0% found this document useful (0 votes)
2 views38 pages

Keerthi Seminar Report

The seminar report discusses the concept of bio-mimicry in aerodynamics, focusing on insights derived from insect wing design. It highlights how nature's evolutionary adaptations, particularly in insect flight, can inspire innovative solutions in engineering and technology. The report covers various aspects of insect wings, their aerodynamics, and the implications for developing bio-inspired micro air vehicles.

Uploaded by

augardencreators
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
2 views38 pages

Keerthi Seminar Report

The seminar report discusses the concept of bio-mimicry in aerodynamics, focusing on insights derived from insect wing design. It highlights how nature's evolutionary adaptations, particularly in insect flight, can inspire innovative solutions in engineering and technology. The report covers various aspects of insect wings, their aerodynamics, and the implications for developing bio-inspired micro air vehicles.

Uploaded by

augardencreators
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 38

UNIVERSITY OF AGRICULTURAL SCIENCES, RAICHUR

COLLEGE OF AGRICULTURE, RAICHUR


DEPARTMENT OF ENTOMOLOGY

SEMINAR REPORT ON

“Bio-mimicry in aerodynamics: Insights from insect wing design”

Submitted To

PG Coordinator

Department of Entomology
College of Agriculture, Raichur

Submitted By

Keerthiraman B
PG22AGR15040
Sr. M.sc (Agri.)

DEPARTMENT OF ENTOMOLOGY
COLLEGE OF AGRICULTURE, RAICHUR
UNIVERSITY OF AGRICULTURAL SCIENCES, RAICHUR
584104
0
UNIVERSITY OF AGRICULTURAL SCIENCES, RAICHUR
COLLEGE OF AGRICULTURE, RAICHURDEPARTMENT OF
ENTOMOLOGY

CERTIFICATE

This is to certify that the seminar entitled “Bio-mimicry in aerodynamics: Insights from insect
wing design” submitted by Mr. Keerthiraman B for the seminar-I AET-591 to the DEPARTMENT
OF ENTOMOLOGY, is a record of the seminar report prepared by him during the period of his study
in the University of Agricultural Sciences, Raichur.

Date: 29/01/2025
Place: Raichur PG Coordinator

1
CONTENTS

Sl. No. Title Page No.


1. Introduction 04

2. Insects wings and flight 07

3. Brief overview of aerodynamics 10

4. Aerodynamics of insect wings 11

5. Bio-mimicry from aerodynamics of insect’s wings 24

6. Conclusion 34

7. Selected references 35

FIGURES

Sl. No. Title Page no.


Natural vs. Engineered Flight: A Comparison of Relative Speed in the 04
1
Spine-Tailed Swift and SR-71 Blackbird
2 Fossil Evidence of Early Insect Flight (Delitzschala bitterfeldensis). 05
06
Relationship Between Lift-to-Drag Ratio and Reynolds Number for
3
Different Surfaces.

4 Reynolds Number and Wing Beat Frequency in Avian and Insect Flight. 06

5 Muscles involved in the wing moments of an insect. 08

6 Wing articulation of insects. 09

7 Wing venation of insect 10


Force Generation & Vortex Wake Formation During a Forward Flight Stroke of 11
8 an Insect Wing and Phases of a Forward Flight Stroke in an Insect: Force
Vectors & Vortex Structures
9 Delayed Stall, Rotational Lift, and Wake Capture in Insect Flight 11

10 Leading edge vortex at different angle of attacks. 12

11 Apparatus for measuring aerodynamic forces on an accelerating wing 13

12 Flow visualization of the model wing. 14


Graph representing the variation of force coefficient with the variation in angle 15
13 of attack.
14 Graph showing the variation of force coefficient with the time. 16

2
Instability and Stability of Leading-Edge Vortices (LEV) on Swept and 18
15
Unswept Wings at Different Reynolds Numbers and Distances Traveled
16 Visual Representation of Rotational Lift Dynamics 19

17 Visual Representation of Rotational Lift and Wake Capture Dynamics. 20


The Dynamics of Clap and Fling: A Step-by-Step Breakdown of Lift 21
18
Production
19 Free flight of a very small chalcid wasp 21
20 The Weis-Fogh Clap-Fling Mechanism in Insect Flight 22

21 Bio-Inspired Robotics: Exploring Wing Variations in Insect Mimics 24

22 Advancements in Micro Air Vehicle Technology: RoboBees 25

23 Mars Exploration with Robobees: A Collaborative Approach 26

24 Beetle Wing Deployment: Natural and Robotic Mimicry 27

25 DelFly: Close-up View of a Flapping-Wing Drone 27

26 Quantifying Flight Maneuvers: A Comparative Study 28

27 Dragonfly inspired blade design geometric characteristics 31

TABLES

Sl.
Title Page no.
No.
1 Contribution of the clap and fling to the average vertical force (Fz) at each 29
stroke period.
2 Contribution of the clap and fling to the average force in the η-direction 30
or average horizontal drag force (Fη) at each stroke period.

3
1 Introduction:
“We must draw our standards from the natural world. We must honor
with the humility of the wise the bounds of that natural world and the
mystery which lies beyond them, admitting that there is something in the
order of being which evidently exceeds all our competence.”
- VÁCLAV HAVEL

Biomimicry (from the Greek words bios, meaning "life," and mimesis, meaning "to imitate")
literally means "to imitate life." It is the study of emulating and mimicking nature to solve human
problems. Biomimicry is a relatively new discipline that examines nature's best ideas and then imitates
these designs and processes to address challenges. (Johnson, 2022)

To make a better understanding of biomimicry, consider the example of Japan’s Shinkansen


Bullet Train, which was redesigned using inspiration from the kingfisher bird. Early bullet trains faced a
significant issue-when exiting tunnels at high speeds, they produced loud noises due to air pressure
build-up, disturbing residents and reducing efficiency. Engineers turned to nature for a solution and
observed how the kingfisher dives into water with minimal splash due to its streamlined beak shape. By
redesigning the train’s nose to mimic the kingfisher’s beak, they successfully reduced noise, improved
aerodynamics, and enhanced energy efficiency. This innovation made the train 10% faster, reduced
electricity consumption by 15%, and eliminated the tunnel boom effect. This example highlights how
biomimicry allows us to solve engineering challenges by learning from nature’s time-tested designs.

Biomimicry suggests that nature is the most influential and reliable source of innovation for
designers due to its 3.85 billion years of evolution. Over this vast period, nature has accumulated
immense experience in solving environmental challenges and supporting life.

