radial centroid
location tends to reduce the initial peak values. From Figure 13, the small dimensionless
radial centroid location cases of 0.4 show notable differences in instantaneous lift and
drag coefficients when compared with r̂1 = 0.5 and 0.6 cases, as fluctuations in the force
coefficients are found at the beginning of the wing translational phase for cases with stroke
Aerospace 2022, 9, x FOR PEER REVIEW 18 of 24
amplitudes greater than 120◦ . However, these fluctuations in lift and drag coefficients are
not observed for the small stroke amplitude case with Φ = 60 . ◦

Figure13.
Figure 13. Variations
Variations of
ofthe
thelift
liftand
anddrag coefficients
drag for for
coefficients different stroke
different amplitudes
stroke and radial
amplitudes cen-
and radial
troid locations. Wing translational phases are indicated by white shading; wing rotational phases
centroid locations. Wing translational phases are indicated by white shading; wing rotational phases
are indicated by grey shading.
are indicated by grey shading.
Theaverage
The average lift/drag
lift/dragcoefficients
coefficientsand
and power
power factor values
factor values for the
for different stroke
the different am-
stroke
plitude and
amplitude andradial
radial centroid
centroidlocation cases
location are are
cases shownshownin Figure
in Figure 14. For
14. the
Forsmall dimen-
the small di-
sionless radial centroid location case 𝑟̂
1 = 0.4, increasing the stroke
mensionless radial centroid location case r̂1 = 0.4, increasing the stroke amplitude first amplitude first in-
creases the
increases theaverage
averagelift liftcoefficient
coefficientduring
duringthethetranslational
translationalphase
phaseuntil untilΦΦ==160°;
160◦however,
; however,
when the stroke amplitude further increases to Φ = 180°,
◦ 𝐶 ̅𝐿𝑇 is slightly reduced. On the
when the stroke amplitude further increases to Φ = 180 , C LT is slightly reduced. On the
otherhand,
other hand,for forr̂ 𝑟̂1==0.5
0.5and
and0.6,
0.6,increasing
increasing the
the stroke
stroke amplitude
amplitude increases
increases the
the average
averagelift lift
1
coefficient during the translational phase. For all radial centroid location cases,increasing
coefficient during the translational phase. For all radial centroid location cases, increasing
thestroke
the strokeamplitude
amplitude decreases
decreases thethe average
average lift
lift coefficients
coefficients duringduring the
the rotational
rotationalphase,
phase,
whichleads
leadsto toaadecreased
decreasedglobal
global average
average lift
lift coefficient,
coefficient, C ̅
𝐶𝐿 ,, with the increase of stroke
which L with the increase of stroke
amplitude.On
amplitude. Onthetheother
other hand,
hand, the
the average
average drag
drag coefficient
coefficient during
duringthe thetranslational
translationalphase,
phase,
𝐶 ̅ , generally decreases with the increase of stroke amplitude, but the variation of 𝐶𝐷𝑇 ̅
C DT , generally decreases with the increase of stroke amplitude, but the variation of C DTisis
𝐷𝑇
smallfor
small forthe
thedifferent
differentstroke
strokeamplitude
amplitudecases.
cases.However,
However,significant
significantdecrease
decreaseininthethe aver-
average
age drag coefficient during the wing rotational phase, 𝐶𝐷𝑅 ̅ , is found for all radial centroid
location cases, which results in reduced global average drag coefficients, 𝐶𝐷̅ , with the in-
crease of stroke amplitude. Finally, the aerodynamic efficiency, 𝑃𝑓 , is found to increase
with the increase of stroke amplitude for all radial centroid locations, due to the larger
decrease in the drag coefficient as the stroke amplitude increases.
Aerospace 2022, 9, 479 18 of 23

drag coefficient during the wing rotational phase, C DR , is found for all radial centroid
location cases, which results in reduced global average drag coefficients, C D , with the
increase of stroke amplitude. Finally, the aerodynamic efficiency, Pf , is found to increase 19 of 2
Aerospace 2022, 9, x FOR PEER REVIEW
with the increase of stroke amplitude for all radial centroid locations, due to the larger
decrease in the drag coefficient as the stroke amplitude increases.

