In this study we compared the wing kinematics of 27 bats representing six pteropodid species ranging more than 40 times in body mass (Mb=0.0278–1.152 kg), to determine whether wing posture and overall wing kinematics scaled as predicted according to theory. The smallest species flew in a wind tunnel and the other five species in a flight corridor. Seventeen kinematic markers on the midline and left side of the body were tracked in three dimensions. We used phylogenetically informed reduced major axis regression to test for allometry. We found that maximum wingspan (bmax) and maximum wing area (Smax) scaled with more positive allometry, and wing loading (Qs) with more negative allometry (bmaxMb0.423; SmaxMb0.768; QsMb0.233) than has been reported in previous studies that were based on measurements from specimens stretched out flat on a horizontal surface. Our results suggest that larger bats open their wings more fully than small bats do in flight, and that for bats, body measurements alone cannot be used to predict the conformation of the wings in flight. Several kinematic variables, including downstroke ratio, wing stroke amplitude, stroke plane angle, wing camber and Strouhal number, did not change significantly with body size, demonstrating that many aspects of wing kinematics are similar across this range of body sizes. Whereas aerodynamic theory suggests that preferred flight speed should increase with mass, we did not observe an increase in preferred flight speed with mass. Instead, larger bats had higher lift coefficients (CL) than did small bats (CLMb0.170). Also, the slope of the wingbeat period (T) to body mass regression was significantly more shallow than expected under isometry (TMb0.180), and angle of attack (α) increased significantly with body mass [α∝log(Mb)7.738]. None of the bats in our study flew at constant speed, so we used multiple regression to isolate the changes in wing kinematics that correlated with changes in flight speed, horizontal acceleration and vertical acceleration. We uncovered several significant trends that were consistent among species. Our results demonstrate that for medium- to large-sized bats, the ways that bats modulate their wing kinematics to produce thrust and lift over the course of a wingbeat cycle are independent of body size.

Body size governs almost every aspect of animal biology. Many fundamental aspects of ontogeny, physiology, ecology and behavior can be largely predicted with little more information than the mass of an animal. These, and many other important aspects of an organism's life, scale predictably with body mass, according to fundamental form–function relationships (Brown and West, 2000; Dial et al., 2008). The influence of body size on locomotion is no less striking, and biomechanical investigations have revealed that just as body shape changes with size, so too do locomotor kinematics (Biewener, 1983; Biewener, 2005; Heglund and Taylor, 1988).

Part of the reason that animals of different sizes move their bodies differently is that the physical demands of locomotion change with body size. As an extreme example, small flying animals moving through the air must deal with high viscous forces relative to inertial forces, whereas for larger flying animals, viscous forces are much smaller than inertial ones. As a result, a <100 μg wasp uses constant clap and fling wing motions to fly whereas a 10 kg Andean condor travels mostly by simply holding its wings outstretched (Ellington, 1999; McGahan, 1973; Miller and Peskin, 2005). Within the range of body sizes between those extremes, animal flight can differ in subtle but important ways. The mechanics of insect flight differ between fruit flies and hawkmoths, and the way a bird flies also varies from hummingbirds to pigeons to vultures (Combes and Daniel, 2003; Dial and Biewener, 1993; Dickinson and Götz, 1996; Sane, 2003; McGahan, 1973; Warrick et al., 2005). Unlike insects and birds, however, bats have largely been assumed to use similar mechanisms of aerodynamic force production in flight, regardless of size (Bullen and McKenzie, 2002; Hedenström et al., 2007; Norberg and Rayner, 1987), even though bats range in body mass over roughly three orders of magnitude, from the ≤0.002 kg bumblebee bat (Craseonycteris thonglongyai) to >1.2 kg flying foxes (Pteropus spp.) (Hill and Smith, 1981; Kunz and Jones, 2000; Surlykke et al., 1993).

The purpose of this study is to examine the influence of body size on wing kinematics in bats. The body shapes and foraging modes of bats vary widely among families, so we focused on a single family, the Pteropodidae. This family consists of ca. 186 species distributed throughout the paleotropics (Wilson and Reeder, 2005) and is characterized by fruit and nectar-feeding, non-echolocating animals that are ecologically and morphologically similar across a broad range of body sizes. No other chiropteran family possesses so wide a range of body sizes as the pteropodids (Nowak, 1994).

We investigated wing posture, wing kinematics and the patterns with which kinematics change with flight velocity. In this study we also examined the influences of horizontal and vertical accelerations on wing kinematics, as a means of studying how the kinematics of bats reflect aeromechanical force production in those directions.

Video recordings and kinematic reconstructions

Flight recordings and calibrations

We recorded the flight kinematics of 27 animals from six species ranging more than 40-fold in body mass, from 0.0278 kg to 1.152 kg (Table 1). Bats of the smallest species, Cynopterus brachyotis, were flown in a 1.4 m×1.2 m×1.2 m (L×W×H) wind tunnel at the Concord Field Station of Harvard University in Bedford MA, USA (Hedrick et al., 2002). Bats belonging to the other five species were flown in a 14.5 m×2.8 m×2.4 m (L×W×H) flight corridor at the Lubee Bat Conservancy in Gainesville FL, USA. Bats were recorded at 1000 Hz using three phase-locked Photron 1024 PCI digital high-speed cameras (Photron USA, Inc., San Diego, CA USA). A fourth phase-locked 1000 Hz camera, a Photron APX, was also used to record corridor flights. All cameras had 1024×1024 pixel resolution.

Before experiments, each bat was anesthetized with isoflurane gas then marked with non-toxic acrylic paint at 17 anatomical landmarks on the fur and skin of the sagittal midline and the left half of the body (Fig. 1A). At least ten flights were recorded from each individual, but only five flights per individual were used for analyses. Those five were selected to sample the broadest range possible of flight velocities for each individual.

The volumes through which bats flew were calibrated using the direct linear transformation (DLT) method (Abdel-Aziz and Karara, 1971), based on a 0.35 m×0.35 m×0.28 m (L×W×H) calibration frame for the wind tunnel trials, and on a series of vertical wires and beads for the corridor experiments, spanning 2.00 m× 1.25 m×2.40 m (L×W×H).

Marker tracking and coordinate systems

The positions of 17 anatomical markers were digitized in each video frame (Fig. 1). Where a marker was visible in two or more cameras at a given instant, its position could be calculated using the DLT method. In those frames where a marker was not visible to at least two cameras, its position was interpolated based on its three-dimensional positions before and after that interval, using an over-constrained least-squares polynomial fit of its trajectory. For contiguous gaps in the data with rich data at the end points, a third-order polynomial was used, and for gaps with sporadic intermediate points, a sixth-order polynomial was used (Riskin et al., 2008). The interpolations were visually checked for all 17 markers in all 135 trials by superimposing the interpolated three-dimensional path on the original movies. The wing membrane was modeled as a series of 18 flat triangles spanning the xyz coordinates of the markers (Fig. 1A).

For the purposes of this study, we define a body-centered coordinate system (xb, yb, zb) and a global coordinate system (xg, yg, zg) (Gatesy and Baier, 2005). Both are right-handed, and in both, x and y are horizontal and z is vertical (Fig. 1B,C). The global coordinates represent the motion of the bat relative to still air, with positive xg in the direction of travel, positive yg to the bat's left and positive zg pointing vertical up. To obtain global coordinates for corridor trials, we calculated the positions of the 17 markers throughout the trial relative to the calibration frame, then rotated the coordinate system so that the bat's anterior sternum marker (a in Fig. 1A) moved forward in the xg direction, beginning and ending the wingbeat cycle at the same yg value. We then subtracted the initial position of the anterior sternum marker from all positions. Thus, every trial began with the anterior sternum marker at (xg, yg, zg)=(0, 0, 0), and ended with the sternum at some positive xg position with yg=0. Global coordinates for wind-tunnel trials were calculated analogously but were adjusted to correct for the velocity and direction of the air so that the global coordinates would also represent the motion of the bat's body relative to still air.