Fig 1. Natural vs. Engineered Flight: A Comparison of Relative Speed in the Spine-Tailed Swift and
SR-71 Blackbird

4
A striking example is the spine-tailed swift, a small bird capable of achieving a relative speed of
150 body-lengths per second. This feat is all the more impressive when compared to the SR-71
Blackbird, a marvel of human engineering, which, despite its impressive absolute speed, achieves a
relative speed of only 32 body-lengths per second.

This significant difference highlights how natural selection has optimized biological systems for
resource efficiency over millions of years, offering valuable lessons for the development of more
sustainable and resource-conscious technologies through biomimicry.

“Humans fly commercially, but animals fly professionally.”

- McMasters & Henderson

Insect wings offer a unique perspective on flight compared to their avian counterparts, making
them a compelling subject of study. As evidenced by the 325-million-year-old fossil Delitzschala
bitterfeldensis, a member of the extinct Paleodictyoptera order, the principles of insect flight have been
established and refined over vast evolutionary timescales. This fossil, one of the oldest known winged
insects, provides valuable insight into the early evolution of insect flight and highlights the long-
standing success of this mode of locomotion. This evolutionary history has resulted in flight strategies
optimized for the unique physical environment experienced by insects.

Fig 2: Fossil Evidence of Early Insect Flight (Delitzschala bitterfeldensis)

A key factor in this environment is the Reynolds number (Re), a dimensionless quantity that
describes the ratio of inertial forces to viscous forces in a fluid flow. Insects operate at low Reynolds
numbers (Re < 10^4), where viscous forces are dominant. The operation of the airfoil (or wing) at
higher chord Reynolds numbers, e.g., higher flow speeds, will improve aerodynamic efficiency because
the inertial effects dominate over the viscous effects. Notice that the Reynolds number can also be

5
increased by increasing the scale (size) of the wing. However, at lower Reynolds numbers, the relative
effects of viscosity are higher, which manifests as higher profile drag and a lower lift-to-drag ratio.

Fig 3: Relationship between Lift-to-Drag Ratio and Reynolds Number for Different
Surfaces.

Insects operate at a much smaller scale, resulting in significantly lower wing inertia and
consequently, faster response times and higher flapping frequencies. Unlike birds, which distribute
actuation across their wings and utilize large active shape deformations for control, insects primarily
actuate at the wing hinge and rely on passive control through shape deformation.

Fig 4: Reynolds Number and Wing Beat Frequency in Avian and Insect Flight.

A key distinction lies in the relationship between timescales: for insects, flapping timescales
are much shorter than those of the overall vehicle, simplifying control strategies. Furthermore, insects
possess a much higher wing-to-body mass ratio (10-20%) compared to birds (1-5%), suggesting a

6
greater role for wings in weight support and potentially, power generation. These highly efficient wings
outperform conventional airfoils at these scales, achieving high lift-to-drag ratios.

The study of insect wings thus offers valuable insights for the development of bio-inspired
micro air vehicles, particularly in areas like agile maneuverability and efficient flight at small scales.

2 Insects wings and flight


Wings are the defining characteristic of pterygotes and a key innovation in the evolutionary
history of insects. Flight enables a wide range of activities; the detailed structure and form of insect
wings reflect their adaptation to this mode of locomotion. Furthermore, flight is only useful if it is stable
and controlled.

Wings are a key feature that sets pterygote insects apart, and they were a major breakthrough
in insect evolution. The ability to fly allows insects to do many different things, and the detailed
structure and shape of their wings clearly show how well they are suited for flight. However, flying is
only helpful if an insect can fly steadily and control its flight. By changing the precise way, they beat
their wings, many insects can move through the air with great agility and control. The way air flows
around insect wings is very different from how air flows around the wings of airplanes. These
differences explain how insects can lift their own weight even with wings that are quite small compared
to their bodies.
Insect wing muscles are categorized into three functional groups: direct, indirect, and
accessory. Direct muscles insert into the wing base, enabling specific movements like flexing and
steering. Indirect muscles are the primary power source for flight, responsible for the upstroke and
downstroke movements. They achieve this by manipulating the wings indirectly through deformations
of the thorax.
Accessory muscles, including pleurosternal muscles controlling lateral thoracic elasticity and
tergopleural muscles affecting thoracic stiffness near the wing hinge, fine-tune wing movements. These
accessory muscles, though not directly attached to the wings, influence the mechanics of the wing hinge
and contribute to precise flight control.
Direct muscles, inserting into the axillary sclerites or the basalar and subalar regions,
contribute to wing twisting, depression, and extension. Odonata insects possess numerous direct
muscles, facilitating precise wing adjustments for efficient aerodynamics and steering. The upstroke in
most insects is powered by the indirect dorso-ventral muscles which pull the tergum down, indirectly
lifting the wings. The downstroke is typically driven by the indirect dorsal longitudinal muscles that pull
the wings down directly.

7
Asynchronous muscle contraction, common in most insects, relies on stretch activation,
highlighting the importance of thoracic and wing base elasticity. Overall, flight mechanics in insects
involve a complex interplay of muscle action, elasticity, and the unique biomechanics of the thorax and
wing base. (Chapman, 1998)

Direct flight muscles

Accessory muscles

Fig 5: Muscles involved in the wing moments of an insect.

Insect wings articulate with the thorax through a complex system of membranes, sclerites, and
muscles, enabling controlled flight. Flexible membranes connect the wing base to the thorax, housing
several key sclerites, including the axillary sclerites, which function as levers to translate thoracic
movements into wing motion.

Typically, three axillary sclerites are involved: the first articulates with the anterior notal wing
process and subcostal vein; the second connects to the pleural wing process and radius base; and the
third articulates with the posterior notal process and anal veins. These sclerites work in concert,
creating multiple articulation points that allow for a wide range of wing movements, including flapping,
rotation, and folding, essential for efficient flight.

8
Muscles, both direct and indirect, attach to these sclerites and the wing base, providing the force
and fine-tuned control necessary for maneuvering during flight.

Fig 6: Wing articulation of insects.

Insect wings are thin, rigid structures extending from the meso- and metathoracic segments,
crucial for flight. These wings consist of a double-layered membrane supported by a network of veins,
which provide both structural integrity and pathways for nerves, tracheae, and hemolymph.

The arrangement of these veins, known as venation, is diverse and plays a significant role in
wing function and insect classification. Basic longitudinal veins include the costa, subcosta, radius
(with radial sector), media, cubitus, and anal veins, which may branch and are connected by cross-veins
(radial, radio-medial, medio-cubital, and humeral), forming distinct cells named after the vein bordering
them anteriorly. Veins can be concave or convex, influencing wing folding patterns.