Figure
Figure 14.14. Comparison
Comparison of the
of the average
average liftdrag
lift and and coefficients
drag coefficients
as well as well asfactor
as power powerforfactor for differen
different
stroke amplitudes and radial centroid locations. Aspect ratio is fixed
stroke amplitudes and radial centroid locations. Aspect ratio is fixed at 4.5. at 4.5.

Figure
Figures15a,b show
15a,b showthe flow evolutions
the flow of the of
evolutions different stroke amplitude
the different cases near
stroke amplitude cases nea
the middle and end of the half-stroke ( t̂ = 4.7
̂ and 4.9) for the cases with r̂
the middle and end of the half-stroke (𝑡 = 4.7 and 4.9) for the cases with 𝑟̂1 = 0.4 and 0.6 1 = 0.4 and
0.6, respectively. The flow evolutions for r̂1 = 0.4 cases show that at t̂̂ = 4.7, the LEV
respectively. The flow evolutions for 𝑟̂1 = 0.4 cases show that at 𝑡 = 4.7, the LEV starts t
starts to detach from the wing surface near the wingtip. Moreover, at t̂ = 4.9, with the
detach from the wing surface near the wingtip. Moreover, at 𝑡̂ = 4.9, with the increase o
increase of stroke amplitude, the detachment/shedding region increases. On the other
stroke
hand, amplitude,
Figure 15b showstheno detachment/shedding
obvious detachment of region
the LEVincreases. Onthe
at t̂ = 4.7 for the
r̂1 other hand, Figur
= 0.6 cases.
15b shows
However, at t̂ no obvious
= 4.9, detachment
a similar of the LEV
trend of increased at 𝑡̂ = 4.7 for the 𝑟̂1 region
LEVdetachment/shedding = 0.6 nearcases.
theHoweve
at 𝑡̂ = 4.9, a similar trend of increased
wingtip is found for larger stroke amplitude cases. LEV detachment/shedding region near the wingti
is found for larger stroke amplitude cases.
Aerospace 2022, 9, 479 19 of 23
Aerospace 2022, 9, x FOR PEER REVIEW 20 of 24

Figure 15. Flow structures for different


different stroke
stroke amplitudes
amplitudes for (a) r̂𝑟̂11==0.4,
for (a) (b)r̂1𝑟̂1= =0.6.
0.4,(b) 0.6.

