ArticlePDF Available

Hovering Energetics and Thermal Balance in Anna's Hummingbirds (Calypte anna)

Authors:

Abstract and Figures

We studied the energetics of hover-feeding Anna's hummingbirds, using three different simultaneous techniques: heat loss as estimated via thermal imaging, metabolic rate as measured at a feeder mask using flow-through respirometry, and aerodynamic power estimated from wingbeat kinematic data. These three methods yielded comparable estimates of power output at ambient air temperatures ranging from 18 degrees to 26 degrees C, whereas heat imbalance at higher air temperatures (up to 34 degrees C) suggested loss by mechanisms other than convection and radiation from the body, such as evaporative cooling and enthalpy rise associated with exhaled air and excreted water and convective heat loss from the patagia. Hummingbirds increased wingbeat frequency and decreased stroke amplitude as air temperature increased, but overall muscle efficiency was found to be approximately constant over the experimental range of air temperatures.
Content may be subject to copyright.
Hovering Energetics and Thermal Balance in Anna’s Hummingbirds ( Calypte anna )
Author(s): DennisEvangelista, MaríaJoséFernández, MadalynS.Berns, AaronHoover, and
RobertDudley
Source:
Physiological and Biochemical Zoology,
Vol. 83, No. 3 (May/June 2010), pp. 406-413
Published by: The University of Chicago Press
Stable URL: http://www.jstor.org/stable/10.1086/651460 .
Accessed: 11/12/2013 11:00
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp
.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.
.
The University of Chicago Press is collaborating with JSTOR to digitize, preserve and extend access to
Physiological and Biochemical Zoology.
http://www.jstor.org
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
406
Hovering Energetics and Thermal Balance in
Anna’s Hummingbirds (Calypte anna)
* Corresponding author; e-mail: devangel@berkeley.edu.
Physiological and Biochemical Zoology 83(3):406–413. 2010. 2010 by The
University of Chicago. All rights reserved. 1522-2152/2010/8303-9091$15.00
DOI: 10.1086/651460
Dennis Evangelista
1,
*
Marı´a Jose´ Ferna´ndez
1
Madalyn S. Berns
2
Aaron Hoover
3
Robert Dudley
1,4
1
Department of Integrative Biology, University of California,
Berkeley, California 94720;
2
Department of Bioengineering,
University of California, Berkeley, California 94720;
3
Department of Mechanical Engineering, University of
California, Berkeley, California 94720;
4
Smithsonian Tropical
Research Institute, P.O. Box 2072, Balboa, Republic of
Panama
Accepted 1/20/2010; Electronically Published 3/26/2010
ABSTRACT
We studied the energetics of hover-feeding Anna’s humming-
birds, using three different simultaneous techniques: heat loss
as estimated via thermal imaging, metabolic rate as measured
at a feeder mask using flow-through respirometry, and aero-
dynamic power estimated from wingbeat kinematic data.These
three methods yielded comparable estimates of power output
at ambient air temperatures ranging from 18to 26C, whereas
heat imbalance at higher air temperatures (up to 34C) sug-
gested loss by mechanisms other than convection and radiation
from the body, such as evaporative cooling and enthalpy rise
associated with exhaled air and excreted water and convective
heat loss from the patagia. Hummingbirds increased wingbeat
frequency and decreased stroke amplitude as air temperature
increased, but overall muscle efficiency was found to be ap-
proximately constant over the experimental range of air
temperatures.
Introduction
Among vertebrates, hummingbirds (Trochilidae, Apodiformes)
are among the smallest endotherms and exhibit extremely high
mass-specific metabolic rates (Weis-Fogh 1972; Suarez 1992).
In addition to their small sizes, hummingbirds are the only
birds capable of sustained hovering flight, an energetically de-
manding form of locomotion that is associated with high levels
of metabolic power input (Bartholomew and Lighton 1986;
Suarez et al. 1990; Suarez 1992) and mechanical power output
(Lasiewski 1963; Wolf and Hainsworth 1971; Epting 1980; Wells
1993; Chai and Dudley 1996). Because of their small size and
costly mode of locomotion, hummingbirds represent an im-
portant taxon with which to evaluate maintenance of endo-
thermic balance in the face of environmental challenge (Miller
1996).
Despite their high energetic cost of flight and exposure to
wide fluctuations in ambient air temperature, hummingbirds
maintain energy balance through varied behavioral and phys-
iological mechanisms such as entering nocturnal torpor and
altering foraging strategies according to environmental con-
ditions (e.g., Hixon and Carpenter 1988; Gass and Garrison
1999; Ferna´ndez et al. 2002). Hummingbirds also economize
by substituting heat generated during flight for that required
for thermoregulation (Berger and Hart 1972; Chai et al. 1998),
although the magnitude of this response may vary with body
size (see Welch and Suarez 2008). Variation in wingbeat ki-
nematics in response to variable air temperature may also alter
efficiency so as to advantageously augment metabolic heat pro-
duction (Chai et al. 1998; see also Zerba and Walsberg 1992).
Although hummingbird hovering energetics are now well
studied, the quantitative extent of heat dissipation has not been
evaluated simultaneously for comparison with estimates of met-
abolic and mechanical work. For starlings in forward flight, the
use of infrared thermography has enabled identification of dif-
ferent modes of heat loss as well as an independent estimate
of flight muscle efficiency (Ward et al. 1999). The specific mech-
anisms of heat retention used by hovering hummingbirds, pos-
sibly enabled by variable conductance and the use of thermal
windows, may be of broader ecological and evolutionary rel-
evance given the historical patterns of trochilid diversification
into colder montane habitats (Altshuler and Dudley 2002;
McGuire et al. 2007). To investigate these mechanisms, we an-
alyze simultaneous measurements of surface temperature, met-
abolic rate, and wingbeat kinematics during hovering flight of
Anna’s hummingbirds over a range of ambient air tempera-
tures. By evaluating temperature-dependent changes in kine-
matic, energetic, and efficiency variables, we seek to determine
whether elevated surface temperatures and heat dissipation are
limiting factors in hot air and whether excess heat loss via
convection limits hovering performance in the cold.
Material and Methods
Anna’s hummingbirds were captured in the wild in Berkeley,
California, and were acclimated to laboratory conditions over
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
Hummingbird Hovering Energetics 407
Table 1: Morphological data for three male Anna’s
hummingbirds (Calypte anna)
Individual Mass (g) Wing Length (m) Aspect Ratio
Adult male 4.6 .3 .050 8.5
Adult male 5.0 .0 .050 9.2
Subadult male 4.7 .2 .059 8.3
Note. Average body mass was 4.8 g. Wing length and aspect ratio were
determined from photographs of the spread wings against a grid. Values rep-
resent mean 1 SD.
Figure 1. Representative thermal images showing behavioral response to increased air temperatures. At low to midrange air temperatures (A),
heat loss is primarily from areas around the head, eyes, and pectoral muscles. At high air temperatures (B), heat loss is evident from the entire
body, including the wings. In addition, the feet are extended.
several days. Birds were maintained in individual mesh cages
(90 cm #90 cm #90 cm) with ad lib. access to a commercial
solution designed for nectar-feeding birds (Nektar-plus, Pforz-
heim, Germany). Morphometric data for the three study in-
dividuals (two adult males and one subadult male) are provided
in Table 1. At the completion of the experiments, all birds were
released into the wild, at the point of capture. Measurements
with individual birds were taken in several 4-h periods that
were spread over two and 10 d. Data were considered to be
from separate flight trials if measurements were separated by
at least 15 min of intermittent flight and perching.
Birds were trained to hover at a feeder suspended within a 90
cm #90 cm #90 cm nonhermetically sealed acrylic chamber.
