Docsity
Docsity

Prepare-se para as provas
Prepare-se para as provas

Estude fácil! Tem muito documento disponível na Docsity


Ganhe pontos para baixar
Ganhe pontos para baixar

Ganhe pontos ajudando outros esrudantes ou compre um plano Premium


Guias e Dicas
Guias e Dicas

Aerodynamics of Hawkmoth Flight: Kinematics, Power Requirements, and Constraints, Notas de estudo de Engenharia de Produção

An analysis of free flight in the hawkmoth manduca sexta, discussing kinematic changes, aerodynamic significance, power requirements, and constraints. It also explores the role of leading-edge vortices and navier–stokes solvers in insect flight. Equations for induced velocity and power requirements.

Tipologia: Notas de estudo

Antes de 2010

Compartilhado em 04/11/2009

igor-donini-9
igor-donini-9 🇧🇷

4.5

(4)

419 documentos

1 / 23

Documentos relacionados


Pré-visualização parcial do texto

Baixe Aerodynamics of Hawkmoth Flight: Kinematics, Power Requirements, and Constraints e outras Notas de estudo em PDF para Engenharia de Produção, somente na Docsity! 2723The Journal of Experimental Biology 200, 2723–2745 (1997) Printed in Great Britain © The Company of Biologists Limited 1997 JEB0994 *Pre MegTHE MECHANICS OF FLIGHT IN THE HAWKMOTH MANDUCA SEXTA II. AERODYNAMIC CONSEQUENCES OF KINEMATIC AND MORPHOLOGICAL VARIATION ALEXANDER P. WILLMOTT* A N D CHARLES P. ELLINGTON Department of Zoology, University of Cambridge, Downing Street, Cambridge CB2 3EJ, UK Accepted 4 August 1997Mean lift coefficients have been calculated for hawkmoth flight at a range of speeds in order to investigate the aerodynamic significance of the kinematic variation which accompanies changes in forward velocity. The coefficients exceed the maximum steady-state value of 0.71 at all except the very fastest speeds, peaking at 2.0 or greater between 1 and 2 m s−1. Unsteady high-lift mechanisms are therefore most important during hovering and slow forward flight. In combination with the wingtip paths relative to the surrounding air, the calculated mean lift coefficients illustrate how the relative contributions of the two halfstrokes to the force balance change with increasing forward speed. Angle of incidence data for fast forward flight suggest that the sense of the circulation is not reversed between the down- and upstrokes, indicating a flight mode qualitatively different from that proposed for lower-speed flight in the hawkmoth and other insects. The mid-downstroke angle of incidence is constant at 30–40 ° across the speed range. The relationship between power requirements and flight speed is explored; above 5 m s−1, further increases in forward velocity are likely to be constrained by available mechanical power, although problems with thrust generation and flight stability may also be involved. Hawkmoth wing and body morphology, and the differences between males and females, are evaluated in aerodynamic terms. Steady-state force measurements show that the hawkmoth body is amongst the most streamlined for any insect. Key words: aerodynamics, hawkmoth, Manduca sexta, lift coefficient, power requirements, morphology. SummaryA detailed analysis of free flight in the hawkmoth Manduca sexta L. has revealed the kinematic changes as speed increases from hovering to fast forward flight (Willmott and Ellington, 1997). In this study, we investigate the aerodynamic significance of the observed kinematic variation, the power requirements for flight at different speeds and the nature of the constraints on maximum flight speed. Insect flapping flight represents an unusual aerodynamic problem because of the inherent ‘unsteadiness’ and the low Reynolds number of the airflow. A large number of models for unsteady animal flight have been formulated, and these have been categorized and evaluated in recent reviews by Spedding (1992), Spedding and DeLaurier (1996) and Smith et al. (1996). The techniques range from those incorporating momentum and blade-element theories to those employing lifting-line or lifting-surface methods, but each requires non- trivial simplifying assumptions. Recent unsteady panel methods, for example, are more advanced than their predecessors in considering trailing wake vorticity and Introductionsent address: Kawachi Millibioflight Project, Japan Science and Te uro-ku, Tokyo 153, Japan (e-mail: sandy@kawachi.jst.go.jp).dynamic effects (Smith et al. 1996), and interactions between wing deformation and the aerodynamic forces (Smith, 1996), but they still have important shortcomings. In particular, they do not yet incorporate the leading-edge vortices that have a profound influence on the flow around hawkmoth wings at all speeds (Willmott et al. 1997) and which are likely to play a major role in insect flight aerodynamics (Ellington et al. 1996). At the low Reynolds numbers and high angles of incidence characteristic of insect flight, viscous effects in the near field are better addressed by Navier–Stokes solvers, which are now beginning to be applied to animal locomotion (e.g. Liu et al. 1996). Advanced numerical techniques will, undoubtedly, become standard in animal flight investigations. Their introduction will, however, be delayed by the paucity of detailed and relevant empirical data for their validation and by the need to investigate fully the inherent errors in the new techniques and their applicability to other conditions (Visbal, 1986; Spedding, 1992). Accurate and reliable modelling of instantaneous forceschnology Corporation (JST), Park Building 3F, 4-7-6 Komaba, 2724 A. P. WILLMOTT AND C. P. ELLINGTONis not an option available at present. Instead, in this report, we concentrate in more general terms on the lift-generating requirements placed on the wings during flight at different speeds and, in particular, on the possible aerodynamic consequences of the kinematic trends described in Willmott and Ellington (1997). The ‘mean coefficients’ method used here is a slightly modified momentum jet/blade-element approach of the type used by Dudley (1990), Dudley and Ellington (1990b), Cooper (1993) and Wakeling and Ellington (1997b). This technique has the advantage of being the most accessible of the current models – using data on the core kinematic parameters – and of being easily modified to incorporate new kinematics or wing morphology. Despite its simplifying assumptions, the mean coefficients model has been shown in a wide range of studies (e.g. Osborne, 1951; Dudley and Ellington, 1990b; Dudley, 1995) to be capable of producing meaningful estimates of aerodynamic forces and power requirements. It has also generated workable hypotheses linking flight kinematics and aerodynamics, thus meeting Spedding’s (1992) definition of a ‘useful’ aerodynamic model. Such models are necessarily refined as and when new techniques become available. The model is used here to determine the mean lift coefficient at each of the flight speeds in the kinematic study and to estimate the associated aerodynamic and inertial power components. The coefficients are calculated under the assumption that they are constant throughout each halfstroke of the wingbeat. They thus represent the lowest possible lift coefficient required from the wings for a given set of wingbeat kinematics; if the coefficients are not constant (a more realistic scenario), then instantaneous values at certain points of the wingbeat cycle must exceed these mean values (Ellington, 1984a). A number of related questions are also addressed, such as the aerodynamic function of the upstroke and the aerodynamic and energetic significance of the asymmetric wingbeat. For comparative purposes, the lifting performance of hawkmoth wings and bodies under ‘steady’ flow conditions was measured. Finally, the morphology of adult Manduca sexta was investigated and the possible aerodynamic significance of the observed differences between the sexes, and between hawkmoths and other insects, is discussed. Materials and methods Measurement of steady-state wing and body forces Steady-state lift and drag were measured using the optoelectronic force transducer described by Dudley and Ellington (1990b). A summary of its design and operation will be given here; details can be found in the earlier paper. The transducer measured the displacement in two orthogonal directions of a stainless-steel tube to which a test object was attached via a mounting pin. The voltage output of the transducer had a linear relationship to the moment applied to the tube, and this was calibrated at the beginning and end of each set of measurements. A range of replaceable spring elements covered forces on the test object of up to 0.016 N.The transducer was positioned at the mouth of the open-jet wind tunnel used in the kinematic study (Willmott and Ellington, 1997). Aerodynamic forces on the body The wings were removed from freshly killed Manduca sexta whose body posture was similar to that seen in free flight. Owing to the problems of orienting the legs and antennae in a realistic position, and to the difficulty in determining such positions from the high-speed videos, these structures were also removed. The body was mounted at the top of a stainless- steel entomological pin which passed