The
The observed
observed trend
trend ofof flow
flow evolution
evolution with
with stroke
stroke amplitude
amplitude is is similar
similar to the different
to the different
aspect ratio cases described in Section 3.2, i.e., for larger stroke amplitudes,
aspect ratio cases described in Section 3.2, i.e., for larger stroke amplitudes, the wing the wing travels
trav-
aelslarger
a larger distance within the same duration, which leads to a more developed LEV, and
distance within the same duration, which leads to a more developed LEV, and
this
this leads
leads to
to aa higher
higher lift
lift coefficient
coefficient during
during thethe wing
wing translational
translational phase
phase for
for larger
larger stroke
stroke
amplitude
amplitude cases. However, this
cases. However, this effect
effect is
is countered
countered by by the
the effect
effect of
of wingtip
wingtip stall
stall which
which
reduces the lift production near the wingtip. It is found that for larger
reduces the lift production near the wingtip. It is found that for larger radial centroid radial centroidlo-
location cases, the larger wing area at the outboard wing serves to sustain a larger LEV
cation cases, the larger wing area at the outboard wing serves to sustain a larger LEV
before its boundary reaches the trailing edge and detach/shed from the wing surface, hence
before its boundary reaches the trailing edge and detach/shed from the wing surface,
supports a larger LEV circulation at higher stroke amplitudes. For the small dimensionless
hence supports a larger LEV circulation at higher stroke amplitudes. For the small dimen-
radial centroid location of r̂1 = 0.4, the small area near the wingtip leads to stronger
sionless radial centroid location of 𝑟̂ = 0.4, the small area near the wingtip leads to
detachment/shedding of the LEV. This1 leads to the fluctuation in lift/drag coefficients for
stronger detachment/shedding of the LEV. This leads to the fluctuation in lift/drag coeffi-
the high stroke amplitude cases, as shown in Figure 13, and also the decreased average lift
cients for the high stroke amplitude cases, as shown in Figure 13, and also the decreased
coefficient during the wing translational phase at the highest stroke amplitude of Φ = 180◦ ,
average lift coefficient during the wing translational phase at the highest stroke amplitude
as shown in Figure 14.
of Φ = 180°, as shown in Figure 14.
4. Conclusions
4. Conclusions
In this study, the combined effect of stroke amplitude and wing planform on the
In this study,
aerodynamic the combined
performance effect
of insect of stroke
wings amplitude and
was investigated. The wing planform
flow around theon the aer-
wings at a
odynamic performance of insect wings was investigated. The flow around the
Reynolds number of 100 was solved using the open-source finite volume solver OpenFOAM. wings at a
Reynolds number
Aerodynamic forceofcoefficients
100 was solved
and flowusing the open-source
structures finite volume
were considered solver
for wing Open-
planforms
FOAM. using
defined Aerodynamic force coefficients
the beta-function and Normal
distribution. flow structures
hoveringwere considered
kinematics for using
defined wing
planforms
simple defined
harmonic using the
functions beta-function
were distribution.
employed within Normal hovering
the simulations. kinematics
The effects de-
of the wing
fined using
planform simple harmonic
parameters of aspectfunctions were employed
ratio (varying betweenwithin the simulations.
three and The effects
six) and dimensionless
of the wing planform parameters of aspect ratio (varying between three and six) and
Aerospace 2022, 9, 479 20 of 23

radial centroid location (varying between 0.4 and 0.6) as well as their interaction effects
with stroke amplitude (varying between 60◦ and 180◦ ) were investigated.
Considering the effect of aspect ratio, it was shown that for all aspect ratios considered,
generally before stroke reversal, the wing sections close to the wing root have continuously
increasing normal force coefficients, whereas the wing sections closest to the wingtip have
continuously decreasing normal force coefficient. Moreover, increasing the aspect ratio
was shown to increase the wingtip stall effects and resulted in stronger LEV detachment
near the wingtip. When the interaction effect of aspect ratio and stroke amplitude was
considered, it was shown that larger stroke amplitudes result in a more developed LEV,
hence higher lift coefficients at the beginning of the translational phase for all aspect ratios.
For aspect ratios of 3 and 4.5, increasing the stroke amplitude leads to increased average
translational lift. However, for the high aspect ratio of 6, the average translational lift only
increases up to a stroke amplitude of 160◦ . These results showed that for higher aspect
ratio wings, a smaller stroke amplitude is required to enable the maximum translational
lift, as further increasing the stroke amplitude leads to reduced translational lift due to the
stronger wingtip stall effects.
Considering the effect of radial centroid location, it was found that for all radial cen-
troid location cases considered, again, generally before stroke reversal, the wing sections
close to the wing root have continuously increasing normal force coefficients, and that the
sections closest to the wingtip have continuously decreasing normal force coefficients. Fur-
thermore, increasing radial centroid location was shown to reduce wingtip stall effects and
resulted in weaker LEV detachment. When the interaction effect of radial centroid location
and stroke amplitude was considered, it was shown that, for the small dimensionless radial
centroid location of 0.4, increasing the stroke amplitude increases the average translational
lift coefficient only up to a stroke amplitude of 160◦ , due to the stronger wingtip stall effects
at the highest stroke amplitude. However, for the larger dimensionless radial centroid
locations of 0.5 and 0.6, increasing stroke amplitude leads to a consistently increasing
average translational lift.
An important outcome of this work is the improved understanding of the distinct
contributions of the translational and rotational phases of the stroke cycle and how these
affect the global average aerodynamic coefficient values. Although the lift and drag
coefficients during the translational phase were influenced by the LEV development and
wingtip stall effects, it was found that the rotational phase plays a significant role in
determining the global force production. When varying the aspect ratio while fixing the
radial centroid location and stroke amplitude, the intermediate aspect ratio of 4.5 was
found to have the highest global average lift coefficient; and the global average drag
coefficient was found to decrease with the increase of aspect ratio, due to the decreased
drag coefficient during the rotational phase. When changing the radial centroid location
while fixing the aspect ratio and stroke amplitude, the intermediate dimensionless radial
centroid location of 0.5 was found to have the highest global average lift coefficient; and
increasing radial centroid location led to decreased global average drag coefficient. Overall,
when the interaction effects were considered, it was found that for all aspect ratios and
radial centroid locations, consistently decreasing global average lift and drag coefficients
were found with the increase in stroke amplitude, due to the decreased average lift and
drag coefficients during the rotational phase. Furthermore, the significant reduction in
average drag coefficients led to a consistently increasing aerodynamic efficiency with the
increase in stroke amplitude.

