You are on page 1of 9

Anatomy

14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of
J. Anat. (2017) 230, pp734--742 doi: 10.1111/joa.12595

Distribution of the characteristics of barbs and


barbules on barn owl wing feathers
Matthias Weger and Hermann Wagner
Institute of Zoology, RWTH Aachen University, Aachen, Germany

Abstract
Owls are known for the development of a silent flight. One conspicuous specialization of owl wings that has
been implied in noise reduction and that has been demonstrated to change the aerodynamic behavior of the
wing is a soft dorsal wing surface. The soft surface is a result of changes in the shape of feather barbs and
barbules in owls compared with other bird species. We hypothesized that as the aerodynamic characteristics of
a wing change along its chordwise and spanwise direction, so may the shape of the barbs and barbules.
Therefore, we examined in detail the shapes of the barbs and barbules in chordwise and spanwise directions.
The results showed changes in the shapes of barbs and barbules at the anterior and distal parts of the wing,
but not at more posterior parts. The increased density of hook radiates at the distalmost wing position could
serve to stiffen that vane part that is subject to the highest forces. The change of pennulum length in the
anterior part of the wing and the uniformity further back could mean that a soft surface may be especially
important in regions where flow separation may occur.
Key words: biomimetics; feather morphology; flow separation; silent flight.

American barn owl (Tyto furcata pratincola) or the eagle


Introduction
owl (Bubo bubo) have elongated pennula on the radiates of
The silent flight of owls (Strigiformes) is a model system for their wing feathers so that neighboring barb shafts are cov-
biomimetic research (Vad et al. 2006; Bachmann et al. 2011; ered (Bachmann et al. 2007; Chen et al. 2012). The surface
Kla€n et al. 2012; Winzen et al. 2012, 2014a,b,c, 2015; Clark of owl wings feels soft. Because this softness is reminiscent
et al. 2014; Winzen & Roidl, 2016). Morphological special- of velvet, the surface of the wings of the owls is often called
izations such as the leading-edge serrations, the trailing- velvet-like. We adopt this terminology in the following.
edge fringes, a soft dorsal surface and a wing shape opti- Bachmann et al. (2007) described several morphological
mized for slow flight are well known, but their exact func- characteristics of the pennula: (i) variation with feather
tion is still a matter of debate and often not well type; (ii) shorter on the outer vane than on the inner vane
understood (Fatio, 1866; Mascha, 1904; Graham, 1934; Sick, on most positions; (iii) variation along the vane. Bachmann
1937; Hertel, 1963). et al. (2007) described these structural properties on three
The surface of bird wings is formed by pennaceous feath- different wing feathers (10th primary feather, 8th secondary
ers. Feathers consist of a shaft (scapus), which is subdivided feather, 1st greater primary covert). What these authors did
into the calamus and the rachis, and barbs. Barbs originate not do, and what is missing so far, is to relate the character-
at the rachis. The barb shafts give off barbules (also called istics of the velvet-like surface to wing position. This is
radiates). The radiates can be interconnected to form a interesting because it was shown that the addition of a vel-
closed vane, as found for example in pennaceous feathers. vet-like surface to an artificial wing reduced flow separation
The radiates consist of a base and a filamentous end called and enabled boundary-layer control (Kla €n et al. 2009;
pennulum. In most bird species, the pennula are short and Winzen et al. 2012, 2014a). Since the aerodynamic charac-
do not overlap neighboring barb shafts. The wing surface in teristics of a wing, such as the boundary layer, Reynolds
these cases feels rough. By contrast, owls such as the number, shear stress and pressure distribution, change in
both the spanwise and chordwise directions (Usherwood
et al. 2005; Kla €n et al. 2009, 2012; Winzen et al. 2012,
Correspondence
2014a, 2015; Winzen & Roidl, 2016), the shapes of the barbs
Matthias Weger, Institute of Zoology, RWTH Aachen University,
Worringerweg 3, D-52074 Aachen, Germany. E: weger@bio2.rwth- and barbules that form the velvet-like surface may do so,
aachen.de too. Thus, here we tested the hypothesis that the character-
Accepted for publication 3 January 2017
istics of the velvet-like surface change along the chordwise
Article published online 3 March 2017 and spanwise directions of the T. furcata pratincola wing.