Specialized features like the pterostigma, a heavier pigmented area on the leading edge in some
groups, contribute to flight control and stability. Wing structure is optimized for managing aerodynamic
forces, with rigid areas (veins) and flexible areas allowing for shape changes during flight and folding
when at rest. Venation patterns can vary significantly, from reduced forms in tiny insects to increased
complexity with accessory or intercalary veins and reticulated networks in larger insects. Additionally,
specialized coverings like microtrichia, macrotrichia, scales (as in Lepidoptera), and patterns contribute
to wing function and appearance. (Chapman, 1998)

9
Fig 7: Wing venation of insect

3 A brief over view of aerodynamics:


Insect flight relies on aerodynamic principles, particularly the generation of lift. Insect wings,
acting as airfoils, create lift through a combination of their shape and angle of attack relative to the
oncoming airflow. The curved upper surface (camber) and rounded leading edge encourage faster
airflow over the top of the wing, a phenomenon directly related to Bernoulli's principle. This principle
states that for an inviscid, incompressible fluid, an increase in flow speed occurs simultaneously with a
decrease in pressure. Consequently, the faster airflow over the curved upper surface results in lower air
pressure compared to the underside of the wing. This pressure differential generates an upward force,
lift, perpendicular to the direction of motion. Simultaneously, a drag force acts parallel to the relative
wind, opposing the wing's movement. The vector sum of lift and drag produces the resultant force. The
center of pressure, the point where the resultant force acts, is crucial for wing stability. (Anon, 2025c)
Understanding these aerodynamic principles, and especially the application of Bernoulli's
principle to explain the pressure difference crucial for lift, is essential for analyzing insect flight
mechanics and the diverse adaptations of their wings.

4 Aerodynamics of insect wings:


Flapping is the exclusive mode of flight for the great majority of insects, and besides enabling
forward flight; it allows many insects to hover or to fly backwards.
Hovering is commonly associated with a particular behavior such as feeding, defending territory,
mating, or oviposition, and is often observed before landing, enabling the insect to land precisely on a
particular spot. Gliding is used by some Odonata, Orthoptera, and Lepidoptera, and ranges from a pause
10
in wing movement lasting only a fraction of a second, to prolonged glides lasting many seconds. Locusts
are able to lock their forewings in an outstretched position, which may facilitate gliding, and it is
suggested that the inability of dragonflies to fold their wings is a secondary adaptation to gliding.
However, even in species that glide, flapping remains the most important mode of flight.

Fig 8: (A) Force Generation & Vortex Wake Formation During a Forward Flight Stroke of an Insect
Wing. (B). Phases of a Forward Flight Stroke in an Insect: Force Vectors & Vortex Structures.

Fig 9: Delayed Stall, Rotational Lift, and Wake Capture in Insect Flight

The aerodynamics of flapping differs in two important ways from the aerodynamics of gliding.
First, whereas the air tends to remain attached on a gliding wing, flowing smoothly over the aerofoil's
surface, air flowing over a flapping wing tends to become entrained in a swirling vortex bound to the
upper surface of the wing. This is known as a separated flow. Second, whereas the attached flow over a
gliding wing looks approximately similar from one moment to the next, the separated flow over a
flapping wing varies continually. This is known as an unsteady flow.

Although once a point of controversy, it is now widely accepted that insects make extensive use
of unsteady separated flow mechanisms to generate far greater aerodynamic forces than would be
possible with steady, or quasi-steady, attached flow.

11
4.1 Leading-edge vortices:

The most ubiquitous, and probably the most important, separated flow mechanism is the leading-
edge vortex. Whereas the flow over a fixed wing remains attached to the wing's surface at low angles of
attack and becomes stalled at high angles of attack, the flow over a flapping wing typically separates at
the leading edge and becomes entrained within a swirling vortex sitting on top of the wing (Fig. 9.25)
The dynamics of vortex formation are govern by a dimensionless number called the Strouhal number,
calculated as wingbeat frequency multiplied by peak-to-peak wing tip excursion divided by air speed.
Other parameters influence the dynamics, but for wings operating at high angles of attack, leading-edge
vortex formation is difficult to avoid when the Strouhal number is much above 0.2.

Accordingly, leading-edge vortices have been observed in most insects for which flow
visualizations are available, including Lepidoptera, Diptera and Odonata. The presence of a leading-
edge vortex on top of the wing results in a local reduction in pressure, which causes an upward-acting
suction force known as vortex lift.

Fig 10: Leading edge vortex at different angle of attacks.

In Orthoptera and Odonata, leading-edge vortex formation may be avoided, at least at times, by
appropriately modifying the angle of attack. This presumably reduces the aerodynamic forces, but may
serve to enhance aerodynamic efficiency. (Ellington, 1995)

4.1.1 Case study 1: Unsteady Aerodynamic Performance of Model Wings at Low Reynolds
Numbers. (Dickinson and Götz 1993)

Objectives:

 To quantify the time-dependent aerodynamic forces (lift and drag) on an insect wing during
rapid acceleration.
 To correlate these force measurements with simultaneous flow visualization to understand the
underlying aerodynamic mechanisms.

12
Materials:

The experiments utilized a 1 mm thick aluminium wing section with a 5 cm chord and a 15 cm
length. The wing profile featured a bluntly rounded leading edge and a sharply tapered trailing edge.
The working fluid was a 54% sucrose solution, the kinematic viscosity of which was measured daily due
to temperature variations, evaporation, and filling inaccuracies inherent in the 200 L aquarium. The
viscosity ranged from 0.22 to 0.25 stokes.

The test chamber was a 100 cm x 40 cm x 60 cm glass aquarium filled with the sucrose solution.
Two edge baffles, placed within 5 mm of the upper and lower edges of the wing profile, were employed
to minimize spanwise flow. A 0.2 mm steel sphere was used in conjunction with Stokes' Law to
determine the kinematic viscosity of the solution each day.

An Isel linear translator, driven by stepper motors, moved a force transducer and wing model,
while a second motor-controlled angle of attack. The transducer measured perpendicular and parallel
forces, used to calculate lift and drag. The wing was submerged in a sucrose solution within an
aquarium with baffles, and flow visualization was achieved using aluminium shavings, a light sheet, and
a mirror.

Fig 11: Apparatus for measuring aerodynamic forces on an accelerating wing.

Methods:

The forces acting on the wing model were measured using a two-dimensional force transducer.
This transducer consisted of orthogonal sets of parallel phosphor-bronze plates with foil strain gauges
(350 Ω, HBM) mounted on both sides and wired in a full bridge configuration. The force transducer was
designed to minimize cross-talk between the two measurement axes. The wing was fixed to the
underside of the force transducer using a 7.5 cm long, 8 mm diameter brass sting. The force transducer

13
and wing assembly were attached to the movable sled of an automated milling device (Isel), driven by
two computer-operated four-phase stepper motors.

