SUMMARY
Organisms that swim or fly with fins or wings physically interact with the surrounding water and air. The interactions are governed by the morphology and kinematics of the locomotory system that form boundary conditions to the Navier–Stokes (NS) equations. These equations represent Newton's law of motion for the fluid surrounding the organism. Several dimensionless numbers,such as the Reynolds number and Strouhal number, measure the influence of morphology and kinematics on the fluid dynamics of swimming and flight. There exists, however, no coherent theoretical framework that shows how such dimensionless numbers of organisms are linked to the NS equation. Here we present an integrated approach to scale the biological fluid dynamics of a wing that flaps, spins or translates. Both the morphology and kinematics of the locomotory system are coupled to the NS equation through which we find dimensionless numbers that represent rotational accelerations in the flow due to wing kinematics and morphology. The three corresponding dimensionless numbers are (1) the angular acceleration number, (2) the centripetal acceleration number, and (3) the Rossby number, which measures Coriolis acceleration. These dimensionless numbers consist of length scale ratios,which facilitate their geometric interpretation. This approach gives fundamental insight into the physical mechanisms that explain the differences in performance among flapping, spinning and translating wings. Although we derived this new framework for the special case of a model fly wing, the method is general enough to make it applicable to other organisms that fly or swim using wings or fins.
INTRODUCTION
The use of dimensionless numbers for understanding complex biological flows has helped us enormously to better understand adaptations for swimming and flying in nature, as well as the corresponding flow phenomena. For example,the Reynolds number, Re, is the ratio of convective acceleration multiplied by density over viscous stress in the fluid (e.g. Tritton, 2005). It not only dictates what kind of propulsive mechanism the organism has at hand (viscous versus inertial) but also determines whether the flow is reversible or irreversible, or whether it is laminar or can become turbulent. Such dimensionless numbers can be interpreted in at least three ways: as ratios of force, time or length (e.g. Tennekes and Lumley, 1983). For example, although the Reynolds number is often described as the ratio of inertial to viscous forces, it can be interpreted as the ratio of convection length (or time) over diffusion length (or time) for a standard transport time (or distance) (e.g. Tennekes and Lumley, 1983). The wide-ranging use of the Reynolds number in biologically relevant flows is well illustrated in Steven Vogel's book `Life In Moving Fluids' (Vogel,1996).
The Strouhal number, St, is another important dimensionless number, which has been used extensively in the biological fluid dynamics literature. It is typically defined as St=fA/U,where f is flapping frequency, A is flapping amplitude and U is mean flow velocity. Its original context was as a measure of dimensionless shedding frequency for a bluff body undergoing von Kármán shedding in a constant flow (e.g. Guyon et al., 2001), but is has additional uses in biological fluid dynamics(Triantafyllou et al., 1993; Taylor et al., 2003). For example, it is proportional to the tangent of maximal induced angle of attack by a flapping wing or fin when the stroke plane is perpendicular to the direction of motion (Taylor et al.,2003; Lentink et al.,2008). This amplitude-based Strouhal number closely resembles the inverse of the advance ratio J=U/2ΦfR as defined by Ellington (Ellington,1984), where Φ is total wing beat amplitude in radians, f is flapping frequency and R is root-to-tip wing length. Note that 2ΦR is actually the total wingtip excursion in the stroke plane (downstroke plus upstroke), whereas U/f measures wingbeat wavelength λ, the distance traveled during one stroke cycle. The advance ratio is therefore a measure of the pitch of a flapping wing; very much like the pitch of a propeller (and the pitch of a screw) provided that the stroke plane is normal to body speed. Dickinson(Dickinson, 1994) and Wang(Wang, 2000b) defined a chord-based Strouhal number Stc=fc/U,where c is chord length. This number closely resembles the reduced frequency k defined by Daniel and Webb(Daniel and Webb, 1987) for swimming k=πfc/U, which is usually defined as a ratio of velocity due to flapping to velocity due to forward motion. It should be noted that, in engineering fluid dynamics, Strouhal numbers are often reserved for their original purpose – to describe natural vortex-shedding processes (e.g. Green,1995) – whereas reduced frequencies (or the analogous dimensionless wavelength) are more appropriate for forced vibrations, such as flapping wings (e.g. Tobalske et al.,2007).
