This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Scaling laws for the sound generation of bio-inspired flapping wings during hovering flight

Li Wang\aff1    Xueyu Ji\aff1    John Young\aff1    Hao Liu\aff2       Fang-Bao Tian\aff1 \corresp [email protected] \aff1School of Engineering and Technology, University of New South Wales, Canberra, ACT 2600, Australia \aff2Graduate School of Engineering, Chiba University, Chiba 263-8522, Japan
Abstract

Bio-inspired flapping wings have been extensively studied for their remarkable aerodynamic performance. Recently, their noise emission has attracted growing interest, but a careful analysis of scaling laws for their sound generation is missing. This work presents scaling laws for the sound generation of bio-inspired flapping wings during hovering flight based on the quasi-steady flow theory and Ffowcs Williams-Hawkings analogy. Direct numerical simulations considering a range of parameters including the Reynolds number, Mach number and wing kinematics confirm that the proposed scaling laws capture the major physics involved and their predictions agree well with the numerical results. The scaling laws can be used as a powerful tool for engineers in the design of micro-aerial vehicles considering both aerodynamics and acoustics performances simultaneously.

1 Introduction

Bio-inspired flapping wings, as an important way of aquatic and aerial animals moving through air and water, have been shown to have superior aerodynamic and hydrodynamic performance in comparison to the stationary wings in the low Reynolds number (O(103)O(10^{3}) or lower) regime (see e.g., Lighthill, 1975; Sane, 2003; Wang, 2005; Wu, 2011; Shyy et al., 2013; Eldredge & Jones, 2019; Liu et al., 2024). The superior performance is achieved through several unique mechanisms such as delayed stall of leading edge vortex (LEV) (Ellington et al., 1996), wake capture (Dickinson et al., 1999; Bomphrey et al., 2017), rapid pitching rotation (Bennett, 1970; Dickinson et al., 1999), unsteady wing-tip vortex (Shyy et al., 2010), synchronised passive pitching deformation (Shyy et al., 2010; Nakata & Liu, 2012a, b; Dai et al., 2012; Tian et al., 2013; Huang et al., 2023), and clap and fling (Weis-Fogh, 1973; Lighthill, 1973). For aerial animals and bio-inspired aerial vehicles, the noise generated by flapping wings is also an important factor measuring the aerodynamic performance, as discussed in a few recent studies (Geng et al., 2017; Wang & Tian, 2019, 2020; Jaworski & Peake, 2020; Clark, 2021). The study of the sound generation of bio-inspired flapping wings would shed light on the sound control of bio-inspired aerial vehicles.

One of the most fascinating features of bio-inspired flapping wings is the high lift generation at low Reynolds numbers (e.g. O(102)O(10^{2}) to O(103)O(10^{3})) by using unsteady flow mechanisms (Shyy et al., 2010, 2013; Eldredge & Jones, 2019; Liu et al., 2024). The unsteady flow mechanisms are realised through the complicated kinematics which could be one or the combination of stroke, pitching, deviating, twisting and revolving components (Bomphrey et al., 2017). Different components contribute to the aerodynamic force generation in different ways (Shyy et al., 2010; Eldredge & Jones, 2019; Liu et al., 2024). For example, the stroke motion (also known as the translational motion for 2D cases) creates an intense LEV undergoing a delayed stall to generate high lift which is not possible according to conventional laws of aerodynamics (Ellington et al., 1996; Kweon & Choi, 2010). The stabilised LEV observed for low Reynolds number insects/birds flight has been further elaborated by using experimental measurements (Poelma et al., 2006) and direct numerical simulations (Liu et al., 1998; Liu & Aono, 2009), showing that the attached LEV generates low pressure regions enhancing the lift generation during translation. The pitching (rotational) motion of the wing generates the lift during stroke reversal through the Kramer effect and the wake capture (Dickinson et al., 1999; Sane & Dickinson, 2002). The wing revolving/rotating enhances the lift generation by stabilising the LEV through the spanwise pressure gradient, centripetal force, and Coriolis force (Ozen & Rockwell, 2012; Jardin & David, 2015). In addition, the dynamic bending of flexible wings delays the breakdown of LEV near the wing-tip augmenting the production of stroke-averaged vertical force (Nakata & Liu, 2012a).