The body-centered coordinate system was calculated from the global coordinates by subtracting the position of the anterior sternum marker in each time step from the positions of all markers in that time step. The net effect was that the anterior sternum marker remained at (xb, yb, zb)=(0, 0, 0) throughout the trial and the head faced positive xb (Fig. 1B).

We restricted our analyses from each flight to a single wingbeat cycle. Where possible (88 of 135 trials), the upper reversal points of the wrist marker in the zb direction were used to denote the beginning and end of the wingbeat cycle (Fig. 1B). For some flight corridor trials, a complete wingbeat cycle occurred within the calibrated volume, but not spanning those endpoints. In those cases, we used, in order of preference, the lower reversal points of the wrist (38 of 135), the upper reversal points of the wingtip (6 of 135), or the lower reversal points of the wingtip (3 of 135).

Measurements of body shape and posture

To determine how wing shape changed with body mass (Mb), we extracted information about three-dimensional wing conformation from the kinematic recordings. Whereas morphometric measurements are typically taken from specimens with wings extended on a flat surface, our methods capture posture, one of the most important aspects of wing form for understanding flight mechanics (Nudds, 2007). The following parameters were calculated from each trial.

Maximum wingspan (bmax): two times the maximum distance of the wingtip marker from the mid-sagittal plane (yb=0) in a trial. This always occurred during the downstroke.

Minimum wingspan (bmin): for each time step we found the yb-value of the wing marker furthest from the yb=0 plane. The bminvalue for a trial was the smallest of those values across the entire wingbeat cycle, multiplied by two. This always occurred during the upstroke.

Maximum wing chord (cmax): the greatest two-dimensional distance in a trial between the wrist and the tip of digit V (e and n in Fig. 1A, respectively), using their xb and zb position data only, and ignoring yb.

Maximum wing area (Smax): the left wing was divided into 18 triangular surfaces (Fig. 1A). The areas of those triangles were summed, then multiplied by two, to arrive at total wing area (S). The result is a coarse three-dimensional mesh which approximates the wetted area, not a two-dimensional projection. This was done for each time step over the course of the wingbeat cycle, and Smax was the maximum value of S for a trial. We chose this method because it uses our high-fidelity measurements to arrive at a more detailed index of wing conformation than would a two-dimensional projected area.

Wing loading (Qs): defined as:
formula
(1)
where g is the acceleration of gravity (9.81 m s–2).
Aspect ratio (AR): defined as:
formula
(2)

Measurements of velocity and acceleration

Parameters were calculated as follows.

Horizontal velocity (Vhoriz): the horizontal distance traveled by the anterior sternum marker over the course of the wingbeat cycle, divided by the duration of the wingbeat cycle.

Vertical velocity (Vvert): the vertical distance traveled by the anterior sternum marker over the course of the wingbeat cycle, divided by the duration of the wingbeat cycle.

Horizontal acceleration (Ahoriz): the change in forward velocity between the beginning of the wingbeat cycle and the end of the wingbeat cycle, divided by the duration of the wingbeat cycle. Both estimates of forward velocity were calculated using the slope of a linear fit of the anterior sternum marker's xg position over time, using a 4 ms window. Since the posture of the bat was roughly the same at the beginning and end of a wingbeat cycle, inertial effects can be neglected, so the change in speed of the sternum markers approximates the change in the speed of the center of mass.

Vertical acceleration (Avert): calculated analogously to Ahoriz, but using zg position over time instead of xg.

Measurements of wing kinematics

The following parameters were obtained.

Wingbeat period (T): the time taken to complete the wingbeat cycle.

Downstroke duration (Tdown): the duration of the downstroke, based on the motion of the wrist in zb dimension.

Downstroke ratio (τ): the proportion of the wingbeat cycle duration occupied by the downstroke, calculated as:
formula
(3)

Wing stroke amplitude (φ): the maximum three-dimensional angle between any two positions of the wrist relative to the shoulder within a wingbeat cycle.

Strouhal number (St): a dimensionless descriptor of flapping motion, calculated using the distance traveled by the wingtip in the zb direction:
formula
(4)

Stroke plane angle (β): the trajectory of the wingtip in the xbzb plane was fitted to a linear function using orthogonal regression. The angle between that regression line and horizontal was multiplied by –1 to calculate β, which is negative by convention (Norberg, 1990) (Fig. 1B).

Wing camber at maximum wingspan: in the time step (camera frame) corresponding to bmax, which always occurred at some time during the downstroke, a parasagittal (xgzg) cross section of the wing at the yg value of the wrist was made. The straight line from the wrist to trailing edge was defined as the chord line. Of the 18 triangular sections of the modeled wing, exactly six always intersected that plane at the time of bmax (Fig. 2A), resulting in seven intersections of a triangle border and the plane. The first term of a sine series was fitted to those seven intersection points to create a curved line between the wrist and the trailing edge of the wing that came as close to those seven points as possible. To calculate wing camber, we divided the maximum distance to the chord line from that curve by the length of the chord line (Fig. 2B). Our estimate of camber is an instantaneous value for a dynamically changing parameter, and although it might not represent the maximum or even average camber over the whole wingbeat cycle, it is a value that can be defined in a clear, unambiguous manner for ease of comparison among wingbeat cycles.

Angle of attack (α): the angle of the wing chord relative to the incoming flow of air was calculated at the instant of maximum wingspan, in the same time step as wing camber was calculated. The chord line was defined as a straight line between the wrist and the intersection of the trailing edge with the xgzg plane of the wrist (Fig. 2C). The angle of the chord line above horizontal was defined as α1. The angle velocity vector of the wrist below horizontal in the xgzg plane, was defined as α2. Angle of attack (α) was calculated as α12. Again, this gives an instantaneous recording that does not capture changes over the course of the wingbeat cycle, but facilitates comparison among wingbeat cycles. Note that this estimate does not account for relative air movement resulting from the induced velocity.

Coefficient of lift (CL): this was estimated based on the overall vertical acceleration of the bat for the entire wingbeat cycle, and the bat's posture and wrist velocity at the time of maximum wingspan, using the equation:
formula
(5)
where ρ is the density of air (1.2 kg m–3), Vwrist is the instantaneous velocity of the wrist in the xgzg plane at the time of maximal wingspan, and wing area (S) is calculated for the same instant in time as Vwrist. Calculated this way, CL is not intended to represent an absolute measure, but rather an assessment of relative aerodynamic effectiveness that can be applied across species and flight speeds.

Statistical analyses

Scaling of body shape

For tests of allometry in body shape parameters, calculations of body shape were made for all 135 flights. Because our species sample size for estimation of scaling relationships was six, we used averages for all individuals within each species. For each parameter, the calculated values from the five trials of an individual were pooled to determine an individual median, then the individual medians within a species were pooled to determine the species median. Those six species medians were used in a log–log orthogonal (reduced major axis or RMA) regression against median body mass for each species (LaBarbera, 1989). To account for statistical non-independence of data resulting from the shared phylogenetic history of the six species (Felsenstein, 1985), we repeated that analysis accounting for phylogeny using a generalized least squares (GLS) model. We used a consensus tree of several recent pteropodid phylogenies (Giannini and Simmons, 2005; Jones et al., 2002; O'Brien et al., 2009), with branch lengths scaled using the method of Pagel (Pagel, 1992). GLM analyses with phylogeny were carried out using REGRESSIONv2 (Lavin et al., 2008) in Matlab. The slope of each regression was compared with that expected under isometry using two-tailed t-tests with four degrees of freedom. For ease of comparison with previous studies, we also present the results of RMA analyses that do not account for phylogeny. The expected relationships under isometry are: bmaxMb1/3, bminMb1/3, cmaxMb1/3, SmaxMb2/3, QsMb1/3 and AR∝Mb0 (Norberg, 1990).