through the left and right wingbases. The bodies were left to ‘set’ for 1 day before use in order to fix the abdomen in flight position. Lift and drag measurements were made on six bodies for airspeeds from 1 to 5 m s−1, and at the following angles of attack: −15 °, 0 °, 15 °, 25 °, 35 °, 45 ° and 60 °. The 0 ° angle was set with the longitudinal axis of the body parallel to the airflow; positive angles of attack refer to the body being pitched head-up with respect to the horizontal. In order to standardize the Reynolds numbers based on body length Reb for the different bodies, the slight differences in body length were corrected for by slight modifications in the airspeed. The Reynolds numbers used were 3150, 6040, 9150, 12 270 and 15 220, corresponding to speeds of approximately 1, 2, 3, 4 and 5 m s−1, respectively. Aerodynamic forces on the wings Before the wings were removed from the bodies of freshly killed moths, the left wing couple was moved into a realistic flight orientation. A small quantity of cyanoacrylate gel was applied between the dorsal surface of the hindwing and the ventral surface of the forewing, close to the wingbase, to hold the wings in the correct alignment. The wing couple was then detached carefully from the body by cutting through the wingbases as close as possible to the body. At this point, the anojugal flap folded under the hindwing. It was not possible to return this region to its correct position in fresh wings, and so the flap was removed by cutting along its fold line. The wing was mounted on a stainless-steel pin with the longitudinal axis of the wing aligned with the shaft of the pin: the head of the pin was removed and its final 5 mm bent perpendicular to the main shaft to form a short hook which was then attached, using a small drop of beeswax, near to the base of the underside of the wings. The Manduca sexta wing is not a planar surface: a spanwise gradient of wing twist is inherent in its structure. The wing was mounted so that the chosen angle of attack corresponded to the inclination of the wing halfway between the wingbase and wingtip, close to the radius of the first moment of wing area which is where the steady-state aerodynamic force is assumed to act. The angle of attack in the proximal area of the wing was up to approximately 5 ° more positive than this value, whilst the angle in the distal region was up to 5 ° more negative. Lift and drag measurements were made at 10 ° angle of attack intervals from −50 ° to +70 °, and at Reynolds numbers based on mean 2727The aerodynamics of hawkmoth flightCalculation of mechanical power requirements Induced power The induced power Pind was calculated using the expression: Pind = w0(mg − BzCD,pro) , (12) where Bz is the sum of Bz,d and Bz,u. This is a departure from previous models, where the induced power was simply w0mg (but see also Wakeling and Ellington, 1997b). The modification recognises that the vertical component of profile drag cannot be ignored when considering the energetic cost of generating weight support. Three potential cases exist under which induced power must be calculated: for no net vertical component of profile drag, for a net upward force, and for a net downward force. Equation 12 is common to all three. In the first case, when BzCD,pro is zero, the momentum jet must generate a vertical force equal to the entire weight of the insect, mg. The induced power associated with the jet is the product of the thrust and the appropriate induced velocity w0, calculated from equations 1 and 2. This special case is equivalent to the model used in earlier studies. When BzCD,pro is positive, however, profile drag contributes to weight support, and the thrust requirement for the momentum jet is reduced accordingly. A new, lower induced velocity should be calculated by replacing mg in equation 2 with the reduced value for thrust, and the thrust and induced velocity iterated until the induced power converges. This would be a time-consuming process for what would only be a second-order correction to Pind. The first-order correction given by equation 12 is pragmatic wherever the contribution of profile drag to the vertical force balance is small. The validity of this assumption will be discussed below. In the final case, profile drag contributes a net downwards vertical component, BzCD,pro is negative, and the momentum jet thrust must be increased for net weight support. From equation 12, the induced power also increases. The total power requirement for weight support is higher here than in the first two cases (in which the total is identical). The additional power produces no extra useful aerodynamic force. Instead, it represents energy wasted by the conflicting flows of the net upwash created by air dragged along with the wings and the downwash from the momentum jet. Profile power The instantaneous profile power Ppro required to overcome the profile drag is the summation along the wing of the products of sectional profile drag and relative velocity. The total profile power over the wingbeat is calculated from a similar summation to that used for profile drag: Parasite power The parasite power Ppar required to overcome parasite drag is given by the product of the parasite drag and the forward flight speed: Ppar = VDpar . (14) pro D,proP C U∑ r it = ∑12 ρdr c 3 . (1Inertial power Inertial power Pacc is required to accelerate the wings early in each halfstroke. The mean value over the wingbeat is the sum of the kinetic energy imparted to the wings during the up- and downstrokes divided by the cycle period: where n is the wingbeat frequency, ω is the radian frequency and I is the moment of inertia of the wings and their virtual mass. The values in parentheses are the maximum angular velocities during the upstroke and downstroke, respectively, resulting from changes in both sweep and elevation angle: Equation 15 gives the mean value for inertial power over the two halfstrokes. In the asymmetric Manduca sexta wingbeat, the mean inertial power during the longer downstroke acceleration will be lower than this, while the mean inertial power during the upstroke will be higher. During the deceleration phase of each halfstroke, kinetic energy is released by the wing. This energy can be used for aerodynamic work, and any excess may either be stored as elastic energy or dissipated as heat and sound (Ellington, 1984e). Total mechanical power required from the flight muscle The aerodynamic power Paero is the sum of the induced, parasite and profile powers. The mean total mechanical power required from the flight muscles over the course of the wingbeat must include the fourth component, the inertial power. If the metabolic cost of doing negative work during wing deceleration is assumed to be negligible compared with the cost of doing positive work, then the mean positive mechanical power depends upon the extent to which any excess negative work during deceleration can be stored in elastic structures in the thorax/flight muscle system. Ellington (1984e) derived expressions for the total mean positive mechanical power requirements Ppos for hovering flight under the two extreme conditions of perfect and zero elastic storage. For forward flight, parasite power must be added to Ellington’s (1984e) equations which, using the terminology adopted in the present paper, then become: Ppos = Paero (17) for perfect elastic storage, and for zero elastic storage. The mechanical power requirements and their components were all normalized to body-mass- and muscle-mass-specific values, to facilitate comparison of the values among the different moths, and between Manduca sexta and other species. Wing and body morphology A range of morphological parameters were measured for six P P Ppos aero acc= +( )12 (18) d d d d d d ω φ θ t t t =    +     2 2 . (16) P nI t tacc max,u 2 max,d 2d d d d =    +           ω ω , (15) 2728 A. P. WILLMOTT AND C. P. ELLINGTONmale and seven female moths (including the three moths from the kinematic study; Willmott and Ellington, 1997). These parameters are defined in the Symbols list, and the methods used to determine them were those described by Ellington (1984a), with the exception of the radii of wing mass which were found by the strip-cutting approach he used for smaller insects rather than the compound pendulum technique. The wing area (and associated virtual mass) parameters were measured from the video images obtained in Willmott and Ellington (1997), with each wing divided into 50 strips of equal spanwise width. Results Steady-state aerodynamic characteristics of the body and wings The mean lift and drag coefficients for the Manduca sexta bodies are presented as a polar diagram in Fig. 1. The coefficients were calculated using the planform area of the bodies, and each point is the mean of measurements from six moths. Standard error bars for the means are not shown since most were smaller than the symbol size: the mean standard errors for the drag and lift coefficients were 0.018 and 0.021, respectively. The plotted lines show how the coefficients change with angle of attack at a constant Reynolds number Reb: the curves are very similar as far as an angle of attack of 