Even though the unsteady kinematics could enhance the force production significantly, the force variations are also larger compared with that of the steady wings. The force variations may have significant consequences on the acoustics generated by the wings especially the Gutin sound (Clark, 2021), which have not been well explored (Nedunchezian et al., 2019; Wang & Tian, 2020). The flapping-wing induced sound has drawn the attention from zoologists in very early stages. For example, Ewing & Bennet-Clark (1968) found that the complex sound produced by drosophila has a similar frequency to the wing beat frequency during flight. One of the earliest rigorous study on the aeroacoustics of flapping wing can be originated to the experimental study conducted by Sueur et al. (2005) who measured the acoustic outputs of flies in tethered flight, finding that the first frequency component is associated with the wing beat frequency dominating the front of the flyer and the second harmonic is dominate at two sides. The two-dimensional numerical study conducted by Bae & Moon (2008) identified two different mechanisms of sound generation by flapping wings, where the frequency of the primary dipole tone is the same as the frequency of the transverse wing, while the vortex edge scattering during a tangential motion is responsible for other higher frequency dipole tones. Bae & Moon (2008) also pointed out that the wing-vortex effects are more prominent in forward flight resulting in a broader high frequency noise band. To further elaborate the sound generation by flapping wing in real world, three dimensional numerical studies have been conducted by Inada et al. (2009); Geng et al. (2017); Wang & Tian (2020). Specifically, the pressure variation on the wing surface is the major source of the acoustics generated by the flapping wing (Inada et al., 2009), which is consistent with the classical acoustics analogy theory. Geng et al. (2017) adopted a hydrodynamic/acoustic splitting approach to numerically study the sound generated by a forward-flying cicada. By using a prescribed wing kinematics considering the wing deformation, they found that the flexibility is beneficial in the noise reduction in all directions. Inspired by the numerical simulations considering fully flexible flapping wings which demonstrated the enhanced aerodynamic performance by appropriate flexibility, Wang & Tian (2019) considered the fluid–structure-acoustics interaction of flapping wings and found that an aerodynamics-dominated flexible wing experiences sound reduction and aerodynamic performance enhancement, while the flexibility of inertia-dominated wings strengths the sound outputs. Nedunchezian et al. (2019) numerically studied the sound generation by flapping wings with various kinematics with Ffowcs Williams-Hawkings (FW-H) equations and found that the high lift and the low sound are found with the large-amplitude flap and pitch, and the delayed pitching motion. Nedunchezian et al. (2019) also pointed out that the delayed pitching motion of the wing may aim to decrease the sound generation instead of enhancing the lift force. The three-dimensional fluid-structure-acoustics interaction considering three-dimensional flapping wings in hovering flight by Wang & Tian (2020) confirmed that the aerodynamics-dominated wing could experience noise reduction and lift enhancement simultaneously.

Even though scaling laws for the aerodynamics of flapping-wing have been developed (Dewey et al., 2013; Kang & Shyy, 2013; Lee et al., 2015; Quinn et al., 2014; Floryan et al., 2017; Sum Wu et al., 2019; Senturk & Smits, 2019; Ayancik et al., 2019), they have not been considered for the acoustics except for two recent works (Wang & Tian, 2020; Seo et al., 2019). Specifically, a simple scaling law for the noise output from flapping wings versus their aerodynamic power requirements was developed in Wang & Tian (2020). Seo et al. (2019) reported that the sound power efficiency is proportional to f2f^{2}. The present work is to develop scaling laws for the aeroacoustics of flapping wings during hovering flight based on the quasi-steady flow theory and Ffowcs-Williams-Hawkings analogy. Specially, we first consider purely translating wings in two-dimensional domain. We then consider two-dimensional flapping wings with a combined translating and pitching motion. Finally, we consider three-dimensional flapping wings during hovering flight. The rest of this work is organised as follows: the physical formation of flapping wing considered in this study is introduced in § 2; the numerical method used for the direct numerical simulation of the aerodynamics and aeroacoustics of flapping wing is introduced in § 3; the results and discussions are then presented in § 4; and a final conclusion is given in § 5.