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Barb characteristics on barn owl wings, M. Weger and H. Wagner 735

The investigations were performed independent of feather Furthermore, for reasons of comparability between different
type. The data presented in the following demonstrate an samples, the following deviations from the strict geometrical deter-
influence of wing position on the characteristics of the vel- mination of the position examined were applied.

vet-like surface at anterior and distal wing positions. • In the case of several overlapping feathers, the dorsal-most
feather at a given measurement position which was exposed
to the airflow, was investigated.
Materials and methods • If at a selected position a covered part occurred (as indi-
cated by its whitish coloration), we interpreted this as an
The investigation was carried out on eight wings of T. furcata prat-
artifact of the preparation, and chose the closest uncovered
incola. The wings were obtained from animals of the colony at the
feather part in spanwise position at the given chordwise
Institute of Biology II at RWTH Aachen University that had been sac-
position for analysis. This was done because major differ-
rificed at the end of other experiments under a local authority per-
ences were observed in the quality of the velvet-like surface
mit [Landespra €sidium fu € r Natur, Umwelt und Verbraucherschutz
between the covered and uncovered feather parts (Sick,
Nordrhein-Westfalen, Recklinghausen, Germany (LANUV)]. The
1937; Bachmann et al. 2007). We chose only to investigate
wings were prepared as they appear in gliding flight. We chose this
the velvet-like surface that occurs at the brownish, uncov-
wing form because in gliding flight the wing is fully extended and
ered feather part because this feather part is exposed to
because it offered the possibility to compare our data with earlier
the airflow, and because the covered, whitish part does not
work (Bachmann et al. 2007, 2011).
occur in all feathers.
• Samples were investigated at the central or proximal part of a
Selection of positions barb. This took into account that the radiates and pennula
tend to be reduced at the barb tips (Bachmann et al. 2007).
To obtain samples at distinct wing positions, a wing was subdi-
vided lengthwise into different sampling positions. The wings
were positioned according to the description in Bachmann et al. Preparation of samples
(2011; Fig. 1). A two-dimensional coordinate system was estab-
Preparation of samples included the following steps (Fig. 2). First, a
lished in the plane formed by wingspan and wing chord. The
sample containing a short rachis length and vane area and several
chordwise axis was set as the x-axis with the leading edge being
barbs was cut off the rachis of a given feather and fixed on a small
set to 0 and the trailing edge being set to 1. The y-axis was set
aluminum block with double-sided adhesive tape so that the area
parallel to the wingspan, with the position of the root chord
of the vane was horizontal. Care was taken that the feather vane
being set to 0 and the wing tip being set to 1. Four spanwise
was closed.
positions were investigated in this study, at 0.2, 0.4, 0.6 and 0.8
of the y-axis. At each spanwise position, 12 measurement posi- Afterwards, images were taken from a horizontal perspective
tions were investigated in the chordwise direction. Note that the with a microscope camera (Scope Tek DCM500, Hangzhou Scopetek
origin of the x-axis was always at the leading edge at the respec- Opto-Electric Co., Hangzhou, Zhejiang Province, China) fixed to a
tive spanwise position, whereas 100% of the x-axis was at the stereomicroscope (Nikon SMT 10-A, Nikon Corp., Tokyo, Japan) at a
trailing edge. Thus, the absolute length of the x-axis differed for magnification of 259.
the different spanwise positions. The first six measurement posi-
tions had a distance of 0.05 in the normalized x-axis, starting at
0.05 and ending at 0.3 of the normalized x-axis. This was done
because most effects of the flow field, including boundary layer
separation at artificial owl-based wing models, were found in this
particular region (Kla €n et al. 2009, 2012; Winzen et al. 2012,
2014a). Between 0.4 and 0.9 of the normalized x-axis, the mea-
surement positions were set in steps of 0.1 of the x-axis. This
resulted in 48 measurement positions per wing. In the following
we will use a combination of the respective values along the
x-axis and the y-axis to describe the measurement positions
on the wing. For example, the position x05y20 describes the
chordwise position (x-axis) 0.05 at 0.2 of the wing length (y-axis).
It is important here to remind the reader that different feather
types such as marginal coverts, lesser, median and greater sec- Fig. 1 Measurement positions. The velvet-like surface was investi-
ondary coverts, primary coverts, alular coverts, alular feathers, pri- gated at four different parts of the wing (0.2, 0.4, 0.6 and 0.8 of
mary remiges and secondary remiges occur on a bird wing. Most of wing length) which is indicated by the corresponding lines. Twelve
these feathers can be numbered by their position, which holds chordwise measurement points were set at each investigated wing
especially for remiges. We decided to categorize the different part. The first six measurement positions had a distance of 0.05 in the
feather types and formed three categories (Fig. 1). The first feather normalized x-axis, between 0.05 and 0.3 of the normalized x-axis.
category includes only the marginal coverts. The second category Measurement points between 0.4 and 0.9 were set in steps of 0.1 of
includes lesser, median and greater secondary coverts, primary cov- the x-axis. Note that samples were only taken from feather parts that
erts and alular coverts. For reasons of simplicity, in the following we exhibited the uncovered velvet-like surface. The feathers that occur at
shall refer to this category as covert feathers. The third category these wing parts include marginal coverts, covert feathers (primary
comprises the flight feathers, which consist of primary remiges, and secondary coverts) and flight feathers, depending on wing part
secondary remiges and alular feathers. and wing position.