Scaling of flight velocity

The speed at which a bat flies is expected to increase with the size of the bat, and that prediction results because wing loading (Qs) scales ∝Mb1/3 under isometry. Rearrangement of Eqns 1 and 5 reveals that as Qs increases, Avert will decrease, unless there is an associated increase in the product of CL and :
formula
(6)

Assuming that CL does not increase with body mass, and assuming Vwrist is proportional to Vhoriz, large animals should fly faster than small ones in order to generate the lift required to fly. Specifically, preferred flight velocity is predicted to scale ∝Mb1/6 (Pennycuick, 1975).

We expected bats flown in the corridor to fly with speeds that scaled with Mb1/6, but since bats in the wind tunnel could not choose their own velocities, we did not expect those individuals to conform to that pattern. We thus performed a linear least squares regression of log(Vhoriz) versus log(Mb) for the median Vhoriz values of individuals flown in the corridor (N=22). A GLS model that accounted for phylogenetic effects was used, and the slope of that regression was compared with 1/6 using a two-tailed t-test with eight degrees of freedom. For ease of comparison with previous studies, we also include the slope estimate based on an ordinary RMA regression.

Scaling of wing kinematics

To test for systematic changes in the values of kinematic parameters with body size, we performed linear least squares regressions of individual median values against log(Mb). Because bats flown in the wind tunnel (C. brachyotis) had different velocity distributions than corridor-flown bats (see Results), we excluded them from kinematic scaling analyses.

To account for phylogenetic effects, we performed all interspecific regression analyses using a GLS model, as described above, with reduced degrees of freedom to account for soft polytomies in the pteropodid tree (Purvis and Garland, 1993; Garland and Díaz-Uriarte, 1999). Parameters included in these analyses were T, Tdown, τ, φ, St, β, wing camber, α, α1, α2 and CL.

We expected T and Tdown to scale with Mb1/3 (Norberg, 1990), and all other regressions were expected not to scale with body mass, in other words, were expected to scale ∝Mb0.

We used log–log regression for the majority of tests but could not use that method for angles (φ, β, α, α1 and α2), since many angles were negative (Smith, 1984). Instead, we used log–linear regression for angles, which assumes that if there is an effect of body size on the angle, it scales ∝Mb1.

For ease of comparison with previous studies, we repeated linear regressions of kinematic parameters against body mass using more classical RMA methods that do not account for phylogeny.

Changes of kinematics with velocity and acceleration

To determine how wing kinematics change with Vhoriz, Ahoriz and Avert, we used multiple regression. This method permits examination of how each of these three variables correlates with changes in wing kinematics, while correcting for the influence of the other two. Each regression had a single kinematic variable as the dependent variable, and Vhoriz, Ahoriz and Avert as model effects. To take into account variability among individuals, we also included individual bat as a random effect (Gelman and Hill, 2007). We performed these regressions separately for each species.

We conducted regression analysis for bmax, bmin, T, Tdown, τ, φ, β, St, α, α1, α2, wing camber and CL. Thus, thirteen multiple regression analyses were performed per species, for a total of 78. In each, we tested for significance of partial regression coefficients for each of Vhoriz, Ahoriz and Avert using two-tailed t-tests. Degrees of freedom for each multiple regression test were equal to Nk, where N is the number of flight trials, and k is the number of estimated parameters. Estimated parameters were Vhoriz partial regression slope, Ahoriz partial regression slope, Avert partial regression slope, and one intercept per individual bat. This resulted in nine degrees of freedom for R. aegyptiacus regressions, 13 for those of P. vampyrus, and 17 degrees of freedom for those of all other species.

Statistical analyses were performed using custom-scripts in Matlab R2008b (MathWorks Inc., Natick, MA, USA), then verified using JMP IN 8.0 (SAS Institute, Cary NC, USA). For models with random effects we used the expected mean squares (EMS) method. To account for possibly inflated family-wise type I error rate resulting from performing multiple statistical tests on closely related data (Curran-Everett, 2000; Curran-Everett and Benos, 2004), we carried out a positive false discovery rate (pFDR) analysis on P-values from the data in supplementary material Table S2A–C using the qvalue package (Version 1.20) (Storey, 2002) for R (Version 2.10.1) (R Development Core Team, 2009) with a pFDR rate of 5% (allowing 5% of ‘significant’ results to be false). From that analysis we found that a more conservative alpha level for significance is 0.034, rather than 0.05, and used the more conservative value as the criterion for rejection of the null hypothesis (supplementary material Table S2A–C).

A complete table of summary statistics for all 135 trials used in this study is provided in supplementary material Table S1. In the body of this paper, we report phylogenetically corrected regression statistics, but for ease of comparison with previous studies, we include results without phylogenetic correction in tables as well.

Scaling of body shape

The scaling of maximum wingspan to body mass was positively allometric (bmaxMb0.423 compared with Mb0.333, P=0.045; Table 2, Fig. 3A). Minimum wingspan scaled isometrically to body mass (bminMb0.366 compared with Mb0.333, P=0.509; Table 2, Fig. 3B). Wing chord scaled isometrically with body mass (cmaxMb0.357 compared with Mb0.333, P=0.457; Table 2, Fig. 3C). Wing area scaled with positive allometry (SmaxMb0.768 compared with Mb0.666, P=0.047; Table 2, Fig. 3D) and wing loading increased more gradually with body size than would be expected under isometry (QsMb0.233 compared with Mb0.333, P=0.024; Table 2, Fig. 3E). Since span increased with positive allometry but chord increased isometrically, aspect ratio increased with Mb, with the slope of that regression approaching significant allometry (AR∝Mb0.072 compared with Mb0.000, P=0.068; Table 2, Fig. 3F).

Flight velocities and accelerations

The horizontal velocities of bats (4.98±0.09 m s–1) were much greater than vertical velocities (0.12±0.03 m s–1), so flight paths were close to horizontal (1.36±0.36 deg above horizontal). The flight speeds of bats of all body sizes overlapped greatly. Bats in the wind tunnel (C. brachyotis) generally flew faster for their size than bats in the corridor did (Fig. 4A). With wind tunnel flights excluded from analysis, the preferred velocities of animals increased with Mb with a slope not significantly different from that expected under isometry (VhorizMb0.005 compared with Mb0.167; P=0.056; Table 3) but much closer to Mb0.000 (P=0.948). We found considerable variability in Ahoriz and Avert among trials (Fig. 4B,C), and no trial showed zero net acceleration.

Scaling of wing kinematics

The regression slope of wingbeat period to body mass was significantly lower than that expected under isometry (TMb0.180 compared with Mb0.333; P=0.039; Table 3, Fig. 5A), but downstroke duration was not significantly different from predicted (TMb0.213 compared with Mb0.333, P=0.140; Table 3, Fig. 5B). Downstroke ratio also did not change significantly with body size (τ∝Mb0.036 compared with Mb0.000; P=0.162; Table 3, Fig. 5C).

Wing stroke amplitude at the wrist did not significantly change with log body mass [φ∝(logMb) · (–6.058); P=0.257; Table 3, Fig. 5D], nor did stroke plane angle [β∝(logMb)8.974; P=0.107; Fig. 5E]. Strouhal number also did not change significantly with log body mass (StMb–0.088 compared with Mb0.000; P=0.532; Table 3, Fig. 5F).

