Conference PaperPDF Available

Abstract

Historically, aviation has benefited from the study of avian flight. Effective biologically informed design requires knowledge of quantifiable biological principles that researchers can integrate into the development of future aircraft. The study of avian flight is especially applicable to the design of uncrewed aerial vehicles (UAVs) due to their similarity in size, speed, and flight environments. However, extracting relevant biological principles is challenging due to the number and variety of bird species, complicating the selection of a species that would best inspire a specific design objective. Proper selection requires information on both overall wing shape and airfoil properties. Additionally, each bird species experiences different natural selective pressures that constrain the evolution of their morphology over time depending on their flight style, life history, and ecological niche. To study the aerodynamic traits of various birds, we developed a systematic method to scan bird wings, extract their airfoils, and quantify their aerodynamic performance. This tool allowed us to test our hypothesis that birds that preferentially perform gliding and soaring flight have airfoil aerodynamic properties that are unique from flapping species. To test this hypothesis, we scanned prepared bird wings from multiple species. We adapted a slicer algorithm from 3D printing software to extract the airfoil shape at known locations along the span. Next, we used XFOIL to estimate the aerodynamic properties at a Reynolds number of 1.5×10 5. We found a slight, but statistically insignificant, increase in the aerodynamic efficiency of airfoils from gliding species compared to airfoils from flapping species. Our results suggest that the morphology of avian airfoils may account for species' flight styles, although future work is required to build a larger dataset for a more robust model. Moreover, we found our highest-performing avian airfoils achieved a similar aerodynamic efficiency, if not higher, than that of human-engineered airfoils (FX63-137, S1223, and NACA 4415) at the analyzed Reynolds number. In all, our results show promise in bio-inspired airfoil design and ultimately advance the fundamental understanding of bird airfoil aerodynamic characteristics.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
Analysis of bird wing airfoil aerodynamic
efficiency
Rowan Glenn1, Lucas B. Dahlke1, Andrew Engilis, Jr.2, Christina
Harvey3
University of California Davis, Davis, CA, 95616, USA
Historically, aviation has benefited from the study of avian flight. Effective biologically informed
design requires knowledge of quantifiable biological principles that researchers can integrate into
the development of future aircraft. The study of avian flight is especially applicable to the design
of uncrewed aerial vehicles (UAVs) due to their similarity in size, speed, and flight environments.
However, extracting relevant biological principles is challenging due to the number and variety
of bird species, complicating the selection of a species that would best inspire a specific design
objective. Proper selection requires information on both overall wing shape and airfoil
properties. Additionally, each bird species experiences different natural selective pressures that
constrain the evolution of their morphology over time depending on their flight style, life history,
and ecological niche. To study the aerodynamic traits of various birds, we developed a systematic
method to scan bird wings, extract their airfoils, and quantify their aerodynamic performance.
This tool allowed us to test our hypothesis that birds that preferentially perform gliding and
soaring flight have airfoil aerodynamic properties that are unique from flapping species. To test
this hypothesis, we scanned prepared bird wings from multiple species. We adapted a slicer
algorithm from 3D printing software to extract the airfoil shape at known locations along the
span. Next, we used XFOIL to estimate the aerodynamic properties at a Reynolds number of
1.5×105. We found a slight, but statistically insignificant, increase in the aerodynamic efficiency
of airfoils from gliding species compared to airfoils from flapping species. Our results suggest
that the morphology of avian airfoils may account for speciesflight styles, although future work
is required to build a larger dataset for a more robust model. Moreover, we found our highest-
performing avian airfoils achieved a similar aerodynamic efficiency, if not higher, than that of
human-engineered airfoils (FX63-137, S1223, and NACA 4415) at the analyzed Reynolds
number. In all, our results show promise in bio-inspired airfoil design and ultimately advance the
fundamental understanding of bird airfoil aerodynamic characteristics.
I. Introduction
By studying bird morphology, we can identify the most advantageous flight strategies that birds have
developed to drive the development of novel bio-informed aircraft [1]. Early studies of bird wing morphology used
prepared specimens to establish how properties like camber and wing twist affect wings’ lift and drag coefficients [2].
Studies on bird wing morphology have also identified how barn owls and eagles adjust to turbulent conditions [3, 4],
how gulls use their elbow angle to negotiate between flight efficiency and stability [5], and have found that the majority
of bird species can shift between more maneuverable and more stable configurations by morphing their wings [6].
Such studies have also expanded our understanding of how birds cope with rapid changes in the angle of attack [7],
which has led engineers to successfully implement flexible flaps on the surface of aircraft wings to improve
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
aerodynamic performance [8]. Additionally, a 2023 study scanned over 1,000 bird wings and found evidence that the
bird wing is composed of two modular units: the handwing and the armwing. [9]
Bird wings vary along their span because their skeletal structure extends only partially into the wingspan,
and near the tip, the airfoil is composed of only feathers (Fig. 1). It has been established that the interior of bird wings
have a more cambered airfoil than the wing tip [10]. Understanding the distribution of airfoils along a bird wing and
how this varies between species will be critical to establishing generalized models of bird flight. Due to the complexity
of bird wings and the lack of data on avian airfoils, previous studies have relied on assuming a “typical” wing profile
[11] or have used historical data that has been acquired for a few individuals of a few species [2,6,9]. However, the
aerodynamic role of each wing section has not yet been quantified. Often researchers must select bird airfoils based
on estimations and limited empirical data. Our method for analyzing bird wing morphology introduces a systematic
way to scan and analyze airfoil data from bird wings. This approach, and others like it [9], builds a more
comprehensive dataset of avian airfoil shapes.