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
736 Barb characteristics on barn owl wings, M. Weger and H. Wagner

B C
Fig. 2 Preparation of measurement samples.
(A) Flight feather of Tyto furcata pratincola.
(B) Cut off part of the outer vane. The arrow
from B to A points to the feather part that
was investigated. Note that the barb next to
the left outer barb was partially cut to obtain
a better view on complete barb. (C) Image of
barbs of the investigated vane part. The
arrow from C to B points to the part of the
sample that was used for the recording.

Quantification of velvet-like surface


The images were analyzed in IMAGEJ (National Institutes of Health,
Bethesda, MD, USA). Several parameters were extracted from the
photographs (Fig. 3). The length of a pennulum was measured
from the base to the tip of the pennulum. The length of the hook
radiates was measured from the basis of the barb to the origin of
the pennulum. The length of the pennulum was divided by the
base of the hook radiate to estimate the pennulum–base ratio.
The overlap (O) of the pennula (Sick, 1937; Bachmann et al. 2007)
describes the extent to which pennula reach over neighboring
barbs, determined by the formula:

rHR
O¼ Fig. 3 Overview of measured parameters. The investigated barb
dbarb
parameters are presented in the sketch. The overlap was estimated as
with the range of the hook radiates (rHR) being measured as the the ratio between the range of hook radiates and the distance
distance between the barb and the tip of the pennula, normal to between two barbs. The density of the hook radiates was determined
the barb axis and dbarb being measured as the distance between as the number of hook radiates per mm. h.r., hook radiates; b.r., bow
two adjacent barbs. The dbarb was measured in regions where two radiates.
adjacent barb ran in parallel, which occurred throughout the major
part of barb length, but not close to the distal end. Finally, the den- in structure. Thirdly, all 48 measured spanwise and chordwise posi-
sity of the pennula was determined as the number of barbules per tions were compared (1128 test cases) to find possible differences
mm for hook radiates. between wing positions.

Statistical analysis Results


The data obtained were analyzed statistically with EXCEL (Microsoft
Occurrence of feathers at investigated wing parts
Corporation, Redmond, WA, USA) and MATLAB (The MathsWorks,
Inc., Natick, MA, USA). The data of the T. furcata pratincola speci- Eight wings were examined in this study. We chose to sub-
men were tested statistically using a multiway ANOVA in MATLAB. A
divide the occurring feathers in three categories: marginal
Tukey HSD test was used for post-hoc analysis. The multiway ANOVA
coverts, covert feathers, and flight feathers. The distribution
investigated three factors. First, the differences between all individ-
uals were checked. This resulted in 28 test cases. In a second step, of the different feather types was similar for the investi-
the difference among the three different feather types was tested gated owl specimen (Fig. 4). Note, however, that due to
(three test cases). This was done to determine possible differences our position-based method of analysis it sometimes