We found that α increased significantly with log body mass [α∝(logMb)7.738; P=0.014; Table 3, Fig. 5G], and that the overall change occurred as a result of changes in α1 but not α21∝(logMb)7.542; P=0.032; Fig. 5H; and α2∝(logMb) · (–0.461); P=0.866; Table 3, Fig. 5I]. Wing camber at maximum span did not change significantly with log body mass (wing camber∝Mb0.070; P=0.714; Fig. 5J). Coefficient of lift increased significantly with body mass (CLMb0.170; P=0.042; Table 3, Fig. 5K).

Kinematic changes with velocity and acceleration

We observed several changes in the kinematics of bats with changes in Vhoriz, Ahoriz and Avert. Some changes that were significant in some species were not significant in others, but in almost no cases did different species show opposing trends; where a partial regression slope was significant for multiple species, it almost always had the same sign, positive or negative, for all other species. These are reported fully in supplementary material Table S2, and summarized in Table 4.

With increases in flight velocity, holding the influences of Ahoriz and Avert constant, we observed decreased maximum wingspan, increased wingbeat period, increased downstroke duration, increased downstroke ratio, decreased stroke plane angle, decreased angle of attack, decreased wing camber, decreased Strouhal number, and decreased lift coefficient. We observed no significant change in amplitude, and observed mixed results among species for changes in minimum wingspan (Table 4, supplementary material Table S2A).

With increases in horizontal acceleration, we observed decreases in minimum wingspan and stroke plane angle, and increases in maximum wingspan, amplitude and angle of attack. Strouhal number and lift coefficient also both increased. We saw no significant changes in wingbeat period, downstroke duration, downstroke ratio or wing camber (Table 4, supplementary material Table S2B).

With increases in vertical acceleration, we observed an increase in maximum wingspan, angle of attack, wing camber, Strouhal number and lift coefficient, along with a decrease in wingbeat period. We observed no significant changes in minimum wingspan, downstroke duration, downstroke ratio, wing stroke amplitude, or stroke plane angle (Table 4, supplementary material Table S2C).

The influence of body size on wing shape and kinematics

Our results provide experimental evidence that for pteropodid bats many aspects of wing kinematics vary with body size, but that the ways kinematics change with velocity and acceleration are relatively consistent across body sizes. Additionally, we found that the scaling relationships for maximum wingspan, maximum wing area and minimum wing loading in pteropodid bats, based on measurements from actual wing form as employed during flight behavior, differ from the scaling relationships measured from outstretched preserved specimens (Table 2). Importantly, we uncovered a significant positive allometry between body mass and wing area during the downstroke, which may help to offset the consequences of higher wing loading that accompany increased body size. Furthermore, large bats had higher coefficients of lift during flight than small bats did. This result highlights the importance of wing posture as a confounding variable for hypotheses about ecological function based solely on the two-dimensional shape of an outstretched wing.

Body size and wing shape: the importance of posture

A bat's wing comprises highly compliant skin membranes that interconnect a jointed skeleton capable of many degrees of freedom (Riskin et al., 2008). By its very morphological structure, the area of a bat wing is highly variable throughout every wingbeat cycle. As a result, measurements of wing area for bat specimens can vary substantially compared with those for insects or birds, depending especially on the degree to which the membrane is stretched before preservation or measurement. It has been extremely valuable to make comparisons of wing area among bat species in a two-dimensional perspective, but our high-fidelity measurements of wing shape permitted the use of actual three-dimensional wing conformation in flight.

It has long been known that the bodies of pteropodid bats do not scale isometrically (Norberg, 1981; Norberg, 1990; Norberg and Rayner, 1987), but several scaling relationships that we observed in this study differ from those previously reported based on measurements of dead specimens measured outstretched on a flat surface. We hypothesize that those discrepancies arise because wing posture during the downstroke does not match the posture in which bats are typically held for morphometric measurements, and the difference between in-flight conformation and measurement conformation varies non-randomly with body size. In other words, we hypothesize that the way bats position their wings during downstroke varies with body size. If true, the actual scaling trends in the mid-flight shapes of bat wings would present themselves only when wing measurements were taken during flight. Given that wing measurements are widely used to predict behavior and foraging ecology for bat species (e.g. Bullen and McKenzie, 2001; Kingston et al., 2000), our hypothesis, if true, would have wide-reaching influence on our understanding of bat evolution and ecology.

In their analysis of wing form in bats, Norberg and Rayner (Norberg and Rayner, 1987) and Norberg (Norberg, 1981) found that wingspan in pteropodids scaled with Mb0.350 and Mb0.362, respectively, both only slightly higher than the Mb0.333 expected under isometry. The scaling relationship uncovered in this study (bmaxMb0.423) was around 20% higher than theirs and suggests that large bats extend the wing membrane more fully during the downstroke than small bats do. Similarly, the scaling coefficient of wing loading to body mass from this study, QsMb0.233, was around 27% lower than the coefficients Mb0.327 and Mb0.309 reported by Norberg and Rayner (Norberg and Rayner, 1987) and Norberg (Norberg, 1981). Here again, posture is a probable contributor to the discrepancy. The area of a bat's wing changes throughout the wingbeat cycle and depends greatly on the positions of the carpus and elbow and the degree of extension and abduction of the digits. Whereas specimens in previous studies have been measured in a fully outstretched and flattened posture, our methods capture the three-dimensional shape of the wing with the degree of wing extension that is biologically relevant.

Interestingly, our results suggest that large bats extend the wing more fully on the downstroke, but not on the upstroke, since the scaling relationship of bmin to body mass (bminMb0.366) did not differ from the relationship expected under isometry (bminMb0.333). This makes sense, since the majority of lift production occurs on the downstroke.

Since wing loading increases with body mass, it is hypothesized to limit the body sizes of the largest flying animals (Greenewalt, 1975). Norberg (Norberg, 1981) and Norberg and Rayner (Norberg and Rayner, 1987) found that wing loading for pteropodid bats scaled isometrically with body mass (QsMb0.333), but we found evidence for negative allometry (QsMb0.233). Our result suggests that the largest bats may effectively reduce wing loading, compensating for their large size, by employing different wing postures than small bats. In his study of allometric scaling of bird wings, Nudds commented on the possible influence of body-size-dependent variation in elbow angle among birds on scaling relationships (Nudds, 2007). Whether or not the relationship we report here for pteropodid bats is also true for birds, or indeed even other bats, is an important topic for future study.

An alternative explanation for the difference between our body size regression slopes and those of Norberg and Rayner is that because we sampled fewer species, six in this study compared with >50 by Norberg and Rayner, we simply got different slopes by chance. To assess the influence of sample bias, we performed orthogonal regressions of log(Mb) versus log(bmax) and log(Mb) versus log(Smax) using Norberg and Rayner's data for the closest species with complete measurements to the six used in this experiment: Cynopterus brachyotis (0.0265 kg), Rousettus aegyptiacus (0.140 kg), Pteropus sp. (0.210 kg), Eidolon helvum (0.274 kg), Pteropus sp. (0.347 kg) and Pteropus vampyrus (1.179 kg). Using Norberg and Rayner's data, we obtained regression slopes of bmaxMb0.331±0.05 and SmaxMb0.651±0.109. Both of those are closer to the regression slopes of Norberg and Rayner's entire pteropodid dataset (bmaxMb0.350 and SmaxMb0.715) than to our regression slopes (bmaxMb0.423 and SmaxMb0.768). This supports our hypothesis that the difference between their results and ours is our use of mid-flight wing shape measurements, and not the species sample used. Further investigations of species-specific differences in posture will help elucidate the aerodynamic relevance of those differences.