Fig 12: Flow visualization of the model wing.

This setup allowed for precise control of the wing's translation and rotation within the sucrose
tank. In all experiments, the wing was moved impulsively from rest for a distance of 37.5 cm (7.5 chord
lengths). The transition from rest to the fixed velocity was achieved using a constant acceleration of 62.5
cm/s². Aluminium particles were suspended in the sucrose solution to visualize the flow patterns. A red-
light source, consisting of light pipes, illuminated a horizontal plane of the fluid, which was recorded by
a CCD camera (Philips).

The lift and drag signals from the force transducer were simultaneously recorded on the audio
channels of the video recording. Each experimental run consisted of 23 different angles of attack,
ranging from -9° to 90° in 4.5° increments. The Reynolds number was varied by adjusting the
translational speed of the wing from 4 to 12 cm/s. To account for inertial forces from the force
transducer and sting, force measurements were also conducted without the wing attached, and these
values were subsequently subtracted from the experimental data.

The analogue force signals were filtered online using a two-pole Butterworth low-pass filter with
a 5 Hz cut-off frequency. Following the inertial transients, the signals were digitized, averaged over four
trials for each angle of attack, and stored for further analysis. A second digital filtering step, with cut-off
frequencies ranging from 1 to 5 Hz depending on the wing's velocity, was also applied. Lift and drag
coefficients were calculated from the filtered force measurements.

14
Results:

Fig 13: Graph representing the variation of force coefficient with the variation in angle of attack.

A conventional polar plot of the aerodynamic force coefficients (C L and CD) shows that at angles
of attack below 13.5˚, there is little difference between the early and late performance of the wing
model. However, above 13.5˚, significant deviations are observed, with the late CL reaching a plateau of
1.5 by 36˚, while the early values exceed 2.0. A broad lift plateau is observed between 27˚ and 54˚, but
the slopes of the polar diagrams differ within this range: the early C L continues to increase gradually,
while the late values show a slight decline. Beyond 54˚, both the early and late curves exhibit a sharp
drop in CL.

The ratio of lift to drag coefficients (CL/CD) is also compared at early and late stages of
translation. For angles above 13.5˚, the airfoil performs better early in translation, whereas, at lower
angles (0˚ to 13.5˚), the ratio is higher later in translation due to the Wagner effect, which causes a delay
in lift generation. As suggested by Vogel (1981), the absolute production of lift, rather than the lift-to-
drag ratio, may be the more important performance metric at low Reynolds numbers, meaning the
greatest benefit of unsteady flow comes from the significant lift augmentation at higher angles of attack,
rather than the increase in the lift-to-drag ratio.

Comparing the measured CL data with thin airfoil theory, it is clear that the theory overestimates
CL. The measured value of dCL/dα at low angles of attack is 4.5, compared to the theoretical value of 2π.
Additionally, a small but consistent "stutter" is observed in the late C L growth between 9˚ and 13.5˚.
This phenomenon, reminiscent of steady-state thin airfoil stall, likely reflects the transition from a stable
short separation bubble to a longer, less stable one.

15
Fig 14: Graph showing the variation of force coefficient with the time.

Unlike steady-state thin airfoil stall, however, the performance of these flat-plate models at low
Reynolds numbers involves unsteady behavior, with the separation bubble quickly growing into a
leading-edge vortex, contributing to vortex shedding patterns typical of a von Karman street.

4.1.2 Case study 2: Unsteady Aerodynamic Performance of Model Wings at Low Reynolds
Numbers. (Lentink and Dickinson, 2009)

Objectives:

 To understand the physical mechanisms responsible for the stability of leading-edge vortices
(LEVs) on flapping fly wings.
 To determine which of the three accelerations (angular, centripetal, and Coriolis) mediate LEV
stability.

Materials and Methods:

To investigate the aerodynamic mechanisms responsible for the stability of leading-edge vortices
(LEVs) on flapping insect wings, a dynamically scaled model of a Drosophila melanogaster wing was
constructed. This model was fabricated from a 2.0mm thick sheet of clear acrylic, ensuring both rigidity
and optical clarity for flow visualization. The wing was designed with a single wingspan (bs) of 0.187
meters and a surface area (S) of 0.0167 square meters, resulting in a mean chord width (c) of S/bs. It's
important to note that all wing parameters refer to a single wing, not a bilateral pair.

The wing was then affixed to a custom-built force transducer, which was connected in series
with a 3-degree-of-freedom actuator. This actuator, in turn, was attached to a translating robot arm. The
entire assembly was immersed in a large, transparent tank measuring 1m x 1m x 2m. This tank was
16
filled with one of three fluids to achieve varying Reynolds numbers (Re): thick mineral oil (ρ=840 kg
m–3; ν=140 x 10–6 m2 s–1) for Re=110, thin mineral oil (ρ=830 kg m–3; ν=11.0 x 10–6 m2 s–1) for
Re=1400, and deionized water (ρ=998 kg m–3; ν=1.004 x 10–6 m2 s–1) for Re=14,000. The Reynolds
number was calculated as Re = (c * Ug) / ν, where c is the average chord length, Ug is the average
velocity at the radius of gyration (Rg), and ν is the kinematic viscosity.

For the flapping motion experiments, the wing was programmed to follow a sinusoidal pattern
for stroke position and a smoothed trapezoidal pattern for the angle of attack, mimicking the kinematics
observed in freely hovering fruit flies. A stroke amplitude of 70° was used, based on previous kinematic
studies. While real insect wing kinematics include features like advanced wing rotation and a U-shaped
stroke plane, these were intentionally omitted to isolate the effects of the fundamental flapping motions.
The geometric angle of attack (α0) was systematically varied from 0° to 90° in 4.5° increments. The
wing was flapped for six complete periods at average frequencies of 0.22 Hz (Re=110), 0.23 Hz
(Re=1400), and 0.20 Hz (Re=14,000).

In addition to flapping, experiments were conducted with revolving and translating wing
motions. These unidirectional motions consisted of a constant velocity stroke with constant acceleration
and deceleration phases at the beginning and end, each lasting 10% of the stroke duration. The angle of
attack was again varied from 0° to 90° in 4.5° steps. The revolving wing swept an arc of 320°, while the
translating wing moved over a distance equivalent to the revolving wing's travel at the radius of
gyration.

To investigate the effect of the Rossby number (Ro) on LEV stability, the robot arm's length was
adjusted to create a range of Ro values at a constant Reynolds number (Re=1400). Specifically, the arm
was elongated by factors of 1.27 and 1.53, resulting in Ro values of 3.6 and 4.4, respectively.
Unidirectional and reciprocating translating motions, corresponding to Ro=∞, were achieved by setting
the stroke amplitude to zero and moving the robot arm's stage. Stroke amplitudes for the Ro=3.6, 4.4,
and ∞ cases were calculated to maintain a constant actuator disc area.

