Computational Analysis of Hovering Hummingbird Flight

Similar documents
Computational Analysis of Hovering Hummingbird Flight

Analysis of a Hinge-Connected Flapping Plate with an Implemented Torsional Spring Model

Unsteady aerodynamic forces of a flapping wing

Two-Dimensional Aerodynamic Models of Insect Flight for Robotic Flapping Wing Mechanisms of Maximum Efficiency

A flow control mechanism in wing flapping with stroke asymmetry during insect forward flight

NUMERICAL SIMULATION OF SELF-PROPELLED FLYING OF A THREE-DIMENSIONAL BIRD WITH FLAPPING WINGS

SENSITIVITY ANALYSIS OF THE FACTORS AFFECTING FORCE GENERATION BY WING FLAPPING MOTION

When vortices stick: an aerodynamic transition in tiny insect flight

(This is a sample cover image for this issue. The actual cover is not yet available at this time.)

The wings and the body shape of Manduca sexta and Agrius convolvuli are compared in

A COMPUTATIONAL FLUID DYNAMICS STUDY OF CLAP AND FLING IN THE SMALLEST INSECTS. Laura A. Miller* and Charles S. Peskin**

Lift and power requirements of hovering flight in Drosophila virilis

International Journal of Micro Air Vehicles

Smart wing rotation and ingenious leading edge vortex control modulate the unconventional forces during insects flights

Effects of unsteady deformation of flapping wing on its aerodynamic forces

RESEARCH ARTICLE Aerodynamic effects of corrugation in flapping insect wings in hovering flight

I nsect flight has long been considered a paradox, defeating the expectations of steady state aerodynamics. We

A computational fluid dynamics of clap and fling in the smallest insects

Many of the smallest flying insects clap their wings together at the end of each upstroke

Lift Enhancement by Dynamically Changing Wingspan. in Forward Flapping Flight (09/10/2013)

Vortex Dynamics in Near Wake of a Hovering Hawkmoth

A Biologically Inspired Computational Study of Flow Past Tandem Flapping Foils

Near-Hover Dynamics and Attitude Stabilization of an Insect Model

The Interaction of Wings in Different Flight Modes of a Dragonfly

An Experimental Investigation on the Wake Flow Characteristics of Tandem Flapping Wings

Dynamic flight stability of a hovering bumblebee

Three-dimensional flows around low-aspect-ratio flat-plate wings at low Reynolds numbers

Fig. 1. Bending-Torsion Foil Flutter

Flexible clap and fling in tiny insect flight

COMPUTATIONAL STUDY ON THE INFLUENCE OF DYNAMIC STALL ON THE UNSTEADY AERODYNAMICS OF FLAPPING WING ORNITHOPTER

ABSTRACT. travels through the wing reducing the drag forces generated. When the wings are in

On the Aerodynamic Performance of Dragonfly Wing Section in Gliding Mode

COMPUTATIONAL MODELING OF UNSTEADY AERODYNAMICS IN HUMMINGBIRD FLIGHT

Optimization of Flapping Airfoils for Maximum Thrust and Propulsive Efficiency I. H. Tuncer, M. Kay

Modeling of Instantaneous Passive Pitch of Flexible Flapping Wings

Dynamic pitching of an elastic rectangular wing in hovering motion

Wing Kinematics in a Hovering Dronefly Minimize Power Expenditure

Rotational lift: something different or more of the same?

PIV and force measurements on the flapping-wing MAV DelFly II

We are IntechOpen, the world s leading publisher of Open Access books Built by scientists, for scientists. International authors and editors

Dynamic flight stability of a hovering model insect: lateral motion

Effect of Pivot Point on Aerodynamic Force and Vortical Structure of Pitching Flat Plate Wings

Modeling of Pitching and Plunging Airfoils at Reynolds Number between and

THE CONTROL OF FLIGHT FORCE BY A FLAPPING WING: LIFT AND DRAG PRODUCTION

Jun 22-25, 2009/San Antonio, TX

APPLICATION OF ARTIFICIAL NEURAL NETWORK IN MODELING OF ENTOMOPTER DYNAMICS

Effects of Unequal Pitch and Plunge Airfoil Motion Frequency on Aerodynamic Response