Fig. 1 Approximate wing anatomy demonstrating the expected structural differences between proximal
(inboard) airfoils and distal (outboard) airfoils. Adapted from [1].
To build the airfoil database, we scanned 18 prepared bird wings, adapted a 3D printing slicer algorithm, and
normalized the data to extract the airfoil shape along each specimen’s wing. This standardized approach allows for
quick and accurate analysis of wing airfoil shapes across various individuals and species. Furthermore, by analyzing
our extracted airfoils in XFOIL [13], we quantified the aerodynamic characteristics of bird airfoils and determined
their maximum lift-to-drag ratio.
This dataset enabled us to compare the aerodynamic characteristics of various bird species based on their
primary flight style. We hypothesized that gliding and soaring birds face similar selective pressures that constrain their
airfoil shape and resulting aerodynamic properties. Despite different evolutionary histories, we predicted that if gliding
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
birds had similar maximum lift-to-drag ratios and geometries, there may exist some empirical ideal airfoil shapes for
efficient gliding flight. We expected these properties to differ compared to the wings of flapping birds. Flapping birds
regularly encounter highly unsteady aerodynamic forces, which may lead to different evolutionary pressures for their
airfoil shape and wing size. Indeed, recent work has found evidence that flapping flight aerodynamics drove the
evolution of the avian wing geometry [10].
II. Methodology
A. Wing Selection
Our method adapts existing structured-light 3D scanning technology to extract information about the airfoils
of various bird wings at a series of standardized locations. For this study, we scanned the geometry of prepared wing
specimens from nine species of birds. The species include prairie falcon (Falco mexicanus), turkey vulture (Cathartes
aura), red-tailed hawk (Buteo jamaicensis), red-shouldered hawk (Buteo lineatus), band-tailed pigeon (Patagioenas
fasciata), Laysan albatross (Phoebastria immutabilis), western gull (Larus occidentalis), yellow-billed magpie (Pica
nuttalli), and black scoter (Melanitta americana). These species were selected to cover a broad range of the avian
phylogeny and account for different flying styles in different clades of birds. We selected specimens from species that
either flap or glide as their modus operandi. Gliding species include prairie falcons, turkey vulture, both hawk species,
western gull, and Laysan albatross, and the flapping species include band-tailed pigeon, yellow-billed magpie, and
black scoter. These selected species span a wide range of avian mass. The smallest bird, the magpie, weighs
approximately 0.16 kg, whereas the largest bird, the Laysan albatross, weighs approximately 3.3 kg.
We collaborated with the UC Davis Museum of Wildlife and Fish Biology to obtain the wing specimens to
scan. All wings were previously prepared and follow a standardized style where the wing is preserved, flattened from
the elbow out, with the elbow joint, arm, phalanges, and flight feathers outstretched, similar to shapes observed by
birds in soaring flight. Our initial round of data collection involved scanning multiple wings for each of our nine
species, resulting in a total of 18 wing models. Three wing scans were not usable due to low scanning resolution and
precision, reducing our final total to 15 models.
B. Wing Scanning
We anchored the wings to a flat surface with a clamp to prevent wing movement during the scan. The wings
were clamped and scanned in two configurations. In the first configuration, we positioned the wings with the wing tip
pointing directly upwards and the wing root anchored. In the second configuration, we clamped the wing's leading
edge at the wrist, with the primary feathers pointing upwards. We covered the wings with infrared markers at a density
of approximately one marker per in2. The wings were scanned using a 3D infrared scanner (Revopoint POP2) with a
precision of 0.05 mm (0.002 in). Each wing was scanned up to six times to ensure that the complete geometry was
captured with the maximum possible resolution. These included scans to capture the geometry and details of the
leading and trailing edges of each wing, as well as detailed images of the dorsal and ventral wing faces.
C. Importing and Processing 3D Files
We used RevoScan 5’s point cloud editing capabilities to align these scans using common morphological
features as a reference. This allowed us to crop the clamps out of the scan while maintaining the accuracy of the scan.
These scans were converted into a mesh (.stl), a format commonly used in additive manufacturing. To prepare these
point clouds for aerodynamic analysis, we edited our scans to linearly interpolate across gaps in the scan, such as those
left by the infrared markers (Fig. 2). Additionally, we removed background objects that were picked up by the scanner.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
Fig. 2 Revopoint editing process. a) Initial point cloud, b) mesh conversion, and c) final edited scan of the albatross
specimen, WFB-17490.