Flight experiments were conducted at each of five nominal air
temperatures (mean SD: , , ,18⬚Ⳳ224.0⬚Ⳳ0.926⬚Ⳳ2
, and C), with the order of presentation30.2⬚Ⳳ0.233.5⬚Ⳳ0.5
chosen randomly. To obtain air temperatures above 24C, a
small convection heater was used to regulate chamber tem-
perature. Ice baths on the chamber floor were used to cool the
chamber below 24C. Chamber air was mixed regularly via the
bird’s periodic flight bouts; air temperature was measured at
the height of the feeder mask, where metabolic and surface
temperature measurements were also obtained. Through the
use of moving visual cues, birds were regularly stimulated to
fly in order to obtain longer hover-feeding bouts and to reduce
transient changes in body temperature at the start of hovering.
Relative humidity during our experiments was 56% 9%
(mean SD); air temperature effects on humidity were non-
significant ( ). Air temperature was monitored duringPp0.17
experiments and varied no faster than 1Cin10min.
Rates of oxygen consumption during hover-feeding were ob-
tained with an open-respirometry system (see Bartholomew
and Lighton 1986; Chai and Dudley 1995). Expired air was
pulled from the nares at a rate of 0.8 L min
1
through a mod-
ified syringe (attached to the feeder) that functioned as a res-
pirometry mask. The air was then drawn through a column of
desiccant (Drierite, Xenia, OH) to remove water vapor. Oxygen
concentration of the airstream was recorded with a portable
oxygen analyzer (Foxbox; Sable Systems International, Las Ve-
gas, NV). Oxygen depletion was estimated as the difference
between baseline and minimum equilibrium values of oxygen
partial pressure, incorporating the rate of airflow and the effect
of ambient humidity. The calculated volume of oxygen con-
sumed was divided by the duration of the feeding bout to obtain
metabolic rate. A minimum of five hover-feeding bouts per
individual bird were obtained at each experimental air tem-
perature; feeding bouts of less than 2 s in duration were dis-
carded. A standard conversion factor of 20.1 J mL O
2
1
was
assumed.
To determine the rate of heat loss from hovering birds, in-
frared (IR) thermal images of the hummingbirds were obtained
using a thermal imaging camera (Fluke, Everett, WA) operated
through a window cut in one wall of the flight chamber. Hov-
ering birds were filmed from a lateral perspective while at the
feeder (see Fig. 1); orthogonal perspectives were then obtained
by rotating the feeder about vertical through 90and repeating
the procedure. An emissivity eof 0.95 for feathers was assumed
(Cossins and Bowler 1987). From the thermal images, the re-
gion of interest containing the hummingbird was first identified
manually. The image was then cropped and an optimal thresh-
old was chosen to segment the background from theforeground
(see Otsu 1979). Mean surface temperature was computed as
the area-weighted average of temperature pixels in the cropped
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
408 D. Evangelista, M. J. Ferna´ndez, M. S. Berns, A. Hoover, and R. Dudley
Figure 2. A, Life-sized (4-cm-long) physical model was used to obtain heat transfer coefficients at typical induced velocities and surface
temperatures. Average surface temperatures were determined using the same thermal imaging camera and algorithm as for the live birds. B,
Representative surface temperature profile using infrared thermal imaging (red corresponds to 37C, blue corresponds to 20C). Variation in
the surface temperatures is due primarily to model grazing angle effects, which are identical to those in the live bird because they are in the
same position.
Figure 3. Metabolic rate measurements (in W) for all birds at
˙
Q
metabolic
all temperatures. for all birds is roughly constant at
˙
Q1.1
metabolic
W, or about mL O
2
g
1
h
1
(mixed-mode regression,0.1 42 5
; individual effects not significant, ). For all re-Pp0.17 Pp0.10
maining results, power is expressed in Watts to allow side-by-side
comparison of metabolic rates, rates of heat loss, and mechanical
power.
image. The wings are ignored in the heat-transfer calculations,
as the thermal imaging camera could not reliably resolve the
fast-moving wings. In “Discussion,” we bound the potential
magnitude of heat loss from the wings on the basis of
convection.
The net rate of heat loss to ambient, (in W), was de-
˙
Q
loss
termined from both convective and radiative terms by the fol-
lowing equation (see Incropera and Dewitt 1996; Ward et al.
1999):
44
¯
˙
QphA(TT)jeA[(T273) (T273) ], (1)
loss s a s a
where is the overall average convective heat transfer coeffi-
¯
h
cient, W m
2
K
4
is the Stefan-Boltzmann
8
jp5.67 #10
constant, and T
s
is the surface temperature and T
a
is the ambient
air temperature (both temperatures in C). This formulation
ignores evaporative heat loss and assumes that the surrounding
environmental temperature for radiative heat transfer is the air
temperature. Thermal images suggest that surrounding envi-
ronmental temperatures are close to ambient air temperatures.
In any case, radiative heat transfer makes up only about 5%
of the heat loss calculated with equation (1). To estimate the
heat transfer coefficient , a life-sized physical model of an
¯
h
Anna’s hummingbird (see Fig. 2) was constructed of a solid
piece of polymer clay (Sculpey; Polyform Products, Elk Grove
Village, IL) wrapped with 30-gauge nichrome wire and paper
tape (Johnson and Johnson, Langhorne, PA) characterized by
an emissivity eof 0.95 (Incropera and Dewitt 1996). The model
was heated by connecting a power supply to the nichrome wire,
yielding an estimated surface heat flux that is based on the
model’s surface area and the applied voltage and current within
the wire. Convective flow was imposed on the model using a
large fan operated at air speeds comparable to estimated values
of the induced velocity for hummingbirds (4.4–4.6 m s
1
, cal-
culated according to Ellington 1984b; values from 3 to 6 m s
1
were tested). Flow was measured with a hot-wire anemometer
(Kurz Instruments, Monterey, CA), sampling at 25 Hz, located
four body lengths from the model and parallel within the work-
ing section. Flow direction was effectively downward relative
to the model oriented in an appropriate feeding position (see
Fig. 2B). The aforementioned thermal imaging camera wasthen
used to obtain surface temperatures, with the model placed at
the feeder in the same orientation as a hovering bird. Measured
heat transfer coefficients were within a factor of 2 of that for
a sphere, indicating reasonable estimates for the nonstandard
hummingbird geometry (Incropera and Dewitt 1996; Ward et
al. 1999). The model test results were used to estimate con-
vective heat loss for each hovering trial, using the model’s mean
surface temperature and an induced airflow velocity based on
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
Hummingbird Hovering Energetics 409
Figure 4. A, Average surface temperature, T
s
, determined from thermal
images. B, Rate of heat loss, , for all birds and all temperatures.
˙
Q
loss
T
s
increases significantly with ambient air temperature T
a
(Tp
s
, ; individual effects not significant, ).0.78T9.32 P0.0001 Pp0.17
a
declines significantly with ambient air temperature as the driving
˙
Q
loss
temperature difference, is reduced (mixed-mode regression,TT
sa
, ; individual effects not significant,
˙
Qp0.06T2.05 P!0.0001
loss a
). Blue arrow (at !20C; a color version of this figure isPp0.34
available in the online edition of Physiological and Biochemical Zoology)
indicates onset of feather fluffing while individual is perched, as ob-
served in bird 2. Red arrow (at 130C) indicates onset of wing spread-
ing, mouth gaping, and foot extension, as observed in all birds.Symbols
are as in Figure 3.