35 °,0.2 0.2 0.4 0.6 0.8 0 C D 60° 45° 35° 25° 15° 0° 15° Reb = 3150 6040 9150 12 270 Reb = 15 220 C L 0 0.2 0.4 0.6 0.8 1 Fig. 1. Polar diagram showing the aerodynamic characteristics of Manduca sexta bodies. Each point is the mean from six bodies. The different curves represent the force coefficients at five different Reynolds numbers Reb, which are given at the right of the figure. The angle of attack, in degrees, is shown close to each cluster of points. CL, lift coefficient; CD, drag coefficient.but at the highest angles they diverge slightly with both lift and drag coefficients increasing with Reynolds number. The lift coefficients for the body increased steadily with increasing angle of attack from close to zero at 0 ° to a maximum of 0.6 at 45 ° and the highest Reynolds numbers, before decreasing to approximately 0.4 at 60 °. The drag coefficients increased steadily with increasing inclination from a minimum of 0.08 at 0 ° to reach 0.78 at 60 ° and Reb=15 220. The mean lift force under the maximum lift conditions (45 ° and Reb=15 220) was 22.4 % of body weight, but the body angles observed during free flight at 5 m s−1 were in the region of 15 and 25 °, and the relative body lift at these angles averaged 7.4 and 12.8 %, respectively. The maximum drag force, at 60 ° and Reb=15 220, was 29.5 % of body weight, but this dropped to 6.3 and 10.3 % at 15 and 25 °, respectively. The maximum lift:drag ratio is a measure of the aerodynamic performance of the body. For Manduca sexta bodies, this ratio ranged from 1.18 at Reb=6040 to 1.38 at Reb=3150. The latter value was recorded at an angle of attack of 15 °, but the maximum slope at every other Reynolds number occurred at an angle of 25 °. The mean force coefficients, based on wing planform area, for the wing couples are presented as a polar diagram in Fig. 2. The three curves represent the changes with angle of attack at Reynolds numbers Rec (using the mean chord as the characteristic length) of 1150, 3300 and 5560. The mean standard errors were 0.03 for the lift coefficients and 0.02 for the drag coefficients. The general C-shaped forms of the curves are similar, but at positive angles of attack, the coefficients, especially the lift coefficient, were lower at Rec=1150 than at the two higher speeds. The minimum drag coefficients at each speed, ranging from 0.05 at Rec=5560 to 0.07 at Rec=1150, were recorded at an angle of attack of 0 °. The zero-lift angle of attack was 0 ° at the lowest speed, but it was at a small negative angle (approximately −2 to −3 °) at the higher speeds. The lift coefficients increased sharply between 0 and 10 °, with only a slight increase in drag coefficient, before levelling off to a plateau from 20 to 40 ° of approximately 0.7 at Rec=5560, 0.65 at Rec=3300, but only 0.45 at Rec=1150. The maximum lift coefficients at these Reynolds numbers were 0.71 (at 20 °), 0.67 (at 30 °) and 0.46 (at 40 °), respectively. The lift coefficients decreased steadily at angles of attack greater than 40 °, falling to between 0.24 and 0.34 at 70 °. The drag coefficients increased almost linearly with angle of attack from 10 to 70 °. At angles of attack of 20 ° or more, the direction of the resultant of the lift and drag forces on the wing was always within 2 ° of the normal to the wing chord. The wing performance at negative angles of attack was similar at all three Reynolds numbers and approximated the reflection in the line CL=0 of the curve for positive angles at Rec=1150. The most negative value for lift coefficient was −0.43 at Rec=1150 and an angle of attack of −40 °. At the highest two Reynolds numbers, the lift and drag coefficients were smaller in magnitude than those at the corresponding positive angle of attack. 2729The aerodynamics of hawkmoth flight 0 0.2 0.4 0.6 0.8 1 0.6 0.4 0.2 0.2 0.4 0.6 0.8 0 C L C D Rec = 1150 Rec = 5560 Rec = 3300 50° 30° 10° 0° 10° 30° 50° 70° 70° 50° 30° 10° Fig. 2. Polar diagram for the aerodynamic characteristics of Manduca sexta wings. Each point is the mean from four wing couples. The different curves correspond to measurements at three different Reynolds numbers Rec, which are shown at the right of the figure. The angles of attack, in degrees, are indicated at intervals along the length of the curves. CL, lift coefficient; CD, drag coefficient.The wings were most aerodynamically efficient, indicated by the maximum lift:drag ratio, at 10 ° for all three airspeeds. The CL:CD ratios were 3.4 at Rec=1150, 4.2 at Rec=3300 and 4.0 at Rec=5560. Relative velocities and wingtip paths The forward and induced velocities are given in Table 1, along with the mean flapping velocity at the wingtip U – t (=2ΦnR), the reduced frequency parameters, and the Reynolds numbers for the body and wings. The induced velocity was largest, both in absolute terms and relative to the forward and flapping velocities, at the lower flight speeds. It exceeded the forward velocity at speeds below 2 m s−1, but then declined steadily in importance as flight speed increased. The mean wingtip flapping velocity was more constant because of the low variation in the wingbeat frequency and stroke amplitude. The high amplitude during hovering flight led to a flapping velocity over 5 m s−1, but at all faster speeds the flapping velocity fell within the range 4.2–5 m s−1. The comparative uniformity of the mean flapping velocity resulted in the advance ratio J (=V/U – t) rising linearly withforward velocity, from approximately 0.2 at 1 m s−1 to 1.1 at 5 m s−1. The reduced frequency parameter k (=ωc–/2V) correspondingly fell from 1.5 to 0.3 over the same speed range. The Reynolds numbers for the wings and the body (Rec and Reb, respectively) both tended to increase with flight speed as a result of increases in the appropriate relative velocity; the only exception was a decrease for the wing between 0 and 1 m s−1 as the lower flapping and induced velocities more than offset the increase in forward velocity. All of the wing Reynolds numbers Rec at speeds below 5 m s−1 fell between the values for the steady-state measurements at 3 and 5 m s−1. The wingtip paths relative to the surrounding air for moth F2 are shown in side view in Fig. 3A; the paths for the other moths were very similar. Each curve corresponds to the motion during three successive wingbeats at one of the six different flight speeds. The path starts at the lower left, with the wings at the supination position. The subsequent wing motion follows the path in the direction of the arrow; the relative velocity experienced by the wing is in the opposite direction. The axis of the wingtip path during hovering flight is vertical owing to the absence of any forward component in the relative velocity. The motion during the translatory periods of each halfstroke is inclined at only a small angle to the horizontal; the upstroke path doubles back over the preceding downstroke. As the forward speed increases, the overall axis of the motion becomes less inclined and there is increasing asymmetry between the two halfstrokes, both in terms of their length and their orientation. The direction of the downstroke motion is remarkably constant across the speed range, but its length increases steadily with forward speed. The upstroke, in contrast, varies in direction rather than length, with its inclination becoming less retrograde with increasing forward speed. When the latter reaches 5 m s−1, it is sufficiently large relative to the flapping velocity that the wingtip moves forward relative to the surrounding air during both halfstrokes. Angles of incidence of the wings The orientation of the wing chord relative to the oncoming air in the middle of each halfstroke is shown in Fig. 3B. The curves show the wing path at 0.5R for moth F2 at flight speeds of 0, 0.9, 2.9 and 5 m s−1. Wing chords, with the short perpendicular arm indicating the ventral side of the leading edge, have been drawn to scale at the middle of most halfstrokes; the arrowheads immediately upstream of the chords signify the direction of the relative velocity at these points. No data were available for the wing inclination during the upstroke at 2.9 m s−1. The angle of incidence for each chord is the average over both the outer and inner wing sections. Details of the individual angles of incidence are shown in Table 2, along with the original angles of rotation (see Willmott and Ellington, 1997) and the CD,pro estimates, which were derived from the mean angles of incidence for use in the aerodynamic force analysis. The angles of incidence in the mid-downstroke were all positive, indicating that the air was striking the ventral surface of the wing. At most speeds, air hits the dorsal surface during 2732 A. P. WILLMOTT AND C. P. ELLINGTON Table 2. The angles of rotation αsp and angles of incidence αr for the inner (I) and outer (O) wing sections at mid-downstroke and mid-upstroke Mid-downstroke Mid-upstroke V αsp (degrees) αr (degrees) αsp (degrees) αr (degrees) Moth (m s−1) I O I O Mean CD,pro I O I O Mean CD,pro M1 0 65.2 45.5 32.4 27.4 29.9 0.28 129.4 150.1 −28.4 −17.0 −22.7 0.14 1.0 72.5 54.2 34.0 28.4 31.2 0.30 137.7 154.0 −20.0 −17.8 −18.9 0.11 2.0 64.4 50.0 38.3 31.9 35.1 0.35 − 145.7 − −20.8 − − 2.9 62.1 53.1 30.2 30.8 30.5 0.29 120.1 145.2 −11.3 −15.6 −13.5 0.08 3.9 68.4 57.8 25.7 28.7 27.2 0.23 − − − − − − F1 0 73.7 49.2 49.9 36.1 43.0 0.50 139.8 159.3 −20.9 −15.8 −18.4 0.11 1.1 71.0 53.9 44.8 38.6 41.7 0.48 116.6 141.7 −33.2 −21.0 −27.1 0.20 3.0 64.5 54.2 36.1 35.0 35.6 0.36 − 132.8 − −17.4 − − 4.0 71.5 63.5 35.3 36.3 35.8 0.36 − − − − − − 5.0 73.2 65.2 36.5 37.0 36.8 0.37 − − − − − − F2 0 69.5 52.0 39.2 35.5 37.4 0.39 129.0 148.7 −38.6 −27.2 −32.9 0.22 0.9 61.8 44.7 30.6 25.3 27.9 0.25 126.4 150.1 −27.9 −21.1 −24.5 0.15 2.1 71.2 57.8 37.9 35.3 36.6 0.38 122.0 142.4 −21.8 −28.0 −24.9 0.15 2.9 59.6 48.1 25.9 22.6 24.3 0.20 − − − − − − 3.8 67.4 57.3 28.5 27.9 28.2 0.25 − − − − − − 5.0 67.8 59.0 23.1 23.0 23.1 0.17 118.4 118.9 26.1 3.0 14.6 0.10 The missing mid-upstroke values indicate cases in which the rotation angles could not reliably be determined from the digitized frames (see Willmott and Ellington, 1997). The mean angles of incidence were used to estimate the profile drag coefficient CD,pro. See text for full details. V, flight velocity. Table 3. The relative velocity Ur,b and angle of incidence χr,b of the body for each flight sequence, along with the estimated body force coefficients and the parasite lift L̂par and drag D̂par forces (as percentages of body weight) V Ur,b χ χr,b L̂par D̂par Moth (m s−1) (m s−1) (degrees) (degrees) CL CD (%) (%) M1 0 1.4 40.0 −50.0 −0.48 0.61 −1.90 −1.51 1.0 1.7 29.4 −26.0 −0.36 0.30 −1.97 −0.56 2.0 2.2 21.0 −5.2 −0.04 0.12 −0.65 0.70 2.9 3.0 17.4 3.1 0.05 0.11 0.23 1.70 3.9 3.9 14.4 5.7 0.09 0.13 1.62 3.28 F1 0 1.3 35.1 −54.9 −0.43 0.64 −1.61 −1.08 1.1 1.6 26.5 −21.0 −0.28 0.22 −1.40 −0.21 3.0 3.1 25.8 13.6 0.18 0.17 1.92 2.84 4.0 4.0 15.7 8.6 0.12 0.14 2.45 3.77 5.0 5.0 28.7 24.0 0.37 0.30 12.93 12.28 F2 0 1.5 34.6 −55.4 −0.44 0.65 −1.79 −1.20 0.9 1.7 28.4 −28.9 −0.40 0.34 −1.78 −0.54 2.1 2.4 26.7 −0.1 −0.00 0.11 −0.35 0.68 2.9 3.0 21.1 4.6 0.06 0.12 0.30 1.56 3.8 3.9 19.5 10.0 0.14 0.15 2.07 3.20 5.0 5.0 16.3 10.8 0.16 0.16 4.71 5.65 V, flight velocity; χ, geometric body angle; CL, lift coefficient; CD, drag coefficient. See text for full details. 2733The aerodynamics of hawkmoth flightdrag coefficients C–D,pro were calculated from the Reynolds number for the wings, using equation 11. Replacing the mean profile drag coefficients by the values based on the observed angles of incidence (which were up to five times larger) had little impact on the mean lift coefficient; the difference between the two estimates was never greater than 0.08. The changes with forward speed in the lift coefficient requirements, and the significance of the wingbeat shape and upstroke mode can be seen in Fig. 5. For each of the moths, the variation in mean lift coefficient is shown for the asymmetric wingbeat, for the sinusoidal approximation, and for the asymmetric wingbeat under the assumption of an inactive upstroke. For the former two cases, the required lift coefficient was calculated under both the assumption of a negative and a positive upstroke lift coefficient, but only the lower of the two values is plotted. Open symbols indicate that the lower value resulted from the assumption of a negative upstroke lift coefficient (i.e. a reversed upstroke circulation), and filled symbols that the assumed upstroke coefficient was positive (i.e. for air striking the ventral wing surface during the upstroke as well as the downstroke). A reversed lift coefficient,0 1 2 3 4 5 0 1 2 3 4 5 Sinusoidal Asymmetric, inactive upstroke Asymmetric M1 F1 F2 Flight speed (m s 1) M ea n lif t c oe ff ic ie nt , C L Fig. 5. The mean lift coefficients C–L calculated for three different combinations of assumed upstroke function and wingbeat shape. Data are given for the true, asymmetric wingbeat (squares), for the sinusoidal approximation to the wingbeat (triangles) and for the asymmetric wingbeat under the assumption of an aerodynamically inactive upstroke (no symbols). The first two cases include both a reversed and a non-reversed upstroke circulation; only the lower of the two calculated coefficients is shown. Open symbols indicate that the reversed upstroke circulation resulted in the lower lift requirement, filled symbols that the non-reversed circulation was preferable. The data for the individual moths are distinguished by colour. The horizontal line at 0.71 indicates the maximum lift coefficient obtained from a Manduca sexta wing couple under steady-state conditions.indicating a negative upstroke angle of incidence, was advantageous at low speeds, but positive angles of incidence throughout the beat resulted in lower required lift coefficients at the highest speeds. The transition for each moth occurred at the speed where Az,u changed from negative to positive: for the asymmetric wingbeat this was at 4 m s−1, for the sinusoidal motion it was at 3 m s−1. An initial increase in mean lift coefficient from hovering to a maximum at 1–2 m s−1 was characteristic of all the curves assuming an active upstroke. The maximum values for the true, asymmetric wing motion ranged from 1.98 for F1 to 2.38 for M1. Lift coefficient declined steadily with further increases in flight speed, falling to minimum values at 5 m s−1 of 0.70 and 0.80 for F1 and F2, respectively. These compare with the maximum steady-state value of 0.71 which is indicated by the horizontal line in Fig. 5. The calculated mean lift coefficients for the sinusoidal motion were comparable to, but usually slightly lower than, those for the asymmetric wingbeat; the minimum values for moths F1 and F2 were 0.59 and 0.75 at 5 m s−1. The relative magnitudes of the coefficients for the asymmetric and sinusoidal wingbeats depended upon which upstroke angle of incidence option was assumed. With negative angles, the lift coefficients still decreased with forward speed, but more slowly, and the sinusoidal estimates were higher than the asymmetric wingbeat values. Aerodynamic contributions of the two halfstrokes Qualitative changes in the relative contributions of the two halfstrokes to thrust and weight support were shown in Fig. 3B. The arrows originating at the aerodynamic centre of each wing chord represent the direction and magnitude of the resultant force vector from each halfstroke, as estimated from the length and direction of the halfstroke paths and the calculated mean force coefficients. Both halfstrokes provided weight support during hovering, but the upstroke was more important. By 2.9 m s−1, however, weight support was largely restricted to the downstroke, but with the upstroke still capable of providing significant thrust if the angle of incidence was negative. At 5 m s−1, the downstroke generated substantial weight support and some thrust under both upstroke assumptions, but the magnitude of the forces depended greatly upon the sign of the upstroke lift coefficient. If this was negative, then the upstroke produced thrust but negative weight support, which had to be offset by larger vertical forces during the subsequent downstroke. The mean lift coefficient, and thus the vertical force required from the downstroke, was lower if the upstroke coefficient was positive (as shown in Fig. 3B), but there was little if any net thrust under these conditions. Mechanical power requirements The body-mass-specific aerodynamic and inertial power components for each flight sequence are summarized in Table 4. Aerodynamic power was calculated for three of the cases considered in the mean lift coefficient analysis: for the real, asymmetric wingbeats with mean profile drag coefficients estimated from equation 11, for sinusoidal motion with the 2734 A. P. WILLMOTT AND C. P. ELLINGTON Table 4. Body-mass-specific power components for each flight sequence V P*par P*pro P*pro,s P*pro,ai P*ind P*ind,s P*ind,ai P*aero P*aero,s P*aero,ai P*acc P*acc,s Moth (m s−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) (W kg−1) M1 0 0.00 3.24 3.32 5.47 14.48 14.46 14.82 17.72 17.78 20.29 27.41 27.53 1.0 0.00 2.67 2.49 4.54 14.21 14.13 14.40 16.87 16.62 18.94 20.29 18.97 2.0 0.13 4.34 3.64 − 9.82 9.76 − 14.29 13.54 − 43.13 23.86 2.9 0.48 4.00 4.72 8.91 7.53 7.54 7.33 12.02 12.74 16.72 20.30 23.47 3.9 1.24 5.78 6.20 − 6.05 6.02 − 13.07 13.47 − 24.34 18.25 F1 0 0.00 4.52 4.68 9.39 13.36 13.27 13.67 17.88 17.96 23.06 38.47 37.38 1.1 0.00 3.16 3.47 9.43 12.36 12.29 12.88 15.53 15.76 22.31 21.55 26.25 3.0 0.84 4.38 5.13 − 6.52 6.55 − 11.74 12.52 − 20.28 23.16 4.0 1.48 5.97 6.72 − 5.08 5.10 − 12.54 13.30 − 26.19 17.27 5.0 6.02 7.74 10.26 − 4.14 4.24 − 17.91 20.53 − 17.94 20.80 F2 0 0.00 3.82 3.29 9.98 15.07 15.01 16.00 18.89 18.30 25.98 29.57 23.04 0.9 0.00 2.86 3.00 5.29 14.37 14.30 14.74 17.23 17.31 20.04 18.52 20.34 2.1 0.14 2.85 3.09 7.96 10.87 10.80 11.21 13.86 14.03 19.30 14.60 16.05 2.9 0.44 4.16 4.10 − 8.70 8.70 − 13.30 13.24 − 20.70 17.44 3.8 1.19 4.92 5.69 − 6.50 6.47 − 12.61 13.36 − 16.50 16.69 5.0 2.77 7.91 9.05 13.96 4.92 4.93 4.88 15.60 16.74 21.60 15.87 17.62 Body-mass-specific profile P*pro, induced P*ind and total aerodynamic P*aero power are given for three cases: for the asymmetric wingbeat (no subscript), for the sinusoidal wing motion (subscript ‘s’) and for the asymmetric wingbeat with profile drag