Body size and wing kinematics

Our data suggest that certain information about flight kinematics for any pteropodid species can be assumed to be representative for the group, despite the remarkable range in body sizes in the family. For the kinematic parameters we report here, wing kinematics were similar across the bats surveyed in this study despite more than a 40-fold range in body mass among individuals. Downstroke period scaled as predicted under isometric scaling, and several kinematic variables showed no significant change with body size (τ, φ, β, St, α2, wing camber). However, a few variables did change with increasing body size, namely T, α and CL.

Although we found that the relationship between Vhoriz and Mb did not differ significantly from expected, there was no significant increase in Vhoriz with Mb. Thus, without air moving faster across the wings, the higher wing loading values of large bats should have required them to have higher lift coefficients than small bats did, and this is what we observed. The higher CL values of large bats were most likely achieved by their higher wingbeat frequencies (for their size) and their higher angles of attack, since no other variables changed across body sizes. Importantly, the departure from isometry in Smax did not contribute to this trend, since the CL equation (Eqn 5) accounts for wing area.

Strouhal number can be employed as a dimensionless descriptor of flight speed for flapping swimmers and fliers. Regardless of body size, when animals fly with Strouhal numbers between 0.2 and 0.4, propulsive efficiency is thought to be particularly high (Taylor et al., 2003). The Strouhal numbers of corridor-flown animals in this study were around 0.3 to 0.6, slightly higher than the predicted range, but they did not differ significantly with size. Strouhal numbers for pteropodid bats were published previously by Taylor et al. (Taylor et al., 2003) based on data from Bullen and McKenzie (Bullen and McKenzie, 2002), who used two corridor-flown pteropodid species with body masses within the range we investigated here (Pteropus poliocephalus: 0.7 kg, and P. scapulatus 0.4 kg). Their values were around 0.42, roughly the middle of our range for corridor-flown bats.

Bats flown in the wind tunnel had lower Strouhal numbers than corridor-flown bats did, probably as a result of their relatively high flight velocities. After all, every species decreased St with increases in flight speed (Table 4). Interestingly, the Strouhal numbers of wind-tunnel-flown bats fell exactly within the expected 0.2 to 0.4 range. This suggests that when bats fly at high speed they do not adjust their amplitude and frequency to maintain consistent Strouhal numbers. We hypothesize that because we used a wind tunnel to collect data for C. brachyotis, we artificially decreased Strouhal number. Conversely, Tobalske et al. (Tobalske et al., 1997) found that the wingbeat frequencies of birds in their study were higher at a given speed in the wind tunnel compared with in the wild, which would increase Strouhal number. Similarly, Liechti and Bruderer found that wind-tunnel flown birds showed higher wingbeat frequencies than those observed in the wild (Liechti and Bruderer, 2002). Clearly a survey of Strouhal number across taxa that takes experimental conditions into account would be informative.

Changes in wing kinematics with velocity and acceleration

The results from our regression analyses varied across species, but a noteworthy trend emerged: when the partial regression coefficient between some variable and one of the regression effects was significantly different from zero for any species, other species with significant regression slopes almost always showed the same directional trend (Table 4). For example, not all bats had significant changes in wingbeat period with Vhoriz, but where a significant trend existed, it was always positive. This consistency is remarkable considering that we investigated 13 variables with respect to each of Vhoriz, Ahoriz and Avert for a total of 39 regressions per species, and found non-conflicting results from all six species in every case but one (bminvs Vhoriz). The stability of our results suggests that across a broad range of body sizes, bats modify their kinematics similarly with changes in velocity and acceleration.

Kinematic changes with flight velocity

Several previous studies of bat flight have considered the kinematic changes that correlate with differences in flight velocity for bats flying at steady speeds (e.g. Aldridge, 1986; Aldridge, 1987; Lindhe Norberg and Winter, 2006). As Vhoriz increases, drag increases so that more thrust is required, and air velocity over the wing is increased, enhancing lift. Thus, changes in wing kinematics with increased Vhoriz might be related to increased thrust production, to drag reduction, to CL alleviation, or to some combination of those (Hedrick et al., 2002; Tobalske et al., 2007).

Strouhal number, by definition, is inversely proportional to velocity (Eqn 4), so it is not surprising that Strouhal number decreased with increasing Vhoriz. Similarly, the mid-downstroke velocity of the wrist, correlated with flight velocity, appears in the denominator of the equation for coefficient of lift (Eqn 5), and we see the expected trend of decreased CL with increased Vhoriz. In essence, a bat flying quickly has a greater airspeed across the wings than it does flying slowly, and since it weighs the same under those two circumstances, the coefficient of lift must decrease with speed.

All species showed a decrease in stroke plane angle with increased flight velocity, as has been reported for other bat species previously (Aldridge, 1987). Although the wake patterns generated by bats are complex (Hedenström et al., 2007; Hubel et al., 2009), simple models such as actuator disk theory can capture some of the relevant connection between kinematics and aerodynamic force production. According to these ideas, a decrease in stroke plane angle should direct the induced velocity of the wing motion more rearward, thereby shifting the contribution of induced velocity towards increased thrust and away from lift generation, to simultaneously overcome increased drag and diminish the lift production of the wings (Pennycuick, 1975). This explanation has been applied to changes in stroke plane angle with speed in previous studies (Norberg, 1990; Pennycuick, 1975), and fits our results as well.

Bats showed a trend toward longer wingbeat cycle with increased velocity, and the downstroke phase of the wingbeat cycle was particularly long. Amplitude, however, did not change with speed. This differs from the trends reported for birds, in which wingbeat period increases with speed because of changes in the duration of upstroke, and amplitude does not change (Hedrick et al., 2002; Hedrick et al., 2003). The lengthening of the downstroke by bats may contribute to the reduction in CL that we observed by slowing the downward motion of the wings during downstroke. Angle of attack and wing camber also both decreased with increased flight velocity, and those trends would also both reduce lift.

Only C. brachyotis showed a significant decrease in both maximum and minimum wingspan with increases in flight velocity. We hypothesize that this resulted because C. brachyotis were flown in a wind tunnel, and therefore flew at higher velocities relative to their body sizes than did the corridor-flown bats. Since drag increases exponentially with velocity, C. brachyotis probably encountered relatively larger drag forces than did the other bats. We hypothesize that C. brachyotis decreased overall wingspan with increasing flight velocity to reduce drag and that the difference between C. brachyotis and other species reflects this. We hypothesize that reduction of maximum wingspan with flight speed only occurs for bats at high flight speeds. We predict that maximum wingspan would not change with Vhoriz below some threshold speed for C. brachyotis, and that other bat species would show a negative correlation between wingspan and flight speed at high speeds.

Horizontal accelerations: thrust production and drag reduction

Over the course of a wingbeat cycle, the orientations of aerodynamic thrust, drag and lift change relative to global axes and can vary along the span of the flapping wing (Norberg, 1976). However, if we consider only the summation of forces over the whole wingbeat cycle, and if we neglect deviations from horizontal flight, we can treat horizontally directed net force as the sum of thrust and drag, and vertically directed net force as lift. This approach simplifies comparisons among individuals and flight conditions, and we employ that simplification here.

The net Ahoriz of a flying bat results from the sum of forward thrust and rearward drag over the course of the entire wingbeat cycle, and the relative contributions of each over time cannot be separated in a purely kinematic analysis (Hedrick et al., 2002). Thus, the kinematic correlates of increased Ahoriz might increase thrust, decrease drag, or do both. Still other changes might not influence those aerodynamic forces directly but change as a secondary result of kinematic changes that do.

When bats increased Ahoriz, their stroke plane angles became more negative, thereby making the path of the wingtip more vertical. We hypothesize that a more vertical wing path shifts the orientation of shed vortices rearward, thereby increasing thrust (Pennycuick, 1975). With increases in Ahoriz some species also increased wingspan, which would reduce the magnitude of induced drag slightly and might also increase thrust by sweeping the wings through a greater volume of air.