Figure 5. Overall average heat transfer coefficient, , for the physical
¯
h
model of Calypte anna is shown as a function of surface temperature,
T
s
, and airflow velocity ( , ). The
¯
hp137 3T13.5uPp0.018
sind
star indicates the average of all values used in calculations forhovering
hummingbirds (see text). A color version of this figure is available in
the online edition of Physiological and Biochemical Zoology.
aerodynamic calculations. For all trials, induced velocity based
on kinematics was m s
1
.4.5 0.2
To obtain kinematic data for estimates of mechanical power
output, hover-feeding hummingbirds were filmed ventrally
with a high-speed digital video camera (AOS Technologies, Ba-
den Daettwil, Switzerland) operated at 500 frames s
1
. Video
sequences were analyzed frame by frame to obtain values of
wingbeat frequency and stroke amplitude (see Chai and Dudley
1995, 1996). Air density was determined from measurements
of barometric pressure. Body mass of individual birds was mea-
sured before and after each experimental series; the mean value
was used in aerodynamic calculations. Outstretched wings of
birds were photographed against graph paper and then digitized
using ImageJ (National Institutes of Health, Bethesda, MD) to
obtain morphological parameters relating to wing planform
(see Ellington 1984a).
Kinematic and morphological data were used to calculate the
mechanical power output using a standard model of animal
hovering (Ellington 1984b) modified to incorporate unsteady
drag coefficients as measured on a hummingbird wing in con-
tinuous rotation (Altshuler et al. 2004). Stroke plane angle was
assumed to equal 0, and simple harmonic motion was assumed
for wing movements within the stroke plane (see Chai and
Dudley 1995; Altshuler and Dudley 2003).
The mechanical power output required to hover is the sum
of the power required to overcome profile drag andthe induced
power required for weight support. Net inertial power to ac-
celerate the wings is taken to be 0 on the basis of the assumption
that the hummingbird flight apparatus exhibits full elastic stor-
age of wing inertial energy (Ellington 1984c). Drag on the body
during hovering flight was similarly assumed to be small com-
pared with profile power lost to drag on the fast-moving wings
(Ellington 1984c). Induced power, the power required to sup-
port body weight, follows from momentum balance and was
calculated following Ellington (1984c), as was profile power
using a wing drag coefficient (see Altshuler et
Cp0.139
D, pro
al. 2004).
Heat balance for the hovering hummingbird at thermal equi-
librium is given by
˙˙
˙
QpQW,(2)
metabolic loss mechanical
where positive is the rate of metabolic heat production
˙
Q
metab olic
(as measured via respirometry), positive is the rate of heat
˙
Q
loss
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
410 D. Evangelista, M. J. Ferna´ndez, M. S. Berns, A. Hoover, and R. Dudley
Figure 6. A, Wingbeat frequency decreases with ambient air temper-
ature T
a
(individual effects significant, ). B, Stroke amplitudeP!0.001
increases with ambient air temperature (individual effects significant,
). A color version of this figure is available in the onlineP!0.001
edition of Physiological and Biochemical Zoology.
Figure 7. Mechanical power output required to hover, , is
˙
W
mechanical
roughly constant at W (mixed-mode regression,0.11 0.01 Pp
; individual effects not significant, ). Mass-specific me-0.2611 Pp0.17
chanical power is approximately 4.1 mL O
2
h
1
g
1
. Symbols are as in
Figure 3.
loss to the surrounding air, and positive is the me-
˙
W
mechan ical
chanical work output as estimated aerodynamically. We rep-
resent the overall efficiency h(Josephson et al. 2001) as
˙
W
mechan ical
hp,(3)
1
˙
Q
metab olic
where h
1
is efficiency based on mechanical power estimates and
respirometric measurements of metabolic power. Alternatively,
we compute h
2
as follows using equation (2) for metabolic
power estimates:
˙
W
mechan ical
hp.(4)
2
˙
˙
WQ
mechan ical loss
Purely on the basis of thermodynamics, h
2
can be viewed as
an “efficiency,” but measurements of may miss some heat
˙
Q
loss
loss terms. Discrepancies between these two estimates indicate
supplemental avenues of heat dissipation (such as evaporative
loss) not represented in equation (1).
The effects of variable air temperature on rate of heat loss,
metabolic rate, wingbeat kinematics, mechanical power, heat
balance, and muscle efficiency were evaluated using mixed-
model regression in JMP (SAS Institute, Cary, NC) unless oth-
erwise noted. In all analyses, individual effects and interactions
(e.g., metabolic rate as a function of bird, bird #temperature,
and temperature) were checked by ANOVA. In all cases except
wingbeat frequency and stroke amplitude, which are discussed
below, individual effects were found to be nonsignificant and
so data were pooled in subsequent analyses.
Results
Metabolic rates of hover-feeding hummingbirds averaged
W (mean SD) and were independent of ambient1.1 0.1
air temperature ( ; see Fig. 3; individual effects notPp0.17
significant, ). The measured metabolic rates corre-Pp0.10
spond to a mass-specific metabolic rate of mL O
2
g
1
42 5
h
1
. By contrast, surface temperature (T
s
) of hovering Calypte
anna increased significantly with ambient air temperature T
a
(Fig. 4A; , ; individual effects notTp0.78T9.32 P0.0001
sa
significant, ). Correspondingly, the estimated rate ofPp0.17
heat loss to the environment by convective and radiative heat
transfer declined significantly with air temperature (Fig. 4B;
, ; individual effects not sig-
˙
Qp0.06T2.05 P!0.0001
loss a
nificant, ).Pp0.34
The overall average heat transfer coefficient, , as measured
¯
h
on the physical model of a hummingbird, varied with both
surface temperature and airflow velocity (Fig. 5; ¯
hp137
, , ). The model test results
2
3.0T13.5urp0.63 Pp0.018
sind
were used to estimate convective heat loss for each hovering
trial on the basis of mean surface temperature of the bird and
an induced airflow velocity on the basis of aerodynamic cal-
culations. To determine whether the small changes in un-
¯
h
derlay observed trends, we also calculated convective heat loss
with a constant ; estimates differed slightly but do not alter
¯
h
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
Hummingbird Hovering Energetics 411
Table 2: Heat balance terms as a function of air
temperature
T
a
(C) ˙
Q
metabolic
˙
W
mechanical
˙
Q
loss
18 2 1.0 .1 .11 .01 .9 .3
24.0 .9 1.0 .1 .12 .01 .5 .2
26 2 1.1 .1 .11 .01 .6 .2
30.2 .2 1.1 .1 .11 .01 .5 .1
33.5 .5 1.0 .1 .111 .003 .4 .2
Note. All values are in Watts unless otherwise noted, and are
means SD. In mass-specific form, metabolic rates are approxi-
mately mL O
2
h
1
g
1
and mechanical power estimates are42 5
about mL O
2
h
1
g
1
.4.1 0.5
Table 3: Efficiency estimates as a function of
air temperature
T
a
(C) h
1
h
2
hh
12
18 2 .11 .02 .109 .001
24.0 .9 .12 .02 .194 .074
26 2 .10 .02 .155 .055
30.2 .2 .10 .02 .180 .080
33.5 .5 .10 .01 .217 .117
Note. h
1
efficiencies are computed using hp
1
.h
2
estimates are computed using
˙
˙
W/Qhp
mechanical metabolic 2
.
˙
˙˙
W/(WQ)
mechanical mechanical loss
our conclusions. For all trials, the average heat transfer coef-
ficient was 107 W m
2
K
1
on the basis of average surface
temperatures of C and average induced airflow veloc-
29⬚Ⳳ4
ities of m s
1
.
4.5 0.2
Wingbeat frequencies decreased with ambient air tempera-
ture by about 10% over the experimental range (Fig. 6A;sig-
nificant individual effects, ). By contrast, stroke am-
P!0.001
plitudes increased with ambient air temperature by about 10%
over the same range (Fig. 6B; significant individual effects,
). Mechanical power output was, however, indepen-
P!0.001
dent of air temperature (Fig. 7; , individual effects not
Pp0.26
significant, ), and it averaged W (or, in
Pp0.17 0.11 0.01
mass-specific form, mL O
2
g
1
h
1
).
