Sie sind auf Seite 1von 16

The Aerodynamics of Hovering Insect Flight. I. The Quasi-Steady Analysis Author(s): C. P.

Ellington Source: Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, Vol. 305, No. 1122 (Feb. 24, 1984), pp. 1-15 Published by: The Royal Society Stable URL: http://www.jstor.org/stable/2396072 . Accessed: 20/09/2011 12:02
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact support@jstor.org.

The Royal Society is collaborating with JSTOR to digitize, preserve and extend access to Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences.

http://www.jstor.org

Phil. Trans. Soc.Lond. 305, 1-15 (1984) R. B Printed Great in Britain

[ 1]

THE AERODYNAMICS OF HOVERING INSECT FLIGHT. I. THE QUASI-STEADY ANALYSIS


BY C. P. ELLINGTON Street, Cambridge CB2 3EJ, UK. Department Zoology,University Cambridge, of of Downing F.R.S. - Received March 1983) 28 (Communicated Sir JamesLighthill, by CONTENTS
PAGE

1. INTRODUCTION

2
THEORY

2. 3.

THE

BLADE-ELEMENT GROUPS

3 5 5 6 8
10
11

KINEMATIC

3.1. Horizontal strokeplane 3.2. Inclined strokeplane 3.3. Vertical strokeplane 4. 5.


THE
PROOF-BY-CONTRADICTION OF THE PRESENT STUDY

OUTLINE

6. SYMBOL TABLE
REFERENCES

12 14

invokes the 'quasiThe conventional aerodynamic analysis of flappinganimal flight steady assumption' to reduce a problem in dynamics to a succession of static conditions: it is assumed that the instantaneous forces on a flapping wing are equivalent to those forsteady motion at the same instantaneous velocity and angle ofattack. The validityofthisassumptionand theimportanceofunsteadyaerodynamic effectshave long been controversial topics. Weis-Fogh tested the assumption for are hoveringanimal flight, where unsteady effects most pronounced, and concluded thatmostinsects indeed hoveraccordingto theprinciples ofquasi-steadyaerodynamics. The logical basis for his conclusion is reviewed in this paper, and it is shown that the available evidence remains ambiguous. The aerodynamics of hovering insect flightare re-examined in this series of six papers, and a conclusion opposite to Weis-Fogh's is tentatively reached. New morphological and kinematicdata fora varietyof insectsare presentedin papers II and III, respectively.Paper IV offersan aerodynamic interpretationof the wing aerodynamicmechanisms. kinematicsand a discussionon thepossiblerolesofdifferent A generalized vortex theoryof hovering flightis derived in paper V, and provides a methodofestimating mean lift, the induced power and induced velocityforunsteady as well as quasi-steady flightmechanisms. The new data, aerodynamic mechanisms and vortextheoryare all combined in paper VI foran analysis of the liftand power requirementsand other mechanical aspects of hoveringflight.
Vol. 305 B
II22.

[Published FebruaryI984 24

C. P. ELLINGTON A large number of symbols are needed for the morphological, kinematic and aerodynamic analyses. Most of them appear in more than one paper of the series, and so a singlecomprehensivetable definingthe major symbolsfromall of the papers is presentedat the end of this paper.
1. INTRODUCTION

Research into the aerodynamic principlesof flapping animal flighthas long been hampered by a limited repertoireof investigativetools. Aerodynamics is primarilya practical science, of concentratingon the forcesexperienced by a wing in steady motion. The effects periodic variationsin wing attitudeand velocitysuperimposedon a mean motionhave been investigated them to in flutter analysis, but such analytical treatmentsemploy linearizations that restrict has thus been small amplitude oscillations. The scope fortheoreticalanalyses of animal flight severelylimited, and fromnecessityit is generally assumed that the instantaneous forceson a flapping wing are those correspondingto steady motion at the same instantaneousvelocity Whether or not this assumption is valid for large and attitude, the quasi-steady assumption. and a decisive amplitude, high frequencymotions has been the source of much controversy, test can only be accomplished by comparing measured instantaneous wing forceswith those of predicted by the assumption. The aerodynamic characteristics a wing in steady motion are easily determined,but a direct measurementof the cyclic forcesproduced by flappinganimal wings has only been achieved recently (Cloupeau et al. I979; Buckholtz I98I), and a techniquesis stilllacking. Researchers comprehensivetestofthe assumptionusing thesedifficult have thereforebeen forced to adopt a debatable assumption which could not be tested experimentally. can be testedtheoretically only The validityof thequasi-steady explanation offlappingflight in a proof-by-contradiction. The mean forces generated by the wings during a cycle are the calculated according to the quasi-steady assumption. If theseforcesdo not satisfy net force balance of the flying animal, then the assumptionmust be false. If the mean quasi-steady forces do satisfythe balance, then the only logical conclusion is that the assumption cannot be discounted. The range of forcesgenerated by wings is restrictedby physical considerations: unconventional aerodynamic mechanismsmay produce mean, and even instantaneous,forces very similar to the quasi-steady values. A satisfactoryexplanation of flightbased on the quasi-steady assumption cannot preclude alternativemechanisms. The early quantitative studies on the aerodynamics of flapping flighthave been reviewed were by Weis-Fogh & Jensen (I956). The available data on wing motion, the kinematics, theories on rough incomplete and often inaccurate, so it was usually necessary to base the kinematic approximations. Some studies concluded that the wing forceswere not consistent with a quasi-steady mechanism, but the sources of error were too large for an unqualified acceptance of the conclusions.In a classic seriesofpapers Weis-Fogh andJensen thenpresented for what is still the most complete studyof flappingflight the desert locust Schistocerca gregaria. of and the aerodynamic characteristics the wingsin steady The kinematicsoffasttorwardflight motion were measured (Weis-Fogh I956; Jensen I956). Jensen (I956) then combined these data in a quasi-steady analysis and found that the mean calculated wing forcesagreed well in with those measured fromlocusts flying a wind tunnel. This is a very persuasive argument but it cannot be considered as a proof. forthe quasi-steady explanation of fastforwardflight, indicate that the Indeed, the cyclicforcemeasurementsof Cloupeau etal. ( I 979) on Schistocerca quasi-steady mechanism is not a sufficient explanation in this case.