Strouhal number increased with Ahoriz, probably as a secondary result of the decreased stroke plane angle; because stroke plane angle became more vertical while amplitude either increased or did not change, the maximum vertical distance traveled by the wingtip increased, and thus so did Strouhal number.

Angle of attack did not change significantly with Ahoriz for most species, but its component angles, α1 and α2, did. All species showed decreases in α2 with increasing Ahoriz, probably also because of the decreased stroke plane angle. Since wingbeat frequency did not change with Ahoriz, the decrease in stroke plane angle caused the leading edge of the wing to move more vertically during downstroke relative to the oncoming flow, resulting in an increase in α2 for all species. That increase alone would have increased the angle of attack, but the overall angle of attack did not increase with Ahoriz for any species except P. hypomelanus. This occurred because all species except P. hypomelanus compensated for the increased α2 by holding the wing in a more horizontal plane during downstroke, that is, decreasing α1. The decrease in α1 with Ahoriz was significant for three species, and nearly significant for two (P<0.07). We hypothesize that as the result of those kinematic changes, bats were able to take advantage of the added thrust that resulted from a decreased stroke plane angle without suffering the drag-inducing effects of an increased angle of attack.

A few species showed trends with increases in Ahoriz not seen in other species. Pteropus pumilus increased wing stroke amplitude with Ahoriz, P. hypomelanus decreased the downstroke ratio and increased angle of attack, and P. vampyrus increased its lift coefficient. No bats showed significant changes in wingbeat period, downstroke duration, or wing camber.

Vertical accelerations: lift production

Vertical acceleration results from increased lift, so the kinematic correlates of Ahoriz should cause increased lift, or result from correlations to kinematic changes that do so. Many kinematic parameters, such as Tdown, τ, φ and β, did not change significantly with increased lift production, even though aerodynamic theory would suggest that many of those variables could influence lift if modified (Norberg, 1990). Bat species typically increased lift production by flapping their wings more quickly (decreasing T), though the duration of the downstroke was not correlated with lift production for any species. Bats also increased lift by increasing wingspan and wing camber. Angle of attack increased with lift production for most species, but that trend was only significant for C. brachyotis. Not surprisingly, the lift coefficient, computed with Avert in the numerator (Eqn 5), was positively correlated with Avert.

Wind tunnel versus corridor flights

Wind tunnels provide substantial benefits for scientific investigations over the use of animals in free flight, including that the investigator can control the speed at which the animal carries out locomotion, and can collect data for a considerable period of time without need to move equipment alongside a free-moving animal. However, wind tunnels might induce kinematics that would not otherwise occur. For example, Tobalske et al. and Liechti and Bruderer found that birds flown in a wind tunnel had higher wingbeat frequencies than birds of the same species did in the wild (Liechti and Bruderer, 2002; Tobalske et al., 1997). Similarly, we believe that a number of differences between C. brachyotis and the other species used in this study probably resulted from the use of a wind tunnel for that species. Specifically, C. brachyotis flown in the wind tunnel flew at generally higher velocities and lower Strouhal numbers than would be predicted based on the size–velocity relationship uncovered for corridor-flown bats (Fig. 4A, Fig. 5F). This speed difference probably underlies all other kinematic differences between C. brachyotis and the trends we saw in the corridor-flown bats.

If the best fit lines for the five corridor-flown species are used to predict what C. brachyotis would have done if flown in a corridor, we see that wingbeat period was higher than expected (Fig. 5A), α2 slightly lower than expected (Fig. 5I), and wing camber and coefficient of lift were lower than those of other species (Fig. 5J,K). Interestingly, all of these trends are what would be expected for high-speed flight based on our regression analyses (Table 4). With increasing Vhoriz, bats increased wingbeat period and decreased wing camber and coefficient of lift. Also, when bats increased thrust production to increase Ahoriz, they increased α2, and it is likely that for fast flight, such as that of C. brachyotis in a wind-tunnel, the relatively higher drag necessitated increased thrust. Obviously, complete resolution of flight speed and the use of a wind tunnel on wing kinematics requires detailed investigation of individual bats flown in a corridor and wind tunnel over a comparable range of speeds, but our results provide an intriguing view of biases that may be introduced by wind tunnel experiments in studies of aerial locomotion.

Implications for non-pteropodid bats

Many of the trends reported here are consistent with patterns observed previously for other bats, but the data in this study might not be representative for bats of all species. Most bats are insectivorous, and many require high levels of maneuverability to catch their evasive food in flight. By contrast, all six species in this study are frugivorous bats that fly nightly between their roosts and food resources and this was probably also true of their last common ancestor (Boon and Corlett, 1989; DeFrees and Wilson, 1988; Jones and Kunz, 2000; Kunz and Jones, 2000; Kwiecinski and Griffiths, 1999; Luft et al., 2003; Giannini and Simmons, 2005; Jones et al., 2002; Teeling et al., 2005). To determine whether bats with improved maneuverability show different kinematic trends with speed and acceleration to those reported here is an exciting objective for future investigations.

Finally, no other family of bats approaches the largest body sizes of pteropodids, and there is a tremendous diversity of bats below the 33 g body mass of our smallest individuals. Small bats tend to fly at lower speeds than the bats in this study did (Akins et al., 2007), suggesting that the VhorizMb1/6 trend reported here might be more robust among bats with smaller body sizes than we observed here. Recent work has revealed that small bats generate leading-edge vortices while flying at low speeds, much like flapping insects do (Muijres et al., 2008), and parallel experiments have not yet been performed for larger bats. Thus, the aeromechanics of bats may differ across the lower spectrum of bat body sizes. Our results demonstrate remarkable consistency in flight mechanics for medium to large-sized bats, and we look forward to parallel studies of smaller bats, especially from non-pteropodid families, that test the applicability of our results to bats in general.

We thank Allyce Sullivan, Pere Tiemo, and Sarah Taylor for assistance in data collection, and thank the many undergraduates at Brown University who assisted in digitizing the movies for this project. We thank Yvonne Dzal, Ty Hedrick, David Lee, David Lentink, Crystal Linkletter, members of the Swartz and Breuer lab groups, members of the Morph Group at Brown University, and two anonymous reviewers for helpful discussions around this project. We also thank Allyson Walsh and the Lubee Bat Conservancy for access to bats and facilities for data collection. This study was supported by the United States Air Force Office of Scientific Research (AFOSR) and the National Science Foundation (NSF).