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Barb characteristics on barn owl wings, M. Weger and H. Wagner 737

happened that in different specimens, different feather occurred over 55  2.14% of the chord at 0.2 of wing
types occurred at a given position. length, 42  5.49% of the chord at 0.4 of the wing length
The marginal coverts and covert feathers occurred pre- and 12  9.64% of the chord at 0.6 of the wing length.
dominantly at the frontal part of the wing at 0.2 of the
wing length and to a lesser extent also at 0.4 of the wing
Distribution of the velvet-like surface
length (Fig. 1). The area covered by the marginal coverts
was 8.2% of the total wing area. Barb shape in marginal The barbs of the feathers investigated were compared at
coverts varied a lot. In this feather type, prolonged hook different spanwise and chordwise wing positions (Fig. 1).
radiates were present, but the singular barbules were com- We used six different morphological parameters to describe
paratively thin (Fig. 5A). The coverts were present over the the properties of the velvet-like surface of an owl wing. The
high cambered wing part where the patagium, skeletal variation of these parameters in chordwise and spanwise
parts and adjoining tissue of the wing are present as well direction was analyzed.
(Fig. 1). Coverts were found in 26.3% of the total wing First we compared the data obtained from the eight indi-
area. The flight feathers made up the largest part of the vidual wings. Two of 28 comparisons exhibited significant
total wing area. They were found in the distal and posterior differences (multiway ANOVA followed by a Tukey post hoc
parts of the wing and covered 65.4% of the total wing area. test, P < 0.05). This low number of differences led us to con-
The outer half of the wing was mainly formed by feathers clude that data of all wings may be pooled for further anal-
emerging from the arm and hand parts of the wing. The yses. No differences between feather types were observed.
distal part of the wing (0.8 of wing length) was completely The pennula had a similar length at the majority of the
made up by flight feathers, which were also present at the measured spanwise and chordwise positions in our study,
caudal part of the wing at all other positions. The marginal with an average length of 793  166 lm (Fig. 6). In gen-
coverts and secondary coverts of T. furcata pratincola eral, the pennulum length was constant over the wing and

Fig. 4 Feather types at different measurement positions. The feather types that occurred at the different measurement positions are depicted for
all individual investigated wings. Marginal coverts occur at the frontal part at 0.2 and to a lesser extent at 0.4 of the wings. Secondary coverts
occur mainly at the frontal and median part of 0.2–0.4 of the wings. The flight feathers are found at the distalmost and caudal part of the wings.
Note that some data points are missing which is shown at the empty square. T.f.p: Tyto furcata pratincola; y/c: spanwise measurement position.

A B C

Fig. 5 Feather types. The three different feather types are depicted in this figure. (A) Example of a marginal covert. (B) Example of a covert
feather. (C) Example of a flight feather. A lower barbule density and a resulting higher porosity can be seen for the marginal coverts in A).

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
738 Barb characteristics on barn owl wings, M. Weger and H. Wagner

exhibited no significant difference between the majority of The bases of hook radiates between individuals exhib-
different spanwise and chordwise positions. One conspicu- ited differences in four of 28 tested cases. No differences
ous deviation from this scheme was observed at the four were observed between the three feather types. The
frontal chordwise positions at 0.2 of the length of the wing. length of the base of the hook radiates from where
The length of the pennula at x20y20 was the greatest of all the pennula emerged showed no positional differences
investigated positions (1053  205 lm) (Fig. 6). The pen- despite a length range between 131 and 392 lm
nula at this position were significantly longer than pennula (224  39 lm, Fig. 7).
of several measurement positions (multiway ANOVA followed The ratio between the pennula and the base of hook
by a Tukey post hoc test, see Supporting Information radiates can be used to describe the elongation of the pen-
Fig. S1, vertically and horizontally aligned squares based at nula. The mean ratio between pennula and hook radiates
x20y20). We observed this difference after analysis of five was 3.58  92. No difference between individuals and
wings, which we initially regarded as a high enough sample between feather types were observed for this ratio. The
size. Since the difference occurred only at the medial and ratio between pennula and hook radiates showed few
frontal part of the wing, we were worried that the differ- positional differences (Fig. 8). Only five of 1128 positional
ence might represent a false positive and decided to ana- combinations exhibited significant differences. These differ-
lyze three more wings. We observed the same difference. ences were between x05y20 and x20y20, as well as between
This led us to conclude that the difference was real. In sum- x15y40, x25y60 and x10y80, and between x10y80 and
mary, pennulum length did not change along the major x40y80 (Fig. S2). Since significance changed when only the
part of chordwise and spanwise directions, with the excep- first five wings were considered compared with when all
tion of a change at the leading edge of the most proximal eight wings were considered, some of these differences
investigated wing part, up to a chordwise position of 0.2. may represent false positives.