Active Control of Separated Cascade Flow

Coupling of the wings and the body dynamics enhances damselfly maneuverability

Relationship between Unsteady Fluid Force and Vortex Behavior around a Discoid Airfoil Simulating a Hand of Swimmer

OSCILLATING AERO-WING MODEL IN THE QUASI-STEADY DOMAIN A REFERENCE FOR SUSTAINED ANIMAL FLIGHT AND MICRO AIR VEHICLES

The manipulation of trailing-edge vortices for an airfoil in plunging motion

Unsteady Flow and Aerodynamic Effect of a Dynamic Trailing-Edge Flap in Flapping Flight

IN-LINE MOTION CAUSES HIGH THRUST AND EFFICIENCY IN FLAPPING FOILS THAT USE POWER DOWNSTROKE

arxiv: v1 [physics.flu-dyn] 27 Mar 2014

Research article The implications of low-speed fixed-wing aerofoil measurements on the analysis and performance of flapping bird wings

Enclosure enhancement of flight performance

The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight

Low-dimensional state-space representations for classical unsteady aerodynamic models

Flapping Wing Micro Air Vehicles: An Analysis of the Importance of the Mass of the Wings to Flight Dynamics, Stability, and Control

REPORT DOCUMENTATION PAGE

Some effects of large blade deflections on aeroelastic stability

Energy-minimizing kinematics in hovering insect flight

Title: Aerodynamics characteristics of butterfly flight through measurement of threedimensional unsteady velocity field using TR-PIV system

Fluid transport and coherent structures of translating and flapping wings

STUDY OF THREE-DIMENSIONAL SYNTHETIC JET FLOWFIELDS USING DIRECT NUMERICAL SIMULATION.

Aeroelastic Analysis Of Membrane Wings

Effects of Flexibility on the Aerodynamic Performance of Flapping Wings

A wing characterization method for flapping-wing robotic insects

AN ABSTRACT OF THE THESIS OF

Computational Aerodynamics of Low Reynolds Number Plunging, Pitching and Flexible Wings for MAV Applications

THE MECHANICS OF FLIGHT IN THE HAWKMOTH MANDUCA SEXTA

Study of design parameters of flapping-wings

Motion Kinematics vs. Angle of Attack Effects in High-Frequency Airfoil Pitch/Plunge

The effect of aerodynamic braking on the inertial power requirement of flapping flight: case study of a gull

Computational Fluid-Structure Interaction of a Deformable Flapping Wing for Micro Air Vehicle Applications

Force production and flow structure of the leading edge vortex on flapping wings at high and low Reynolds numbers

Biologically Inspired Design Of Small Flapping Wing Air Vehicles Using Four-Bar Mechanisms And Quasi-steady Aerodynamics

ScienceDirect. Experimental Validation on Lift Increment of a Flapping Rotary Wing with Boring-hole Design

Low-dimensional state-space representations for classical unsteady aerodynamic models

Unsteady Force Generation and Vortex Dynamics of Pitching and Plunging Flat Plates at Low Reynolds Number

Dynamics and Energy Extraction of a Surging and Plunging Airfoil at Low Reynolds Number

Object-Oriented Unsteady Vortex Lattice Method for Flapping Flight

Computation of Inertial Forces and Torques Associated with. Flapping Wings

K. M. Isaac, P. Shivaram + University of Missouri-Rolla Rolla, MO T. DalBello NASA Glenn Research Center

DISSECTING INSECT FLIGHT

INSTITUTE OF AERONAUTICAL ENGINEERING (Autonomous) Dundigal, Hyderabad

Experimental and computational investigations of flapping wings for Nano-air-vehicles

RESEARCH ARTICLE Aerodynamics of tip-reversal upstroke in a revolving pigeon wing