Author Contributions: Conceptualization, H.L. and M.R.A.N.; methodology, H.L. and M.R.A.N.;
investigation, H.L. and M.R.A.N.; writing—original draft preparation, H.L.; writing—review and
editing, M.R.A.N.; supervision, M.R.A.N.; project administration, M.R.A.N.; funding acquisition,
M.R.A.N. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Leverhulme Trust via Research Project Grant RPG-2019-366.
Institutional Review Board Statement: Not applicable.
Aerospace 2022, 9, 479 21 of 23

Informed Consent Statement: Not applicable.


Data Availability Statement: The data presented in this study are available on request from the authors.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ma, K.Y.; Chirarattananon, P.; Fuller, S.B.; Wood, R.J. Controlled Flight of a Biologically Inspired, Insect-Scale Robot. Science 2013,
340, 603–607. [CrossRef] [PubMed]
2. Lentink, D.; Jongerius, S.R.; Bradshaw, N.L. The Scalable Design of Flapping Micro-Air Vehicles Inspired by Insect Flight. In
Flying Insects and Robots; Springer: Berlin/Heidelberg, Germany, 2009; pp. 185–205. ISBN 978-3-540-89392-9. [CrossRef]
3. Keennon, M.; Klingebiel, K.; Won, H. Development of the Nano Hummingbird: A Tailless Flapping Wing Micro Air Vehicle.
In Proceedings of the 50th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition,
Nashville, TN, USA, 9–12 January 2012; American Institute of Aeronautics and Astronautics: Reston, VA, USA, 2012. [CrossRef]
4. Nabawy, M.R.A.; Marcinkeviciute, R. Scalability of Resonant Motor-Driven Flapping Wing Propulsion Systems. R. Soc. Open Sci.
2021, 8, 210452. [CrossRef] [PubMed]
5. Chen, S.; Wang, L.; He, Y.; Tong, M.; Pan, Y.; Ji, B.; Guo, S. Aerodynamic Performance of a Flyable Flapping Wing Rotor with
Passive Pitching Angle Variation. IEEE Trans. Ind. Electron. 2022, 69, 9176–9184. [CrossRef]
6. Zhu, Z.; Zhao, J.; He, Y.; Guo, S.; Chen, S.; Ji, B. Aerodynamic Analysis of Insect-like Flapping Wings in Fan-Sweep and Parallel
Motions with the Slit Effect. Biomim. Intell. Robot. 2022, 2, 100046. [CrossRef]
7. Ellington, C.P. The Aerodynamics of Hovering Insect Flight.3. Kinematics. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 1984, 305,
41–78. [CrossRef]
8. Maxworthy, T. Experiments on the Weis-Fogh Mechanism of Lift Generation by Insects in Hovering Flight Part 1. Dynamics of
the ‘Fling.’ J. Fluid Mech. 1979, 93, 47–63. [CrossRef]
9. Ellington, C.P.; van den Berg, C.; Willmott, A.P.; Thomas, A.L.R. Leading-Edge Vortices in Insect Flight. Nature 1996, 384, 626–630.
[CrossRef]
10. Dickinson, M.H.; Lehmann, F.O.; Sane, S.P. Wing Rotation and the Aerodynamic Basis of Insect Flight. Science 1999, 284, 1954–1960.
[CrossRef]
11. Nabawy, M.R.A.; Crowther, W.J. The Role of the Leading Edge Vortex in Lift Augmentation of Steadily Revolving Wings: A