Fig. 6 Pennulum length. The length of the


pennula was measured at all four wing
positions on 12 different chordwise positions.

Fig. 7 Hook radiate length. The length of the


hook radiates was measured at all four wing
positions on 12 different chordwise positions.

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Barb characteristics on barn owl wings, M. Weger and H. Wagner 739

Barb distance was different in two out 28 cases for The overlap describes the extent to which the pennula
T. furcata pratincola individuals, again leading to the con- reach over neighboring barbs. If the pennula were to
clusion that the data may be pooled. In our investigation exactly reach the adjacent barb, the overlap would be 1.
the distance between adjacent barbs was significantly We observed a mean overlap of 2.1  0.58 (Fig. 10). When
higher in flight feathers (292  42 lm) than in covert comparing different individuals, no differences were
feathers (258  43 lm, multiway ANOVA, P < 0.001). Barb observed. In contrast, differences were observed between
distance varied little (284  43 lm) for most positions on the different feather types. The overlap on flight feathers
the feather. Yet, several positional differences occurred (2.26  0.52) was significantly higher than the overlap on
(Fig. 9). Forty-two of 1128 positions exhibited significant marginal coverts (1.64  0.49, multiway ANOVA, P < 0.001)
positional differences. These differences were especially and covert feathers (1.98  0.55, n-way ANOVA, P < 0.001).
conspicuous between position x05y40 and most (nine of Two of 1128 positional comparisons exhibited significant
12 comparisons) of the chordwise positions at 0.8 of the differences [x05y20 differed from x60y20 (multiway ANOVA,
wing length (Supporting Information Fig. S3, horizontally P = 0.018) and x70y40 (multiway ANOVA, P = 0.02)] Support-
aligned squares in sector y40y80). This suggested that barb ing Information Fig. S4).
distance was increased at 0.8 wing length. Similar differ- The number of hook radiates per mm was the sixth inves-
ences occurred between caudal chordwise positions at 0.8 tigated parameter (Fig. 11). The number of hook radiates
of the wing length and frontal to central chordwise posi- per mm can be used to describe the density of the velvet-
tions at the other spanwise positions (vertically aligned like surface and further indicates the porosity of the feather
squares in sectors y20y80, y40y80 and y60y80). Both obser- vane. The mean value was 25.62  4.81 hook radiates/mm).
vations are consistent with earlier findings (Bachmann No differences between the individuals were observed.
et al. 2007). However, the different feather types exhibited a

Fig. 8 Ratio between pennula and base of


hook radiates. The ratio of pennula to hook
radiates was estimated for all four wing
positions on 12 different chordwise positions.

Fig. 9 Barb distance. The distance of two


adjacent barbs was measured at all four wing
positions on 12 different chordwise positions.

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
740 Barb characteristics on barn owl wings, M. Weger and H. Wagner