Finally, we corrected minor meshing errors and standardized the orientation of all scans using the .stl editing
tool PrusaSlicer [14]. The total number of mesh triangles corrected by PrusaSlicer was minor, comprising less than
1% of the total triangle count of each wing. They resulted from computational artifacts generated during the conversion
of our models from point clouds to meshes. The wing meshes were oriented with the leading edge pointing vertically
upwards. The primary flight feathers, which roughly form a flat plane near the shoulder joint for all specimens, were
aligned horizontally. PrusaSlicer’s internal angle measurement capabilities provided angle guidelines every 5° while
scan orientation was taking place. By visually aligning the leading edge with these markers, we could ensure that the
scan was accurately aligned within 2.5°.
Fig. 3 Slicer algorithm results for the Laysan Albatross wing, specimen WFB-17490. a) With the wing
model and b) without the wing model.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
D. Slicer Algorithm Adaptation
We then imported our scans to a function adapted from an open-source slicing function [15] intended for
generating G-Code used in 3D printing. The function divides the .stl file into evenly spaced slices, determined by user
input. It does this by detecting mesh triangles that intersect with each given slicing plane and outputting these points
as a set of comma-separated values. The points can be graphed as a series of slices for user visualization (Fig. 3), and
the x-y coordinates for each airfoil are saved to an individual spreadsheet for future input into XFOIL. By adjusting
the distribution and number of slices, this function outputs data about the 2D airfoil shape at any specified location.
We sliced each wing at ten evenly spaced locations along the span between the root and top, creating sections that
were each 1/11th of the wingspan.
In the process of creating the final three-dimensional models of each wing, multiple scans from various angles
were combined to capture as much detail as possible. Inaccuracies in the alignment occasionally resulted in duplicate
surfaces within the interior of the wing model. As a result, we occasionally observed incidental points in the center of
the sliced airfoils, as demonstrated in Fig. 4. Because these points do not affect the wing's aerodynamic properties but
would cause errors in our XFOIL analysis, we removed all interior points. Additionally, noise was intermittently
present along the perimeter of the airfoils generated; these came in three varieties, as demonstrated in Fig. 5 below.
Fig. 4 Incidental points in airfoil interior. a) Before and b) after editing.
Fig. 5 No-Go Scenarios. a) Splits down trailing or leading edges, b) irregular airfoils,
or c) secondary bodies along the airfoil perimeter.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
In cases like Fig. 5a) and c), the noise along the leading edge made it impossible to determine an airfoil shape
that accurately represented bird wing geometry, so these airfoils were discarded. In cases like Fig. 5b), the slicer
generated a continuous geometry, but the airfoil was too irregular or jagged to be analyzed in XFOIL. These, along
with any cases where several geometries were created during slicing, were discarded, although future work should
explore these unique shapes.
E. Curve Fitting
We conducted the initial pass to clean up the data with a custom R code that read in the input airfoil
coordinates, aligned the airfoils, and normalized their chord length to one. The data was split at the leading and trailing
edges to extract the points that form the top and bottom surfaces. Then, a fourth-degree B-spline was fitted separately
to the top and bottom surfaces using knots at 1%, 5%, 25%, 50%, 75%, and 90% of the chord. To account for the large
shape change at the airfoil nose, a third constrained B-spline was fit to the front 2.5% of the chord, requiring the end
points to align with the top and bottom B-spline at 2.5% of the chord. To smooth the predicted data, we performed a
second fit on the predicted nose data, expanding to a constrained B-spline with endpoints at 5% of the chord. An
example of the output from this process can be seen in Fig. 6.
Fig. 6 The airfoils underwent a smoothing process. a) All ten albatross airfoils smoothed, rotated, and scaled.
b) The rotated and scaled raw data points in gray are overlaid with the final B-spline. This airfoil is the 11% span
airfoil on the albatross, specimen WFB-17490.
F. Aerodynamic Analysis (XFOIL)
XFOIL is a computational solver that predicts the aerodynamic characteristics of input airfoils in various
flight conditions for either viscous or inviscid flow conditions [16]. Prior to analysis, our geometry was further
modified in XFOIL to revise the number of panel nodes to avoid sharp angles that could cause separation bubbles
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
(locations where the flow separates and reattaches) and further add points to reduce any lingering sharp angles. We
began our set design routine by fixing the number of panel nodes to 500 in the panel editor. Afterward, in the geometry
editor, the maximum panel corner angle was targeted to 10° with five iterations using the arcline spline parameter
over the full length of the airfoil by specifying the interval [-0.1, 1.1], 10% beyond the chordline. This allowed us to
achieve a smoother leading edge while reducing further irregularities throughout the airfoil to increase our likelihood
of convergence.
For our analysis, we estimated the lift, drag, and pitching moment coefficients of our given airfoils under
viscous conditions for a range of angles of attack and a fixed Reynolds number. We varied the angles of attack from
-5 to 10° with a step size of 1° and fixed the Reynolds number to 1.5 x 105. This Reynolds number was selected as it
is a mid-range value used across bird flight; however, it is not necessarily the true flight condition for each species.