     
  • Ahoriz

    net forward acceleration for the wingbeat cycle (m s–2)

  •  
  • AR

    aspect ratio (dimensionless)

  •  
  • Avert

    net vertical acceleration for the wingbeat cycle (m s–2)

  •  
  • bmax

    maximum wingspan (m)

  •  
  • bmin

    minimum wingspan (m)

  •  
  • CL

    coefficient of lift (dimensionless)

  •  
  • cmax

    maximum wing chord (m)

  •  
  • COM

    center of mass

  •  
  • d.f.

    degrees of freedom

  •  
  • DLT

    direct linear transformation

  •  
  • g

    acceleration of gravity (9.81 m s–2)

  •  
  • GLM

    generalized linear model

  •  
  • Mb

    body mass (kg)

  •  
  • Qs

    wing loading (N m–2)

  •  
  • RMA

    reduced major axis

  •  
  • S

    wing area (m2)

  •  
  • Smax

    maximum wing area (m2)

  •  
  • St

    Strouhal number

  •  
  • T

    wingbeat period (s)

  •  
  • Tdown

    downstroke duration (s)

  •  
  • Vhoriz

    forward velocity (m s–1)

  •  
  • Vvert

    vertical velocity (m s–1)

  •  
  • Vwrist

    velocity of the wrist in the xgzg plane at the time of max wingspan (m s–1)

  •  
  • xb

    body-centered x dimension

  •  
  • xg

    global x dimension

  •  
  • yb

    body-centered y dimension

  •  
  • yg

    global y dimension

  •  
  • zb

    body-centered z dimension

  •  
  • zg

    global z dimension

  •  
  • α

    angle of attack at mid-downstroke (deg; α=α12)

  •  
  • α1

    angle of wing chord to horizontal at mid-downstroke (deg)

  •  
  • α2

    angle of wrist trajectory to oncoming flow at mid-downstroke (deg)

  •  
  • β

    stroke plane angle (deg)

  •  
  • ρ

    density of air (1.204 kg m–3)

  •  
  • τ

    downstroke ratio (dimensionless)

  •  
  • φ

    stroke amplitude (deg)