Flow visualization was conducted by releasing small air bubbles near the leading and trailing
edges of the wing in the oil-filled tank (Re=110 and 1400). Air was supplied through a thin tube glued
flush to the wing edge, with holes created using insect pins. The buoyancy of the bubbles meant they
didn't perfectly trace streamlines, but their movement provided a clear indication of vortex formation
and flow direction, particularly in strong vortices with low-pressure cores.

17
Lift and drag forces were measured using the custom-built force sensor. The data was down-
sampled to 300 Hz and filtered using a zero-phase delay low-pass 4-pole digital Butterworth filter, with
a cutoff frequency corresponding to a distance traveled of 0.3 chord lengths at Rg. This filtering
minimized noise while preserving relevant flow features. Force data was averaged over multiple cycles
or stroke periods, and force coefficients (CL and CD) were calculated based on dynamic pressure, using
a blade element method to normalize forces for comparison across different experimental conditions.

Results:

The stable LEVs were consistently observed on revolving wings, regardless of reciprocation,
while translating wings exhibited unstable LEVs. This highlights the crucial role of low Rossby number,
indicative of rotational motion, in LEV stabilization. While Reynolds number (Re) did not affect LEV
stability within the tested range (Re=110 and 1400), it did influence LEV integrity, with spiral bursting
observed at Re=1400 on revolving wings. Reciprocating motion, characterized by a finite A*, did not
independently stabilize LEVs, but at low stroke amplitudes, it helped maintain the LEV's proximity to
the wing during shedding and reformation. These findings, summarized through illustrative cartoons,
demonstrate that a propeller-like revolving motion is essential for stable LEV formation, while
translational motion alone is insufficient.

Fig 15: (A) LEVs are unstable on swept wings, shown for 40° sweep at α=36° and Re=1400. (B) The
LEV is also unstable on the same wing without sweep. (C) Revolving the same wing results in a LEV
that remains stably attached, shown for clarity at Re=110 and s=8c, where s is the distance traveled in
chord lengths.

18
4.2 Rotational lift:

The aerodynamic forces that act upon a translating wing are modified if the wing rotates about a
spanwise axis. The axis of rotation of an insect's wit is located anteriorly, so rotating the leading edge
upward about this axis causes an increase in the wing's effective angle of attack at the point where this
should be measured, three-quarters of the way back from the leading edge.

The aerodynamic forces acting on a wing increase with increasing angle of attack, so the
expected effect of rotating the wind leading-edge upward, as in supination, is to increase the
aerodynamic forces. This is a quasi-steady effect of rotation, which amounts in essence to a proper
accounting of the incident flow. The unsteady effects of wing rotation are less well known for the large
angular amplitudes associated with insect flight, but rotating a wing leading-edge upward delays the
onset of stall and thereby extends the production of useful aerodynamic lift to higher angles of attack.
This unsteady effect is known as the Kramer effect, and may be responsible for the transient lift
enhancement that is observed during wing rotation in insect flight. (Sane, 2003)

Fig 16: Visual Representation of Rotational Lift Dynamics

4.3 Wing-wake interactions:

Wing-wake interactions are likely to be of particular importance in hovering flight, where a wing
has the possibility of re-encountering the wake that it left behind on the previous stroke. Wing-wake
interactions are also important in functionally four-winged insects, where the hindwing operates in the
wake of the forewing. Such interactions are probably responsible for the changes in aerodynamic forces
that result from altering the timing of wing rotation with respect to stroke reversal in hovering Diptera,
and from altering the phasing of the forewing and hindwing pairs in Orthoptera and Odonata. Wing-
wake interactions are the least well understood of the various unsteady aerodynamic mechanisms of
insect flight, because the effect of the interaction depends upon the fine details of the movement of the

19
wing and of the wake it leaves behind. The effects of wing-wake interaction may therefore be quite
specific to the system being considered.

Fig 17: Visual Representation of Rotational Lift and Wake Capture Dynamics.

4.4 Clap and fling mechanism:

The first unsteady aerodynamic mechanism to be described in insects was the "clap and fling."
This refers to the clapping together and subsequent flinging apart of the wings at the top of the stroke.
This mechanism is best known from the chalcid wasp Encarsia (Hymenoptera), which has a wingspan
of about 1.3 mm and a wing beat frequency of about 400 Hz. Large Orthoptera and Lepidoptera have
also been observed to clap their wings dorsally, so the mechanism is not restricted to small insects.
However, it is unclear how important the clap and fling mechanism is under natural conditions. For
example, Drosophila are commonly observed to clap their wings dorsally in tethered flight, but rarely
do so in free flight. This is presumably an artefact of the large stroke amplitudes that are associated with
tethering.

The aerodynamics of the clap and fling are complicated, but are best understood as an example
of wing-wing interaction. As the wings clap together, they squeeze out a jet of air between them, which
the insect can use to augment thrust. Some insects may enhance their manoeuvrability by redirecting this
jet of air. The more important part of the mechanism, however, is the fling. As the wings pronate, their
leading edges are flung apart, while the trailing edges remain in contact. This allows the generation of
rotational lift, the growth of which is enhanced by suction of air into the expanding gap between the
wings, and by a favourable aerodynamic interaction resulting from the proximity of the wings and their
associated vortex systems. (Chapman, 1998)

20
Fig 18: The Dynamics of Clap and Fling: A Step-by-Step Breakdown of Lift Production

4.4.1 Case study 3: Quick Estimates of Flight Fitness in Hovering Animals, Including Novel
Mechanisms for Lift Production. (Weis-Fogh, 1973)
Materials and method:

Fig 19: Free flight of a very small chalcid wasp

The free flight of a very small chalcid wasp Encarsia formosa has been analysed by means of
slow-motion films. At this low Reynolds number (10–20), the high lift coefficient of 2 or 3 is not

21
possible with steady-state aerodynamics and the wasp must depend almost entirely on non-steady
flow patterns.

Results and discussion:

The wings of Encarsia are moved almost horizontally during hovering, the body being
vertical, and there are three unusual phases in the wing stroke: the clap, the fling and the flip. In the
clap the wings are brought together at the top of the morphological upstroke. In the fling, which is a
pronation at the beginning of the morphological downstroke, the opposed wings are flung open like a
book, hinging about their posterior margins. In the flip, which is a supination at the beginning of the
morphological upstroke, the wings are rapidly twisted through about 180°.