2 Physical problem

To develop the scaling laws for hovering flapping wings, the sound generation by flapping wings in both two-dimensional (2D) and three-dimensional (3D) spaces is considered, as shown in Fig. 1. For the simplicity, the wing translating at a constant angle of attack namely π2αm\frac{\pi}{2}-\alpha_{m} in the horizontal direction is first considered, as shown in Fig. 1 (a). The kinematics of the translation is prescribed by the following equation

x=Amsin(2πft),x=A_{m}\sin(2\pi ft), (1)

where AmA_{m} is the amplitude of the translation, ff is the frequency, xx is the position of the leading edge and tt is the time. For the flapping wing with a combined translating and pitching motion, as shown in Fig. 1 (b), the pitching montion is prescribed as

α=αmcos(2πft),\alpha=\alpha_{m}\cos(2\pi ft), (2)

where π/2αm{\pi}/{2}-\alpha_{m} is the maximum angle of attack. The dimensionless parameters governing this problems include translating amplitude, ama_{m}, Reynolds number and Mach number,

Am/c¯,am,Re=ρUc¯μ,M=Ucs,A_{m}/\bar{c},\quad a_{m},\quad Re=\frac{\rho U\bar{c}}{\mu},\quad M=\frac{U}{c_{s}}, (3)

where U=2πfAmU=2\pi fA_{m}, c¯\bar{c} is the mean chord length, csc_{s} is the speed of sound, and ρ\rho and μ\mu are respectively the density and viscosity of the fluid. In this work, low Reynolds numbers and Mach numbers are considered to limit the scope for insect flight, as shown in Tab. 1. Other parameters of kinematics are also selected based on insects. For the 3D flapping wing undergoing stroke and pitching motion, as shown in Fig. 1 (c), the kinematics is governed by

ϕ=ϕmsin(2πft),α=αmcos(2πft),\phi=\phi_{m}\sin(2\pi ft),\quad\alpha=\alpha_{m}\cos(2\pi ft), (4)

where ϕm\phi_{m} and αm\alpha_{m} are respectively the amplitude of stroke and pitching. The flow of the 3D hovering flapping wing is governed by ϕm\phi_{m}, αm\alpha_{m}, Reynolds number and Mach number as defined in Eq. 3 with U=2πfϕmc¯ARU=2\pi f\phi_{m}\bar{c}AR, and the aspect ratio ARAR of the wing. A range of parameters are examined to construct the scaling law for the sound generation, as shown in Tab. 1. Therefore, the fluid flow is considered as weakly-compressible. The lift and drag coefficients are defined as

CL=FL0.5ρU2A,CD=FD0.5ρU2A,C_{L}=\frac{F_{L}}{0.5\rho U^{2}A},\quad C_{D}=\frac{F_{D}}{0.5\rho U^{2}A}, (5)

where FLF_{L} and FDF_{D} are respectively the sum of force exerted by the fluid on the wing in the vertical and horizontal directions which can be directly obtained by using the sum of the Lagrangian force introduced in the immersed boundary method (IBM), and AA is the area of the wing which is c¯\bar{c} for 2D cases. The sound pressure is defined as Δp=pp¯\Delta p=p-\bar{p} with p¯\bar{p} being the time averaged pressure, and the time averaging is conducted using ten cycles after the periodic state (cycle-to-cycle variation in the forces is negligible) is achieved. All the sound pressure presented here are scaled by the flow density and sound speed in quiescent state.