estimated from the observed angles of incidence (subscript ‘ai’). The body-mass-specific parasite power P*par is common to all three cases. P*acc, body-mass-specific inertial power. See the text for full details. V, flight velocity.same drag coefficients, and for the asymmetrical kinematics but using the individual upstroke and downstroke profile drag coefficients estimated from the angles of incidence. The subscripts ‘s’ and ‘ai’ denote the second and third cases, respectively. The power requirements are independent of the sign of the upstroke lift coefficient. The values for parasite power were common to all three cases, but there were differences in the induced and profile powers and, therefore, in the total aerodynamic power. The inertial power required during wing acceleration is given for both the asymmetric and sinusoidal kinematics. The trends in power with increasing forward speed were qualitatively similar for all three cases, and are illustrated in Fig. 6 for moth F2 and the first case. The solid lines describe the trends in aerodynamic power and its three components for the asymmetric wingbeat using the lower estimates of mean profile drag. The dark blue crosses show the markedly increased profile power when the higher profile drag coefficients were used in the model. Estimates for aerodynamic power and its components at 6 m s−1 are also presented; they will be used below as evidence for possible power constraints on maximum flight speed. The estimates were calculated by extrapolation from the kinematic parameters at speeds up to 5 m s−1. The wing motion within the stroke plane was assumed to be the same as at 5 m s−1, but modified values of 63.5 ° and 13.6 ° were assigned to the stroke plane and body angles, respectively. Wingbeat frequency was taken to be 25 Hz.For moths M1 and F2, the aerodynamic power was highest during hovering, with the induced power contributing substantially more than half of the total power requirement (Table 4). Induced power decreased slightly between hovering and 1 m s−1, and thereafter more rapidly with increasing flight speed. Above 2 m s−1, this decrease was partially offset by the increases in parasite and profile power, but the total requirement did not reach its minimum until 3 or 4 m s−1. At higher speeds, the profile and parasite powers increased more steeply, and the total aerodynamic power rose again. The magnitude of the aerodynamic power requirement was dependent upon the value assigned to the profile drag coefficient. For the profile drag coefficients of approximately 0.1, derived from equation 11, the aerodynamic power decreased from 17–19 W kg−1 at hovering to a minimum of 12–13 W kg−1, before increasing to over 15 W kg−1 at 5 m s−1. The markedly higher profile drag coefficients used in the third case resulted in correspondingly higher profile power requirements, raising the aerodynamic power to values between 17 and 26 W kg−1. The net vertical contribution from profile drag was always in the opposite direction to weight support, resulting in induced power estimates at all flight speeds which were greater than the value of w0mg which would have been obtained using the simpler model from previous studies. The increase in induced power ranged from 3 to 5 % for the lower profile drag coefficients, and up to 11 % for the higher coefficients. For the 2737The aerodynamics of hawkmoth flight Table 6. Body and wing morphology for the moths in the kinematic study (Willmott and Ellington, 1997), and means for larger samples of male and female Manduca sexta M1 F1 F2 Male mean ± S.E.M. Female mean ± S.E.M. t-test significance Body parameters (N=6) (N=7) m (mg) 1579 1648 1995 1199±110 1833±112 ** L̂ 0.876 0.811 0.817 0.965±0.025 0.852±0.012 ** l̂ 0.489 0.530 0.540 0.467±0.015 0.529±0.007 ** l̂ 1 0.256 0.311 0.287 0.221±0.010 0.291±0.004 *** l̂ 2 0.344 0.401 0.358 0.337±0.008 0.369±0.006 ** χo 82.9 82.8 76.3 79.6±1.0 79.2±1.1 m̂t 28.4 25.0 17.8 33.0±2.3 21.1±0.9 *** m̂tm 22.8 19.7 16.2 27.0±1.6 17.7±0.5 *** m̂a 54.1 57.8 61.6 42.4±3.4 61.7±1.1 *** Wing parameters (N=4) (N=7) R (mm) 48.5 51.9 52.1 47.3±0.6 51.0±0.9 * pw (N m−2) 8.93 8.53 10.33 7.75±0.78 9.73±0.35 * AR 5.28 5.65 5.52 5.20±0.13 5.53±0.04 * r̂1 (S) 0.445 0.447 0.450 0.449±0.001 0.451±0.001 r̂2 (S) 0.514 0.515 0.518 0.517±0.001 0.518±0.001 r̂3 (S) 0.564 0.565 0.567 0.566±0.002 0.567±0.001 v̂ 1.07 1.08 1.08 1.08±0.00 1.08±0.00 r̂1 (v) 0.413 0.418 0.425 0.419±0.002 0.424±0.002 * r̂2 (v) 0.475 0.479 0.483 0.478±0.001 0.482±0.001 ** eĉ3r̂dr̂ 0.469 0.480 0.492 0.479±0.006 0.495±0.003 * eĉ E∑r̂ T∑dr̂ 0.199 0.201 0.204 0.202±0.001 0.204±0.001 m̂w (%) 5.93 5.79 4.49 6.52±0.59 5.02±0.18 * ĥ (%) 0.081 0.072 0.067 0.080±0.002 0.072±0.001 ** r̂1 (m) 0.293 0.297 0.292 0.310±0.007 0.298±0.003 r̂2 (m) 0.376 0.378 0.379 0.392±0.007 0.383±0.003 See the Symbols list for definition of the parameters. The right-hand column shows the results of t-tests of the difference between the means for male and female moths. Key to significance levels: *P<0.05; **P<0.01; ***P<0.001.also been reported for bumblebees (Dudley and Ellington, 1990b; Cooper, 1993) and hoverflies (Dudley, 1987), the only other studies of individual insects over a range of speeds. The decrease in mean lift coefficient at high speeds was more marked for Manduca sexta than for bumblebees (Dudley and Ellington, 1990b; Cooper, 1993) or hoverflies (Dudley, 1987), even with a negative upstroke lift coefficient as in the earlier studies. The reduction was more dramatic still for Manduca sexta if the wing was assumed to be operating at positive angles of incidence during the upstroke. This change in wing orientation is beneficial only where the upstroke path, relative to the surrounding air, is directed forwards. For the most distal regions of the wing this criterion is not met until 5 m s−1, but for the middle of the wing it is reached before 3 m s−1. For asynchronous species, the large flapping velocities associated with their high wingbeat frequencies mean that the upstroke path is directed backwards at all speeds (Nachtigall, 1966; Dudley and Ellington, 1990b). There are changes, however, in the aerodynamic functions of the two halfstrokes with increasing forward speed; as demonstrated by Ellington’s (1995) analysis of the bumblebee data from Dudley andEllington (1990a). As advance ratios increase, the downstroke will increasingly dominate weight support as the associated wing path becomes longer. The upstroke contributes thrust at all speeds but its contribution decreases with increasing forward speed, a trend that is offset by an increasing thrust component from the downstroke force. The trends in the resultant forces predicted for Manduca sexta flight are similar, at least up to forward speeds of 3–4 m s−1 (advance ratios of 0.8–0.9). The transition to positive upstroke angles of incidence at higher speeds, which is suggested by the data for moth F2 at 5 m s−1, is unexpected. This switch can be viewed, using the data for mean lift coefficient, as a means of reducing the lifting performance required from the wings. Such a change should be regarded as speculative at present, however, because of the paucity of data for angle of incidence, and the fact that the data are subject to errors both in the angle of rotation and in the direction of the relative velocity. If the transition in angle of incidence is genuine then, to the best of the authors’ knowledge, Manduca sexta is the first insect species for which a non-reversed upstroke has been proposed. In other species (including Urania 2738 A. P. WILLMOTT AND C. P. ELLINGTONfulgens at all speeds), the aerodynamic underside of the wing, and thus the sense of circulation, is thought to be reversed between successive halfstrokes (Brodsky and Ivanov, 1984; Dudley and DeVries, 1990; Dudley and Ellington, 1990b). The possible aerodynamic functions of the upstroke have been considered more thoroughly for flying vertebrates (e.g. Rayner, 1979, 1986; Norberg, 1985, 1990; Spedding, 1992). The upstroke is aerodynamically inactive in general, as a result of wing folding, during slow forward flight in birds and bats (Spedding, 1992), although it may generate small amounts of thrust (Norberg, 1985). At intermediate speeds, however, the upstroke becomes active as the wing path changes from a backwards to a forwards direction (e.g. Norberg, 1976; Spedding, 1987). Aldridge (1986) found a gradual change in the kinematics and aerodynamics of forward flight in the greater horseshoe bat Rhinolophus ferrumequinum, with the upstroke changing from thrust generation at low speeds to weight support at intermediate speeds. At the highest speed, 4.8 m s−1 (J=0.93), the upstroke angles of incidence were large and positive, contributing weight support but also negative thrust to the force balance. The morphological limitations on wing span and area reduction placed on insects and bats mean that Manduca sexta is faced with the same aerodynamic options for the upstroke as Rhinolophus ferrumequinum: the wing may be operated at very low angles of incidence in order to minimize the aerodynamic forces, at positive angles of incidence for weight support or at negative angles for thrust production. The final