THE

QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

It is generallybelieved thatthequasi-steady assumptionis valid forfastforward because flight of the low value of the reduced frequencyparameter. This parameter, the ratio of a flapping velocityof the wings to the forwardflight velocity,may be expressed in several ways and was introduced to animal flight studies by Walker (1925). At low values, the steady flight velocity dominates the flow over the wings and reduces the spatial derivatives of the fluctuating aerodynamic parameters. Thus unsteady aerodynamic effectsare small compared with quasi-steady ones. At slowerflight speeds thereduced frequency parameterincreasesand unsteadyeffects should become more important. Hovering flightpresents the extreme condition where the flight velocity is zero and the reduced frequencybecomes infinite.Bennett (i 966) and Weis-Fogh (I972, I973) emphasized the increased significance unsteadyeffects be found in hovering of to flight.The wings accelerate and decelerate in a distance of only a few chord lengths: virtual mass forcesand unsteadycirculatory effects may no longerbe negligible.The amplitude ofwing rotationabout a longitudinal axis during pronation and supination exceeds a rightangle, and the rotational velocity is comparable with the flapping velocity. Under these conditions the aerodynamic behaviour of the wingsmay differ appreciably fromthe quasi-steady assumption. By exaggeratingthe unsteady aerodynamic effects flappingflight, of hoveringthus presentsan ideal case forthe investigationof flightmechanisms. In addition to its advantage in stressing unsteady effects, study of hoveringis usefulfrom a other considerations.The forceand moment balance during flightis simplifiedby the lack of a net horizontal thrust, and the effects body liftand body drag can be ignored. The absence of of a linear flightvelocity acting in combination with the flapping velocity also reduces the complexityof the mathematical treatment.Excluding brisk manoeuvres and climbing flight, the liftcoefficients the wings during hovering must be greater than other types of flight of because the relativevelocityof the wingsis not enhanced by the flight velocity.Thus hovering demands maximum liftcoefficients, which are particularlyuseful when comparing the wing performancewith steady-stateconditions. Finally, the total power expenditureof the animal is greatestin hoveringflight, providing a convenient case forphysiological discussions. Weis-Fogh (I972, I973) applied a quasi-steady analysis to hovering animal flightand concluded that 'most insects performnormal hovering on the basis of the well-established principles of steady-stateflow' (Weis-Fogh I973). The aerodynamics of hovering flightare re-examined in this study using new data and a new theory,and the opposite conclusion is stronglyindicated. Before this work is presented, the previous approaches to the problem will be reviewed. 2.
THE
BLADE-ELEMENT THEORY

The usual aerodynamic treatmentof animal flightis based on the blade-element theory of propellersdeveloped by Drzewiecki. Osborne (I95i) derived the general equations that apply to flappingflight.The fundamentalunit of the analysis is the blade element,or wing element, which is that portion of a wing between the radial distances r and r+ dr fromthe wing base (figure 1a). The aerodynamic forceF' on a wing element can be resolved into a component normal to the flow velocity,called the lift,and a component parallel to the flow, called the profiledrag (figure1 b). The liftL' and profiledrag Dpro fora wing section of spanwise width dr are L'= lpcUr CL, (1)
D;ro = 1P CUr CD ,pro,

(2)
I1-2

C. P. ELLINGTON

where p is the mass densityof air, c is the wing chord, and Ur is the relativevelocitycomponent perpendicular to the longitudinal wing axis. Any spanwise component of the relative velocity on is assumed to have no effect theseforces.CL and CD,pro are liftand profiledrag coefficients, which can be defined for unsteady as well as steady motion. According to the quasi-steady are assumption, though, these coefficients functionsonly of the Reynolds number and angle Equations (1) and (2) are resolved ofattack ofthe relativewind forgiven profilecharacteristics. into vertical and horizontal components,integratedalong the wing length and averaged over a cycle. For hoveringflightthe net forcebalance requires the mean vertical forceto equal the weight of the animal, and the mean horizontal forceto be zero.

(b)

L'*

F'

Dproo

to plane, whichis inclinedat an is confined a stroke duringhovering approximately FIGURE 1. (a) The wingbeat

of parameters a wingelement widthdrand chordc are shown for to The aerodynamic angle/3 thehorizontal. view in a two-dimensional in (b).