AFRL-VA-WP-TM

Viscous investigation of a flapping foil propulsor

Thrust and Efficiency of Propulsion by Oscillating Foils

State-Space Representation of Unsteady Aerodynamic Models

Entomological Analysis of the Popular Press News Article: Scientists Finally Figure Out How Bees Fly

An experimental study of the vortex structures in the wake of a piezoelectric flapping plate for Nano Air Vehicle applications

Masters in Mechanical Engineering Aerodynamics 1 st Semester 2015/16

A STUDY OF THE INSTANTANEOUS AIR VELOCITIES IN A PLANE BEHIND THE WINGS OF CERTAIN DIPTERA FLYING IN A WIND TUNNEL

Numerical Analysis of Unsteady Viscous Flow through a Weis-Fogh-Type Water Turbine

ABSTRACT PARAMETRIC INVESTIGATIONS INTO FLUID-STRUCTURE INTERACTIONS IN HOVERING FLAPPING FLIGHT. Jesse Maxwell, 2013

Stall Suppression of a Low-Reynolds-Number Airfoil with a Dynamic Burst Control Plate

Transcription:

48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition 4-7 January 2010, Orlando, Florida AIAA 2010-555 Computational Analysis of Hovering Hummingbird Flight Zongxian Liang 1 and Haibo Dong 2 Department of Mechanical & Materials Engineering, Wright State University, Dayton, OH 45435 Mingjun Wei 3 Department of Mechanical and Aerospace Engineering, New Mexico State University, Las Cruces, NM 88003 Different from large-size birds, hummingbirds are found to share more common flight patterns with insects, especially in hovering motion. Comparing to hovering insects, hummingbird wing kinematics uses apparent asymmetric motions during downstroke and upstroke. In current study, we present a direct numerical simulation of modeled hummingbird wings undergoing hover flight (Tobalske et al. JEB 2007). 3D wake structures and associated aerodynamic performance are of particular interests in this paper. Computational results are also compared with PIV experiments (Warrick et al. 2005 ). NOMENCLATURE C φ The ratio of time of downstroke to a stroke φ Stroke angle, deg φ m Azimuthal amplitude, deg α c Chord angle, deg γ Stroke plane angle relative to the horizontal plane, deg β Body angle, deg η Pitching angle, deg θ Heaving angle, deg T Time of a whole wing stroke, sec U Average tip speed, m/s c mean chord length, m C L Force coefficient in y direction C X Force coefficient in x direction Force coefficient in z direction C Z Introduction Being a small-scale bird, hummingbird is found to share more common flight patterns with insects (Weis-Fogh 1972) other than large-scale birds especially in hovering motion. The Reynolds number for hummingbird hover flight ranges from 10 3 to 10 4 (Altshuler et al. 2004) and flapping frequency is about 41Hz (Tobalske et al. 2007). Thus, hummingbird is widely thought to employ aerodynamic mechanisms similar to those used by insects in spite of the differences in 1 Ph.D Student, liang.5@wright.edu. 2 Assistant Professor, AIAA Senior Member, haibo.dong@wright.edu. 3 Assistant Professor, AIAA Member, mjwei@nmsu.edu. 1 Paper 2010-0555 Copyright 2010 by the, Inc. All rights reserved.