Future work should consider a range of Reynolds numbers. To increase the likelihood of convergence in XFOIL, we
used up to 500 iterations in the solver, but convergence is challenging for high angles of attack and complex
geometries. As applicable, un-converged angles were omitted from the analysis. XFOIL returns the airfoil sectional
lift and drag coefficients at each converged condition.
With the outputs from XFOIL, we estimated the aerodynamic efficiency as the maximum lift-to-drag ratio
(maximum Cl/Cd) predicted for each airfoil. This prediction used R to fit a linear model to the coefficient drag with a
quadratic function for the coefficient of lift to all airfoils that converged for a minimum of five angles of attack. We
verified that the linear models were capturing the shape of the XFOIL predictions by verifying that the adjusted R2
value was above 0.75. If the fit was below this threshold, we first would remove the data points above an 8° angle of
attack as approaching stall causes outliers in the theoretically expected quadratic curve. Two airfoils were eliminated
for converging for less than five angles (the yellow-billed magpie at 64% and the scoter at 9%). Ten airfoils still had
low adjusted R2 values due to non-quadratic curves after this adjustment and thus were removed from our analysis
(This includes: three gulls, two turkey vultures, two Laysan albatrosses, one prairie falcon, one scoter, and one red-
shouldered hawk). Future work is required to develop a methodology to evaluate the performance of these unique
avian-derived airfoils. From the extracted quadratic models with an adjusted R2 value over 0.75, we estimated the
maximum lift-to-drag ratio (maximum Cl/Cd), which we refer to within this paper as the aerodynamic efficiency of
the airfoil. The intercept of the model also provides an estimate of the zero-lift drag coefficient (Cd, L=0). In total, we
obtained results for 39 of the 51 airfoils with a converging result across eight species. Note that the prairie falcon
airfoils converged for too few angles of attack to allow a linear model to be fit effectively. The eliminated airfoils will
be explored in detail in future work.
III. Results
To quantify the effect of flight style and airfoil position on key aerodynamic performance variables, we fit
two linear models. Both models had flight style and the position of the airfoil as the dependent variables and the
independent variable was either aerodynamic efficiency or zero-lift drag. We found that the aerodynamic efficiency
of the extracted airfoils increased for airfoils from gliding species, although the effect was not statistically significant
(p-value = 0.1). Airfoils from gliding species were estimated to have an average aerodynamic efficiency of 46.3 and
the flapping species had an average of 38.8 which demonstrates a 19.3% increase in the efficiency of the gliding birds
we tested. Note that the statistical significance of this result was highly sensitive to small changes in our sample size,
suggesting that it will be crucial to increase the number of species and overall airfoils to confirm this result. (Fig. 7a).
We found that the position of the airfoil along the span did not have a significant effect on aerodynamic efficiency (p-
value = 0.90).
Next, we fit the second linear model with zero-lift drag as the independent variable. We found that both flight
style and the airfoil position had an insignificant effect on the zero-lift drag for the extracted airfoils (p-value = 0.87
and 0.23, respectively). The insignificant effect of the airfoil span position on both linear models is contradictory to
our expectation that interior airfoils would be more cambered than tip airfoils. However, this shape change may not
be effectively captured through these two aerodynamic parameters alone.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
Fig. 7 Gliding birds (blue) had an insignificantly higher aerodynamic efficiency than flapping birds (pink).
a) Aerodynamic efficiency of flapping vs. gliding birds. b) The sectional lift and drag coefficients for each airfoil.
Both linear models had low adjusted R2 values (0.0217 for the aerodynamic efficiency model and -0.0134
for the zero-lift drag model), suggesting that span position and flight style alone do not effectively capture all the
variance in the data set. We have a relatively low sample size of species (n=8) compared to the diversity of modern
bird species, which has more than 10,000 species. Therefore, caution should be applied when extrapolating these
results to species outside the current data set, though our selected species do broadly span the avian phylogeny.
Fig. 8 Drag polar of top-performing bird and engineered airfoils.
The top two performing airfoils resulted from the red-shouldered hawk between 36 and 45% of its wingspan
and the third highest performer came from the turkey vulture at 45% of its wingspan. The estimated maximum Cl/Cd
values are in the corresponding descending order: 78.8, 64.5, and 60.9 at the investigated Reynolds number of 1.5×105.
The highest aerodynamic efficiency from a flapping bird was 53.5, preceded by 10 gliding birds, for the band-tailed
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
pigeon at approximately 45% of the span. These avian results can be contrasted with engineered airfoils such as the
NACA 4415 used on the Shadow-200 UAV [16], which has a maximum Cl/Cd of 62.4 at the investigated Reynolds
number of 1.5×105. The high-lift airfoil, S1223 [17], exhibits similar flight characteristics intermediate to the two
highest-performing hawk airfoils, with a maximum Cl/Cd of 71.9 at the Reynolds number of 1.5×105. Furthermore,
when we compare the top hawk airfoil with an efficient low-speed airfoil suggested by Shunsun and Zhen, we can see
the engineered FX63-137 [16] achieves a higher lift-to-drag ratio of 76.2 at a Reynolds number of 1.5×105. Note that
the top-performing hawk airfoil outperforms all three of these airfoils at the equivalent Reynolds number. Therefore,
the top avian airfoils show promise as a template for novel UAV airfoil design, as we are already seeing a higher
aerodynamic efficiency than top-performing engineered airfoils at the the Reynolds number of 1.5×105. Additionally,
the geometry has not yet undergone any additional streamlining to optimize performance, showing promise for an
even higher aerodynamic efficiency with future work.