The relativevelocityis the vector sum of the flappingvelocity U and the induced velocities of bound and wake vorticity.The blade-element theory ignores the existence of vortices, however,and so can reveal nothingof the induced velocity.Following Osborne (I 95 I), a mean value of the vertical induced velocity w0 is usually estimated by the Rankine-Froude axial momentum theoryofpropellers. One method of applying the blade-element analysis relies on successive stepwise solutions of equations (1) and (2). Complete kinematicdata are necessaryforthisapproach: the motion of the longitudinal wing axis, the geometrical angle of attack and the section profilemust all of be known as functions time and radial position. The angle of attack of the relative wind ar can then be calculated, and appropriate values of CL and CD,pro selected fromexperimental

THE QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

measurements.This was the method used so effectively byJensen (I956) in his studyofforward flightin the desert locust, but it becomes increasingly unreliable as the hovering state is approached, where momentumconsiderationsshow that the wake induced velocityreaches a maximum. For many hoveringanimals thereare periodswhen the opposite wingsare separated by a distance of less than a chord length,so that the bound vorticity the opposite wing will of also contribute significantlyto the relative velocity of a wing. This effectcan only be approximated and, when coupled with the crude estimate of the wake induced velocity, a considerable inaccuracy can be expected for the calculated value of ar. The liftcoefficient is strongly dependent on a, so the solution to equation (1) is prone to serious error. Osborne (I 95 I) introduced an alternativemethod of analysis which solves formean values ofthe coefficients, and CD, pro, satisfying net forcebalance. This is equivalent to assuming the CL thattheforcecoefficients constantalong thewing and throughout cycle. The assumption are the is not unreasonable, because the wingscould well be operatingat some optimal angle of attack. For small insectslike Drosophila, is nearlyconstantanyway over the range ofar foundduring CL flight(Vogel I967). The mean liftcoefficient particularlyinteresting is because it is also the minimum value compatible withflight;ifCL varies duringthewing beat thensome instantaneous values must exceed CL. The kinematicdetail required forthismethod is greatlyreduced since only the motion of the longitudinal wing axis is needed. The geometrical angle of attack and profile section of the wings are extremelydifficultto measure accurately, and one of the principal advantages of this method is its neglect of them. The double integralsof radius and timeproduced by manipulation ofequations (1) and (2) can be separated when the coefficients are assumed to be constant; the resulting singleintegralsmay thenbe reduced to morphological and kinematic parameters, allowing simple expressions to be derived for CL and CD, pro the satisfying integral constraints. The theoreticalinvestigations hoveringflight of have usually relied on the mean coefficients method (Weis-Fogh I 972, I973; Ellington I975; R. A. Norberg I975; U. M. Norberg I975, I976). The animals studied so far can be divided into three functional groups, for which different approximations can be used in the calculations outlined above. During a cycle the wingsbeat roughlyin a stroke plane,which is inclined at an angle ,/withrespectto the horizontal 1 a); it is the attitude of this strokeplane that distinguishesthe groups. (figure

3.

KINEMATIC

GROUPS

3.1. Horizontal stroke plane The most commonly observed type of hovering is characterized by an approximately horizontal strokeplane (figure2 a). This group was extensively discussed by Weis-Fogh (I972, I973), who called thepatternnormal and hovering, includes thehummingbirds(Stolpe & Zimmer I939; Greenewalt I960) and most insects (Weis-Fogh I973). The amplitude of wing rotation during pronation and supination is large, enabling the wings to operate at an angle of attack favourable forlifton both the morphological downstrokeand the upstroke (figure2a, b). The wake induced velocityis small with respectto the flappingvelocityof the wingsforthisgroup, in resulting a relativevelocityveryclose to the horizontal (Ellington I978, I980). This justifies using the horizontal flapping velocity for Ur in equation (1) and neglectingany small drag contributionsto the net vertical force.A mean liftcoefficient then easily found such that the is

C. P. ELLINGTON

(a)

(b)

W >\

FIGURE

aureoventris viewedfrom side, illustrating a 2. (a). The wingtip path of a hummingbird the Chlorostilbon is on for stroke horizontal plane.The wingattitude indicated thepath. (b) The wingbeat pattern Melanotrochilus above. Adaptedfrom viewedfrom fuscus Stolpe & Zimmer(I939).

net liftgiven by the integral form of equation (1) balances the body mass. A mean drag cannot be determined by the integral constraints,but once CL is calculated an coefficient appropriate CD, pro may be selected fromexperimentalmeasurements. In his comparative survey of normal hovering animals, Weis-Fogh (I973) used simple analytical expressionsfor wing chord and flapping velocity; a semi-ellipticalwing planform was used, and the flappingvelocitywas taken as simpleharmonic motion. These are reasonable a approximations,permitting quick calculation of CL, but increase the error to an estimated 30 %. Table 1 presentssome of Weis-Fogh's resultsand a Reynolds number Re forthe wings; the Reynolds number is based on the chord and the mean flappingvelocityat 0.7 of the wing is length. CL forthe small wasp Encarsiaformosa taken fromEllington (I975). In this case the wing shape and velocitywere not well representedby Weis-Fogh's approximationsand a more detailed treatmentwas necessary. stroke 3.2. Inclined plane I Small passerine birds(ZimmerI94; BrownI963; U. M. Norberg 7), bats (Eisentraut I936; U. M. NorbergI970, 1976) and the 'true' hover-fliesof the subfamily Syrphinae

THE

QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

(Weis-Fogh I973) hover with the strokeplane inclined at a value of , between 300 and 400 (figure3). Dragonflies (Odonata) also hover in this manner, although , is about 600 (R. A. Norberg I975). These animals appear to generate relativelysmall forceson the upstroke.The supinates itswingson the upstroke,making the angle ofattack dragonfly Aeschnajuncea strongly vertebrate fliers anatomically are near zero (R. A.Norberg I 97 5). Excluding thehummingbirds, unable to rotate theirwings to thisextent.The birds and bats partiallyflextheirwings during the upstroke,however, and the individual primaries of birds are also rotated to a negligible angle of attack. Any lifton the upstroke of an inclined strokeplane would produce a large upstrokelift. horizontalthrust component,and thisis probably the explanation forinsignificant The mean forceon the downstrokeis thus primarilyresponsibleformass support.