Fig 20: The Weis-Fogh Clap-Fling Mechanism in Insect Flight

The fling is a hitherto undescribed mechanism for creating lift and for setting up the
appropriate circulation over the wing in anticipation of the downstroke. In the case of Encarsia the
calculated and observed wing velocities at which lift equals body weight agree, and lift is produced

22
almost instantaneously from the beginning of the downstroke and without any Wagner effect. The
fling mechanism seems to be involved in the normal flight of butterflies and possibly
of Drosophila and other small insects. Dimensional and other considerations show that it could be a
useful mechanism in birds and bats during take-off and in emergencies.

The flip is also believed to be a means of setting up an appropriate circulation around the
wing, which has hitherto escaped attention; but its operation is less well understood. It is not confined
to Encarsia but operates in other insects, not only at the beginning of the upstroke (supination) but
also at the beginning of the down stroke where a flip (pronation) replaces the clap and fling
of Encarsia.

A study of freely flying hover-flies strongly indicates that the Syrphinae (and Odonata)
depend almost entirely upon the flip mechanism when hovering. In the case of these insects a
transient circulation is presumed to be set up before the translation of the wing through the air, by the
rapid pronation (or supination) which affects the stiff anterior margin before the soft posterior
portions of the wing. In the flip mechanism vortices of opposite sense must be shed, and a Wagner
effect must be present.

In some hovering insects the wing twistings occur so rapidly that the speed of propagation of
the elastic torsional wave from base to tip plays a significant role and appears to introduce beneficial
effects.

Non-steady periods, particularly flip effects, are present in all flapping animals and they
will modify and become superimposed upon the steady-state pattern as described by the
mathematical model presented here. However, the accumulated evidence indicates that the
majority of hovering animals conform reasonably well with that model.

5 Biomimicry of insect wings:


Nature’s fliers have inspired scientists to produce their replicas through creative engineering
designs. Earlier the birds were the primary source of inspiration for the aircraft designs. The last
two decades were the most active period on this front, and many principles of insect flight such as
leading edge vortexgeneration, rotational circulation, clap-and-fling, and control moment
generation, were realized in various robotic systems. (Phan and Park, 2020)

23
Fig 21: Bio-Inspired Robotics: Exploring Wing Variations in Insect Mimics

5.1 Robobees:
Inspired by the biology of a bee, researchers at the Wyss Institute are developing RoboBees,
manmade systems that could perform myriad roles in agriculture or disaster relief. A RoboBee
measures about half the size of a paper clip, weighs less than one-tenth of a gram, and flies using
“artificial muscles” compromised of materials that contract when a voltage is applied. Additional
modifications allow some models of RoboBee to transition from swimming underwater to flying, as
well as “perch” on surfaces using static electricity. (Anon, 2025b)

The masterminding of the RoboBee was motivated by the idea to develop autonomous
micro-aerial vehicles capable of self-contained, self-directed flight and of achieving coordinated
behavior in large groups. To that end, the RoboBee development is broadly divided into three main
components: the Body, Brain, and Colony. Body development consists of constructing robotic
insects able to fly on their own with the help of a compact and seamlessly integrated power source;
brain development is concerned with “smart” sensors and control electronics that mimic the eyes and

24
antennae of a bee, and can sense and respond dynamically to the environment; the Colony’s focus is
about coordinating the behavior of many independent robots so they act as an effective unit. (Wood
et al, 2013)

Fig 22: Advancements in Micro Air Vehicle Technology: RoboBees

5.2 Marsbees:
Mars exploration has received significant interest from academia, industry, government, and
the general public. Despite continued interest, flying on Mars remains challenging, mainly due to the
ultra-thin Martian atmospheric density. Although the gravitational acceleration on Mars is 38 percent
of Earth's 9.8 meters per second squared, the Martian atmospheric density is only 1.3 percent of the
air density on Earth. The aerodynamic forces are proportional to the ambient fluid density.
Therefore, flying near the surface of Mars has been considered nearly impossible. The proposed
mission architecture (Fig.21) consists of a Mars rover (already existing) that serves as a mobile base
for Marsbees - a deployable swarm of small bio-inspired flapping wing vehicles. In one ConOps
scenario, each Marsbee would carry an integrated stereographic video camera and the swarm could
construct a 3D topographic map of the local surface for rover path planning. These flying scouts
would provide a "third-dimension" to the rover capabilities.
In other scenarios, each part of the swarm of Marsbees could carry pressure and temperature
sensors for atmospheric sampling, or small spectral analyzers for identification of mineral
outcroppings. In each scenario, the rover acts as a recharging and deployment/return station and data
and communication hub. Human exploration of Mars is one of the major objectives of NASA and
commercial entities such as SpaceX and Boeing. The identified innovations unique to the bio-

25
inspired flapping Marsbee provide viable multi-mode flying mobility for Martian atmospheric and
terrain exploration. A swarm of Marsbees provides an enhanced reconfigurable Mars exploration
system that is resilient to individual component failures.

Fig 23: Mars Exploration with Robobees: A Collaborative Approach

These Marsbees can carry sensors and wireless communication devices in combination with
a Mars rover and helicopters. These enhanced sensing and information gathering abilities can
contribute to the following NASA Mars mission objectives: i) "Determine the habitability of an
environment", ii) "Obtain surface weather measurements to validate global atmospheric models",
and iii) "Prepare for human exploration on Mars." Various commercial entities, e.g. SpaceX and
Boeing, are investing in technologies to transport humans to Mars. (Kang et al, 2019)

5.3 KU Beetle-S:
The KU Beetle-S represents a pioneering effort in bio-inspired robotics, drawing direct
inspiration from the Allomyrina dichotoma beetle's remarkable flight capabilities. This flying robot
is specifically designed to replicate the beetle's efficient flight despite its relatively large size and
high wing loading, a characteristic significantly exceeding that of typical insects. Utilizing flapping
wings to generate both lift and control forces, the KU Beetle-S incorporates an active control system
with micro actuators, enabling precise adjustments to the wing stroke-plane and twist for pitch, roll,
and yaw control. Notably, yaw control is achieved through asymmetric wing twist, demonstrating a
sophisticated approach to maneuverability. The robot's design, as illustrated in the slide, emphasizes
the integration of these control mechanisms within a compact and functional structure, showcasing

26
the potential for insect-inspired robotics to achieve stable and efficient flight. (Phan et al, 2017)

Figure 24: Beetle Wing Deployment: Natural and Robotic Mimicry

5.4 DelFly:

Fig 25: DelFly: Close-up View of a Flapping-Wing Drone


The DelFly, a camera-equipped Micro Air Vehicle (MAV), represents a significant advancement
in bio-inspired robotics, developed at the Micro Air Vehicle Lab of the Delft University of Technology.
This fully controllable, flapping-wing robot draws its innovative design directly from the intricate wing
movements of fruit flies, showcasing the potential of biomimicry in engineering. (Anon, 2025a)

27
Notably, the DelFly's tailless configuration reveals a crucial aspect of fruit fly flight: the
utilization of torque coupling during rapid banked turns. By replicating this mechanism, the DelFly
achieves enhanced maneuverability, particularly in complex environments.
The project underscores the importance of studying biological systems to develop
sophisticated robotic solutions, demonstrating the successful integration of biological principles with
advanced engineering. (De Croon et al, 2016)

Fig 26: Quantifying Flight Maneuvers: A Comparative Study

5.4.1 Case study 4: Clap-and-fling mechanism in a hovering insect-like two-winged flapping-


wing micro air vehicle. (Phan et al, 2016)

Objectives:
 To quantify the force enhancement achieved by the clap-and-fling mechanism in a hovering
insect-like micro air vehicle.