Change in Perspective. J. R. Soc. Interface 2017, 14, 20170159. [CrossRef]
12. Nabawy, M.R.; Villamor, G.J.; Li, H. Aerodynamic Modelling of Insect Wings Using Joukowski Transformation. In Proceedings of
the AIAA AVIATION 2021 FORUM, Virtual Event, 2–6 August 2021; American Institute of Aeronautics and Astronautics: Reston,
VA, USA, 2021. [CrossRef]
13. Bomphrey, R.J.; Nakata, T.; Phillips, N.; Walker, S.M. Smart Wing Rotation and Trailing-Edge Vortices Enable High Frequency
Mosquito Flight. Nature 2017, 544, 92–95. [CrossRef]
14. Li, H.; Nabawy, M.R. Wake Effects on Force Production of a Translating-Pitching Flat Plate. In Proceedings of the AIAA
AVIATION 2021 FORUM, Virtual Event, 2–6 August 2021; American Institute of Aeronautics and Astronautics: Reston, VA,
USA, 2021. [CrossRef]
15. Nabawy, M.R.A.; Crowther, W.J. Optimum Hovering Wing Planform. J. Theor. Biol. 2016, 406, 187–191. [CrossRef] [PubMed]
16. Usherwood, J.R.; Ellington, C.P. The Aerodynamics of Revolving Wings-II. Propeller Force Coefficients from Mayfly to Quail. J.
Exp. Biol. 2002, 205, 1565–1576. [CrossRef]
17. Kruyt, J.W.; Van Heijst, G.J.F.; Altshuler, D.L.; Lentink, D. Power Reduction and the Radial Limit of Stall Delay in Revolving
Wings of Different Aspect Ratio. J. R. Soc. Interface 2015, 12. [CrossRef] [PubMed]
18. Broadley, P.; Nabawy, M.R.; Quinn, M.K.; Crowther, W.J. Wing Planform Effects on the Aerodynamic Performance of Insect-like
Revolving Wings. In Proceedings of the AIAA AVIATION 2020 FORUM, Virtual Event, 15–19 June 2020; American Institute of
Aeronautics and Astronautics: Reston, VA, USA, 2020. [CrossRef]
19. Broadley, P.; Nabawy, M.R. Effects of Wing Planform Shape on Low Reynolds Number Revolving Wings. In Proceedings of the
AIAA AVIATION 2021 FORUM, Virtual Event, 2–6 August 2021; American Institute of Aeronautics and Astronautics: Reston,
VA, USA, 2021. [CrossRef]
20. Harbig, R.R.; Sheridan, J.; Thompson, M.C. Reynolds Number and Aspect Ratio Effects on the Leading-Edge Vortex for Rotating
Insect Wing Planforms. J. Fluid Mech. 2013, 717, 166–192. [CrossRef]
21. Phillips, N.; Knowles, K.; Bomphrey, R.J. The Effect of Aspect Ratio on the Leading-Edge Vortex over an Insect-like Flapping
Wing. Bioinspir. Biomim. 2015, 10, 056020. [CrossRef] [PubMed]
22. Phillips, N.; Knowles, K.; Bomphrey, R.J. Petiolate Wings: Effects on the Leading-Edge Vortex in Flapping Flight. Interface Focus
2017, 7, 20160084. [CrossRef]
23. Han, J.S.; Chang, J.W.; Cho, H.K. Vortices Behavior Depending on the Aspect Ratio of an Insect-like Flapping Wing in Hover. Exp.
Fluids 2015, 56, 181. [CrossRef]
24. Luo, G.; Sun, M. Effects of Corrugation and Wing Planform on the Aerodynamic Force Production of Sweeping Model Insect
Wings. Acta Mech. Sin. Xuebao 2005, 21, 531–541. [CrossRef]
Aerospace 2022, 9, 479 22 of 23