From our brief overview it becomes clear that different points of view exist in the biomechanics field on how to best define and use dimensionless numbers to study swimming and flight. Further, these dimensionless numbers are not always simple to interpret or, perhaps more importantly, similarly defined throughout the field. Our goal is to improve this for flapping studies by deriving a new set of dimensionless numbers. These numbers are not only directly linked to the Navier–Stokes (NS) equations but also can be interpreted more easily based on the morphology and kinematics of the wing (or fin). For this we chose to use morphological and kinematic length scale ratios, because they are most easy to interpret and illustrate geometrically.
For simplicity we focus our analysis on fly wings. In the discussion we will indicate how to apply the theory to the wings and fins of other organisms such as insects, birds, fish and samara seeds. First we derive the dimensionless NS equation with respect to the surface of the flapping wing of a forward flying fly. Next we show that the angle between the body velocity vector of a fly and its (approximate) stroke plane can be neglected for estimating the correct order of magnitude of the dimensionless numbers that depend on speed. We then further simplify the NS equation for hover conditions. Next we simplify the NS equations even further for spinning and translating fly wings. These more simplified forms of the NS equation and the corresponding dimensionless numbers are illustrated graphically for 3D wings. Our framework can therefore be readily applied to the design of appropriate parameter spaces for complex fluid dynamic studies of flapping, spinning and translating wings and fins. Finally we discuss how 3D and 2D wing kinematics are related and can potentially mediate the stall characteristics of a wing. Experimental tests of the efficacy of this new approach in characterizing salient features of biologically relevant forces and flows are presented in the accompanying paper (Lentink and Dickinson, 2009).
MATERIALS AND METHODS
NS equation of a fly in forward flapping flight
Coordinate transformation that simplifies the fly's velocity boundary condition
A complementary approach is to first transform the governing equations and boundary conditions (Eqns 1, 2, 3, 4, 5) such that the velocity boundary condition on the surface of interest is simplified. Preferably, the velocity boundary condition becomes zero such that we do not need to track the surface explicitly but implicitly through the coordinate transformation (e.g. Anderson, 1991; White, 1991; Vanyo, 1993; Guyon et al., 2001; Greitzer et al., 2004; Tritton, 2005). Such an approach is standard for studying cars, airplanes and even weather patterns on earth (e.g. Anderson, 1991; Batchelor, 2000). Transformation of the NS equation by placing the reference frame on a moving object such as an airplane (Fig. 2) seems almost trivial through its common use, e.g. in wind tunnels. Transformations of coordinate systems are similarly helpful when studying the flow around a propeller (Fig. 2) or turbine blade (e.g. Du and Selig, 1998; Dumitrescu and Cardos, 2003). Such an approach simplifies the mathematical analysis of the boundary layer flow, which mediates the shear stress and pressure distribution on the surface, and therefore the net aerodynamic force and moment.
Scaling the NS equation in the reference frame attached to the fly wing
RESULTS
Basic kinematic model of a fly wing in forward flight
Visualizing rotational flow accelerations due to wing stroke
For many insects, including flies, the velocity and acceleration due to wing stroke are larger than the velocity and acceleration due to angle of attack variation, becauseΦ 0R>α0c holds. Using Eqns 11, 12, 13 we can draw and interpret the rotational accelerations that result from the wing stroke and act on the fluid near the wing (Fig. 4C). The first component aang is the manifestation of the angular acceleration of the wing around its base, which results locally in a chord-wise acceleration (Fig. 4C). The second term acen represents the centripetal acceleration, which is directed spanwise towards the wing's base(Fig. 4C). The third term aCor represents the Coriolis acceleration; its direction depends on the direction of local fluid velocity uloc(Fig. 4C). The precise directions of aang, acen and aCor for insect kinematics also depend on contributions from wing deviation and rotation. We omitted these additional contributions in Fig. 4C, because they are generally smaller than the contribution from stroke and we wanted to keep the figure simple enough to interpret. Both the centripetal and Coriolis accelerations (acen and aCor) are`quasi-steady' in that they depend on the instantaneous value of the angular velocity Ω of the wing (Eqns 12 and 13), in contrast to the angular acceleration (aang), which depends on the rate of change of angular velocity
Scaling rotational accelerations due to wing stroke
Graphical representation of dimensionless numbers
To better understand how the dimensionless numbers in the NS equation of a flapping wing relate to wing kinematics and morphology we represent them graphically. The advance ratio (Eqn 36) is equal to dimensionless wavelengthλ *=U∞/fc divided by the total wingtip excursion in the stroke plane 4A*(Fig. 5). The fly hovers for J=0 and flies forward (or descends) in an arbitrary direction when J>0. For γ=0°, which approximates fast forward and climbing flight, the advance ratio is a direct measure of the average pitch of a flapping wing. The average pitch is a measure of the average induced angle of attack of the flapping wing (Fig. 7) and determines together with the geometric angle of attack amplitude (with amplitude α0) the average effective angle of attack amplitude (Fig. 7). The effective angle of attack amplitude modulates wing lift and drag. For 0°<γ&≤90° (Fig. 5), the geometric interpretation of J is gradually modified, γ=90° being the extreme case. This case is relevant for slow hovering (Dickson and Dickinson,2004); under such conditions the advance ratio also measures how much the fly moves forward along its flight path compared with its total stroke length. The average induced angle of attack is, however, zero, because it is proportional to cos(γ).