Abdel-Aziz
Y. I.
,
Karara
H. M.
(
1971
).
Direct linear transformation from comparator coordinates into object space coordinates in close-range photogrammetry
. In
Proceedings of the Symposium on Close-Range Photogrammetry
, pp.
1
-
18
.
Falls Church, VA
:
American Society of Photogrammetry
.
Akins
J. B.
,
Kennedy
M. L.
,
Schnell
G. D.
,
Sánchez-Hernández
C.
,
Romero-Almaraz
M. De L.
,
Wooten
M. C.
,
Best
T. L.
(
2007
).
Flight speeds of three species of Neotropical bats: Glossophaga soricina, Natalus stramineus, and Carollia subrufa
.
Acta Chiropterol.
9
,
477
-
482
.
Aldridge
H. D. J. N.
(
1986
).
Kinematics and aerodynamics of the greater horseshoe bat, Rhinolophus ferrumequinum, in horizontal flight at various speeds
.
J. Exp. Biol.
126
,
479
-
497
.
Aldridge
H. D. J. N.
(
1987
).
Body accelerations during the wingbeat in six bat species: the function of the upstroke in thrust generation
.
J. Exp. Biol.
130
,
275
-
293
.
Biewener
A. A.
(
1983
).
Allometry of quadrupedal locomotion: the scaling of duty factor, bone curvature and limb orientation to body size
.
J. Exp. Biol.
105
,
147
-
171
.
Biewener
A. A.
(
2005
).
Biomechanical consequences of scaling
.
J. Exp. Biol.
208
,
1665
-
1676
.
Boon
P. P.
,
Corlett
R. T.
(
1989
).
Seed dispersal by the lesser short-nosed fruit bat (Cynopterus brachyotis, Pteropodidae, Megachiroptera)
.
Malay. Nat. J.
42
,
251
-
256
.
Brown
J. H.
,
West
G. B.
(
2000
).
Scaling in Biology
.
Oxford
:
Oxford University Press
.
Bullen
R. D.
,
McKenzie
N. L.
(
2001
).
Bat airframe design: flight performance, stability and control in relation to foraging ecology
.
Aust. J. Zool.
49
,
235
-
261
.
Bullen
R. D.
,
McKenzie
N. L.
(
2002
).
Scaling bat wing beat frequency and amplitude
.
J. Exp. Biol.
205
,
2615
-
2626
.
Combes
S. A.
,
Daniel
T. L.
(
2003
).
Into thin air: contributions of aerodynamic and inertial-elastic forces to wing bending in the hawkmoth Manduca sexta
.
J. Exp. Biol.
206
,
2999
-
3006
.
Curran-Everett
D.
(
2000
).
Multiple comparison: philosophies and illustrations
.
Am. J. Physiol. Reg. Integr. Comp. Physiol.
279
,
R1
-
R8
.
Curran-Everett
D.
,
Benos
D. J.
(
2004
).
Guidelines for reporting statistics in journals published by the American Physiological Society
.
J. Appl. Physiol.
97
,
457
-
459
.
DeFrees
S. L.
,
Wilson
D. E.
(
1988
).
Eidolon helvum. Mamm. Sp.
312
,
1
-
5
.
Dial
K. P.
,
Biewener
A. A.
(
1993
).
Pectoralis muscle force and power output during different modes of flight in pigeons (Columba livia)
.
J. Exp. Biol.
176
,
31
-
54
.
Dial
K. P.
,
Greene
E.
,
Irschick
D. J.
(
2008
).
Allometry of behavior
.
Trends Ecol. Evol.
23
,
394
-
401
.
Dickinson
M. H.
,
Götz
K. G.
(
1996
).
The wake dynamics and flight forces of the fruit fly, Drosophila melanogaster
.
J. Exp. Biol.
199
,
2085
-
2104
.
Ellington
C. P.
(
1999
).
The novel aerodynamics of insect flight: applications to micro-air vehicles
.
J. Exp. Biol.
202
,
3439
-
3448
.
Felsenstein
J.
(
1985
).
Phylogenies and the comparative method
.
Am. Nat.
125
,
1
-
15
.
Garland
T.
Jr
,
Díaz-Uriarte
R.
(
1999
).
Polytomies and phylogenetically independent contrasts: an examination of the bounded degrees of freedom approach
.
Syst. Biol.
48
,
547
-
558
.
Gatesy
S. M.
,
Baier
D. B.
(
2005
).
The origin of the avian flight stroke: a kinematic and kinetic perspective
.
Paleobiology
31
,
382
-
399
.
Gelman
A.
,
Hill
J.
(
2007
).
Data Analysis Using Regression and Multilevel/Hierarchical Models
.
Cambridge
:
Cambridge University Press
.
Giannini
N. P.
,
Simmons
N. B.
(
2005
).
Conflict and congruence in a combined DNA-morphology analysis of megachiropteran bat relationships (Mammalia: Chiroptera: Pteropodidae)
.
Cladistics
21
,
411
-
437
.
Greenewalt
C. H.
(
1975
).
The flight of birds: the significant dimensions, their departure from the requirements for dimensional similarity, and the effect of flight aerodynamics of that departure
.
Trans. Am. Phil. Soc.
65
,
1
-
67
.
Hedenström
A.
,
Johansson
L. C.
,
Wolf
M.
,
von Busse
R.
,
Winter
Y.
,
Spedding
G. R.
(
2007
).
Bat flight generates complex aerodynamic tracks
.
Science
316
,
894
-
897
.
Hedrick
T. L.
,
Tobalske
B. W.
,
Biewener
A. A.
(
2002
).
Estimates of circulation and gait change based on a three-dimensional kinematic analysis of flight in cockatiels (Nymphicus hollandicus) and ringed turtle-doves (Streptopelia risoria)
.
J. Exp. Biol.
205
,
1389
-
1409
.
Hedrick
T. L.
,
Tobalske
B. W.
,
Biewener
A. A.
(
2003
).
How cockatiels (Nymphicus hollandicus) modulate pectoralis power output across flight speeds
.
J. Exp. Biol.
206
,
1363
-
1378
.
Heglund
N. C.
,
Taylor
C. R.
(
1988
).
Speed, stride frequency and energy cost per stride: how do they change with body size and gait?
J. Exp. Biol.
138
,
301
-
318
.
Hill
J. E.
,
Smith
S. A.
(
1981
).
Craseonycteris thonglongyai
.
Mamm. Sp.
160
,
1
-
4
.
Hubel
T. Y.
,
Hristov
N. I.
,
Swartz
S. M.
,
Breuer
K. S.
(
2009
).
Time-resolved wake structure and kinematics of bat flight
.
Exp. Fluids
46
,
933
-
943
.
Jones
D. P.
,
Kunz
T. H.
(
2000
).
Pteropus hypomelanus
.
Mamm. Sp.
639
,
1
-
6
.
Jones
K. E.
,
Purvis
A.
,
MacLarnon
A.
,
Bininda-Emonds
O. R. P.
,
Simmons
N. B.
(
2002
).
A phylogenetic supertree of the bats (Mammalia: Chiroptera)
.
Biol. Rev.
77
,
223
-
259
.
Kingston
T.
,
Jones
G.
,
Zubaid
A.
,
Kunz
T. H.
(
2000
).
Resource partitioning in rhinolophid bats revisited
.
Oecologia
124
,
332
-
342
.
Kunz
T. H.
,
Jones
D. P.
(
2000
).
Pteropus vampyrus
.
Mamm. Sp.
642
,
1
-
6
.
Kwiecinski
G. G.
,
Griffiths
T. A.
(
1999
).
Pteropus hypomelanus
.
Mamm. Sp.
611
,
1
-
9
.
LaBarbera
M.
(
1989
).
Analyzing body size as a factor in ecology and evolution
.
Annu. Rev. Ecol. Syst.
20
,
97
-
117
.
Lavin
S. R.
,
Karasov
W. H.
,
Ives
A. R.
,
Middleton
K. M.
,
Garland
T.
Jr
(
2008
).
Morphometrics of the avian small intestine compared with that of nonflying mammals: a phylogenetic approach
.
Physiol. Biochem. Zool.
81
,
526
-
550
.
Liechti
F.
,
Bruderer
L.
(
2002
).
Wingbeat frequency of barn swallows and house martins: a comparison between free flight and wind tunnel experiments
.
J. Exp. Biol.
205
,
2461
-
2467
.
Lindhe Norberg
U. M.
,
Winter
Y.
(
2006
).
Wing beat kinematics of a nectar-feeding bat, Glossophaga soricina, flying at different flight speeds and Strouhal numbers
.
J. Exp. Biol.
209
,
3887
-
3897
.
Luft
S.
,
Curio
E.
,
Tacud
B.
(
2003
).
The use of olfaction in the foraging behaviour of the golden-mantled flying fox, Pteropus pumilus, and the greater musky fruit bat, Ptenochirus jagori (Megachiroptera: Pteropodidae)
.
Naturwissenshaften
90
,
84
-
87
.
McGahan
J.
(
1973
).
Gliding flight of the Andean condor in nature
.
J. Exp. Biol.
58
,
225
-
237
.
Miller
L. A.
,
Peskin
C. S.
(
2005
).
A computational fluid dynamics of ‘clap and fling’ in the smallest insects
.
J. Exp. Biol.
208
,
195
-
212
.
Muijres
F. T.
,
Johansson
L. C.
,
Barfield
R.
,
Wolf
M.
,
Spedding
G. R.
,
Hedenström
A.
(
2008
).
Leading-edge vortex improves lift in slow-flying bats
.
Science
319
,
1250
-
1253
.
Norberg
U. M.
(
1976
).
Kinematics, aerodynamics and energetics of horizontal flapping flight in the long-eared bat (Plecotus auritus)
.
J. Exp. Biol.
65
,
179
-
212
.
Norberg
U. M.
(
1981
).
Allometry of bat wings and legs and comparison with bird wings
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
292
,
359
-
368
.
Norberg
U. M.
(
1990
).
Vertebrate Flight
.
Berlin
:
Springer-Verlag
.
Norberg
U. M.
,
Rayner
J. M. V.
(
1987
).
Ecological morphology and flight in bats (Mammalia; Chiroptera): wing adaptations, flight performance, foraging strategy and echolocation
.
Philos. Trans. R. Soc. Lond. B. Biol. Sci.
316
,
335
-
427
.
Nowak
R. M.
(
1994
).
Walker’s Bats of the World
.
Baltimore
:
Johns Hopkins University Press
.
Nudds
R. L.
(
2007
).
Wing-bone length allometry in birds
.
J. Avian Biol.
38
,
515
-
519
.
O’Brien
J.
,
Mariani
C.
,
Olson
L.
,
Russell
A. L.
,
Say
L.
,
Yoder
A. D.
,
Hayden
T. J.
(
2009
).
Multiple colonisations of the western Indian Ocean by Pteropus fruit bats (Megachiroptera: Pteropodidae): the furthest islands were colonised first
.
Mol. Phylogenet. Evol.
51
,
294
-
303
.
Pagel
M. D.
(
1992
).
A method for the analysis of comparative data
.
J. Theor. Biol.
156
,
431
-
442
.
Pennycuick
C. J.
(
1975
).
Mechanics of flight
. In
Avian Biology
, Vol.
5
(ed.
Farner
D. S.
,
King
J. R.
), pp.
1
-
75
.
New York
:
Academic Press
.
Purvis
A.
,
Garland
T.
, Jr (
1993
).
Polytomies in comparative analyses of continuous characters
.
Syst. Biol.
42
,
569
-
575
.
R Development Core Team
(
2009
).
R: A Language and Environment for Statistical Computing
.
Vienna, Austria
:
R Foundation for Statistical Computing
.
Riskin
D. K.
,
Willis
D. J.
,
Iriarte-Díaz
J.
,
Hedrick
T. L.
,
Kostandov
M.
,
Chen
J.
,
Laidlaw
D. H.
,
Breuer
K. S.
,
Swartz
S. M.
(
2008
).
Quantifying the complexity of bat wing kinematics
.
J. Theor. Biol.
254
,
604
-
615
.
Sane
S. P.
(
2003
).
The aerodynamics of insect flight
.
J. Exp. Biol.
206
,
4191
-
4208
.
Smith
R. J.
(
1984
).
Allometric scaling in comparative biology: problems of concept and method
.
Am. J. Physiol.
246
,
R152
-
R160
.
Storey
J.D.
(
2002
).
A direct approach to false discovery rates
.
J. R. Stat. Soc. B
64
,
479
498
.
Surlykke
A.
,
Miller
L. A.
,
Møhl
B.
,
Andersen
B. B.
,
Christensen-Dalsgaard
J.
,
Jørgensen
M. B.
(
1993
).
Echolocation in two very small bats from Thailand: Craseonycteris thonglongyai and Myotis siligorensis
.
Behav. Ecol. Sociobiol.
33
,
1
-
12
.
Taylor
G. K.
,
Nudds
R. L.
,
Thomas
A. L. R.
(
2003
).
Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency
.
Nature
425
,
707
-
711
.
Teeling
E. C.
,
Springer
M. S.
,
Madsen
O.
,
Bates
P.
,
O’Brien
S. J.
,
Murphy
W. J.
(
2005
).
A molecular phylogeny for bats illuminates biogeography and the fossil record
.
Science
307
,
580
-
584
.
Tobalske
B. W.
,
Olson
N. E.
,
Dial
K. P.
(
1997
).
Flight style of the black-billed magpie: variation in wing kinematics, neuromuscular control and muscle composition
.
J. Exp. Zool.
279
,
313
-
329
.
Tobalske
B. W.
,
Warrick
D. R.
,
Clark
C. J.
,
Powers
D. R.
,
Hedrick
T. L.
,
Hyder
G. A.
,
Biewener
A. A.
(
2007
).
Three-dimensional kinematics of hummingbird flight
.
J. Exp. Biol.
210
,
2368
-
2382
.
Warrick
D. R.
,
Tobalske
B. W.
,
Powers
D. P.
(
2005
).
Aerodynamics of the hovering hummingbird
.
Nature
435
,
1094
-
1097
.
Wilson
D. E.
,
Reeder
D. M.
(
1993
).
Mammal Species of The World: a Taxonomic and Geographic Reference
.
Washington
:
Smithsonian Institute Press
.