IV. Limitations
Several potential sources of error should be addressed before this method is used in future studies. First, the
factory resolution of the POP2 scanner is 0.05 mm. However, bird feathers present a unique challenge to 3D scanning
technology. We expect there will be small variations in the quantified airfoil coordinates due to the accuracy of the
3D scanner and because the leading edge is covered in small, easily deformed feathers (known as coverts). Our method
of scanning bird airfoils and generating models is being developed to be both efficient and highly accurate. As
translucent, dark objects with a near-zero thickness along some edges, they can be extremely challenging for the
scanner to resolve at given angles. As a result, the trailing edge thickness of some wings may have been overpredicted
in the airfoils in this study.
A major limitation of this methodology is our inability to scan wings in motion. Our 3D scanner requires
complete stability to produce an accurate scan, so out of necessity, we have scanned preserved wings in a series of
standardized positions. However, these taxidermy standards do not necessarily conform to the wing configurations
used by living birds. Because birds can morph their wing shapes during flight, the airfoil shape data collected in this
research cannot capture the variety of airfoils used in flight. In addition, our scans were conducted at atmospheric
pressure in a still room with no significant airflow. Bird wings are not rigid, and in normal flight, we expect the
feathers to deform under aerodynamic loading [18], which could lead to different morphology and aerodynamic
characteristics. Additionally, the aerodynamic analysis software we used, XFOIL, comes with its own limitations.
Foremost, it has a limited ability to model complex shapes and restricts us to a 2D analysis, which ignores the 3D
effects of flight. Furthermore, our work is limited to a steady analysis, which is appropriate for gliding flight, but it
challenges our ability to extend this analysis to flapping flight conditions. In future work, we could investigate the
implementation of unsteady panel codes. As we continue our study, we intend to compile an open-source database of
airfoils from a large selection of bird species and individuals. This work uses a broad grouping for gliding flight,
where both gliding and soaring species have been included in the same group. Future work should investigate the
specific role of these different modes of bird flight. This will allow us to expand our knowledge of how biological
attributes can be used in aircraft design and will allow for direct comparison to engineered airfoils.
V. Conclusions
In this study, we scanned multiple prepared bird wing specimens and adapted a 3D printer slicing algorithm
to extract airfoils along the length of their wings. We focused on species that are known to primarily flap and those
that primarily glide to determine if there were functional differences between the airfoils used within these flight
styles. Using XFOIL to analyze the smoothed airfoil shapes revealed that gliding birds employed airfoils that tended
to have a higher aerodynamic efficiency than flapping species, although the effect was not statistically significant.
Future work is required to expand on and confirm this result, as it was sensitive to the sample size. However, we found
that the top-performing airfoils in this study performed at a similar level, if not higher, than engineered airfoils used
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
on modern UAVs, suggesting that future UAVs could benefit from exploring the use of avian-inspired airfoils in low-
speed conditions.
VI. Acknowledgments
The authors would like to thank the Museum of Wildlife and Fish Biology’s Collections Manager, Irene Engilis, for
her support, use of specimens, and contributions to the project. Thank you to Kiran Weston, Kaleb B. Bordner, and
other members of the BIRD lab for their helpful discussions and input on the manuscript.
References
[1] C. Harvey, L. L. Gamble, C. R. Bolander, D. F. Hunsaker, J. J. Joo, and D. J. Inman, “A review of avian-
inspired morphing for UAV flight control,” Prog. Aerosp. Sci., vol. 132, p. 100825, Jul. 2022, doi:
10.1016/j.paerosci.2022.100825.
[2] P. C. Withers, “An Aerodynamic Analysis of Bird Wings as Fixed Aerofoils,” J. Exp. Biol., vol. 90, no. 1, pp.
143162, Feb. 1981, doi: 10.1242/jeb.90.1.143.
[3] J. A. Cheney et al., “Bird wings act as a suspension system that rejects gusts,” Proc. R. Soc. B Biol. Sci., vol.
287, no. 1937, p. 20201748, Oct. 2020, doi: 10.1098/rspb.2020.1748.
[4] K. V. Reynolds, A. L. R. Thomas, and G. K. Taylor, “Wing tucks are a response to atmospheric turbulence in
the soaring flight of the steppe eagle Aquila nipalensis,” J. R. Soc. Interface, vol. 11, no. 101, p. 20140645,
Dec. 2014, doi: 10.1098/rsif.2014.0645.
[5] C. Harvey, V. B. Baliga, P. Lavoie, and D. L. Altshuler, “Wing morphing allows gulls to modulate static pitch
stability during gliding,” J. R. Soc. Interface, vol. 16, no. 150, p. 20180641, Jan. 2019, doi:
10.1098/rsif.2018.0641.