(a)

(c)

(d)

(e)

(f)

(g:

(h

FIGURE

an stroke plane during hovering bat auritus, illustrating inclined 3. The wingbeat thelong-eared Plecotus of U. flight. Adaptedfrom M. Norberg(1970).

The relative velocity Ur on the downstrokeshould not be approximated by the flapping velocity; because of the inclination of the strokeplane, the induced velocity will reduce the essential value of Or well below U2. An accurate estimate of the induced velocityis therefore in calculating the magnitude and direction of the relative velocity. It is likely that previous estimatesof the induced velocityare substantiallytoo small (see papers V and VI), which led to an incorrectconsiderationofdownstrokeprofiledrag by Ellington (I980). From the results of R. A. Norberg (I975) and estimatesbased on papers V and VI, the value of CL required of should lie between 3 and 4. U. M. Norberg forweightsupporton thedownstroke Aeschnajuncea (I975) calculated that CL mustequal 5.3 forthe net verticalforceon the downstroketo balance In the body mass of the pied flycatcherFicedulahypoleuca. paper VI, I estimate that CL for Ficedulais about 6.0, so a value between 5 and 6 can be expected. It should be noted that in both cases the morphological and kinematicdata were not fromthe same animals, and so an

unKnown

inaccuracy

iS present.

C. P. ELLINGTON

were assumptions about upstroke forceson the mean coefficients The effectsof different evaluated by U. M. Norberg (1976) in a thoroughinvestigationof hoveringin the bat Plecotus occur when no liftis generated on auritus. Her resultsdemonstratethat minimum coefficients the upstroke,which then requires CL= 3.1 on the downstroke;my estimateof CL is about 4.3 (paper VI), and so a value between 3 and 4 seemsprobable. She also commentedon Weis-Fogh's were invalid. This is also noting that his procedural simplifications (I973) analysis of Plecotus, true of his calculations forother animals using an inclined or vertical strokeplane.

(a)

ol(b)

(C)

(d)

FIGURE

and The plane is vertical, thewingmotion take-off Pieris brassicae. stroke by 4. The downstroke a vertical of is and the resulting createdby thedownstroke indicated, pattern is perpendicular the chord.The vortex to in section (d). ringis shownin vertical vortex large-cored

stroke 3.3. Vertical plane the In addition to theinsects coveredin paper III, I have filmed Large Cabbage Whitebutterfly L. Pierisbrassicae in freeflight.A unique kinematic pattern, an approximately vertical stroke and hovering. Figure 4 shows several stages of the downplane, is oftenseen during take-off froma small platform.The wings are clapped together strokewhich initiatesa vertical take-off dorsallyat the startof the downstroke,and then 'fling' open (Weis-Fogh I973) as in figure4 a. The wings move with the chord perpendicular to theirmotion (figure4b, c) and nearly clap togetherat the end ofthe downstroke (figure4d). The body pitches nose-up and the wings supinate during the upstroke,which is not shown here, producing an angle of attack strongly near zero. Thus littleforceis generated on the upstroke,as forthe inclined strokeplane. The sustainingvertical force obviously resultsfromthe pressuredrag on the wings during the downstroke. A drag mechanism of flighthas been suggested several times for flightat Reynolds numbersbelow about 100 (Horridge I956; R. A.Norberg 1972 a, b; Bennett I973),

THE

QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

where thickboundary layersreduce the steady-state coefficients wings.The drag is mainly lift of due to skinfriction low Re, however,and is not verysensitiveto ar. A drag mechanismbased at on a differential velocityof the half-strokes was therefore proposed (Bennett 1973), such that the downstrokevelocity and drag were greater than the upstroke. However, the small wasp Encarsiawas foundto use a horizontalstrokeplane and a circulatory mechanism (Weis-Fogh lift 1973; Lighthill I973; Ellington 1975). I have also filmedthe small fringe-winged insect Thrips physapus and the fruit fly Drosophila melanogaster (Ellington 1983); the kinematics and aerodynamics are verysimilar to Encarsia.Ironically the drag mechanism does not operate for small insects,forwhich it was predicted,but ratherfora large one at higherReynolds number (about 2800). The pressuredrag during the downstrokeis much greater than the skinfriction is drag of the upstrokeat high Re, and so the need fora velocitydifferential eliminated.

TABLE

1.