option is the usual case for insects, but in Manduca sexta it would generate thrust at the expense of considerable negative weight support. The second case provides additional weight support, thus decreasing the required mean lift coefficient, but it also generates negative thrust. If this is the case for Manduca sexta at 5 m s−1 then, as can be seen in Fig. 3B, the insect is struggling to produce a net positive thrust. A final observation concerning quasi-steady aerodynamic force production is that the asymmetric wingbeat does not appear to provide an aerodynamic advantage over sinusoidal motion if positive upstroke angles of incidence are an option. However, if the upstroke angles of incidence remain negative, then the mean lift coefficients for the sinusoidal wingbeat are consistently higher than those for the asymmetric wingbeat at high flight speeds. It must be noted that any benefits of the asymmetric wingbeat may not be revealed by a quasi-steady analysis, which assumes that force production is proportional to the square of relative velocity. Lift from unsteady mechanisms, such as a leading-edge vortex, may not scale with relative velocity in the same way. Without further information, no conclusions can be drawn on the aerodynamic significance of an asymmetric wingbeat. The profile power estimates (Table 4) provide some evidence that there may be a slight energetic advantage to the asymmetric wingbeat at fast forward speeds: at speeds of 4 m s−1 and above, the profile power for the asymmetric wingbeat was lower, but typically by less than 15 %, than the value for the corresponding sinusoidal motion.Aerodynamic power for flight The U-shaped relationship between forward speed and specific aerodynamic power for Manduca sexta is consistent with power curves drawn for a wide range of insect (reviewed in Cooper, 1993) and vertebrate (reviewed in Norberg, 1990) fliers. The precise shape does vary, from the J-shape found by Cooper (1993) for bumblebees to the deep U-shape obtained by Rayner (1979) for a range of birds, but these represent two extreme points on a continuum of shapes derived from the same basic trends in the aerodynamic power components. The increases in aerodynamic power at the two ends of the speed range result from different components (Pennycuick, 1968), and much of the variation between species is explicable and to be expected from such diverse morphologies and kinematics. The initial decrease in flight power from hovering to the minimum power speed results from a decrease in induced power. Parasite power is negligible at these speeds, and the profile power, whilst not negligible, is fairly constant. The location of the minimum power speed depends upon how rapidly the sum of profile and parasite power increases relative to the decrease in induced power. Profile power increases as the cube of the relative velocity (equation 13), not the forward velocity as stated by Spedding (1992). This difference is considerable for animals such as bumblebees, whose flight is characterized by low advance ratios. The flapping velocity dominates the relative velocity and, where frequency and amplitude vary little with forward speed, the profile power is approximately constant over almost all the speed range, as clearly seen with bumblebee flight (Dudley and Ellington, 1990b; Cooper, 1993). The advance ratios for Manduca sexta flight increase more rapidly with forward speed, and this is reflected in the steeper curve for profile power. In contrast, the parasite power remains small and accounts for less than half of the increase in the sum of profile and parasite power. The bumblebee body is less streamlined, however, and the parasite power contributes in large part to the increase in aerodynamic power at high forward speeds (Cooper, 1993). The relative importance of parasite power is also very sensitive to the chosen body angle, as can be seen in the discrepancy between the results at high flight speeds for moth F1 and moths M1 and F2 (Table 4), and between the different curves for pigeons derived by Pennycuick (1968) and Rayner (1979), for which different body angles were assumed (Norberg, 1990). For any species, the curve for total aerodynamic power becomes increasingly steep at high speeds. Induced power is still decreasing, but this is offset by the ever more rapid increase in the profile and parasite components. The correction to the traditional expression for induced power (equation 12), which results from consideration of the vertical contribution from profile drag, ranged from 3 to 11 % for Manduca sexta, depending upon the selected profile drag coefficient. The same increase for dragonfly and damselfly flight averaged 12 % (Wakeling and Ellington, 1997b). The magnitude of this correction is therefore comparable to those from previous modifications to the simple momentum jet 2739The aerodynamics of hawkmoth flightmodel to incorporate temporal and spatial effects (Ellington, 1984d), which have been adopted as standard in subsequent studies. A net downwards force from profile drag is probably typical of flight in most insects. If the profile drag coefficients are similar during the two halfstrokes, and the vertical components of the flapping velocities are approximately equal and opposite, then this net downwards force is a consequence of the induced velocity. Flight in which there is a net upwards force from profile drag, whilst less common, is not unknown. The generation of large vertical forces by pressure drag during the downstroke has been identified by Ellington (1980) and Sunada et al. (1993) for take-off flight in butterflies from the genus Pieris. Weight support of this kind requires a near- vertical stroke plane with high angles of incidence during the downstroke, and low angles during a ‘feathered’ upstroke. Similar kinematics have been reported by Betts and Wootton (1988) as one of the flight modes used during both hovering and fast forward flight in butterflies. Profile drag will have a substantial vertical component under these conditions, and second-order corrections to Pind should be incorporated. Total mechanical power requirements Estimates of the total mechanical power required for flight can be determined by two independent methods: from data for aerodynamic and inertial power, and from the product of muscle efficiency and the metabolic cost of flight (Casey, 1981). Comparing the two estimates can shed light on the extent of elastic storage in the thorax and/or wings. Casey (1981), for example, found good agreement between a metabolic estimate for the total body-mass-specific mechanical power for Manduca sexta hovering (47.4 W kg−1, assuming 20 % muscle efficiency) and the estimated sum of aerodynamic power (14.6 W kg−1, modified from Weis-Fogh, 1973) and inertial power (36.4 W kg−1). He concluded that the inertial costs of hovering flight were not significantly reduced by an elastic system in the thorax, in contrast to Weis-Fogh’s (1972) finding for the closely related privet hawkmoth Sphinx ligustri. Recalculating the metabolic estimate of total mechanical power from Casey’s (1976) measured body-mass-specific metabolic cost of flight (237 W kg−1) and the muscle efficiency of 10 % determined experimentally by Stevenson and Josephson (1990) for Manduca sexta dorsoventral muscle, the available body-mass-specific power is only 23.7 W kg−1. The inertial power component has been greatly reduced, but not completely removed. Similarly, Wilkin and Williams (1993) concluded that elastic storage was effective at a flight speed of 3.36 m s−1, but that some power was still available, if needed, to damp the wing oscillations. The specific mechanical power requirements obtained in the present study do not provide an unequivocal conclusion about the extent of elastic storage. The highest performance during flight was most likely to have been reached in moth F2, which had both the highest wing loading and the lowest percentage of flight muscle. The total muscle-mass-specific mechanical power for this moth, assuming perfect elastic storage, reached117 W kg−1 during hovering, and over 160 W kg−1 if the higher estimates of profile power were used. These compare with a mean specific power output of 90 W kg−1 of flight muscle and a maximum of 130 W kg−1 at muscle temperatures and cycle frequencies which corresponded well with free-flight values (Stevenson and Josephson, 1990). The inertial power estimates for Manduca sexta were not much higher than the aerodynamic power, however, and most of the estimates for total mechanical power with no elastic storage were also within the range observed in the earlier study. The potential for considerable elastic storage has been demonstrated indirectly by the observation that, in a Manduca sexta with the thoracic nervous system removed, a downstroke generated by electrical stimulation of the depressor muscle can be followed by a nearly normal, passive upstroke (S. Dierkes, personal communication). The possibility remains that the storage is far from perfect, but the percentage of elastic storage cannot be quantified accurately at present. The capacity for elastic storage does not need to be considerable. Dickinson