4.1 0.5
At low air temperatures, metabolic heat generation was ap-
proximately equal to the summed rate of heat loss via radiation
and convection and the estimated mechanical power output
(Table 2). At midrange and high temperatures, however, the
terms did not balance, demonstrating additional modes of heat
loss, such as evaporative cooling, enthalpy rise of excreted water,
and avenues of convective heat loss not included in equation
(1). Overall efficiencies calculated for hover-feeding on the basis
of either aerodynamic power estimates and respirometry (h
1
)
or aerodynamic power and radiative and convective heat loss
(h
2
) at low temperatures were about 10% (Table 3). The h
1
efficiencies of 10% are comparable to results from other studies
of hovering hummingbirds (9%–11%; see Wells 1993; Chai et
al. 1998). At high air temperatures, h
2
estimates (15%–20%)
are somewhat higher than efficiencies based on aerodynamic
power (h
1
), indicating unaccounted modes of heat dissipation
in the overall energy balance (eq. [1]).
We observed several behaviors corresponding to the loss of
the driving temperature differential at higher air temperatures.
At air temperatures of 18–24C, the main areas of heat loss
were found around the head, eye, and pectoral muscles (Fig.
1A). Above 30C, these regions expanded to include the entire
body, as seen in thermal images (see Fig. 1B), potentially cor-
responding to an increase in blood perfusion over the entire
body. Above 30C, all birds exhibited bill gaping while perched
and one bird exhibited wing spreading while perched. Birds
were also observed to extend their feet while hovering at high
air temperatures (see Fig. 1B).
Discussion
At low air temperatures, application of thermal imaging, aero-
dynamic estimates, and respirometric measurements yielded
consistent results for the energetics of hovering flight. By con-
trast, use of infrared thermography at higher ambient air tem-
peratures underestimated heat lost to the environment because
of unaccounted modes of heat transfer. At higher air temper-
atures, the primary mechanisms available for dissipating heat
are evaporative cooling and enthalpy rise in either exhaled air
or excreted water (Lasiewski 1964; Powers 1992; Lotz et al.
2003). In addition, heat loss from the wings, which have rel-
atively large surface areas and higher convective heat transfer
coefficients due to high flapping velocities, may become rela-
tively more important at higher temperatures.
As ambient air temperatures approach body temperature, the
driving thermal gradient ( ) for convection and radiativeTT
sa
loss is progressively diminished. At high ambient air temper-
atures, flight performance may then become thermally limited,
as is suggested by estimated rates of convective and radiative
heat loss that extrapolate to zero heat loss at approximately
40C (Fig. 4). This diminished shedding of heat is further com-
pounded by a slight decline in the heat transfer coefficient with
increased air temperature (Fig. 5). At temperatures higher than
its thermal neutral zone, Calypte anna may become slightly
hyperthermic, which could augment heat transfer by increasing
the gradient (Powers 1992).TT
sa
Whereas our calculations do not include such heat transfer
modes as evaporative cooling, enthalpy rise, and convection
from the wings, the measurements taken here provide an in-
direct way to estimate these terms. At low temperatures, these
additional heat loss terms, taken as ˙˙
QW
metabolic mechanical
, are negligible ( W). At the highest air temper-
˙
Q0.0 0.3
loss
ature, these additional terms would have to be W,0.5 0.2
almost half of the metabolic rate. It is interesting to consider
how this could be split among evaporative water loss, enthalpy
rise of ingested liquid, and wing convection. Powers (1992) and
Lasiewski (1964) give evaporative water loss rates for hum-
mingbirds of around 20 mg g
1
h
1
. Assuming that 2,257 kJ is
needed to vaporize 1 kg of water, the associated evaporative
heat loss would be about 0.06 W, corresponding to 6% of the
metabolic rate measured here. The energy needed to warm
ingested nectar could also be significant (Lotz et al. 2003).
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
412 D. Evangelista, M. J. Ferna´ndez, M. S. Berns, A. Hoover, and R. Dudley
Ferna´ndez et al. (2002) give fluid ingestion rates of up to 0.03
g min
1
for a 6-g green-backed firecrown hummingbird (Se-
phanoides sephanoides). The energy required to raise the tem-
perature of such nectar volumes from 20C to typical body
temperatures would be about 0.04 W, or 4% of the measured
metabolic rate for Anna’s hummingbirds. To estimate wing
convection, we assume a convective heat transfer coefficient
for the wings of 200 W m
2
K
1
, on the basis of forced con-
vection in air on a flat plate at a local velocity of 15 m s
1
(Incropera and Dewitt 1996). The patagia are the main vas-
cularized areas of the wings, and the heat transfer surface areas
for the dorsal and ventral surfaces of both patagia are approx-
imately m
2
. For wing temperatures equal to the ob-
4
2.6 #10
served body surface temperatures, convective heat loss from
the wings would be about 0.4 W. As with the body, heat loss
from the wings will become restricted as ambient air temper-
ature increases.
Limited heat dissipation in high air temperatures may impose
constraints on hovering performance. Hummingbirds in hot,
humid conditions in the tropics may raise their body temper-
atures in order to increase convective heat transfer when evap-
orative cooling is limited (see Powers 1992), although further
ecological and behavioral data are needed. Field observations
of the giant Andean hummingbird (Patagona gigas) indicate
decreased activity on hot days (M.J. Ferna´ndez, personal ob-
servation). Mechanisms of heat dissipation will be further lim-
ited under humid conditions, as enthalpy rise associated with
exhalation declines with increasing water content of inhaled
air. For hummingbirds with relatively high ventilatory fre-
quencies, the magnitude of respiratory cooling is potentially
substantial. The fraction of metabolic heat dissipated by evap-
oration in Anna’s hummingbirds is lower than that of other
birds in dry air, but it exceeds that of other birds at high
humidities when air temperature is less than 33C (Powers
1992). Experimental manipulation of both air temperature and
relative humidity for hovering hummingbirds would defini-
tively test the role of thermal constraints on flight performance.
Convection is expected to be the dominant mode of heat
loss at low and midrange air temperatures. However, variation
in wingbeat frequency and stroke amplitude at low ambient air
temperature (see Fig. 6) does not significantly alter either in-
duced velocity or wing relative velocity, which otherwise might
change the convective heat transfer coefficient. Variation in
estimated rates of heat loss with air temperature (Fig. 4B) sug-
gests that metabolic rates must increase to compensate for in-
creased heat loss below 15C, as is observed for ruby-throated
hummingbirds hovering at 5–15C (Chai et al. 1998). At even
lower air temperatures, below the thermal neutral zone, heat
loss becomes large compared with rates of metabolic heat pro-
duction. For example, we observed substantial feather fluffing
by perched birds in very cold air. Studies of otherhummingbird
species have documented increases in foraging and food intake
to maintain energy balance at low air temperatures (Gass et al.
1999). At high elevations in the Andes, hummingbirds often
fly and forage at near-freezing air temperatures; associated
physiological responses may have been an important factor
influencing montane colonization by this lineage.
In spite of significant changes in wingbeat frequency and
stroke amplitude over the experimental range of air tempera-
tures (Fig. 6), we observed no significant changes in hovering
metabolic rates or mechanical power expenditure at the lowest
temperatures used in our experiments (Fig. 3; Table 2). Altered
kinematics could potentially generate more heat at the expense
of mechanical efficiency (Ivanov 1989; Full et al. 1998; Jo-
sephson et al. 2001; Bicudo et al. 2002). Some indication of
this effect was evident at the lowest tested air temperatures, for
which heat loss approached metabolic rate (see Figs. 3, 4B; see
also Chai et al. 1998). Flight efficiency, however, remained re-
markably constant at about 10% (see Table 3). These efficiency
estimates are comparable to those derived in other studies of
hovering hummingbirds (9%–11%; see Wells 1993; Chai et al.