[6] C. Harvey, V. B. Baliga, J. C. M. Wong, D. L. Altshuler, and D. J. Inman, “Birds can transition between stable
and unstable states via wing morphing,” Nature, vol. 603, no. 7902, pp. 648653, Mar. 2022, doi:
10.1038/s41586-022-04477-8.
[7] A. C. Carruthers, S. M. Walker, A. L. R. Thomas, and G. K. Taylor, “Aerodynamics of aerofoil sections
measured on a free-flying bird,” Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng., vol. 224, no. 8, pp. 855864,
Aug. 2010, doi: 10.1243/09544100JAERO737.
[8] Y. Murayama, T. Nakata, and H. Liu, “Flexible Flaps Inspired by Avian Feathers Can Enhance Aerodynamic
Robustness in low Reynolds Number Airfoils,Front. Bioeng. Biotechnol., vol. 9, p. 612182, May 2021, doi:
10.3389/fbioe.2021.612182.
[9] J. A. Rader and T. L. Hedrick, “Morphological evolution of bird wings follows a mechanical sensitivity
gradient determined by the aerodynamics of flapping flight,” Nat. Commun., vol. 14, no. 1, p. 7494, Nov. 2023,
doi: 10.1038/s41467-023-43108-2.
[10] W. Thielicke and E. J. Stamhuis, “The influence of wing morphology on the three-dimensional flow patterns of
a flapping wing at bird scale,” J. Fluid Mech., vol. 768, pp. 240260, Apr. 2015, doi: 10.1017/jfm.2015.71.
[11] G. K. Ananda and M. S. Selig, “Design of Bird-Like Airfoils,” in 2018 AIAA Aerospace Sciences Meeting,
Kissimmee, Florida: American Institute of Aeronautics and Astronautics, Jan. 2018. doi: 10.2514/6.2018-0310.
[12] T. Liu, K. Kuykendoll, R. Rhew, and S. Jones, “Avian Wing Geometry and Kinematics,” AIAA J., vol. 44, no.
5, pp. 954963, May 2006, doi: 10.2514/1.16224.
[13] M. Drela, “XFOIL: Subsonic Airfoil Development System.” Massachusetts Institute of Technology. [Online].
Available: web.mit.edu/drela/Public/web/xfoil
[14] “PrusaSlicer.” Prusa Research.
[15] S. Bhandari, “slice_stl_create_path(triangles,slice_height).” [Online]. Available:
https://www.mathworks.com/matlabcentral/fileexchange/62113-slice_stl_create_path-triangles-slice_height
[16] W. Shunshun and G. Zheng, “Design, optimization and application of two-element airfoils for tactical UAV,”
Adv. Mech. Eng., vol. 14, no. 11, p. 168781322211370, Nov. 2022, doi: 10.1177/16878132221137027.
[17] M. S. Selig and J. J. Guglielmo, “High-Lift Low Reynolds Number Airfoil Design,” J. Aircr., vol. 34, no. 1, pp.
7279, Jan. 1997, doi: 10.2514/2.2137.
[18] B. Klaassen Van Oorschot, R. Choroszucha, and B. W. Tobalske, “Passive aeroelastic deflection of avian
primary feathers,” Bioinspir. Biomim., vol. 15, no. 5, p. 056008, Jul. 2020, doi: 10.1088/1748-3190/ab97fd.
This is a preprint of: Glenn et al., “Analysis of bird wing airfoil aerodynamic efficiency,” AIAA SciTech
2024 Forum. The published article may differ from this preprint.
Appendix
Table 1. Wing measurements.
Species
ID
Longest
Primary
(mm)
Shoulder to
Longest
Primary (mm)
Shoulder
Chord Length
(mm)
Wrist Chord
Length
(mm)
Shoulder to
Wrist
(mm)
Turkey Vulture
Cathartes aura
WFB-7375
500
672
283
295
177
EM-20
468
663
250
285
192
Red-Shouldered Hawk
Buteo jamaicensis
WFB-8469
332
389
175
220
125
Laysan Albatross
Phoebastria immutabilis
WFB-17490
509
749
171
154
245
WFB-13973
516
744
162
159
240
Western Gull
Larus occidentalis
FAV016
432
575
198
189
155
JC-07
406
545
181
179
147
Band-tailed Pigeon
Patagioenas fasciata
WFB-13607
214
277
115
133
77
Yellow-billed magpie
Pica nuttalli
WFB-9793
187
230
118
130
65
WFB-6155
199
236
140
140
55
Black Scoter
Melanitta americana
WFB-15829
222
295
102
114
73
WFB-15865
214
262
130
111
68
WFB-15836
213
275
138
109
59
Red-tailed Hawk
Buteo jamaicensis
WFB-6901
402
490
195
235
155
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The physical principles that govern the function of biological structures also mediate their evolution, but the evolutionary drivers of morphological traits within complex structures can be difficult to predict. Here, we use morphological traits measured from 1096 3-dimensional bird wing scans from 178 species to test the interaction of two frameworks for relating morphology to evolution. We examine whether the evolutionary rate (σ²) and mode is dominated by the modular organization of the wing into handwing and armwing regions, and/or the relationship between trait morphology and functional output (i.e. mechanical sensitivity, driven here by flapping flight aerodynamics). Our results support discretization of the armwing and handwing as morphological modules, but morphological disparity and σ² varied continuously with the mechanical sensitivity gradient and were not modular. Thus, mechanical sensitivity should be considered an independent and fundamental driver of evolutionary dynamics in biomechanical traits, distinct from morphological modularity.