and Lighton (1995) found that in Drosophila hydei, another species in which aerodynamic power and inertial power are very similar in magnitude, the cost of flight can be minimized with an elastic storage capacity as low as 10 % of the inertial power. The remaining 90 % of the inertial energy can be usefully converted to aerodynamic work during wing deceleration. The equivalent storage capacity to minimize costs for Manduca sexta is approximately 30 % of the inertial power. Limits to maximum flight speed The discussion of the kinematics and aerodynamics of fast forward flight can be extended to consider the question of what limits the maximum speed of Manduca sexta. For such a large insect, the maximum observed speed is surprisingly low. The highest value recorded by Stevenson et al. (1995) was 5.3 m s−1 while their calculations, based on body size, suggested that Manduca sexta should be capable of speeds of at least 7–10 m s−1. One potential constraint in their study and in the current one was the use of flight chambers. This will undoubtedly affect flight behaviour, as it did for a range of butterflies whose typical flight speeds during insectary studies were significantly lower than those observed in the wild (Srygley and Dudley, 1993). Other observations, however, suggest that flight chambers and their size may not be so important in determining maximum speed. Cooper (1993) obtained significantly higher flight speeds from bumblebees in a wind tunnel whose closed working area was much smaller than the largest arena (an ice rink) used by Stevenson et al. (1995). Given sufficient motivation for flight, chamber size may not limit flight speed. There was clearly no shortage of motivation in the moths filmed in the present study: several moths tried repeatedly to reach the feeder at 5 m s−1, but they appeared to have great difficulty in flying at these very fast speeds. The available mechanical power is often considered to be the major constraint on maximum forward speed (e.g. 2742 A. P. WILLMOTT AND C. P. ELLINGTONand the accuracy of the latter has been queried because the measured minimum drag coefficient is lower than theoretically possible (Ellington, 1984d). The results for Manduca sexta, the odonatans and the locust match the conclusions of Spedding (1992) and Brodsky (1994) that the highest lift:drag ratios are found at the high Reynolds numbers characteristic of large, fast-flying insects whose wings operate under flow conditions where inertial forces dominate. Maximum lift coefficient is relatively independent of Reynolds number, lying in the range from approximately 0.7 to 0.9 for most species studied, and the variation in the lift:drag ratio results from the friction drag component of profile drag scaling as Re−0.5 (Blasius, 1908). The relative contribution of friction drag is highest at the low angles of attack where the maximum lift:drag ratios occur. The minimum value of the drag coefficient for Manduca sexta wings was 0.05, for example, compared with 0.25 and 0.13 for a bumblebee worker at Reynolds numbers of 460 and 1240, respectively (Dudley and Ellington, 1990b), and 0.54 and 0.29 for cambered Drosophila virilis wings at Reynolds numbers of 65 and 260, respectively (Vogel, 1967). As the Reynolds number decreases, the maximum CL:CD ratio also decreases, but the wings display a very gentle stall. These characteristics can be seen in the polar diagrams for the mayfly Ephemera vulgata (CL:CD=3.3; Brodsky, 1970), the cranefly Tipula oleracea (CL:CD=2.4; Nachtigall, 1977) and the bumblebee Bombus terrestris (Dudley and Ellington, 1990b) whose CL:CD ratio was approximately 2.5 at Reynolds numbers from 460 to 1520. At even lower Reynolds numbers, the increasing importance of viscous effects reduces the lift:drag ratio further but leads to the virtual absence of stall, which was seen with the wings of Drosophila virilis at Reynolds numbers between 65 and 260 (Vogel, 1967). The details of the airflow around insect wings under steady- state conditions are not well understood, but a number of structural features that influence aerodynamic performance have been identified, and some of these can be seen in the hawkmoth wing. The most widespread feature is the beneficial effect of wing camber in increasing lift coefficients at positive angles of attack. Vogel (1967) found that the polar diagram for flat Drosophila virilis wings was nearly symmetrical about the 0 ° angle-of-attack position, but that the addition of camber resulted in a positive shift in lift coefficients which was most pronounced for positive angles of attack. Similarly, the convex profile seen in locust forewings during late downstroke increased the lift:drag ratio of the wing (Jensen, 1956). The small, but consistent, positive camber inherent in the Manduca sexta wing structure probably contributed to the better performance of the wing at positive than at negative angles of attack, and it is likely to enhance lift production during the long downstroke of flapping flight when the wing profile is similar to that of the isolated wing couples in this study. The same wing shape is, however, unlikely to be such an accurate model of the wings during the upstroke, and this may, in part, explain the asymmetry between the lift coefficients at positive and negative angles of attack. The wing twist and camber during the upstroke have not been as well quantifiedas those of the downstroke (Willmott and Ellington, 1997) but, qualitatively, the wing shape appears to be both more complicated and more variable during the former. In addition, the direction of camber (at least in the regions posterior to the rigid leading-edge spar) may be the reverse of that used here. The lift and drag coefficients measured at negative angles of attack in this study may, therefore, underestimate the true steady-state values for the hawkmoth wing couple configuration appropriate to upstroke periods when the wing is acting at negative angles of incidence. In terms of the aerodynamic analysis undertaken here, however, any discrepancies in these force coefficients would only have had a noticeable impact on the values for profile drag and profile power estimated from the observed angles of incidence. An important role for leading-edge separation bubbles in shaping wing performance through the avoidance of abrupt stall and the enhancement of lift production has been proposed (Ellington, 1984c; Spedding, 1992). Leading-edge vorticity had a strong influence on the flow around the wings during Manduca sexta flapping flight at all speeds (Ellington et al. 1996; Willmott et al. 1997), and it may also explain the gradual stall seen in Fig. 2 and the marked difference in performance of the Manduca sexta wing between a Reynolds number of 1150 and the two higher values. This discrepancy has not been seen before for insect wings, but it may represent the increase in lift production which accompanies the formation of a strong leading-edge separation bubble at Rec=3300 and Rec=5560. The wing performance at negative angles of attack is similar for all three Reynolds numbers, and it mirrors the curve for positive angles at the lowest Reynolds number, suggesting that any separation bubbles are comparable under these flow conditions. Visualization of the flow around the wings is needed in order to confirm these hypotheses. The possibility cannot be excluded from these data that the slight delay between the force measurements at Rec=1150 and at Rec=3300 and 5560 might have contributed to the differences in the force coefficients. A number of observations suggest, however, that the differences probably represent a genuine aerodynamic phenomenon. First, the Manduca sexta wing is a rigid structure, even when fresh, and there is little qualitative change in this over 24 h, or over much longer periods. Second, the discrepancy was only in the coefficients at positive angles of attack. At negative angles, the performance of the wings was similar at all speeds. Finally, the unusually low coefficients were obtained on the first day, with the later results being closer to the coefficients recorded for other species. The maximum lift coefficients at Rec=3300 and Rec=5560 were, for example, very similar to those obtained from the wing couple of a noctuid moth at 5 m s−1 (Lendle, 1981, cited in Nachtigall, 1989). Directions for future studies This study has identified the manner in which the mean lift coefficient varies with changes in wing kinematics. The requirement for unsteady aerodynamic mechanisms is clear, but their precise nature and contribution cannot be ascertained 2743The aerodynamics of hawkmoth flightfrom the mean coefficients model. Future models must incorporate new techniques which can accurately simulate both the flow around the wings and the instantaneous forces. Such models will allow full investigation of the impact of structures such as leading-edge vortices: the benefits of unsteady mechanisms in terms of lift production have often been cited, but there is also likely to be an associated cost through, for example, increased profile drag. Symbols AR Aspect ratio Ax, Az Summations used in calculating the horizontal and vertical components