RESULTS FROM THE QUASI-STEADY ANALYSIS

(References:1, Weis-Fogh (1972); 2, Weis-Fogh (I973); 3, Ellington (I975); 4, R. A. Norberg (I975); 5, U. M. Norberg(1975); 6, U. M. Norberg(1976); 7, present study.) species Amaziliafimbriatafluviatilis Melolontha vulgaris sexta Manduca Bombus terrestris Apismellifera Eristalis tenax Tipulasp. virilis Drosophila Encarsiaformosa Plecotus auritus Ficedula hypoleuca Aeschnajuncea Pieris brassicae
CL

Re 6100 3900 5400 3600 1600 1600 630 210 23 15000 11000 1900 2800

ref. (1) (2) (2) (2) (2) (2) (2) (1) (3) (6,7) (5,7) (4,7) (7)

horizontal stroke plane 1.8 0.6 1.2 1.2 0.8 0.9 0.8 0.8 1.6

inclined stroke plane 3-4 5-6 3-4 vertical stroke plane 0

Equation (2) of the blade-element analysis could be applied to the Large Cabbage White but a constant value of CD pro on the downstrokeis unlikely because of pressure butterfly, fromthe opposite wings. A different interference aerodynamic approach has been developed, based on the vortex pattern created by the wing motion. Air rushes into the gap therefore, between the wings as they flingopen at the beginning of the downstroke,producing a sink on motion bounded by the vorticity the wing surfacesand the freevortexlines connectingthe leading edges and the trailing edges (figure 4 a). This flow pattern is maintained and strengthenedduring the downstroke,resultingin a large-core vortex ring (figure4d). The reaction force and kinetic energy associated with this ring have been estimated from flow mechanismis conceptually quite simple, many visualization photographs.Although thisflight subtle modificationsto the flow are effectedby wing twistingand flexibility, and a detailed analysis will be presentedelsewhere (Ellington I983).

10
4.
THE

C. P. ELLINGTON
PROOF-BY-CONTRADICTION

The mean lift coefficient been chosen in all ofthehoveringanalyses to testthequasi-steady has assumption. If CL is greater than the maximum steady-stateliftcoefficient max' then the CL, assumptionis contradictedand unsteadyflight mechanismsmustbe invoked. When CL is about equal to CL, max the testis not conclusive: some variation of the liftcoefficient likelyduring is a cycle, and instantaneousvalues may then exceed the mean. Values of CL,max are not known for the animals in table 1. For the insectswe may note a value of 1.1 at Re about 2000 for the flatforewingof a desert locust, which rises to 1.3 when the vannal region is depressed as a flap (Jensen 1956). Nachtigall (I977) measured a smaller CL, max of 0.85, however, for a crane-fly(Tipula oleracea)wing at a similar Re of 1500. Vogel (i 967) found a value of0.6 fora flatDrosophila virilis wing at a lower Re of 200, which increases to 0.8 fora cambered wing; his resultsare very similar to those of Thom & Swart (1940) for a flat-bottomedaerofoil in this low Re range. Of the insects in table 1, Encarsia and Aeschna are the only examples where CL definitely exceeds CL, max for the relevant Re, and unsteady effects are therefore indicated. The flightmechanism of Pierishas already been outlined and must also be considered as unsteady. Values of CL forthe remaininginsectsare uncomfortably close in general to the expected values of CL max) and several lie in the region of uncertainty between Jensen's and Nachtigall's results. Because of this and the large margin of error in Weis-Fogh's (I973) analysis, no firmconclusions can be drawn concerning these insects. The values of CL,max forbird and bat wings are more controversialthan those forinsects. The wings are thin and highlycambered, and generally operate at Re between 104 and 105. Experimentson similar aerofoil sections have found CL,max up to 1.6; a very strongcamber or an effective increase in camber resultingfromleading edge and trailingedge flaps can raise
the value to about 2.5 (Schmitz 1957; Abbott & Doenhoff 1959; Mises 1959; Tucker & Parrott
1970;

Hoerner & Borst 1975).

Which of these sections best respresents avian and bat wings

is notclear. Primaryseparationon thedownstroke as ofbirdsmayfunction slottedor mul'ti-slotted flaps (Graham 1932), which can produce values of CL,max in excess of 3. The alula of bird wings is commonly raised in slow and hoveringflight, increasing the liftsome 25 % by acting as a Handley-Page leading edge slot (Nachtigall & Kempf 197 1). These comparisonsindicate that CL,max may be about 1.6 forthe wings, and possiblymuch more. Measurements on bird wings and wing models suggestthat these values are too high. Even is with the alula raised, CL, max for the wing of a house sparrow Passer domesticus only 1.1 at it Re 16000, and fora European blackbird Turdusmerula is only 0.8 at Re 24000 (Nachtigall & Kempf I 971). Measurementsfrommany otherbird wingsover an Re range of 10000-50 000 are much the same: 0.7-1.2 (Withers I98I). There are problemsof course, to mountinga wing in a realisticattitudein a wind tunnel. Studies on wing models of the pigeon Columba liviashow a similar range of values, however, from0.8 to 1.2 (Nachtigall 1979). For gliding birds and bats values around 1.5 are found for the maximum lift coefficientbased on wing area (Pennycuick I968, 1971; Tucker & Parrott 1970), but when the tail area is included as a lifting surface CL max drops to about 1.3 for the pigeon (Pennycuick I968). The values of CL max actually measured for bird wings and wing models are thus well below those that might be
expected.