The geometric representation of the dimensionless amplitude A* (Eqn 37)is shown in Fig. 5 and its geometric interpretation is simple. An equivalent dimensionless total amplitude Λ=2A*, has been defined by Ellington(Ellington 1984). The geometric interpretation of the single-wing aspect ratio(Eqn 38) is also straightforward and can be inferred from Fig. 3B by noting ARs=R/c=Rbs/S(in which bs is the single-wing span). Finally, Re can be interpreted as the ratio of convective versusdiffusive transport length for a fixed time interval. It measures how strongly the velocity boundary condition at the wing surface is diffused into the flow and is a measure of boundary layer thickness (e.g. Schlichting, 1979; Tennekes and Lumley, 1983)(Fig. 2).
How do scale factors of the rotational accelerations in Eqn 35 behave? The angular acceleration scales with 1/(J2+1)A*,which increases for decreasing A* at constant Jand increases for decreasing J at constant A*. When A*=0 a careful analysis of the product is needed,which shows that it will become zero (non-singular) provided that U≠0; there is flow. The analysis holds for the subsequent terms discussed below; they are also non-singular provided that U≠0. The centripetal and Coriolis accelerations scale with 1/(J2+1)ARs, which increases for decreasing ARs at constant J and increases for decreasing J at constant ARs.
Hovering flight
From flapping to spinning and translating fly wings
The single difference between propeller and turbine kinematics is that propellers operate at positive angles of attack generating forward pointing lift, which costs power, while turbines operate at negative angles of attack,which results in backward pointing lift and allows for the harvesting of power from wind (and water currents). The comparison of Eqn 41 and 35 shows that the dimensionless numbers and accelerations involved in the aerodynamics of flappers, propellers and turbines are indeed similar, provided that unsteady effects measured by A* do not dominate over `quasi-steady' rotational effects measured by ARs. By noting that A*=Φ0ARs (combining Eqns 37 and 38) for a fly wing and thatΦ 0 is typically close to one for insects(Ellington, 1984), we find that unsteady accelerations measured by 1/A* and`quasi-steady' accelerations measured by 1/ARs are of the same magnitude. This might explain the physical analogy between flapping and spinning insect wings and, possibly, propellers and turbines that operate at much higher Reynolds numbers. Propellers that operate at zero advance ratio, J=0, operate under hover conditions, such as hovering insects, which further simplifies Eqn 41.
2D pitch and heave wing kinematics
Simple vibrating fly wings
DISCUSSION
We derived a dimensionless form of NS equations for a 3D flapping fly wing to identify the dimensionless numbers that scale the underling physical mechanisms. This derivation shows that flapping wings induce three rotational accelerations: angular, centripetal and Coriolis in the air near to the wing's surface, which diffuse into the boundary layer of the wing. Next we simplified these equations incrementally using increasingly more restrictive assumptions. These simplifications allow us to easily interpret the dimensionless numbers geometrically for conditions that approximate both forward flight and hovering. In subsequent steps we derived the NS equation for spinning and translating 3D fly wings and for 2D flapping and vibrating wings.
Dimensionless template for parametric flapping wing studies
Whilst J and A* are measures of the wing's kinematics, ARs is a measure of single-wing morphology. Again we note that the effect of large values of these dimensionless numbers is that it diminishes the corresponding accelerations. The effect of forward flight (J>0) is, therefore, that it reduces the rotational accelerations. The rotational accelerations increase for smaller stroke amplitudes and single-wing aspect ratios at constant advance ratio. The extreme case is hovering flight (J=0) for which the rotational accelerations are maximal. Using the dimensionless scale variables J,A* and ARs we can now systematically vary the influence of rotational accelerations in parametric studies of the aerodynamics of fly wings in forward and hovering flight. There are, however,alternative ways to combine or split up the principal dimensionless numbers Cang, Ccen and Ro into other dimensionless factors, because they all depend on the same set of scale variables (length, time and mass). Alternatives are, for example, factors that more intuitively represent ratios of force or time scales. We chose to build up our three principal dimensionless numbers (Eqns 59, 60, 61) such that they correspond with the easy to interpret geometrical scales presented in Figs 5, 7 and 8. This approach facilitates an intuitive design of numerical and experimental studies of flapping wings with a direct link to the NS equations.