Article
Full-text available
Owning to the requirements for longer endurance, larger payloads as well as simpler mechanical structure of tactical UAV, the two-element airfoil technology is introduced. Firstly, the aerodynamic characteristics of five common low-speed airfoils are calculated and compared. The airfoil with good aerodynamic performance after segmentation is selected as initial airfoil. Then, the influence of different segmentation types on the aerodynamic characteristics of airfoils is analyzed, and the segmentation type which brings the best aerodynamic benefits is selected. The slot parameters are optimized by NSGA-II optimization algorithm under cruising, take-off, and a large angle of attack states to obtain a high-performance self-adaption tactical UAV airfoil. Finally, the new airfoil is applied to Sparrow-hawk tactical UAV, the CL 1.5 /CD of acceleration state increases 15.5%, the CL of large angle of attack increases 20.4% and the CL 1.5 /CD of cruising state after enlarge the aspect ratio increase12.3%.
Article
Full-text available
The impressive maneuverability demonstrated by birds has so far eluded comparably sized uncrewed aerial vehicles (UAVs). Modern studies have shown that birds' ability to change the shape of their wings and tail in flight, known as morphing, allows birds to actively control their longitudinal and lateral flight characteristics. These advances in our understanding of avian flight paired with advances in UAV manufacturing capabilities and applications has, in part, led to a growing field of researchers studying and developing avian-inspired morphing aircraft. Because avian-inspired morphing bridges at least two distinct fields (biology and engineering), it becomes challenging to compare and contrast the current state of knowledge. Here, we have compiled and reviewed the literature on flight control and stability of avian-inspired morphing UAVs and birds to incorporate both an engineering and a biological perspective. We focused our survey on the longitudinal and lateral control provided by wing morphing (sweep, dihedral, twist, and camber) and tail morphing (incidence, spread, and rotation). In this work, we discussed each degree of freedom individually while highlighting some potential implications of coupled morphing designs. Our survey revealed that wing morphing can be used to tailor lift distributions through morphing mechanisms such as sweep, twist, and camber, and produce lateral control through asymmetric morphing mechanisms. Tail morphing contributes to pitching moment generation through tail spread and incidence, with tail rotation allowing for lateral moment control. The coupled effects of wing-tail morphing represent an emerging area of study that shows promise in maximizing the control of its morphing components. By contrasting the existing studies, we identified multiple novel avian flight control methodologies that engineering studies could validate and incorporate to enhance maneuverability. In addition, we discussed specific situations where avian-inspired UAVs can provide new insights to researchers studying bird flight. Collectively, our results serve a dual purpose: to provide testable hypotheses of flight control mechanisms that birds may use in flight as well as to support the design of highly maneuverable and multi-functional UAV designs.
Article
Full-text available
Birds morph their wing shape to accomplish extraordinary manoeuvres 1–4 , which are governed by avian-specific equations of motion. Solving these equations requires information about a bird’s aerodynamic and inertial characteristics ⁵ . Avian flight research to date has focused on resolving aerodynamic features, whereas inertial properties including centre of gravity and moment of inertia are seldom addressed. Here we use an analytical method to determine the inertial characteristics of 22 species across the full range of elbow and wrist flexion and extension. We find that wing morphing allows birds to substantially change their roll and yaw inertia but has a minimal effect on the position of the centre of gravity. With the addition of inertial characteristics, we derived a novel metric of pitch agility and estimated the static pitch stability, revealing that the agility and static margin ranges are reduced as body mass increases. These results provide quantitative evidence that evolution selects for both stable and unstable flight, in contrast to the prevailing narrative that birds are evolving away from stability ⁶ . This comprehensive analysis of avian inertial characteristics provides the key features required to establish a theoretical model of avian manoeuvrability.
Article
Full-text available
Unlike rigid rotors of drones, bird wings are composed of flexible feathers that can passively deform while achieving remarkable aerodynamic robustness in response to wind gusts. In this study, we conduct an experimental study on the effects of the flexible flaps inspired by the covert of bird wings on aerodynamic characteristics of fixed-wings in disturbances. Through force measurements and flow visualization in a low-speed wind tunnel, it is found that the flexible flaps can suppress the large-scale vortex shedding and hence reduce the fluctuations of aerodynamic forces in a disturbed flow behind an oscillating plate. Our results demonstrate that the stiffness of the flaps strongly affects the aerodynamic performance, and the force fluctuations are observed to be reduced when the deformation synchronizes with the strong vortex generation. The results point out that the simple attachment of the flexible flaps on the upper surface of the wing is an effective method, providing a novel biomimetic design to improve the aerodynamic robustness of small-scale drones with fixed-wings operating in unpredictable aerial environments.