For the hummingbird CL is 1.8, which Weis-Fogh (I973) accepted as possible under quasi-steady conditions.Judging by the resultsabove, I thinkthat thisis unlikelyand unsteady

THE QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

11

CL will then be significant.For Plecotus is 3-4, which should be greater than CL,max effects fora bat wing. The quasi-steady assumption failsforFicedulaas well. The resultsof the analysis can now be summarized forthe kinematic groups. The animals mechanisms,as concluded by using an inclined or vertical strokeplane relyon unsteady flight even as an approximation to the Weis-Fogh (I973); the quasi-steady mechanismis insufficient aerodynamic forces. Of the animals using a horizontal stroke plane, Encarsia must use an unsteady mechanism,and thisis probably true forthe hummingbirdas well. The proofis not conclusive forthe remaining animals, but theirvalues of CL are very close to CL, max, From such resultsWeis-Fogh (I973) concluded that mostanimals hoveringwitha horizontal strokeplane rely on the quasi-steady flightmechanism. However, threepoints must be borne these results. If CL does not exceed CL, max) then the proof-byin mind when interpreting that thequasi-steady assumptioncannot be discounted; it does contradictiononly demonstrates not prove that the assumption is true. Furthermore,some instantaneous values of the lift are coefficient likelyto be greater than the mean CL over the cycle. And finally,the mean lift in unsteady motion can be substantiallysmaller than the quasi-steady estimatebecause of the so Wagner effect, the wings cannot actually achieve the maximum liftduring hovering that is predicted by CL max (paper IV). In the light of these last two points, the close agreement between CL and CL, max couldindicate that the quasi-steady mechanism is notable to provide sufficient forhovering. The example of the hummingbirdcould also lead to a conclusion lift to contrary Weis-Fogh's. The kinematicsofmostinsectsare similarto thoseofthehummingbird, flightmechanisms were involved; if the quasi-steady and it would be surprisingif different explanation is indeed inadequate forthe hummingbird,then it mightnot apply to the insects either. Much more research is obviously required before Weis-Fogh's conclusion can be substantiatedor disproved.

5.

OUTLINE

OF

THE

PRESENT

STUDY

are The aerodynamicsofhoveringinsectflight re-examinedin thisseriesofpapers, to evaluate the rolesofquasi-steady and unsteadyaerodynamic mechanisms.Any aerodynamic studymust be based on accurate kinematicand morphological data, and the previous studieshave usually in been unsatisfactory this respect: the amplitude of wing motion was oftendeterminedonly by visual estimation,some of the data used by Weis-Fogh (1973) were frominsects in slow forwardflight, and the morphological and kinematicdata were not fromthe same animals in Ficedulaand Amazilia. Papers II and III presentcomplete data sets for the analyses of Aeschna, a varietyof insects to be used in this study. Weis-Fogh (I973) proposed two unsteady flight mechanisms for animals where the of quasi-steady assumptionfails,based on an aerodynamic interpretation the wing kinematics. aerodynamic This approach is followed in paper IV to discuss the possible roles of different mechanisms. The flightmechanisms cohsidered in paper IV cannot be directlyincorporated into the developed in paper V, blade-element theory.A more generalized vortex theoryis therefore as well as quasi-steady mechanisms. The theory which calculates the mean liftfor unsteady also provides more accurate estimates for the induced velocity and induced power than previouslypossible. data, aerodynamic mechanismsand vortextheoryare all combined in paper The new flight

12

C. P. ELLINGTON

VI foran analysis of the flightmechanisms,theirpower requirements,and other mechanical considerationsof hoveringflight. This study would not have been possible without the help, encouragement and guidance of three men: the late ProfessorT. Weis-Fogh, who inspired the initial lines of research, Dr of K. E. Machin, whose many talentsand enthusiasticsupport were vital to the fruition much work, and Mr G. G. Runnalls, whose photographic expertisewas invaluable. To these three, I extend my warmest gratitude.

6.

SYMBOL

TABLE

the of in all This tablecontains ofthemajorsymbols thisseries papers.Exceptfora fewgeneralsymbols, paper and thisis indicated formally equations, by is is are whereeach symbol defined givenbelow.Some symbols defined withinthatpaper. A 'hat' denotesa by XX.xx, whereXX is thepaper numberand xx is theequationnumber a rules unless specialdefinition to obeythefollowing Othermodifications symbols form non-dimensional ofa symbol. and minimum values,a bar over the symbol to 'max' and 'min' refer maximum below: subscripts is presented mechanical represents a per denotes force unitspan,and an asterisk meanvalue,a prime indicates time-averaged a powerper unitbodymass. paper or equationno. V V V.10 V V 11.4 V V II 11.5 V.34 IV.7 11.17 III VI VI.32 V VI V 11.11 V V.5 V II

symbol a A Ao Ar Al ? b bo c

definition in axial separation vortex of rings thefarwake area cross-sectional ofthefarwake area oftheactuatordisc sheetproducedby a half-stroke area ofthevortex sheet area of thatvortex rolled-up aspectratioofthewingpair in rings thefarwake radiusofvortex at radiusofvortex rings theactuatordisc chordofthewing mean chord of ringvelocity self-induced component thevortex drag coefficient CD fora flatplate parallelto theflow profile drag coefficient lift coefficient of mean diameter body/body length to of ratioofduration downstroke upstroke of drag vector magnitude themean profile D/mg of impulses frequency lift f/wingbeat frequency force a wing on aerodynamic theforce from momentum a jet gravitational acceleration length mean wingthickness/wing circulatory impulse lift non-dimensional form I of for theimpulserequired weight support moment inertia of for moment inertia animalbody of moment inertia thevirtualmassofa wingpair of for for moment inertia themassofa wingpair of
profiledrag c/c