significantly different number of hook radiates. The flight


feathers (26.98  4.11 per mm) had a significantly higher
Discussion
number of hook radiates per mm compared with covert The velvet-like surface on wings of T. furcata pratincola was
feathers (23.51  6.03 per mm, multiway ANOVA, P = 0.0013) investigated with stereomicroscopy at preassigned wing
and marginal coverts (19.2  3.95 per mm, multiway ANOVA, positions to evaluate whether the characteristics of the sur-
P < 0.001). Covert feathers had a significantly higher num- face depend on wing position. We observed a change in
ber of hook radiates per mm compared with marginal cov- pennulum length at medial and anterior wing positions
erts (multiway ANOVA, P < 0.001). In marginal coverts the and an increase in the density of hook radiates at the distal-
hook radiates were widely spread. This results in a higher most wing position. In the following, we shall discuss the
porosity. The differences in between the three feather types results in the context of the existing literature and attempt
also resulted in a higher number of positional differences. to draw conclusions with regard to biological aspects and
Of 1128 positional combinations, 158 were significantly dif- technical applications.
ferent (Supporting Information Fig. S5). As demonstrated
by the thick horizontal stripe of squares in Fig. S5, the dif-
Distribution of the velvet-like surface on an owl wing
ferences were present between the first three chordwise
positions (905, 910, 915) at 0.2 of the wing length (y20) in Several attempts have been made qualitatively to describe
comparison with the majority of the other positions. Addi- (Sick, 1937) or quantitatively to compare the occurrence of
tionally, as indicated by the vertical and horizontal stripes the elongated pennula that are fundamental for the forma-
of squares based at x05y80, the hook radiate density was tion of a velvet-like feather surface (Bachmann et al. 2007;
especially high at the first two chordwise positions at 0.8 of Chen et al. 2012). The feather-type related analysis of Bach-
wing length. mann et al. (2007) detected a lower number of barbules

Fig. 10 Overlap. The overlap of the pennula


in regard to the adjoining barb shafts was
measured at all four wing positions on 12
different chordwise positions.

Fig. 11 Number of hook radiates. The


number of the hook radiates per mm was
measured at all four wing positions on 12
different chordwise positions.

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Barb characteristics on barn owl wings, M. Weger and H. Wagner 741