musculoskeletal system. Using particle image velocimetry (PIV) techniques, researchers have attempted to provide visual details into the essential of unsteady fluid dynamics of hummingbird (Warrick et al. 2005, Altshuler et al. 2009). It is found that hovering hummingbird produces 75% of total lift in downstroke and 25% in upstroke of wing motion. Whereas in insects hover flight, the lift is commonly found to be produced approximately evenly in both strokes (Dickinson et al. 1999). Moreover, leading-edge vortex (Ellington et al. 1996), which is observed in both downstroke and upstroke of insect hover flight (Dickinson et al. 1999, Sane and Dickinson 2001), however, is not seen either at the beginning or the end of the upstroke but seen in the full downstroke (Warrick et al. 2005). One explanation is that hummingbirds have higher angle of attack at mid-downstroke than during mid-upstroke and only have partially inversion of wing camber at distal portion of wing during upstroke (Warrick et al. 2005). However, it is still lack of comprehensive computational or experimental work to reveal the 3D unsteady vortex formations of this complicated motion. In this paper, we describe a direct numerical simulation of modeled hummingbird wings undergoing hover flight based on datum from Tobalske et al. 2007, to investigate the complete details of three-dimensional vortex topologies and associated aerodynamic performance. Results will also be compared with Warrick et al. 2005. Results A second-order finite-difference based immersed-boundary solver (Dong et al. 2006, Mittal et al. 2008) has been developed which allows us to simulate flows with complex immersed 3-D moving bodies. The method employs a second-order central difference scheme in space and a second-order accurate fractional-step method for time advancement. Code validations and details of numerical methodologies can be found in Mittal et al. 2008. In this section, a sequence of numerical simulations that explore the wake structures and associated aerodynamic performance of modeled hummingbird wings in hovering flight are presented. Model configuration and kinematics (Tobalske et al. 2007) approach are discussed first. Following this, the wake topology due to prescribed kinematics and aerodynamic force production are examined. A. Wing Configuration and Prescribed Wing Kinematics stroke plane (a) (b) (c) Figure 1: (a): wing shape outline; (b): sketch of chord angle from lateral view, where dash-dot line is lateral midline of hummingbird body, solid line is stroke plane, and the dot-head solid line is wing chord during downstroke; (c): coordinate system for controlling wing motions, where XYplane is stroke plane. A half elliptic wing analogical to hummingbird wing as shown in Figure 1(a) is employed 2

in this paper. Wing aspect-ratio is 9.1, which is comparable to average hummingbird wing aspectratio 9 (Tobalske et al. 2007). The pitching axis is through quarter chord and perpendicular to minor axis of the wing as shown in Figure 1(c). Simplified kinematics described in Tobalske et al. 2007 for hover flight was implemented and controlled by the parameters in Figure 1(b) and 1(c). Especially, the stroke angle is defined as πt φm cos( ) downstroke CT φ φ() t = tt Cφ φm cos( π( + 1)) upstroke 1 Cφ φ m where = 54.5 o is the azimuthal amplitude, C φ = 0.48 is the ratio of the time spending in downstroke to that of a full stroke. Following the definitions in Tobalske et al. 2007, stroke plane angle and body angle relative to the horizontal plane are γ and β (Figure 1), respectively. Chord angle (α c ) is reconstructed from the first eight terms of Fourier series of chord angle curve (Tobalske et al. 2007) in the form of 8 k = 1 ( ) ( ) α = ak ( )cos 2 πkt + bk ( )sin 2πkt c Where a(k) and b(k) are Fourier coefficients. As shown in Figure 1(b), pitching angle in laboratory coordinate system is defined as α c +γ+β. The Reynolds number was defined as Re = Uc ν (with average tip speed U, mean chord length c, and kinematic viscosity ν of the fluid). In the simulation, Re number is set to 3000. Strouhal number, St = fa U (Taylor et al. 2003), which is based on wingtip excursion A ( A= bsinφ m, where b is span), wing stroke frequency, and average tip speed, is 0.214 for this hover flight. Figure 2: Left: Downstroke; Center: Upstroke; Right: Stroke and chord angles as a function of dimensionless time (t/t), where α c follows the definition in [16], shadow part for upstroke. The wingbeat begins with downstroke then upstroke. The downstroke is nearly sinusoidal shape (Figure 2), whereas the upstroke consists of initial pitching-down rotation with translational acceleration, followed by translation with chord angle gradually decreasing, and finally pitchingup rotation with translational deceleration. Based on domain independence studies and grid size independent studies, a domain size of 30 30 30 and a 201 121 237 grid has been finally chosen for all simulations. 3