C
CD

II

CD,f

d/u D 1^

CD, pro CL

f f

Dpro

F g I

or

f
I
Ib

or

Iv 1w

~~II
II

THE QUASI-STEADY
J k K
I
11 12

ANALYSIS

OF HOVERING

FLIGHT

13

or L L m mw rAw n Pd Pw Pa Pace
Pind
mk

111.23 V.31 V.33 11.18 II 11.20 II II VI II II 11.8 V VI VI.39 VI VI

advance ratio inducedpowerfactor specific in of due to otherrings thewake component thevortex ringvelocity end distance from anterior ofbodyto centre mass/body of length of distance from base axis to centre bodymass/body forewing length radiusofgyration thebody/body for length bodylength bodylength/wing length lift mean lift/mg bodymass massofwingpair mw/m of thekthmoment wingmassabout thewingbase wingbeat frequency disc loading wingloading aerodynamic power powerrequired acceleratemassand virtual to massofthewings mechanical poweroutputofmuscle of Pmper unitweight muscle
profilepower induced power

VI.20

Pm

Ppro

VI.29

PRF

VI.21 11.9 11.7 11.16 VI.28 V.36 11.3 II~1.6

f fk(m)
fic(S) fk (V)

R Re s S

t t U Ur Ut v v 0 V V w wo
Vk

Sk

or

V.8 III V 11.13 11.14 II.15 III III V V


V

Rankine-Froude estimate inducedpower of radial position along thewing r/wing length non-dimensional radiusofthekthmoment wingmass of non-dimensional radiusofthekthmoment wingarea of non-dimensional radiusofthekthmoment virtual of mass winglength Reynolds number spacingparameter wingarea kthmoment wingarea of time t/wingbeat period flapping velocity thewing of relative velocity thewing of flapping velocity thewingtip of axial velocity vortex of ring virtual massofa wingpair non-dimensional virtual massofa wingpair kthmoment virtual of mass flight velocity = V/nR number winglengths of travelled wingbeat per axial wake velocity thefarwake in axial wake velocity theactuatordisc = inducedvelocity at Rankine-Froude estimate w of Rankine-Froude estimate wo of distance from leadingedge to axis ofwingrotation xO/c geometrical angle ofattack engle ofincidence zero-lift angle ofattack effective angle ofattack effective angle ofincidence half-angle between chords duringfling

WRF

X0 xO ao ar ar. ,3 /,
a at

WO,RF

or

V IV IV IV IV IV IV IV IV

I3o
,dr

III ~~VI.4

valueoffiin 'true'hovering relative stroke planeangle

stroke plane angle

14
y

C. P. ELLINGTON
VI.13
v

r
j2

r
a

Fo
e
e

v V V VI.34 or

IV.14

rotational coefficient lift circulation non-dimensional circulation (=r/ ..) mean ofr overarea ofvortex sheet mean ofr2overarea ofvortex sheet circulation a pulsedFroudeactuatordisc of magnitude themean profile of drag/mean themagnitude theprofile of of drag
A

downwash angle

60 y1 6 A A A p
Va

V V III VI.53 111.22 IV IV. 17 IV. 18 III II II V.43 V.44 111.24

6
Pb

Pw a'
X

x Xo w 6

III III III III

coreradiusofvortex ringin farwake core radiusofvortex disc ringat actuator rollangle aerodynamic efficiency of angleofelevation wingwithrespect stroke to plane distancetravelled wingelement/chord by value ofA overa half-stroke mean distancetravelled wingovera half-stroke/mean by chord kinematic viscosity air of angle between mean flight path and thehorizontal massdensity air of massdensity thebody of massdensity thewing of factor inducedpower for spatialcorrection correction factor inducedpower for temporal positional angle ofwingin thestroke plane non-dimensional form qJ of stroke angle angle between longitudinal bodyaxis and thehorizontal freebodyangle angularvelocity and supination duringpronation mean angularvelocity/wingbeat frequency

REFERENCES

New Abbott, H. & Doenhoff, E. von 1959 Theory wing I. A. of sections. York: Dover N.Y. Bennett, I966 Insect L. aerodynamics: vertical sustaining force near-hovering in flight. Science, 152,1263-1266. and of Ann. Soc.Am.66, 1187-1190. ent. Bennett, I973 Effectiveness flight smallinsects. L. Brown, H. J. I963 The flight birds.Biol. Rev.38, 460-489. R. of in Buckholtz, H. I 98 I Measurements unsteady R. of periodic forces generated theblowfly by flying a windtunnel. J. exp.Biol.90, 163-173. Cloupeau,M., Devilliers, F. & Devezeaux,D. 1979 Directmeasurements instantaneous in desert J. of lift locust; with on comparison Jensen's experiments detachedwings.J. exp.Biol. 80, 1-15. Eisentraut, 1936 Beitragzur Mechanikdes Fledermausfluges. wiss. M. Z. Zool. 148, 159-188. Ellington, P. 1975 Non-steady-state C. aerodynamics theflight Encarsiaformosa. of of In Swimming in andflyingNature vol. 2, pp. 783-796. New York: PlenumPress. (ed. T. Y. Wu, C.J. Brokaw& C. Brennen), - water, C. In Ellington, P. 1978 The aerodynamics ofnormal hovering flight: three approaches. Comparativephysiology ionsandfluidmechanics K. Schmidt-Nielsen, Bolis & S. H. P. Maddrell), pp. 327-345. Cambridge L. (ed. Press. University In C. and Ellington, P. I 980 Vortices hovering flight. Instationdre anschwingenden Effekte (ed. Tierfliigeln W. Nachtigall), pp. 64-101. Wiesbaden:Franz Steiner. C. Ellington, P. I983 Paper in preparation. Graham,R. R. 1932 Slotsin thewingsofbirds.Jl R. aeronaut. 36, 598-600. Soc. Greenewalt, H. I960 Hummingbirds. York: Doubleday. C. New Hoerner, F. & Borst, V. 1975 Fluid-dynamic Bricktown, S. H. lift. NewJersey: HoernerFluid Dynamics. G. Horridge, A. 1956 The flight verysmallinsects. of Nature, Lond.178, 1334-1335. M. 1956 Biologyand physics locustflight. The aerodynamics locustflight. of III. of Phil. Trans. Soc. R. Jensen, B Lond. 239, 511-552. Lighthill, J. 1973 On theWeis-Fogh M. mechanism lift of generation. FluidMech.60, 1-17. J. New Mises,R. von 1959 Theory offlight. York: Dover. Nachtigall, W. 1977 Die aerodynamische Polaredes Tipula-Flugels und eine Einrichtung halbautomatischen zur In Polarenaufnahme. Thephysiology movement; of biomechanics W. Nachtigall),pp. 347-352. Stuttgart: (ed. Fischer.