25. Shahzad, A.; Tian, F.-B.; Young, J.; Lai, J.C.S. Effects of Wing Shape, Aspect Ratio and Deviation Angle on Aerodynamic
Performance of Flapping Wings in Hover. Phys. Fluids 2016, 28, 111901. [CrossRef]
26. Bhat, S.S.; Zhao, J.; Sheridan, J.; Hourigan, K.; Thompson, M.C. Aspect Ratio Studies on Insect Wings. Phys. Fluids 2019, 31,
121301. [CrossRef]
27. Sane, S.P.; Dickinson, M.H. The Control of Flight Force by a Flapping Wing: Lift and Drag Production. J. Exp. Biol. 2001, 204,
2607–2626. [CrossRef] [PubMed]
28. Phillips, N.; Knowles, K. Reynolds Number and Stroke Amplitude Effects on the Leading-Edge Vortex on an Insect-like Flapping
Wing. In Proceedings of the International Powered Lift Conference, Philadelphia, PA, USA, 5–7 October 2010; pp. 365–374.
29. Altshuler, D.L.; Dickson, W.B.; Vance, J.T.; Roberts, S.P.; Dickinson, M.H.; Altshuler, D.L.; Dickinson, M.H. Short-Amplitude
High-Frequency Wing Strokes Determine the Aerodynamics of Honeybee Flight. Proc. Natl. Acad. Sci. USA 2005, 102, 18213–18218.
[CrossRef]
30. Lentink, D.; Dickinson, M.H. Rotational Accelerations Stabilize Leading Edge Vortices on Revolving Fly Wings. J. Exp. Biol. 2009,
212, 2705–2719. [CrossRef]
31. Ellington, C.P. The Aerodynamics of Hovering Insect Flight.2. Morphological Parameters. Philos. Trans. R. Soc. Lond. Ser. B Biol.
Sci. 1984, 305, 17–40. [CrossRef]
32. Broadley, P.; Nabawy, M.R.A.; Quinn, M.K.; Crowther, W.J. Dynamic Experimental Rigs for Investigation of Insect Wing
Aerodynamics. J. R. Soc. Interface 2022, 19, 20210909. [CrossRef] [PubMed]
33. Zhao, L.; Huang, Q.; Deng, X.; Sane, S.P. Aerodynamic Effects of Flexibility in Flapping Wings. J. R. Soc. Interface 2010, 7, 485–497.
[CrossRef]
34. Du, G.; Sun, M. Effects of Wing Deformation on Aerodynamic Forces in Hovering Hoverflies. J. Exp. Biol. 2010, 213, 2273–2283.
[CrossRef]
35. Toomey, J.; Eldredge, J.D. Numerical and Experimental Study of the Fluid Dynamics of a Flapping Wing with Low Order
Flexibility. Phys. Fluids 2008, 20, 073603. [CrossRef]
36. Eldredge, J.D.; Toomey, J.; Medina, A. On the Roles of Chord-Wise Flexibility in a Flapping Wing with Hovering Kinematics. J.
Fluid Mech. 2010, 659, 94–115. [CrossRef]
37. Yin, B.; Luo, H. Effect of Wing Inertia on Hovering Performance of Flexible Flapping Wings. Phys. Fluids 2010, 22, 111902.
[CrossRef]
38. Kang, C.; Shyy, W. Scaling Law and Enhancement of Lift Generation of an Insect-Size Hovering Flexible Wing. J. R. Soc. Interface
2013, 10, 20130361. [CrossRef]
39. Nakata, T.; Liu, H. Aerodynamic Performance of a Hovering Hawkmoth with Flexible Wings: A Computational Approach. Proc.
R. Soc. B Biol. Sci. 2012, 279, 722–731. [CrossRef] [PubMed]
40. Shahzad, A.; Tian, F.-B.; Young, J.; Lai, J.C.S. Effects of Flexibility on the Hovering Performance of Flapping Wings with Different