B. Analysis of 3D Flapping Flight LEV TPV TV VR2 VR1 (a) Force coefficients history, shadow part (b) Vortex structure at the end of upstroke for upstroke Figure 3. Force coefficients history and vortex structure at the end of upstroke The force coefficients (defined by C = F 2, where S ( 1 ) wing is the wing area ρu S 2 wing and F is the force) history for the fourth cycle, at which the flow field is thought to be steady, is shown in Figure 3(a) and average forces in three directions for half-stroke and entire stroke are tabulated in Table 1. It can be found in Table 1 that the amount of lift generated during downstroke is 2.95 times of that generated during upstroke. This matches with the results of Warrick et al. 2005, which concluded that the amount of lift generation in downstroke is 3 times of that in upstroke for hummingbird hover flight. Moreover, the average lift is one order of magnitude larger than the average forces in other two directions. Table 1: Comparison of C L, C X and C Z of right wing C L C X C Z Downstroke 0.932 0.387 0.273 Upstroke 0.316-0.264-0.120 Total Avg. 0.623 0.060 0.076 Figure 3(b) shows a 3-D perspective view of the wake topologies at the end of upstroke ui by using isosurfaces of the eigenvalue imaginary part of the velocity gradient tensor. Two x j vortex rings, VR1, VR2 are circularly generated by trailing edge of the wing and the wing tip vortex during downstroke. These are similar to the discussion of Figure 9 in Altshuler et al. 2009. Strong downwash is expected to be induced by these two vertically oriented rings. Leading edge vortex (LEV), tip vortex (TPV), and trailing edge vortex (TV) are also clearly shown in the plot and connected to the vortex rings. In order to qualitatively understand the wake structures from computational results, we use two cross-section views to compare with the PIV results in Warrick et al. 2005. 4

U1 U2 U TPV D1 D2 LEV D a. Dorsal cross-section b. Lateral cross-section Figure 4. Two cross-section views of Figure 3(b) Figure 4 shows two cross-sections of wake structures in Figure 3(b). Figure 4(a) is a slice cut in YZ-plane parallel to the X-axis near the root of wing and Figure 4(b) is a cross-section cut near the half-span length of wing. Comparing to Figure 2 in Warrick et al. 2005, vortices U1 and D1 in Figure 4(a) are in accordance with the vortices U and D in Warrick et al. 2005 respectively. Vortex U2 and D2 can also be found in Warrick et al. 2005. Here, vortices D1 and D2 form the downstroke vortex ring VR2 as seen in Figure 3(b). In Figure 4(b), both tip vortex (TPV) and slice of vortex ring VR2 can be found in the Figure 2 of Warrick et al. 2005. To better understand how the leading edge vortex develops and sheds during the wing flapping, several key frames are selected at the corresponding time points marked in Figure 3(a) (P A to P H ). The time frame P A and P B are the time frames before and after the mid-downstroke. The leading edge vortex can be clearly seen at these two time frames (Figure 5(a) and Figure 5(b)). During this period, the lift coefficient keeps increasing along with the geometrical angle of attack until the stall angle is reached at 62, where the delayed stall is stopped. The frame P C is the inflection point from downstroke to upstroke. After this point, the wing starts to move upward and the shed trailing edge vortex can be seen at P D. Meanwhile, the strength of the leading edge vortex is weaken and totally diminishes at P E. This agrees with the observation in Warrick et al. The frame P F (t/t=0.8) is right after the mid-upstroke. The leading edge vortex begins to form again due to the increase of the geometrical angle of attack. As a result, lift coefficient begins to increase in the upstroke. The wing begins the reversal rotation again at P F. This causes local minimum point between P F and P G in the lift coefficient curve as shown in Figure 3(a). The continuously enhancing leading edge vortex delayed the stall (Figure 6) but finally the stall dominates the tendency of the curve until the wings are almost perpendicular in laboratory coordinate system and next downstroke is about to begin. Summary A direct numerical simulation has been conducted to investigate wake structures of hummingbird hovering flight and associated aerodynamic performance. It's found that the amount 5 VR2