1998). Kinematic variation without reduction in efficiency of
hovering may also be of benefit when hummingbirds visit dif-
ferent floral morphologies (Wells 1993).
The thermal limits on flight performance suggested here
likely influence hummingbird evolution across altitudinal gra-
dients (McGuire et al. 2007), and they may also impinge on
size-based trade-offs in flight performance (see Chai and Dud-
ley 1999; Altshuler and Dudley 2002). The most thermally chal-
lenging conditions for trochilids are likely to involve hovering
in hot and humid air (resulting in insufficient heatdissipation),
as well as forward flight in cold, dry air (which may result in
excess heat loss). Fieldwork assessing flight behaviors in relation
to microclimatic data would enable assessment of the specific
role of thermal limits in influencing ecological distributions of
hummingbirds. Similarly, comparative work on thermal re-
sponses by differently sized species would assess the allometry
of heat production and loss during hovering flight.
Acknowledgments
We thank C. Clark for capturing birds and providing a skin
specimen for the construction of the thermal model. We also
thank R. Full, M. Koehl, Y. Munk, S. Sponberg, and T. Libby
for assistance, equipment, and comments as a part of Berkeley’s
Mechanics of Organisms Lab class (IB 135L) within the Center
for Integrative Biomechanics Education and Research (CIBER).
Several anonymous reviewers provided comments that were
helpful. D.E. was supported by a National Science Foundation
(NSF) graduate fellowship. M.J.F. was supported by a Fulbright
fellowship. Y. Munk provided useful comments on the man-
uscript. Laboratory support was provided by NSF grant DEB-
0543556.
Literature Cited
Altshuler D.L. and R.A. Dudley. 2002. The ecological and evo-
lutionary interface of hummingbird flight physiology. J Exp
Biol 205:2325–2337.
———. 2003. Kinematics of hovering hummingbird flight
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
Hummingbird Hovering Energetics 413
along simulated and natural elevational gradients. J Exp Biol
206:3139–3147.
Altshuler D.L., R. Dudley, and J.A. McGuire. 2004. Resolution
of a paradox: hummingbird flight at high elevation does not
come without a cost. Proc Natl Acad Sci USA 101:17731–
17736.
Bartholomew G.A. and J.R.B. Lighton. 1986. Oxygen con-
sumption during hover feeding in free-ranging Anna’s hum-
mingbirds. J Exp Biol 123:191–199.
Berger M. and J.S. Hart. 1972. Die Atmung beim Kolibri Ama-
zilia fimbriata wa¨hrend des Schwirrfluges bei verschiedenen
Umgebungstemperaturen. J Comp Physiol A 81:363–380.
Bicudo J.E.P.W., A.C. Bianco, and C.R. Vianna. 2002. Adaptive
thermogenesis in hummingbirds. J Exp Biol 205:2267–2273.
Chai P., A. Chang, and R.A. Dudley. 1998. Flight thermogenesis
and energy conservation in hovering hummingbirds. J Exp
Biol 201:963–968.
Chai P. and R. Dudley. 1995. Limits to vertebrate locomotor
energetics suggested by hummingbirds hovering in heliox.
Nature 377:722–725.
———. 1996. Limits to flight energetics of hummingbirds hov-
ering in hypodense and hypoxic gas mixtures. J Exp Biol
199:2285–2295.
———. 1999. Maximum flight performance of hummingbirds:
capacities, constraints and trade-offs. Am Nat 153:398–411.
Cossins A.R. and K. Bowler. 1987. Temperature Biology of An-
imals. Chapman & Hall, London.
Ellington C.P. 1984a. The aerodynamics of hovering insect
flight. II. Morphological parameters. Philos Trans R Soc B
305:17–40.
———. 1984b. The aerodynamics of hovering insect flight. V.
A vortex theory. Philos Trans R Soc B 305:115–144.
———. 1984c. The aerodynamics of hovering insect flight. VI.
Lift and power requirements. Philos Trans R Soc B 305:145–
181.
Epting R.J. 1980. Functional dependence of the power for hov-
ering on wing disc loading in hummingbirds. Physiol Zool
53:347–357.
Ferna´ndez M.J., M.V. Lo´pez-Calleja, and F. Bozinovic. 2002.
Interplay between the energetics of foraging and thermoreg-
ulatory costs in the green-backed firecrown hummingbird
Sephanoides sephanoides. J Zool 258:319–326.
Full R.J., D. Stokes, A. Ahn, and R. Josephson. 1998. Energy
absorption during running by leg muscles in a cockroach. J
Exp Biol 201:997–1012.
Gass C.L. and J.S.E. Garrison. 1999. Energy regulation by trap-
lining hummingbirds. Funct Ecol 13:483–492.
Gass C.L., M.T. Romich, and R.K. Suarez. 1999. Energetics of
hummingbird foraging at low ambient temperature. Can J
Zool 77:314–320.
Hixon M.A. and F.L. Carpenter. 1988. Distinguishing energy
maximizers from time minimizers: a comparative study of
two hummingbird species. Integr Comp Biol 28:913–925.
Incropera F.P. and D.P. Dewitt. 1996. Fundamentals of Heat
and Mass Transfer. Wiley, New York.
Ivanov K.P. 1989. Thermoregulatory chemical metabolism and
muscle work efficiency. J Therm Biol 14:1–18.
Josephson R.K., J.G. Malamud, and D.R. Stokes. 2001. The
efficiency of an asynchronous flight muscle from a beetle. J
Exp Biol 204:4125–4139.
Lasiewski R.C. 1963. Oxygen consumption of torpid, resting,
active and flying hummingbirds. Physiol Zool 36:122–140.
———. 1964. Body temperatures, heart and breathing rate,
and evaporative water loss in hummingbirds. Physiol Zool
37:212–223.
Lotz C.N., C. Martı´nez del Rio, and S.W. Nicholson. 2003.
Hummingbirds pay a high cost for a warm drink. J Comp
Physiol B 173:455–461.
McGuire J.A., C.C. Witt, D.L. Altshuler, and J.V. Remsen. 2007.
Phylogenetic systematics and biogeography of humming-
birds: Bayesian and maximum likelihood analyses of parti-
tioned data and selection of an appropriate partitioning strat-
egy. Syst Biol 56:837–856.
Miller P.J., ed. 1996. Miniature Vertebrates: The Implications
of Small Body Size. Oxford University Press, Oxford.
Otsu N. 1979. A threshold selection method from gray level
histograms. IEEE Trans Syst Man Cybern 9:62–66.
Powers D.R. 1992. Effect of temperature and humidity on evap-
orative water loss in Anna’s hummingbird (Calypte anna).
J Comp Physiol B 162:74–84.
Suarez R.K. 1992. Hummingbird flight: sustaining the highest
mass-specific metabolic rates among vertebrates. Experientia
48:565–569.
Suarez R.K., J.R.B. Lighton, C.D. Moyes, G.S. Brown, C.L.Gass,
and P.W. Hockachka. 1990. Fuel selection in rufous hum-
mingbirds: ecological implications of metabolic biochemis-
try. Proc Natl Acad Sci USA 87:9207–9210.
Ward S., J.M.V. Rayner, U. Mo¨ller, D.M. Jackson, W. Nachtigall,
and J.R. Speakman. 1999. Heat transfer from starlings Stur-
nus vulgaris during flight. J Exp Biol 202:1589–1602.
Weis-Fogh T. 1972. Energetics of hovering flight in humming-
birds and in Drosophila. J Exp Biol 56:79–104.