Comparing 3D and 2D wing kinematics
How important are rotational accelerations for understanding the aerodynamics of a fly? Here we present an alternative approach to compare 2D and 3D stroke kinematics, which proved to be pivotal for designing our experiment to test how important rotational accelerations (due to stroke) are for the stability of a fly's LEV (Lentink and Dickinson, 2009). This vortex allows the fly to generate high lift with its wing at angles of attack at which helicopter blades and airplane wings stall. There exists, however, an intriguing parallel between the lift augmentation due to the presence of a stable LEV on a fly wing and the lift augmentation found near the hub of wind turbine blades. Such blades are said to undergo `3D stall' or `stall delay' near their hub, which increases lift,whereas they undergo `2D stall' near the blade tip, which decreases lift (e.g. Tangler, 2004)(Fig. 9A). For wind turbines 3D stall is not observed beyond `local aspect ratios' of three(r/c>3, in which r is the local radius; see Fig. 9A) (e.g. Tangler, 2004). Three is approximately the value of a fruit fly's wing aspect ratio(Fig. 9), which might explain why flapping and spinning fly wings do not seem to stall and generate extraordinary high lift like wind turbines and propellers(Himmelskamp, 1947) do near their hub. To gain insight into the possible effect of rotational accelerations we gradually transform rotational stroke kinematics into translational stroke kinematics (Fig. 9B). We do this by letting Φ0→0 and R→∞ such that the wing amplitudeΦ 0R=A remains constant within the limit. In doing so we do not change the wing's geometry, we simply place the same wing farther outward as shown in Fig. 9. Because the wing's radial distance from the center of rotation R goes to infinity (R→∞), the single-wing aspect ratio based on this radial distance ARs=R/c also goes to infinity. The NS equations of a flapping (Eqn 35)and spinning wing (Eqn 41) show that the centripetal and Coriolis acceleration go down with aspect ratio. We have performed exactly this experiment to show that rotational accelerations mediate LEV stability in hovering insect flight(Lentink and Dickinson, 2008). The similar importance of rotational accelerations (e.g. for LEV stability)for flapping and spinning wings are confined to the midstroke phase of the flapping cycle when the flapping wing revolves with a propeller-like swing at nearly constant angular acceleration. This phase corresponds with the maximum dynamic pressure due to wing stroke and therefore dominates lift production for most insects. In practice this analogy is complicated by the effects of significant stroke acceleration, wing deviation and wing rotation at the beginning and end of the stroke (e.g. Dickinson et al., 1999; Altshuler et al., 2005). These significant complicating factors do not, however, modify the analogy between the flow physics and forces we (Lentink and Dickinson, 2009) and others (e.g. Dickinson et al., 1999; Usherwood and Ellington, 2002)found at midstroke for flapping and spinning insect wings.
Rotational accelerations due to wing stroke versus angle of attack kinematics
The angular velocity of flapping wings consists of two significant components. (1) Angular velocity due to wing stroke and (2) angular velocity due to geometric angle of attack variation. Angular velocity due to wing stroke is approximately maximal at midstroke, while angular velocity due to wing angle of attack variation is approximately maximal at stroke reversal. The resulting velocity magnitudes are proportional toΦ 0Rf for stroke and α0cf for angle of attack. The ratio of the two velocities isΦ 0R/α0c<R/cbecause Φ0<α0 for most insects. For aspect ratios significantly larger than one we can assume, therefore, that rotational accelerations due to stroke dominate those due to angle of attack variation. One can test experimentally whether this holds by doing the experiments with and without rotational stroke kinematics (reciprocating revolving versus reciprocating translating wings) for all relevant angle of attack amplitudes, which lie in the range 0°≤α0≤90°(Lentink and Dickinson,2009).