Shapes and Aspect Ratios. J. Fluids Struct. 2018, 81, 69–96. [CrossRef]
41. Dai, H.; Luo, H.; Doyle, J.F. Dynamic Pitching of an Elastic Rectangular Wing in Hovering Motion. J. Fluid Mech. 2012, 693,
473–499. [CrossRef]
42. Wang, L.; Tian, F.-B. Numerical Study of Sound Generation by Three-Dimensional Flexible Flapping Wings during Hovering
Flight. J. Fluids Struct. 2020, 99, 103165. [CrossRef]
43. Sridhar, M.; Kang, C. Aerodynamic Performance of Two-Dimensional, Chordwise Flexible Flapping Wings at Fruit Fly Scale in
Hover Flight. Bioinspir. Biomim. 2015, 10, 036007. [CrossRef]
44. Kang, C.-K.; Aono, H.; Cesnik, C.E.S.; Shyy, W. Effects of Flexibility on the Aerodynamic Performance of Flapping Wings. J. Fluid
Mech. 2011, 689, 32–74. [CrossRef]
45. Sun, M.; Tang, J. Unsteady Aerodynamic Force Generation by a Model Fruit Fly Wing in Flapping Motion. J. Exp. Biol. 2002, 205,
55–70. [CrossRef]
46. Nabawy, M.R.A.; Crowther, W.J. Aero-Optimum Hovering Kinematics. Bioinspir. Biomim. 2015, 10, 044002. [CrossRef]
47. Li, H.; Nabawy, M.R. The Combined Effect of Wing Planform and Stroke Kinematics on Aerodynamics of Flapping Insect Wings.
In Proceedings of the AIAA SCITECH 2022 Forum, San Diego, CA, USA & Virtual Event, 3–7 January 2022; American Institute of
Aeronautics and Astronautics: Reston, VA, USA, 2022. [CrossRef]
48. Patankar, S.V.; Spalding, D.B. A Calculation Procedure for Heat, Mass and Momentum Transfer in Three-Dimensional Parabolic
Flows. Int. J. Heat Mass Transf. 1972, 15, 1787–1806. [CrossRef]
49. Issa, R.I. Solution of the Implicitly Discretised Fluid Flow Equations by Operator-Splitting. J. Comput. Phys. 1986, 62, 40–65.
[CrossRef]
50. Li, H.; Nabawy, M.R.A. Wing Planform Effect on the Aerodynamics of Insect Wings. Insects 2022, 13, 459. [CrossRef] [PubMed]
51. Hunt, J.C.R.; Wray, A.A.; Moin, P. Eddies, Streams, and Convergence Zones in Turbulent Flows. Center for Turbulence
Research, Proceedings of the Summer Program 1988. Document ID: 19890015184. pp. 193–208. Available online: https:
//www.researchgate.net/publication/234550074_Eddies_streams_and_convergence_zones_in_turbulent_flows (accessed on
2 July 2022).
52. Cheng, X.; Sun, M. Wing-Kinematics Measurement and Aerodynamics in a Small Insect in Hovering Flight. Sci. Rep. 2016,
6, 25706. [CrossRef]
Aerospace 2022, 9, 479 23 of 23

53. Dillon, M.E.; Dudley, R. Surpassing Mt. Everest: Extreme Flight Performance of Alpine Bumble-Bees. Biol. Lett. 2014, 10, 20130922.
[CrossRef]
54. Ellington, C.P. The Aerodynamics of Hovering Insect Flight.4. Aerodynamic Mechanisms. Philos. Trans. R. Soc. Lond. Ser. B Biol.
Sci. 1984, 305, 79–113. [CrossRef]
55. Bhat, S.S.; Zhao, J.; Sheridan, J.; Hourigan, K.; Thompson, M.C. Evolutionary Shape Optimisation Enhances the Lift Coefficient of
Rotating Wing Geometries. J. Fluid Mech. 2019, 868, 369–384. [CrossRef]

You might also like