 To validate a computational fluid dynamics (CFD) model against experimental measurements


of the forces generated by the clap-and-fling motion.

Material and methods:


The experiment utilized both numerical and experimental methodologies to investigate the
force generation of a two-winged flapping mechanism designed to mimic the wings of a rhinoceros
beetle.

The flapping mechanism, designed using a combination of a four-bar linkage and pulley-string
system, was constructed from carbon/epoxy panels and driven by a DC motor. Wings, with a wingspan
of 70mm and a mean chord of 25mm, were crafted with carbon strip veins and polyethylene

28
terephthalate membranes. Three synchronized high-speed cameras, operating at 2000 frames per second,
captured the movement of markers on the wings. This data was then used to calculate the flapping angle
and the rotation angles, further analyzed using computational fluid dynamics (CFD).

A 6-axis load cell was employed to measure the forces generated by the flapping wings at a
frequency of 20 Hz. The flapping-wing system was operated for 100 cycles, and the data was averaged
over 10 experiments. A separate experiment was conducted to examine the forces without the clap-and-
fling effect, using an extended distance between the wings. The flapping motion of the wings was
captured using three synchronized high-speed cameras at 2000 frames per second. The time history of
the flapping angle of the leading edge was then obtained from the coordinates of the dots during the
flapping motion. The flapping amplitude was measured to be approximately 192°.

A computational fluid dynamics (CFD) model was constructed based on the three-dimensional
wing kinematics to estimate the force generation. The computed forces were then validated by the
measured forces using a 6-axis load cell.

Results:

The average vertical forces (Fz) obtained by the numerical simulation are about 3.2% and 7.5%
larger than the measured vertical forces for the cases with and without the clap-and-fling effect,
respectively. This proved that the numerical simulation could be used to properly estimate the average
forces generated by the FW-MAV.

Table 1. Contribution of the clap and fling to the average vertical force (Fz) at each stroke
period.

29
Table 2. Contribution of the clap and fling to the average force in the η-direction or average horizontal
drag force (Fη) at each stroke period.

The horizontal forces were less than 3 mN and less than 3% of the vertical forces. The simulated
horizontal forces are approximately 17.2% and 7.7% smaller than the measured horizontal forces for the
cases with and without the clap-and-fling effect, respectively. This was reasonable given that the minor
asymmetric flapping motion generated a small amount of the horizontal force. The results from the CFD
simulation indicated that the average horizontal forces in a complete cycle generated by the flapping-
wing mechanisms with and without the clap-and-fling effect were the same. However, for the
measurement, the difference between the average horizontal forces in the two cases was 11.5%.

In reality, this increment was a result of the asymmetric contribution to the average horizontal
force of two clap-and-fling mechanisms at the dorsal and ventral stroke reversals, which produced
forces in the opposite direction, and was not a result of the clap-and-fling effect. This difference may not
significantly affect the horizontal force in the y-direction as the average horizontal force in the y-
direction was small and close to the resolution of the load cell.

The average horizontal drag force (Fη) generated by the flapping-wing system with the clap-and-
fling effect (5.1 mN) was 41.7% higher than that of the system with a minimized clap-and-fling effect
(3.6 mN). This enhancement was not caused by the clap-and-fling effect because of the opposite
motions of the wings during the downstroke and upstroke, and the presence of the clap-and-fling
mechanisms at both stroke reversals. Instead, in a manner similar to the CFD situation for the horizontal
force in the y-direction, the difference was caused by the asymmetric contribution of the clap-and-fling
effects at the dorsal and ventral stroke reversals.

30
Conclusion:

The clap-and-fling mechanism in a two-winged insect-like MAV boosts vertical and horizontal
forces, with fling contributing more to lift, but its effectiveness declines beyond a critical wing spacing.
These insect-mimicking structures can also serve as tools to study and understand the operational
mechanics of insect flight.

5.5 Dragonfly inspired turbine blade design:

Dragonflies have a unique and superior flight performance compared to most other insect and
bird species. They can glide for extended periods of time as well as hover and change direction swiftly.
The dragonflies are mainly notable for their agility and impressive aerodynamics associated with low
wind speeds. Therefore, since this study's main focus is improving the aerodynamic performance of
wind turbines at low-wind speed conditions, it is of utmost interest to explore the dragonfly design for
wind turbines that are subject to wind speeds not exceeding 4 m/s. Their fast-flying ability is due to
pairs of individually regulated forewings and hind wings. Veins and membranes, a common component
of Nano composite materials, make up the majority of the wings of dragonflies. Within the wing, the
membranes and veins have a complicated architecture that results in whole-wing properties. Figure 25,
shows a real dragonfly wing.

Fig 27: Dragonfly inspired blade design geometric characteristics: (a) Geometry of wing as found in
nature, (b) airfoil profile, and (c) bio inspired designed turbine.

31
Another interesting trait characterizing the wing geometry of the dragonflies is the wing cross
section. The surfaces of dragonfly wings are not smooth or simply cambered as is the case for most
other birds. The wing's cross-sectional camber of dragonflies has a well-defined thin corrugated pattern.
This design is crucial to the ultra-lightweight and stability characterizing the dragonflies wings. The
same cross-section, obtained by connecting the maxima and minima of the corrugated cross-section
pattern, as well as similar geometric attributes were used in the CAD development of the dragonfly
bioinspired design depicted in figure 25 (c) The golden eagle and dragon fly blade shapes are
innovatively being used for small-scale turbine blades. (Yossri et al., 2023)

Advantages of insect biomimicry:


The field of bio-mimicry, specifically in the context of insect flight, offers numerous advantages.
Designs inspired by insects demonstrate enhanced energy efficiency, leading to reduced power
consumption. They also exhibit superior maneuverability, enabling agile hovering and rapid directional
shifts. Furthermore, these designs provide valuable aerodynamic insights, improving performance in
turbulent conditions. The development of insect-inspired robots promotes eco-friendly and sustainable
engineering practices, harmonizing technology with nature. Lastly, these bio-mimetic systems serve as a
powerful tool to study and understand the intricate operational mechanics of insect flight, furthering
our knowledge of both biology and engineering.