Scaling the NS equation more accurately
Relationship between existing and present dimensionless numbers
Some of the dimensionless numbers that we derived here can be related to existing dimensionless numbers. The advance ratio J is already in use(Ellington, 1984). It is equivalent to the inverse of the amplitude-based Strouhal number as discussed in the Introduction. We prefer the advance ratio because it is readily interpretable geometrically and commonly used as such in aeronautics and biological fluid dynamics. Further, the Strouhal number is perhaps best reserved for its original purpose – characterizing natural shedding frequencies. The dimensionless amplitude A* is a normal dimensionless variable, 2A* represents the total dimensionless amplitude Λ introduced by Ellington(Ellington, 1984); we prefer A*, because it represents the mathematical definition of amplitude. We further prefer the dimensionless wavelength λ*over the inverse, the reduced frequency k, becauseλ * is a length scale ratio that can easily be drawn and interpreted graphically, e.g. Fig. 5, while the time scale ratio k cannot. The importance of the dimensionless single-wing aspect ratio for calculating rotational accelerations is new in the field of insect flight. The corresponding Rossby number is, however, commonly used in the analyses of rotating fluids(Rossby, 1936; Vanyo, 1993; Greitzer et al., 2004; Tritton, 2005). The inverse of the single-wing aspect ratio, c/r, is in use in the wind turbine literature (e.g. Lindenburg,2004), but we prefer r/c because it corresponds to the single-wing aspect ratio which is easier to interpret geometrically for animal flight, and because r/c corresponds with the definition of Rossby number, which is in use in the much more elaborate literature on rotational flows (compared with wind turbines). Finally, there is the Reynolds number; our definition has the advantage that it works continuously from hovering flight to fast forward flight(Eqn 39)(Lentink and Gerritsma,2003).
Application of dimensionless numbers in wing and fin studies
Our derivation of the dimensionless NS equation for flapping, revolving and translating fly wings and airfoils represent the various 3D and 2D insect flight models in the literature; from 3D flapping wings to 2D vibrating wings. By comparing Eqn 35 for a 3D flapping wing and Eqn 47 for a 2D flapping wing we conclude that the significant rotational accelerations due to the flapping motion in 3D fly aerodynamics are neglected in 2D models. The possible importance of such differences is amplified by the experimental observation that 3D fly wings that either spin (e.g. Usherwood and Ellington, 2002)or flap (e.g. Birch et al.,2004; Ellington et al.,1996) around their base generate a stable LEV, while flapping 2D airfoils (e.g. Dickinson and Goetz,1993; Dickinson,1994; Lentink et al.,2008) do not. We further note that 2D vibrating insect wing models neglect all rotational accelerations (e.g. Wang, 2000b; Lentink and Gerritsma, 2003). The above insect flight models therefore increasingly incorporate the rotational accelerations induced in the flow due to the rotational kinematics of the wing motion.
2D and 3D models similar to those of insect flight have also been used to study the aerodynamics of birds (Hubel,2006) and the hydrodynamics of the fins of fish(Triantafyllou et al., 1993; Bandyopadhyay et al., 2008). The dimensionless NS equation we derived for a flapping fly wing also represents such models, provided that care is taken that the assumptions used to derive the various equations hold. For completeness we repeat three non-trivial assumptions: (1) the fluid behaves in a fully Newtonian way like water and air, (2) the fluid does not cavitate, and (3) the amplitudes of undulations in the body are small compared with stroke amplitude. The flow can, however, be turbulent.
Finally we note that even the abstract problem of spinning fly wings, Eqn 41, is of direct relevance for biological fluid dynamics studies, because autorotating seeds spin with exactly such kinematics as they swirl down to earth(Azuma and Yasuda, 1989). While doing so, autorotating seeds extract energy from the flow very much like wind turbines harvest energy from wind at much higher Reynolds numbers. This intriguing example illustrates that our theoretical approach has the potential to provide a valuable link between the fluid dynamics of translating, rotating and flapping wings and fins in nature and technology.