of lift produced during downstroke is about 2.95 times of that produced in upstroke. Two parallel vortex rings are formed at the end of the upstrokes. There is no obvious leading edge vortex can be observed at the beginning of the upstroke. Results are in good agreement with PIV experiments in Warrick et al. 2005 and Altshuler et al. 2009. a. P A, t/t=0.2 b. P B, t/t=0.3 c. P C, t/t=0.5 d. P D, t/t=0.6 e. P E, t/t=0.7 f. P F, t/t=0.8 Figure 5. Slice views of iso-surface for 6 frames, following time sequence 6

a. P F, t/t=0.8 b. P G, t/t=0.9 c. P H, t/t=0.975 Figure 6. Perspective views of vortex structure during the upstroke, following time sequence Acknowledgement This is work is supported under DAGSI Student Fellowship and Wright State University Research Challenge grant. References 1 Altshuler D. L., Dudley, R., and Ellington, C. P., (2004), Aerodynamic forces of revolving hummingbird wings and wing models, J. Zool., Lond. 264, 327 332 2 Altshuler D. L., Princevac M., Pan H., Lozano J., (2009), Wake patterns of the wings and tail of hovering hummingbirds, Exp Fluids 46:835 846 3 Aono, H., Liang, F. and Liu, H. (2007). Near- and far-field aerodynamics in insect hovering flight: an integrated computational study. J. Exp. Biol. 211, 239-257. 4 Azuma, A., (2006), The Biokinetics of Flying and Swimming, Published by AIAA 5 Dickinson, M.H., Lehman, F. O. and Sane, S.P. (1999) Wing rotation and the aerodynamic basis of insect flight, Science, 284(5618): 1954-1960. 6 Dong, H., Mittal, R., and Najjar, F., 2006 Wake Topology and Hydrodynamic Performance of Low Aspect-Ratio Flapping Foils, Journal of Fluid Mechanics, 566:309-343, 2006. 7 Ellington, C. P. (1984), The Aerodynamics of Hovering Insect Flight. IV. Aerodynamic Mechanisms. Phil. Trans. R. Soc. Lond. B Vol. 305, pp 79-113. 8 Ellington, C. P. (1999), The novel aerodynamics of insect flight: Applications to micro-air vehicles. The Journal of Experimental Biology. Vol. 202, pp 3439-3448. 9 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. 10 Lehmann, F. O., Sane, S. P., and Dickinson, M. H. (2005), The aerodynamic effects wing-wing interaction in flapping insect wings, Journal of Experimental Biology, 208: 3075-3092. 7

11 Liang Z., Dong H., Wei M., 2009, Unsteady Aerodynamics and Wing Kinematics Effect in Hovering Insect Flight, AIAA-2009-1299 12 Mittal, R., Dong, H., Bozkurttas, M., Najjar, F. M., Vargas, A., and von Loebbecke A., (2008), A versatile sharp interface immersed boundary method for incompressible flows with complex boundaries, Journal of Computational Physics, 227:4825-4852. 13 Poelma C., Dickson, W.B. and Dickinson, M. H. (2006), Time-resolved reconstruction of the full velocity field around a dynamically-scaled flapping wing, Experiments in Fluids, 41: 213-225. 14 Sane, S. P., Dickinson, M. H., (2001), The control of flight force by a flapping wing: lift and drag production, The Journal of Experimental Biology 204, 2607-2626. 15 Taylor, G.. K., Nudds, R. L., and Thomas, A. L. R., 2003, Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency, Nature Vol. 425. 16 Tobalske B. W., Warrick D. R., Clark C. J., Powers D. R., Hedrick T. L., Hyder G. A. and Biewener A. A., 2007, Three-dimensional kinematics of hummingbird flight, The Journal of Experimental Biology 210, 2368-2382 17 Warrick, D. R., Tobalske, B. W., and Powers, D. R. (2005) Aerodynamics of the hovering hummingbird, Nature, Vol. 435, No. 23, pp. 1094-1097. 18 Weis-Fogh, T. (1972). Energetics of Hovering Flight in Hummingbirds and in Drosophila. J. Exp. Biol. 56, 79-104. 19 Weis-Fogh, T. (1973). Quick estimates of flight fitness in hovering animals, including novel mechanisms for lift production. J. Exp. Biol. 59, 169-230. 8