Article
Full-text available
Musculoskeletal systems cope with many environmental perturbations without neurological control. These passive preflex responses aid animals to move swiftly through complex terrain. Whether preflexes play a substantial role in animal flight is uncertain. We investigated how birds cope with gusty environments and found that their wings can act as a suspension system, reducing the effects of vertical gusts by elevating rapidly about the shoulder. This preflex mechanism rejected the gust impulse through inertial effects, diminishing the predicted impulse to the torso and head by 32% over the first 80 ms, before aerodynamic mechanisms took effect. For each wing, the centre of aerodynamic loading aligns with the centre of percussion, consistent with enhancing passive inertial gust rejection. The reduced motion of the torso in demanding conditions simplifies crucial tasks, such as landing, prey capture and visual tracking. Implementing a similar preflex mechanism in future small-scale aircraft will help to mitigate the effects of gusts and turbulence without added computational burden.
Article
Full-text available
A gliding bird's ability to stabilize its flight path is as critical as its ability to produce sufficient lift. In flight, birds often morph the shape of their wings, but the consequences of avian wing morphing on flight stability are not well understood. Here, we investigate how morphing the gull elbow joint in gliding flight affects their static pitch stability. First, we combined observations of freely gliding gulls and measurements from gull wing cadavers to identify the wing configurations used during gliding flight. These measurements revealed that, as wind speed and gusts increased, gulls flexed their elbows to adopt wing shapes characterized by increased spanwise camber. To determine the static pitch stability characteristics of these wing shapes, we prepared gull wings over the anatomical elbow range and measured the developed pitching moments in a wind tunnel. Wings prepared with extended elbow angles had low spanwise camber and high passive stability, meaning that mild perturbations could be negated without active control. Wings with flexed elbow angles had increased spanwise camber and reduced static pitch stability. Collectively, these results demonstrate that gliding gulls can transition across a broad range of static pitch stability characteristics using the motion of a single joint angle.
Conference Paper
Full-text available
Article
Full-text available
The effect of airfoil design parameters, such as airfoil thickness and camber, are well understood in steady-state aerodynamics. But this knowledge cannot be readily applied to the flapping flight in insects and birds: flow visualizations and computational analyses of flapping flight have identified that in many cases, a leading-edge vortex (LEV) contributes substantially to the generation of aerodynamic force. In flapping flight, very high angles of attack and partly separated flow are common features. Therefore, it is expected that airfoil design parameters affect flapping wing aerodynamics differently. Existing studies have focused on force measurements, which do not provide sufficient insight into the dominant flow features. To analyse the influence of wing morphology in slow-speed bird flight, the time-resolved three-dimensional flow field around different flapping wing models in translational motion at a Reynolds number of 22 000 < Re < 26 000 was studied. The effect of several Strouhal numbers (0.2 < St < 0.4), camber and thickness on the flow morphology and on the circulation was analysed. A strong LEV was found on all wing types at high St. The vortex is stronger on thin wings and enhances the total circulation. Airfoil camber decreases the strength of the LEV, but increases the total bound circulation at the same time, due to an increase of the 'conventional' bound circulation at the inner half of the wing. The results provide new insights into the influence of airfoil shape on the LEV and force generation at low Re. They contribute to a better understanding of the geometry of vertebrate wings, which seem to be optimized to benefit from LEVs in slow-speed flight.
Article
Bird feathers are complex structures that passively deflect as they interact with air to produce aerodynamic force. Newtonian theory suggests that feathers should be stiff to effectively utilize this force. Observations of flying birds indicate that feathers respond to aerodynamic loading via spanwise bending, twisting, and sweeping. These deflections are hypothesized to optimize flight performance, but this has not yet been tested. We measured deflection of isolated feathers in a wind tunnel to explore how flexibility altered aerodynamic forces in emulated gliding flight. Using primary feathers from seven raptors and a rigid airfoil, we quantified bending, sweep, and twisting, as well as ∝ (attack angle) and slip angle. We predicted that 1) feathers would deflect under aerodynamic load, 2) bending would result in lateral redirection of force, 3) twisting would alter spanwise ∝ "washout" and delay the onset of stall, and 4) flexural stiffness of feathers would exhibit positive allometry. The first three predictions were supported by our results, but not the fourth. We found that bending resulted in the generation of lateral forces towards the base of the feather on the order of ~10% of total lift. In comparison to the airfoil which stalled at ∝=13.5°, all feathers continued to increase lift production with increasing angle of attack to the limit of our range of measurements (α=27.5°). We observed that feather stiffness exhibited positive allometry (∝ mass1.1±0.3), approximately consistent with previous research showing that stiffness does not scale as predicted by geometric similarity (∝ mass1.67). These findings demonstrate that feather flexibility may provide passive roll stability and delay stall by twisting to reduce local ∝ at the feather tip. Our findings are the first to measure forces due to feather deflection under aerodynamic loading and can inform future models of avian flight as well as biomimetic morphing-wing technology.