APPENDIX I
Contribution of body versus wing velocity to Rossby number
APPENDIX II
Derivation of dimensionless numbers
3D kinematics
2D kinematics
LIST OF ABBREVIATIONS
- α
wing angle of attack
- \({\dot{{\alpha}}}\)
1st time derivative of wing angle of attack
- \({\ddot{{\alpha}}}\)
2nd time derivative of wing angle of attack
- α0
wing angle of attack amplitude
- αeff
effective angle of attack
- αind
induced angle of attack
- αgeo
geometric angle of attack
- β
stroke plane angle
- γ
angle between the stroke plane and direction of flight
- Δr
width spanwise wing section
- λ
wingbeat wavelength
- λ*
dimensionless wavelength
- λ*0
natural von Kármán vortex shedding wavelength
- Λ
dimensionless total stroke amplitude
- μ
dynamic viscosity
- ν
kinematic viscosity
- ξ
flight path angle with respect to horizon
- ρ
fluid density
- \({\varphi}\)
wing deviation with respect to stroke plane
- ϕ
wing stroke angle
- \({\dot{{\phi}}}\)
1st time derivative of wing stroke angle
- \({\ddot{{\phi}}}\)
2nd time derivative of wing stroke angle
- Φ
total wing amplitude in radians
- Φ0
wing stroke amplitude (half the total stroke amplitude Φ)
- Ω
absolute time-averaged angular velocity amplitude
- \({\dot{{\Omega}}}\)
absolute time-averaged angular acceleration amplitude
- Ω
angular velocity of the rotating frame, and wing
- \({\dot{{\Omega}}}\)
angular acceleration of the rotating frame, and wing
- Ωwing
amplitude of vector Ωwing
- Ωwing
angular velocity of the fly wing
- *
dimensionless variable scaled with its order of magnitude
\({\nabla}=\left[\begin{array}{c}{\partial}{/}{\partial}\mathrm{x}\\{\partial}{/}{\partial}y\\{\partial}{/}{\partial}z\end{array}\right]\)gradient (del) operator - aang
angular acceleration
- acen
centripetal acceleration
- aCor
Coriolis acceleration
- ainert
acceleration with respect to inertial coordinate system
- aloc
acceleration with respect to local coordinate system
- awing
linear acceleration of an airfoil
- A
flap amplitude
- A*
stroke amplitude
- ARs
single-wing aspect ratio
- bs
single-wing span
- c
average wing (or foil) chord length
- Cang
angular acceleration number
- Ccen
centripetal acceleration number
- CFD
computational fluid dynamics
- D*
dimensionless circular distance moved during one full period(propeller)
- D/Dt
total differential operator:
\(\frac{D}{Dt}=\frac{{\partial}}{{\partial}t}+\mathbf{u}.{\nabla}\) - Eu
Euler number
- f
flap frequency
- g
gravitational vector
- J
advance ratio
- k
reduced frequency
- LEV
leading edge vortex
- n
multiple (of λ*0)
- NS
Navier–Stokes
- O
order of magnitude operator
- p
pressure
- p0
ambient atmospheric pressure
- r
magnitude of radius vector
- r
position of a fluid particle in the rotating frame
- R
wing radius
- Rg
wing radius of gyration
- Re
Reynolds number
- Reb
Reynolds number component due to body speed
- Res
Reynolds number component due to wings stroke,
- Ro
Rossby number
- s*
dimensionless linear distance moved through air during one full period(propeller)
- sg
number of chord lengths traveled at Rg during a full stroke (Fig. 9)
- \({\ddot{s}}_{\mathrm{wing}}\)
magnitude of linear acceleration of the wing
- S
single-wing area
- Sfb
outer surface of fly's body
- Sob
outer boundary surface
- St
Strouhal number
- Stc
chord-based Strouhal number
- t
time
- u
velocity vector
- ubody
velocity center of gravity of fly
- ufly
velocity of fly at its outer surface
- uinert
velocity with respect to inertial coordinate system
- uloc
velocity in local coordinate system
- un
component of wingtip speed normal to flight direction
- ut
component of wingtip speed in flight direction
- uα
linear velocity distribution due to angle of attack variation
- uϕ
linear velocity distribution due to stroke
- u\({\varphi}\)
linear velocity distribution due to deviation
- U
characteristic speed; absolute time-averaged speed of the wingtip
- Ubody
amplitude of vector ubody
- U̇ wing
absolute time-averaged linear acceleration of the wing
- U∞
forward flight speed (arbitrary direction with respect to gravity)
- V
velocity along wing radius (Fig. 9)
- x
position vector
- \({\ddot{x}}_{\mathrm{wing}}\)
x-component of awing
- (x, y, z)
local coordinate system
- (X, Y, Z)
inertial coordinate system
- \({\ddot{x}}_{\mathrm{wing}}\)
y-component of awing
FOOTNOTES
We gratefully acknowledge Will Dickson, Koert Lindenburg and John Dabiri for proof reading preliminary versions of the manuscript. We thank Thomas Daniel for proof reading the final manuscript. We thank Johan van Leeuwen and GertJan van Heijst for hearty support, encouragement and proof reading of the various versions of the manuscript. This research is supported by NWO-ALW grant 817.02.012to D.L. and by NSF grant IBN-0217229 to M.H.D.