Welch K.C. and R.K. Suarez. 2008. Altitude and temperature
effects on the energetic cost of hover-feeding in migratory
rufous hummingbirds, Selasphorus rufus. Can J Zool 86:161–
169.
Wells D. 1993. Muscle performance in hovering hummingbirds.
J Exp Biol 178:39–57.
Wolf L.L. and F.R. Hainsworth. 1971. Time and energy budget
of territorial hummingbirds. Ecology 52:980–988.
Zerba E. and G.E. Walsberg. 1992. Exercise-generated heatcon-
tributes to thermoregulation by Gambel’s quail in the cold.
J Exp Biol 171:409–422.
This content downloaded from 152.2.15.253 on Wed, 11 Dec 2013 11:00:44 AM
All use subject to JSTOR Terms and Conditions
... T s − T a = 5-8°C in hummingbirds; [9]) and other areas of the body. Three main HDAs have been identified during flight in birds: (i) the head, in particular the unfeathered hot spot eye area, (ii) the feet, and (iii) the proximal wing with the hot spot around the shoulder area [9][10][11]. In addition, it appears that in species with relatively large bills, the bill is an important HDA after flight, accounting for 35-60% of total heat exchange in toco toucans (Ramphastos toco; [12]), 10-18% in tufted puffins (Fratercula cirrhata; [13]) and 1.4-19.9% in southern yellow-billed hornbills (Tockus leucomelas; [14]). ...
... Birds were flown at up to two randomly selected speeds per day (selected from 5,6,7,8,9,10,11,12,13,14 and 15 m s −1 which covered speeds both below and above the minimum power speed (A Lewden, CM Bishop and GN Askew 2018, unpublished data)), for a mean duration of 179 ± 9 s (± s.e. with a minimum flight duration of 144 s and a maximum flight duration of 547 s). Birds performed between 2 and 16 flights at different speeds with a mean of 11 flights per bird and there were no differences in flight speeds between birds (ANOVA; p = 0.791) and no individual flew longer than the other (ANOVA; p = 0.115). ...
Article
Full-text available
Animal flight uses metabolic energy at a higher rate than any other mode of locomotion. A relatively small proportion of the metabolic energy is converted into mechanical power; the remainder is given off as heat. Effective heat dissipation is necessary to avoid hyperthermia. In this study, we measured surface temperatures in lovebirds (Agapornis personatus) using infrared thermography and used heat transfer modelling to calculate heat dissipation by convection, radiation and conduction, before, during and after flight. The total non-evaporative rate of heat dissipation in flying birds was 12× higher than before flight and 19× higher than after flight. During flight, heat was largely dissipated by forced convection, via the exposed ventral wing areas, resulting in lower surface temperatures compared with birds at rest. When perched, both before and after exercise, the head and trunk were the main areas involved in dissipating heat. The surface temperature of the legs increased with flight duration and remained high on landing, suggesting that there was an increase in the flow of warmer blood to this region during and after flight. The methodology developed in this study to investigate how birds thermoregulate during flight could be used in future studies to assess the impact of climate change on the behavioural ecology of birds, particularly those species undertaking migratory flights.
... Low mechanical efficiency of flight muscles results in substantial heat generation during flight in birds [1,2] that must be dissipated to avoid hyperthermia [3,4]. Dissipating excess heat produced during flight is challenging for birds because feathers provide an insulative layer that restricts heat loss, particularly at slow flight speeds when forced convection is low [3][4][5][6]. The degree to which plumage restricts evaporative heat dissipation during flight is unknown, but will probably not be a substantial barrier [7]. ...
... The degree to which plumage restricts evaporative heat dissipation during flight is unknown, but will probably not be a substantial barrier [7]. However, we know that birds passively dissipate heat (radiation, conduction and convection) through specific heat dissipation areas (HDAs) around the eyes, shoulder and feet/legs where plumage density is low, thereby exposing the skin [3,4,6]. In calliope hummingbirds (Selasphorus calliope), these HDAs allow sufficient passive heat dissipation for maintenance of heat balance at a moderate environmental temperature (21°C) across flight speeds in the range 0-12 m s −1 [3]. ...
Article
Full-text available
At high temperature (greater than 40°C) endotherms experience reduced passive heat dissipation (radiation, conduction and convection) and increased reliance on evaporative heat loss. High temperatures challenge flying birds due to heat produced by wing muscles. Hummingbirds depend on flight for foraging, yet inhabit hot regions. We used infrared thermography to explore how lower passive heat dissipation during flight impacts body-heat management in broad-billed (Cynanthus latirostris, 3.0 g), black-chinned (Archilochus alexandri, 3.0 g), Rivoli’s (Eugenes fulgens, 7.5 g) and blue-throated (Lampornis clemenciae, 8.0 g) hummingbirds in southeastern Arizona and calliope hummingbirds (Selasphorus calliope, 2.6 g) in Montana. Thermal gradients driving passive heat dissipation through eye, shoulder and feet dissipation areas are eliminated between 36 and 40°C. Thermal gradients persisted at higher temperatures in smaller species, possibly allowing them to inhabit warmer sites. All species experienced extended daytime periods lacking thermal gradients. Broad-billed hummingbirds lacking thermal gradients regulated the mean total-body surface temperature at approximately 38°C, suggesting behavioural thermoregulation. Blue-throated hummingbirds were inactive when lacking passive heat dissipation and hence might have the lowest temperature tolerance of the four species. Use of thermal refugia permitted hummingbirds to tolerate higher temperatures, but climate change could eliminate refugia, forcing distributional shifts in hummingbird populations.
... The avian flying muscles introduce the most energetically expensive muscle work with the highest mass-specific metabolic rates in vertebrates, in comparison to exercising mammals, flapping is energetically more costly than running [6,9]. To cope with these demands, several mechanisms have been described. ...
Article
Full-text available
Background The currently known homing pigeon is a result of a sharp one-sided selection for flight characteristics focused on speed, endurance, and spatial orientation. This has led to extremely well-adapted athletic phenotypes in racing birds. Methods Here, we identify genes and pathways contributing to exercise adaptation in sport pigeons by applying next-generation transcriptome sequencing of m.pectoralis muscle samples, collected before and after a 300 km competition flight. Results The analysis of differentially expressed genes pictured the central role of pathways involved in fuel selection and muscle maintenance during flight, with a set of genes, in which variations may therefore be exploited for genetic improvement of the racing pigeon population towards specific categories of competition flights. Conclusions The presented results are a background to understanding the genetic processes in the muscles of birds during flight and also are the starting point of further selection of genetic markers associated with racing performance in carrier pigeons.
... To the best of our knowledge, the only studies measuring the effect of temperature on kinematics in mammals or birds focus on hummingbirds. These studies find that despite maintaining similar flapping frequencies, hummingbirds decrease stroke amplitude at cold temperatures (Chai et al., 1998;Evangelista et al., 2010). By nature of regulating internal temperatures, endothermic animals are less likely to demonstrate temperature-dependent locomotion. ...
Article
Locomotor biomechanics faces a core trade-off between laboratory-based and field-based studies. Laboratory conditions offer control over confounding factors, repeatability, and reduced technological challenges, but limit the diversity of animals and environmental conditions that may influence behavior and locomotion. This article considers how study setting influences the selection of animals, behaviors and methodologies for studying animal motion. We highlight the benefits of both field- and laboratory-based studies and discuss how recent work leverages technological advances to blend these approaches. These studies have prompted other subfields of biology, namely evolutionary biology and ecology, to incorporate biomechanical metrics more relevant to survival in natural habitats. The concepts discussed in this Review provide guidance for blending methodological approaches and inform study design for both laboratory and field biomechanics. In this way, we hope to facilitate integrative studies that relate biomechanical performance to animal fitness, determine the effect of environmental factors on motion, and increase the relevance of biomechanics to other subfields of biology and robotics.