THE QUASI-STEADY

ANALYSIS

OF HOVERING

FLIGHT

15

W. 1979 Der Taubenflugel Gleitflugstellung: in Geometrische Kenngrossen Flugelprofile der und Nachtigall, J. Orn., Lpz. 120, 30-40. Luftkrafterzeugung. zur Funktion Daumendes Untersuchungen Flugbiologischen Nachtigall, & Kempf, 1971 Vergleichende W. B. bei Physiol. 326-341. 71, fittichs (Alulaspuria) Vogeln.Z. vergl. L. Alucita pentadactyla and Orneodes L. hexadactyla of Norberg, A. 1972 a Flightcharacteristics twoplumemoths, R. Zool.Scr.1, 241-246. (Microlepidoptera). of of Zool.Scr.1, 247-250. Norberg, A 1972 b Evolution flight insects. R. In L., and flight thedragonfly of Aeschnajuncea kinematics aerodynamics. Swimming Norberg, A. 1975 Hovering R. vol. in (ed. andflyingNature T. Y. Wu, C. J. Brokaw& C. Brennen), 2, pp. 763-781. New York: PlenumPress. of L. flight Plecotus auritus Bijdr.Dierk40, 62-66. (Proc. 2nd int.Bat Res. Conf.) Norberg, M. 1970 Hovering U. in hypoleuca). Swimming flying Nature In and of (Ficedula Norberg, M. 1975 Hoveringflight the pied flycatcher U. vol. 2, pp. 869-881. New York: PlenumPress. (ed. T. Y. Wu, C. J. Brokaw& C. Brennen), in bat auritus. exp. Biol.65, 459-470. of flight thelong-eared Plecotus Norberg, M. 1976 Aerodynamics hovering U. J. J. of flight withapplicationto insects. exp.Biol.28, 221-245. Osborne,M. F. M. 195I Aerodynamics flapping in Biol.49, 509-526. livia.J. exp. study gliding of flight thepigeonColumba Pennycuick, J. I 968 A wind-tunnel C. in J. bat observed a wind-tunnel. exp. aegyptiacus of Pennycuick, J. 197 I Glidingflight thedog-faced Rousettus C. Biol.55, 833-845. des F. Duisburg:Carl Lange Verlag. Schmitz, W. 1957 AerodynamikFlugmodells. J. K. des Lpz. 87, 136-155. Stolpe,M. & Zimmer, 1939 Der Schwirrflug Kolibriim Zeitlupenfilm. Orn., Soc. at on Thom,A. & Swart,P. 1940 The forces an aerofoil verylow speeds.Jl R. aeronaut. 44, 761-770. in of flight a falconand otherbirds.J. exp.Biol. 52, G. Tucker,V. A. & Parrott, C. 1970 Aerodynamics gliding 345-367. of III. characteristics flywingsand wingmodels.J. exp.Biol. Vogel,S. I967 Flightin Drosophila. Aerodynamic 46, 431-443. Soc. G. of flight birds.JI R. aeronaut. 29, 590-594. Walker, T. 1925 The flapping of II. of T. Weis-Fogh, 1956 Biologyand physics locustflight. Flightperformance thedesertlocust(Schistocerca gregaria). Phil. Trans. Soc.Lond.B 239, 459-510. R. in and of T. flight hummingbirds in Drosophila.J. exp.Biol.56, 79-104. Weis-Fogh, 1972 Energetics hovering in for of fitness hovering animals,includingnovel mechanisms lift Weis-Fogh, 1973 Quick estimates flight T. J. production. exp.Biol. 59, 169-230. in A of I. T. M. 1956 Biology and physics locustflight. Basic principles insect flight. critical Weis-Fogh, & Jensen, R. B Phil. Trans. Soc.Lond. 239, 415-458. review. aerofoils. exp.Biol.90, 143-162. J. of P. analysis birdwingsas fixed Withers, C. I98I An aerodynamic J. Lpz. 91, 371-387. (Cinnyris). Orn., Zimmer, 1943 Der Flug des Nektarvogels K.

Das könnte Ihnen auch gefallen