Challenges in Bio-mimicry of Insect Flight:

1. Complex Mechanics: Replicating the intricate and highly coordinated wing movements of
insects is a significant engineering hurdle. The precise control required for flapping, twisting,
and other aerodynamic manipulations is difficult to achieve with current technology.
2. Material Limits: Developing lightweight yet durable materials that can withstand the stresses of
repeated flapping and maneuvering is challenging and often costly. Materials must be strong
enough to maintain structural integrity while being light enough to enable flight.
3. Energy Sources: Finding compact and efficient power sources that can provide sufficient
energy for sustained flight in small-scale robots is a major limitation. Existing battery technology
often falls short in terms of energy density and power output.
4. Environmental Sensitivity: Adapting insect-inspired designs to function effectively in diverse
and unpredictable environments is difficult. Insects have evolved to thrive in specific niches, but
replicating this adaptability in robots is a complex task.

32
6 Conclusion:
Insects are extraordinary creatures, particularly when it comes to their aerodynamic abilities and
flight control. Their wings, with their unique shape, movement, and surface structures, provide
invaluable insights into the principles of efficient and versatile flight. For instance, insects can adjust
their wing movements with remarkable precision, allowing them to perform complex maneuvers like
hovering, rapid turns, and even flying backward. These characteristics are not only impressive in nature
but also present a significant opportunity for engineering and design, particularly in the development of
flying machines. By studying the intricate mechanics of insect wings, engineers and designers can learn
how to enhance the performance of aerial vehicles, making them more agile and efficient. This has led
to the advancement of technologies such as micro-aerial vehicles and robotic flyers, which mimic the
flight patterns and structures of insects.

The aerodynamics of insect wings offers significant scope for future research, especially in
understanding how wing shapes and surface structures affect flight kinetics. These factors play a crucial
role in how insects achieve high levels of maneuverability, lift, and stability, and further exploration
could lead to breakthroughs in aerodynamics that can be applied to both nature-inspired and
conventional flight systems. However, replicating the complex, flexible wing movements of insects
remains a significant challenge. Insects use highly specialized, fine-tuned movements that allow for
maximum efficiency, and accurately reproducing these in a mechanical system requires overcoming
substantial engineering hurdles. Additionally, developing lightweight materials that can withstand the
demands of flight, while also being durable and cost-effective, remains an ongoing challenge. To
address these issues, a robust digitization of insect flight dynamics is essential, which could be
accomplished through collaborative efforts between entomologists and computer scientists.

Furthermore, another difficulty lies in creating compact and efficient energy sources to power
such machines. Nevertheless, the potential applications for insect-inspired designs are vast. These could
range from creating sustainable technologies, such as environmentally friendly drones for monitoring
ecosystems, to advancing exploration tools capable of reaching areas that are too difficult or dangerous
for traditional aircraft. With continued research, the field promises to unlock new breakthroughs that can
not only push the boundaries of engineering but also deepen our understanding of the fascinating world
of insects.

33
7 Selected References:

Anonymous, 2025 a. Available at: https://www.delfly.nl/home/ [Accessed 28 January 2025].


Anonymous, 2025 b. Available at: https://wyss.harvard.edu/technology/autonomous-flying-microrobots-
robobees/ [Accessed 28 January 2025].
Anonymous, 2025 c. Available at: https://www.nasa.gov/learning-resources/for-kids-and-students/
[Accessed 28 January 2025].
Chapman, R. F., 1998. The insects: structure and function. Cambridge university press.
Dickinson, M. H. and Götz, K. G., 1993. Unsteady aerodynamic performance of model wings at low
Reynolds numbers. J. Exp. Biol., 174(1): 45-64.
De Croon, G. C. H. E., Perçin, M., Remes, B., Ruijsink, R. and De Wagter, C., 2016. The
delfly. Dordrecht: Springer Netherlands, 10: 978-94.
Ellington, C., 1995, January. Unsteady aerodynamics of insect flight. In Symposia of the society for
experimental biology (Vol. 49: 109-129).
Hasan, J., Roy, A., Chatterjee, K. and Yarlagadda, P. K., 2019. Mimicking insect wings: The roadmap
to bioinspiration. ACS Biomater. Sci. Eng., 5(7): 3139-3160.
Johnson, J., 2022. Application of biomimicry in product design: bridging between classical and
biomimetic approaches.
Kang, C. K., Fahimi, F., Griffin, R., Landrum, D. B., Mesmer, B., Zhang, G., Lee, T., Aono, H., Pohly,
J., McCain, J. and Sridhar, M., 2019. Marsbee-Swarm of Flapping Wing Flyers for Enhanced
Mars Exploration: NASA Innovative Advanced Concepts (NIAC)-Phase I (HQ-E-DAA-
TN67472).
Lentink, D. and Dickinson, M. H., 2009. Rotational accelerations stabilize leading edge vortices on
revolving fly wings. J. Exp. Biol., 212(16): 2705-2719.
Phan, H. V., Au, T. K. L. and Park, H. C., 2016. Clap-and-fling mechanism in a hovering insect-like
two-winged flapping-wing micro air vehicle. R. Soc. Open Sci., 3(12): 160746.
Phan, H. V., Kang, T. and Park, H. C., 2017, February. Controlled hovering flight of an insect-like
tailless flapping-wing micro air vehicle. In 2017 IEEE International Conference on
Mechatronics (ICM) (74-78).
Phan, H. V. and Park, H. C., 2020. Mimicking nature’s flyers: a review of insect-inspired flying
robots. Curr. Opin. Insect Sci., 42: 70-75.
Sane, S. P., 2003. The aerodynamics of insect flight. J. Exp. Biol., 206(23): 4191-4208.

34
Weis-Fogh, T., 1973. Quick estimates of flight fitness in hovering animals, including novel mechanisms
for lift production. J. Exp. Biol., 59(1): 169-230.
Wood, R., Nagpal, R. and Wei, G. Y., 2013. Flight of the robobees. Scientific American, 308(3): 60-
65.
Yossri, W., Ayed, S. B. and Abdelkefi, A., 2023. Evaluation of the efficiency of bioinspired blade
designs for low-speed small-scale wind turbines with the presence of inflow turbulence
effects. Energy, 273:127210.

35
37

You might also like