... Open-flow, positive pressure Non-BMR conditions (field and laboratory) Lasiewski 1963a, b;Withers 1977a, b;Schuchmann 1979a, b;Schuchmann et al. 1979;Schuchmann and Schmidtmarloh 1979a, b;Powers 1991Powers , 1992Prinzinger et al. 1992;Chaui-Berlinck et al. 2002;Powers et al. 2010;Dick et al. 2020 Oxygen consumption and carbon dioxide production (including respiratory exchange rate) Bartholomew et al. 1957;Lasiewski 1963b;Hainsworth and Wolf 1970;Hainsworth and Wolf 1978;Krüger et al. 1982;Hiebert 1990Hiebert , 1992Hiebert 1993a;Bucher and Chappell 1997;Powers et al. 2003;Eberts et al. 2019;Shankar et al. 2020a Total evaporation rate Open-flow, negative pressure, mask Allometry, energy cost, energy metabolism, metabolic fuel use, thermogenesis (field and laboratory) Berger and Hart 1972;Berger 1974a, b;Epting 1980;Bartholomew and Lighton 1986;Suarez et al. 1991;Wells 1993a;Chai et al. 1998;Altshuler et al. 2001;Welch Jr et al. 2006Evangelista et al. 2010;Fernández et al. 2011a, b;Fernandez and Suarez 2011;Chen and Welch Jr 2014;Kim et al. 2014;Groom et al. 2018;Shankar et al. 2020a Oxygen consumption and carbon dioxide production 2 Open-flow, positive pressure, metabolic chamber Energy cost (laboratory) Lasiewski 1963b; Schuchmann and Schmidtmarloh 1979a Respiratory evaporation rate 2 Open-flow, negative pressure, mask Water loss, heat dissipation (field and laboratory) Berger and Hart 1972;Powers et al. 2012 Field metabolic rate 5 Doubly labeled water Daily energy expenditure (laboratory) Powers and Nagy 1988;Weathers and Stiles 1989;Powers and Conley 1994;Shankar et al. 2019Shankar et al. , 2020a Daily energy intake ...
Book
Full-text available
Research on hummingbirds over the decades has provided insights into their evolution, migration, physiology, and numerous other areas, including conservation biology. Their small size, energy demands, and high metabolic rates are some of the challenges researchers face when obtaining research samples and biologic materials from live hummingbirds. This manuscript summarizes the established literature dealing with basic methods that scientists have used when capturing, handling, and otherwise researching hummingbirds. Based on the authors’ experience, best practices for working with live hummingbirds are presented, including permitting requirements for studying live hummingbirds, trapping and marking, handling techniques, safe collection of tissue samples, first-aid measures, and euthanasia of hummingbirds, as well as processing of hummingbird specimens (e.g., necropsy and preservation).
... Bee TnT, PBC29029.1. which produces the most energetically expensive muscle work with the highest mass-specific metabolic rates in vertebrates (Evangelista et al., 2010). ...
Article
Full-text available
The powered flight of animals requires efficient and sustainable contractions of the wing muscles of various flying species. Despite their high degree of phylogenetic divergence, flight muscles in insects and vertebrates are striated muscles with similarly specialized sarcomeric structure and basic mechanisms of contraction and relaxation. Comparative studies examining flight muscles together with other striated muscles can provide valuable insights into the fundamental mechanisms of muscle contraction and energetic efficiency. Here, we conducted a literature review and data mining to investigate the independent emergence and evolution of flight muscles in insects, birds, and bats, and the likely molecular basis of their contractile features and energetic efficiency. Bird and bat flight muscles have different metabolic rates that reflect differences in energetic efficiencies while having similar contractile machinery that is under the selection of similar natural environments. The significantly lower efficiency of insect flight muscles along with minimized energy expenditure in Ca²⁺ handling is discussed as a potential mechanism to increase the efficiency of mammalian striated muscles. A better understanding of the molecular evolution of myofilament proteins in the context of physiological functions of invertebrate and vertebrate flight muscles can help explore novel approaches to enhance the performance and efficiency of skeletal and cardiac muscles for the improvement of human health.
... The scaling of mechanochemical efficiency (mechanical work output divided by metabolic energy input) may also have implications upon flight performance, as it varies as a function of body size [7], at least among cursorial vertebrates. However, the few studies that have examined mechanochemical efficiency in hovering hummingbirds concluded that efficiency does not change with increasing mechanical demand (such as variation in body mass or air density) [8], is consistently 10% both within and among species [9][10][11], and is thought to be constant across body sizes [12]. However, the species examined in these studies were of similar body mass, and any scaling of efficiency may not be readily apparent across such a narrow size range. ...
Article
Full-text available
Wing kinematics and morphology are influential upon the aerodynamics of flight. However, there is a lack of studies linking these variables to metabolic costs, particularly in the context of morphological adaptation to body size. Furthermore, the conversion efficiency from chemical energy into movement by the muscles (mechanochemical efficiency) scales with mass in terrestrial quadrupeds, but this scaling relationship has not been demonstrated within flying vertebrates. Positive scaling of efficiency with body size may reduce the metabolic costs of flight for relatively larger species. Here, we assembled a dataset of morphological, kinematic, and metabolic data on hovering hummingbirds to explore the influence of wing morphology, efficiency, and mass on hovering metabolic rate (HMR). We hypothesize that HMR would decline with increasing wing size, after accounting for mass. Furthermore, we hypothesize that efficiency will increase with mass, similarly to other forms of locomotion. We do not find a relationship between relative wing size and HMR, and instead find that the cost of each wingbeat increases hyperallometrically while wingbeat frequency declines with increasing mass. This suggests that increasing wing size is metabolically favourable over cycle frequency with increasing mass. Further benefits are offered to larger hummingbirds owing to the positive scaling of efficiency.
Article
Full-text available
In this paper, the interpretation of hovering flight for hummingbirds is firstly presented and discussed from a hummingbird morphology perspective (muscle and skeleton) including weight distribution, followed by a discussion of hovering aerodynamics. Next, by studying the scale laws, geometry similarity, and statistical analysis on wing parameters, the parametric relation between wing performances and weight is achieved before applying in wing design of flapping wing micro autonomous drones (FWMADs). The efficiency of designed wing based on the scaling law is verified by flying test. Designed wings based on different materials and methods are summarized. Last, the morphology of bird’s tails is presented, and then the designed tails inspired by hummingbirds are introduced before tail performances are also discussed simply. The results show that the tail could be predicted to apply to the stability of hovering twin-wing FWMADs. The current studies provide a simple but powerful guideline for biologists and engineers who study the morphology of hummingbirds and design FWMADs.
Article
Full-text available
Lift production is constantly a great challenge for flapping wing micro air vehicles (MAVs). Designing a workable wing, therefore, plays an essential role. Dimensional analysis is an effective and valuable tool in studying the biomechanics of flyers. In this paper, geometric similarity study is firstly presented. Then, the Pw-AR ratio is defined and employed in wing performance estimation before the lumped parameter is induced and utilized in wing design. Comprehensive scaling laws on relation of wing performances for natural flyers are next investigated and developed via statistical analysis before being utilized to examine the wing design. Through geometric similarity study and statistical analysis, the results show that the aspect ratio and lumped parameter are independent on mass, and the lumped parameter is inversely proportional to the aspect ratio. The lumped parameters and aspect ratio of flapping wing MAVs correspond to the range of wing performances of natural flyers. Also, the wing performances of existing flapping wing MAVs are examined and follow the scaling laws. Last, the manufactured wings of the flapping wing MAVs are summarized. Our results will, therefore, provide a simple but powerful guideline for biologists and engineers who study the morphology of natural flyers and design flapping wing MAVs.