and longer pennula on T. furcata pratincola wings in com- member of the family of diurnal raptors. Feathers of
parison with pigeon (Columba livia domestica) wings, but B. bubo exhibited better sound-absorbing properties in
no differences between different primary and secondary flight than the feathers of the buzzard (by approximately
barn-owl feathers. Our position-dependent analysis demon- 10 dB for frequencies below 1 kHz). Clark et al. (2014)
strated an influence of wing position on the characteristics demonstrated that an artificial fabric based on a velvet-like
of the velvet-like surface for anterior positions. These differ- surface exhibited noise reduction capabilities up to a maxi-
ences were missed in earlier studies. This shows that it is mum of about 30 dB, depending on the fabric model, sur-
important to take feather position into account when it face structure, flow velocity and frequency. Jaworski (2016)
comes to the characteristics of the velvet-like surface. suggested a model in which filaments of the velvet surface
Apart from the frontal positions, our results suggest a act as inclined fibers that achieve noise reduction by the
high uniformity of the specialized barb modifications on interaction of fiber motion and a moving line vortex.
owl wings. We interpret this uniformity such that an evolu- Although these data demonstrate the noise-reducing
tionary pressure for the creation of a position-dependent capabilities of the velvet-like surface, recently a further func-
morphology existed only for frontal positions. tion has been described. Wind tunnel studies showed
One interesting finding of this study was the change of reduced flow separation for artificial wings covered with
the pennulum length at the most proximal wing positions €n et al. 2009; Winzen et al.
artificial velvet-like surfaces (Kla
investigated, reaching a maximum at 0.15 or 0.2 chord 2012, 2014a,b). Thus, the velvet-like surface has also an aero-
length. This was paralleled by an increase in the number of dynamic effect. Whereas these authors used a uniform vel-
hook radiates for the most medial lengthwise position. vet-like surface on the whole wing, we show here that at
Interestingly, the start of the boundary layer on an owl- anterior and proximal wing positions the pennula are
based wing model was found slightly behind this chordwise shorter. They reach their full length only in wing regions
position (Winzen et al. 2012, 2013, 2014a,b). This leads to where flow separation may occur. Thus, it might be interest-
the speculation that the presence of a velvet-like surface ing to test wings with a velvet-like structure that resembles
may be especially important in regions where flow separa- the positional distribution on natural wings. This suggestion
tion may occur (see also below). is also based on the findings of Winzen et al. (Winzen et al.
Another conspicuous deviation is the increased density of 2014b, 2015; Winzen & Roidl, 2016) that demonstrate an
hook radiates at the frontal part at the distalmost wing even better aerodynamic behavior of natural owl wings
position. This observation reflects the high number of hook than of artificial wings equipped with a velvet-like surface.
radiates that were observed earlier (Bachmann et al. 2007)
on 10th primary feathers. One function of this increased
Conclusions and outlook
density could probably be to stiffen that vane part most
likely subject to the highest forces. Future experiments will Several questions relating to the evolution and function of
have to clarify whether this conclusion is correct. the velvet-like surface remain unanswered. First, if the major
Our findings of the pennulum and hook radiate length advantage of the velvet-like surface was a reduction of
on flight and covert feathers correspond well to the obser- flight noise in combination with passive hearing, why did it
vations reported in Bachmann et al. (2007), which can now evolve in nightjars (Sick, 1937), which mainly hunt insects
be extended to the major part of the wing. that are not expected to perceive flight noise? To our best
We chose to restrict our investigation to uncovered knowledge, nothing is known about the situation in other
feather parts because of large differences in the characteris- night-active birds such as frogmouths, potoos and owlet
tics of the pennula in uncovered and covered areas. For nightjars. Secondly, if the major advantage was a reduction
example, the shape of the pennula in the brownish, uncov- of flow separation, why did the velvet-like surface not
ered feather parts differed from the shape in the covered, evolve in other bird species? A more thorough comparative
whitish feather parts (Bachmann et al. 2007). study may help to answer part of these questions.
Within the owl clade, it is known among ornithologists
that the small, diurnal pygmy owl (Glaucidium passerinum)
Functional aspects of the velvet-like surface
does not possess a velvet-like surface (H-H Bergmann, pers.
The velvet-like surface on owl wings has been related to comm.). We have investigated feathers of the diurnal little
reduction in flight noise. Lilley (1998), for example, pro- owl (Athene noctua) and the nocturnal B. bubo and
posed that the downy surface helps in dampening the aero- observed longer pennula in the three owl species than in C.
dynamic effects that create flight noise. Vad et al. (2006) livia domestica, but shorter pennula in A. noctua than in
demonstrated noise reduction in the human hearing range T. furcata pratincola and B. bubo (M. Weger & H. Wagner,
in an RAF-6E profile with an artificial downy surface that unpubl. obs.). Thus, it would be helpful to collect data on
was inspired by the surface of owl feathers. Chen et al. more owl species.
(2012) conducted noise measurements on feathers of It would also be interesting to make a further comparison
B. bubo, and the common buzzard (Buteo buteo), a of the velvet-like surface on covered and uncovered feather

© 2017 Anatomical Society


14697580, 2017, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/joa.12595 by INASP/HINARI - INDONESIA, Wiley Online Library on [30/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
742 Barb characteristics on barn owl wings, M. Weger and H. Wagner

parts. The velvet-like surface of covered feather parts exhi- Jaworski JW (2016) Vortex sound generation from flexible
bits qualitative differences to this surface and has been fibers. AIAA/CEAS 22nd Aeroacoustics Conference, Lyon,
related to friction noise reduction (Bachmann, 2010) and France. 2014 (p. 2752).
Kla€n S, Bachmann T, Klaas M, et al. (2009) Experimental analysis of
was probably the inspiration for the ‘canopy’ structure used
the flow field over a novel owl based airfoil. Exp Fluids 46, 975–989.
in the experiments of Clark et al. (2014). The differences Kla€n S, Burgmann S, Bachmann T, et al. (2012) Surface structure
between the two different types of surfaces have been and dimensional effects on the aerodynamics of an owl-based
mentioned in previous studies, but a quantitative compar- wing model. Eur J Mech B – Fluids 33, 58–73.
ison and a possible relation to function is missing and Lilley GM (1998) A study of the silent flight of the owl. AIAA
should be examined further in future studies. Paper 4, 1–6.