of lift, respectively. Bx, Bz Summations used in calculating the horizontal and vertical components of profile drag, respectively. c Wing chord ĉ Wing chord/mean chord c– Mean wing chord CD Drag coefficient CD,pro Profile drag coefficient C – D,pro Mean profile drag coefficient CL Lift coefficient C – L Mean lift coefficient D Drag Dpar Parasite drag D̂par Parasite drag/body weight Dpro Profile drag Fx Horizontal force required from the wings to overcome parasite drag Fz Vertical force required from the wings to complete the force balance g Gravitational acceleration ĥ Mean thickness of the wing/wing length i Wing strip number I Moment of inertia J Advance ratio (V/U – t) k Reduced frequency parameter or correction factor for spatial and temporal variation in induced velocity l Characteristic length for calculating wing or body Reynolds numbers l̂ Distance from anterior tip of body to centre of mass/body length l̂ 1 Distance from forewing base axis to centre of mass/body length l̂ 2 Radius of gyration for the body/body length L Lift or body length L̂ Body length/wing length Lpar Parasite lift: vertical force acting on the body L̂par Parasite lift/body weight m Body mass m̂a Mass of the abdomen/body mass m̂t Mass of the thorax/body mass m̂tm Mass of the thoracic muscle/body mass m̂w Wing mass/body massn Wingbeat frequency pw Wing loading P* Body-mass- or muscle-mass-specific values of the power components Pacc Inertial power Paero Total aerodynamic power Pind Induced power Ppar Parasite power Ppos Mean total positive mechanical power Ppro Profile power r Radial position along the wing r̂ Radial position along the wing/wing length r̂k(m) Radius of the kth moment of wing mass/wing length r̂k(S) Radius of the kth moment of wing area/wing length r̂k(v) Radius of the kth moment of wing virtual mass/wing length R Wing length Re Reynolds number Reb Reynolds number for the body, based on body length Rec Reynolds number for the wings, based on mean chord S Wing area t Time U Flapping velocity Un Normal to the relative velocity Ur Relative velocity Ur,b Relative velocity of the body Ut Flapping velocity at the wingtip U – t Mean flapping velocity at the wingtip v̂ Non-dimensional virtual mass of a wing pair (see Ellington, 1984b) V Flight velocity w0 Induced velocity at the actuator disc w0,RF Rankine–Froude estimate of induced velocity at the actuator disc α Angle of attack αr Angle of incidence αr′ Effective angle of incidence αsp Angle of rotation relative to the stroke plane θ Angle of elevation of the wing with respect to the stroke plane υ Kinematic viscosity of air ρ Mass density of air ρw Mass density of the wings φ Sweep angle of the wing in the stroke plane Φ Stroke amplitude χ Body angle χr,b Angle of incidence of the body χ0 Free body angle ω Radian frequency Subscripts: ai Using profile drag coefficients estimated from the angles of incidence 2744 A. P. WILLMOTT AND C. P. ELLINGTONd For the downstroke s For a sinusoidal approximation to the wingbeat u For the upstroke x, z Component in the x (horizontal) or z (vertical) direction. We are very grateful to Dr J. M. Wakeling, Dr S. Sunada and Dr R. D. Stevenson for their help in many ways during the course of this study, and to Mr S. Ellis for manufacturing parts for the force transducer. This work was supported by grants from the BBSRC (A.P.W.), the SERC and the Hasselblad Foundation (C.P.E.). References ALDRIDGE, H. D. J. N. (1986). Kinematics and aerodynamics of the greater horseshoe bat, Rhinolophus ferrumequinum, in horizontal flight at various flight speeds. J. exp. Biol. 126, 479–497. BETTS, C. R. AND WOOTTON, R. J. (1988). Wing shape and flight behaviour in butterflies (Lepidoptera: Papilionoidea and Hesperioidea): a preliminary analysis. J. exp. Biol. 138, 271–288. BLASIUS, H. (1908). Grenzschichten in Flüssigkeiten mit kleiner Reibung. Z. Math. Phys. 56, 1–37. BRODSKY, A. K. (1970). On the role of wing pleating in insects. Zh. evol. biochem. physiol. 6, 470–471. BRODSKY, A. K. (1994). The Evolution of Insect Flight. Oxford: Oxford University Press. BRODSKY, A. K. AND IVANOV, V. D. (1974). Aerodynamic peculiarities of the flight of insects. II. Smoke spectra. Vestnik Leningr. Univ. (Biology) 3, 16–21. BRODSKY, A. K. AND IVANOV, V. D. (1984). The role of vortices in insect flight. Zool. Zhurn. 63, 197–208. BUNKER, S. J. (1993). Form, flight pattern and performance in butterflies (Lepidoptera: Papilionoidea and Hesperioidea). PhD thesis, University of Exeter. CASEY, T. M. (1976). Flight energetics of sphinx moths: power input during hovering flight. J. exp. Biol. 64, 529–543. CASEY, T. M. (1981). A comparison of mechanical and energetic estimates of flight cost for hovering sphinx moths. J. exp. Biol. 91, 117–129. CASEY, T. M. AND STEVENSON, R. D. (1989). A comparative analysis of wing morphology and frequency in moths of the family Sphingidae. Am. Soc. Zool. 29, 82A. CHANCE, M. A. C. (1975). Air flow and the flight of a noctuid moth. In Swimming and Flying in Nature, vol. 2 (ed. T. Y. Wu., C. J. Brokaw and C. Brennen), pp. 829–843. New York: Plenum Press. COOPER, A. J. (1993). Limitations of bumblebee flight performance. PhD thesis, University of Cambridge. DICKINSON, M. H. AND LIGHTON, J. R. B. (1995). Muscle efficiency and elastic storage in the flight motor of Drosophila. Science 268, 87–90. DUDLEY, R. (1987). The mechanics of forward flight in insects. PhD thesis, University of Cambridge. DUDLEY, R. (1990). Biomechanics of flight in neotropical butterflies: morphometrics and kinematics. J. exp. Biol. 150, 37–53. DUDLEY, R. (1995). Extraordinary flight performance of orchid bees (Apidae: Euglossini) hovering in heliox (80 % He/20 % O2). J. exp. Biol. 198, 1065–1070. DUDLEY, R. AND DEVRIES, P. J. (1990). Flight physiology of migratingUrania fulgens (Uraniidae) moths: kinematics and aerodynamics of natural free flight. J. comp. Physiol. A 167, 145–154. DUDLEY, R. AND ELLINGTON, C. P. (1990a). Mechanics of forward flight in bumblebees. I. Kinematics and morphology. J. exp. Biol. 148, 19–52. DUDLEY, R. AND ELLINGTON, C. P. (1990b). Mechanics of forward flight in bumblebees. II. Quasi-steady lift and power requirements. J. exp. Biol. 148, 53–88. ELLINGTON, C. P. (1980). Vortices and hovering flight. In Instationäre Effekte an schwingenden Tierflügeln (ed. W. Nachtigall), pp. 64–101. Wiesbaden: Franz Steiner. ELLINGTON, C. P. (1984a). The aerodynamics of hovering insect flight. I. The quasi-steady analysis. Phil. Trans. R. Soc. Lond. B 305, 1–15. ELLINGTON, C. P. (1984b). The aerodynamics of hovering insect flight. II. Morphological parameters. Phil. Trans. R. Soc. Lond. B 305, 17–40. ELLINGTON, C. P. (1984c). The aerodynamics of hovering insect flight. IV. Aerodynamic mechanisms. Phil. Trans. R. Soc. Lond. B 305, 79–113. ELLINGTON, C. P. (1984d). The aerodynamics of hovering insect flight. V. A vortex theory. Phil. Trans. R. Soc. Lond. B 305, 115–144. ELLINGTON, C. P. (1984e). The aerodynamics of hovering insect flight. VI. Lift and power requirements. Phil. Trans. R. Soc. Lond. B 305, 145–181. ELLINGTON, C. P. (1991). Aerodynamics and the origin of insect flight. Adv. Insect Physiol. 23, 171–210. ELLINGTON, C. P. (1995). Unsteady aerodynamics of insect flight. In Biological Fluid Dynamics (ed. C. P. Ellington and T. J. Pedley). Symp. Soc. exp. Biol. 49, 109–129. Cambridge: Company of Biologists. ELLINGTON, C. P., VAN DEN BERG, C., WILLMOTT, A. P. AND THOMAS, A. L. R. (1996). Leading-edge vortices in insect flight. Nature 384, 626–630. ENNOS, A. R. (1989). The kinematics and aerodynamics of the free flight of some Diptera. J. exp. Biol. 142, 49–85. HERTEL, H. (1963). Biologie und Technik. Mainz: Krausskopf. HOCKING, B. (1953). The instrinsic range and speed of flight of insects. Trans. R. ent. Soc. Lond. 104, 223–345. JENSEN, M. (1956). Biology and physics of locust flight. III. The aerodynamics of locust flight. Phil. Trans. R. Soc. Lond. B 239, 511–552. LENDLE, K. (1981). Untersuchungen zur Erzeugung stationärer Auftriebs-und Widerstandskräfte an Insektenflügeln mittels einer umschaltbaren aerodynamischen Einkomponentenwaage. Staatsexamensarbeit, Universität des Saarlandes, Saarbrücken. LIU, H., WASSERSUG, R. J. AND KAWACHI, K. (1996). A computational fluid dynamics study of tadpole swimming. J. exp. Biol. 199, 1245–1260. NACHTIGALL, W. (1964). Zur Aerodynamik des Coleopteren flugs: Wirken die Elytren als Tragflügel? Verh. dt. zool. Ges. 52, 319–326. NACHTIGALL, W. (1966). Die Kinematik der Schlagflügelbewegungen von Dipteren. Methodische und Analytische Grundlagen zur Biophysik des Insektenflugs. Z. vergl. Physiol. 52, 155–211. NACHTIGALL, W. (1977). Die aerodynamische Polare des Tipula- Flügels und eine Einrichtung zur halbautomatischen Polarenaufnahme. In The Physiology of Movement; Biomechanics (ed. W. Nachtigall), pp. 347–352. Stuttgart: Fischer. NACHTIGALL, W. (1989). Mechanics and aerodynamics of flight. In Insect Flight (ed. G. J. Goldsworthy and C. H. Wheeler), pp. 1–29. Boca Raton, FL: CRC Press. NACHTIGALL, W. AND HANAUER-THIESER, U. (1992). Flight of the
Docsity logo



Copyright © 2024 Ladybird Srl - Via Leonardo da Vinci 16, 10126, Torino, Italy - VAT 10816460017 - All rights reserved