Conference Paper
Full-text available
Wing, as one of important components of a flapping wing Micro Air Vehicle (MAV), is directly related to aerodynamic performance such as lift force and torque around body. Therefore, design of efficient flapping wing has become an interesting research topic recently. In this paper, the aerodynamic performance of flexible micro-membrane flapping wing under the condition of hovering was investigated to ascertain the best combination of wing geometric parameters. For this purpose, a flapping wing mechanism and an experiment setup were built to measure the aerodynamic forces and power usage of the flapping motion. Then four different families of wings were fabricated. The lift and power usage was measured at different flapping frequencies with different shapes of wings. In order to investigate the effects of wing geometry, only one parameter of wing was varied at a time. The results indicate that the aspect ratio and wing surface have a critical impact on the force production and wing efficiency. The best performance was obtained with a trapezoidal wing with a straight leading edge; both its shape and aspect ratio () are similar to a typical hummingbird wing.
Article
Full-text available
Biologists have traditionally focused on a muscle's ability to generate power. By determining muscle length, strain and activation pattern in the cockroach Blaberus discoidalis, we discovered leg extensor muscles that operate as active dampers that only absorb energy during running. Data from running animals were compared with measurements of force and power production of isolated muscles studied over a range of stimulus conditions and muscle length changes. We studied the trochanter-femoral extensor muscles 137 and 179, homologous leg muscles of the mesothoracic and metathoracic legs, respectively. Because each of these muscles is innervated by a single excitatory motor axon, the activation pattern of the muscle could be defined precisely. Work loop studies using sinusoidal strains at 8 Hz showed these trochanter-femoral extensor muscles to be quite capable actuators, able to generate a maximum of 19-25 W kg-1 (at 25°C). The optimal conditions for power output were four stimuli per cycle (interstimulus interval 11 ms), a strain of approximately 4 %, and a stimulation phase such that the onset of the stimulus burst came approximately half-way through the lengthening phase of the cycle. High-speed video analysis indicated that the actual muscle strain during running was 12 % in the mesothoracic muscles and 16% in the metathoracic ones. Myographic recordings during running showed on average 3-4 muscle action potentials per cycle, with the timing or the action potentials such that the burst usually began shortly after the onset of shortening. Imposing upon the muscle in vitro the strain, stimulus number and stimulus phase characteristic of running generated work loops in which energy was absorbed (-25 W kg-1) rather than produced. Simulations exploring a wide parameter space revealed that the dominant parameter that determines function during running is the magnitude of strain. Strains required for the maximum power output by the trochanter-femoral extensor muscles simply do not occur during constant, average-speed running. Joint angle ranges of the coxa-trochanter-femur joint during running were 3-4 times greater than the changes necessary to produce maximum power output. None of the simulated patterns of stimulation or phase resulted in power production when strain magnitude was greater than 5%. The trochanter-femoral extensor muscles 137/179 of a cockroach running at its preferred speed of 20 cm s-1 do not operate under conditions which maximize either power output or efficiency. In vitro measurements, however, demonstrate that these muscles absorb energy, probably to provide control of leg flexion and to aid in its reversal.
Article
Full-text available
Summary The flight energetics of hovering hummingbirds was examined by simultaneous collection of metabolic and kinematic data followed by a morphometric analysis of wing characteristics. These data were then used for an aerodynamic analysis of the power output required to generate sufficient lift; this, together with the metabolic power input, allowed an estimate of the flight efficiency. The use of two closely related species demonstrated common design features despite a marked difference in wing loading. Considerations of the inertial power costs strongly suggest that hummingbirds are able to store kinetic energy elastically during deceleration of the wing stroke. This analysis predicts that hummingbirds hover with a muscle power output close to 100-120W kg21 at 9-11% mechanochemical efficiency.
Article
As the smallest homeotherms, hummingbirds suffer from low thermal inertia and high heat loss. Flapping flight is energetically expensive, and convective cooling due to wing and air movements could further exacerbate energy drain. Energy conservation during flight is thus profoundly important for hummingbirds. The present study demonstrates that heat produced by flight activity can contribute to thermoregulatory requirements in hovering hummingbirds. The rate of oxygen consumption, as an indicator of metabolic cost, was measured during hover-feeding and compared with that during perch-feeding. In hover-feeding, oxygen consumption increased only moderately between 35 and 5 degrees C in contrast to the sharp increase during perch-feeding over the same temperature range. This result suggests that heat produced by contraction of the flight muscles substituted for regulatory thermogenesis to accommodate for heat loss during temperature. With declining air temperatures, the mechanical power requirements of hovering decreased slightly, but metabolic costs increased moderately. As a result, the mechanical efficiency of the muscle in converting metabolic power to mechanical power was I educed. Changes in wingbeat kinematics also accompanied the reduction in muscle efficiency. Wingbeat frequency increased but stroke amplitude decreased when hovering in the cold, suggesting thermoregulatory roles for the flight muscles. Hovering hummingbirds modulated their wingbeat frequency within a narrow range, reflecting the physical constraints of tuning to a natural resonant frequency with an elastic restoring force. We hypothesize that, by forcing the resonant system of the wings and thorax to oscillate at different frequencies, muscle contraction in the cold generates more heat at the expense of mechanical efficiency. This mechanism of modulating the efficiency of muscle contraction and heat production allows flying hummingbirds to achieve energy conservation at low air temperatures.
Article
The mass-specific power input (metabolic cost) of hovering is independent of body mass but is proportional to the 0.5 exponent of wing disc loading for the sample considered here. The efficiency of converting energy metabolism to aerodynamic power is approximately constant at 6.23% in hummingbirds with a wide range of body mass and wing disc loading. Actuator wing disc theory of helicopter aerodynamics appears adequate for comparing the energetic cost of hovering flight among hummingbirds. Differences in the power input for hovering due to differences in wing disc loading are greater for larger hummingbirds than for smaller ones. Impaired integrity of the outer portion of the wing disc caused by molt or breaking primaries 8, 9, or 10 greatly increases the required power for hovering. The unusual nonsequential pattern of primary-feather molt in hummingbirds may be of adaptive value in that it results in the most rapid restoration of the integrity of the wing disc.
Article
Flight performance trade-offs and functional capacities of ruby-throated hummingbirds (Archilochus colubris L.) were studied using an integrative approach. Performance limits were measured by noninvasively challenging birds with two strenuous forms of flight: hovering in low-density gas mixtures (a lift assay for the capacity to generate vertical force) and fast forward flight in a wind tunnel (a thrust assay for the capacity to generate horizontal force). Functional capacities during hovering were measured by simultaneously col- lecting metabolic data using respirometry and information on wing- beat kinematics for aerodynamic analysis. Intraspecific differences in flight capacity, presumably reflecting diverse selective forces because of sexual dimorphism, migration, and plumage renewal, were then compared. Birds with shorter wings (adult males) or with increased body weight displayed a reduced hovering capacity, although their maximum flight speed was unaffected by such morphological changes. Birds undergoing molt of their flight feathers exhibited a diminished performance during both hovering and forward flight. Hovering capacities in relation to variation in wing morphology and body mass were congruent with aerodynamic predictions, whereas performance capacities in fast forward flight differed from theoretical models. Kinematically, hovering hummingbirds operate within a nar- row range of wing-beat frequencies, and modulation of aerodynamic forces and mechanical power is achieved primarily through variation in wing-stroke amplitude. Although differing in hovering perform- ance, both sexes of nonmolt birds demonstrate similar mechanical and metabolic capacities, whereas molting inflicts high energetic costs. Relatively invariant physiological capacities may thus ultimately constrain the extent of intraspecific trade-offs between morphology and performance, providing mechanistic insights into the multilevel functional design of the hummingbird flight system.