Mascha E (1904) Uber die Schwungfedern. Z Wiss Zool 77, 606–651.
Sick H (1937) Morphologisch-funktionelle Untersuchungen u € ber
Acknowledgements die Feinstruktur der Vogelfeder. J Ornithol 85, 206–372.
Usherwood JR, Hedrick TL, McGowan CP, et al. (2005) Dynamic
We would like to thank J. Fischoeder for the help on data acquisi- pressure maps for wings and tails of pigeons in slow, flapping
tion and S. Brill for the help with the statistical analysis. flight, and their energetic implications. J Exp Biol 208, 355–369.
Vad J, Koscso G, Gutermuth M, et al. (2006) Study of the aero-
acoustic and aerodynamic effects of soft coating upon airfoil.
Author contributions JSME Int J 49, 648–656.
M.W. and H.W. conceived and designed the experiments. Winzen A, Roidl B (2016) Combined particle-image velocimetry
and force analysis of the three-dimensional fluid-structure
H.W. contributed reagent, materials and analysis tools.
interaction of a natural owl wing. Bioinspir Biomim 11, 026005.
M.W. performed the experiments. M.W. and H.W. analyzed €n S, Klaas M, et al. (2012) High-speed PIV mea-
Winzen A, Kla
the data and wrote the paper. surements of the influence of artificial surface structures on
the near-wall flow field of 3D wing models based on an owl
geometry. 16th Int Symp on Applications of Laser Techniques
Conflict of interest to Fluid Mechanics, Lisbon, Portugal.
The authors declare no conflict of interest. Winzen A, Klaas M, Schro € der W (2013) High-speed PIV measure-
ments of the near-wall flow field over hairy surfaces. Exp Flu-
ids 54, 1–14.
References Winzen A, Klaas M, Schro € der W (2014a) High-speed particle
image velocimetry and force measurements of bio-inspired
Bachmann T (2010) Anatomical, Morphometrical, and Biome- surfaces. J Aircr 52, 471–485.
chanical Studies of Barn Owls’ and Pigeons’ Wings. Aachen: Winzen A, Klaas M, Schro € der W (2014b) PIV measurements com-
RWTH. paring natural and model owl wings. Exp Fluids 46, 975–989.
Bachmann T, Kla €n S, Baumgartner W, et al. (2007) Morphomet- Winzen A, Roidl B, Kla €n S, et al. (2014c) Particle-image
ric characterisation of wing feathers of the barn owl Tyto alba velocimetry and force measurements of leading-edge serra-
pratincola and the pigeon Columba livia. Front Zool 4, 23. tions on owl-based wing models. J Bionic Eng 11, 423–438.
Bachmann T, Mu € hlenbruch G, Wagner H (2011) The barn owl Winzen A, Roidl B, Schro € der W (2015) Particle-image velocime-
wing: an inspiration for silent flight in the aviation industry? try investigation of the fluid-structure interaction mechanisms
In: SPIE Smart Structures and Materials+ Nondestructive Evalu- of a natural owl wing. Bioinspir Biomim 10, 056009.
ation and Health Monitoring. International Society for Optics
and Photonics; 2011. pp. 79750N
Chen K, Liu Q, Liao G, et al. (2012) The Sound Suppression Char- Supporting Information
acteristics of Wing Feather of Owl (Bubo bubo). J Bionic Eng
Additional Supporting Information may be found in the online
9, 192–199.
version of this article:
Clark IA, Devenport W, Jaworski JW, et al. (2014) The Noise
Fig. S1. Statistical differences of the pennulum length between
Generating and Suppressing Characteristics of Bio-Inspired
different measurement positions.
Rough Surfaces. AIAA/CEAS 20th Aeroacoustics Conference,
Fig. S2. Statistical differences of the pennula-base ratio between
Atlanta, GA. 2014.
different measurement positions.
Fatio V (1866) Des diverses modifications dans les formes et la
Fig. S3. Statistical differences of the barb distance between dif-
coloration des plumes. Geneva (Switzerland): Ramboz and
ferent measurement positions.
Schuchardt.
Fig. S4. Statistical differences of the overlap between different
Graham RR (1934) The silent flight of owls. J R Aeronaut Soc 38,
measurement positions.
837–843.
Fig. S5. Statistical differences of the number of hook radiates
Hertel H (1963) Struktur, Form, Bewegung. Mainz: Otto Kraus-
between different measurement positions.
kopf-Verlag.

© 2017 Anatomical Society

You might also like