Refer to caption

Figure 1: A schematic of flapping wings: (a) 2D translating wing at a constant angle of attack, (b) 2D translating and pitching wing, and (c) 3D flapping wing in hovering flight.
2D MM Re Am/c¯A_{m}/\bar{c} αm\alpha_{m}
0.04-0.1 200-800 0.8-1.4 30o60o30^{o}-60^{o}
3D MM Re ϕm\phi_{m} αm\alpha_{m} AR
0.04-0.1 200-800 30o60o30^{o}-60^{o} 45o60o45^{o}-60^{o} 2.0-4.0
Table 1: Parameters for the 2D and 3D flapping wings.

3 Numerical methods

The unsteady aerodynamics of bio-inspired flapping wings is directly solved by using an immersed boundary–lattice Boltzmann method (IB-LBM), where the weakly compressible flow is solved by the LBM and the no-slip boundary condition is implemented by an IBM. Without loss of generality, here we use 2D IB-LBM as an example to introduce the method. In the multi-relaxation-time (MRT)-based LBM, the particle distribution function gig_{i} is given as (Coveney et al., 2002; Guo & Zheng, 2008; Xu et al., 2018)

gi(𝒙+𝒆iΔt,t+Δt)=gi(𝒙,t)Ωi(𝒙,t)+ΔtGi,i=1,2,9,g_{i}(\boldsymbol{x}+\boldsymbol{e}_{i}\Delta t,t+\Delta t)=g_{i}(\boldsymbol{x},t)-\Omega_{i}(\boldsymbol{x},t)+\Delta tG_{i},\quad i=1,2,...9, (6)

where Δt\Delta t denotes the time step, 𝒆i\boldsymbol{e}_{i} is the discrete velocity, subscript ii indicates the ii-th direction, 𝒙\boldsymbol{x} is the space coordinator, Ωi\Omega_{i} the collision operator, and GiG_{i} is the effect of the body force. For the MRT LBM, Ωi\Omega_{i} and GiG_{i} are given as,

Ωi\displaystyle\Omega_{i} =\displaystyle= (M1SM)ij[gj(𝒙,t)gjeq(𝒙,t)],\displaystyle-(M^{-1}SM)_{ij}[g_{j}(\boldsymbol{x},t)-g_{j}^{eq}(\boldsymbol{x},t)], (7)
Gi\displaystyle G_{i} =\displaystyle= [M1(IS/2)M]ijFj,\displaystyle[M^{-1}(I-S/2)M]_{ij}F_{j}, (8)

where MM is a transform matrix with a dimension of 9×99\times 9 for the 2D nine discrete velocity (D2Q9) model and 19×1919\times 19 for 3D nineteen discrete velocity (D3Q19), SS is the collision matrix, II is the identity matrix, gieqg_{i}^{eq} is the equilibrium distribution function, and FjF_{j} is the force effect term. Details of MM and SS can be found in Coveney et al. (2002). Here gieqg_{i}^{eq} and FjF_{j} can be calculated by

gieq=ωiρ[1+𝒆i𝒖cs2+𝒖𝒖:(𝒆i𝒆ics2𝑰)cs4],\displaystyle g_{i}^{eq}=\omega_{i}\rho[1+\frac{\boldsymbol{e}_{i}\cdot\boldsymbol{u}}{c_{s}^{2}}+\frac{\boldsymbol{u}\boldsymbol{u}:(\boldsymbol{e}_{i}\boldsymbol{e}_{i}-c_{s}^{2}\boldsymbol{I})}{c_{s}^{4}}], (9)
Fj=ωj[𝒆j𝒖cs2+𝒆j𝒖cs4𝒆j]𝒇,\displaystyle F_{j}=\omega_{j}[\frac{\boldsymbol{e}_{j}-\boldsymbol{u}}{c_{s}^{2}}+\frac{\boldsymbol{e}_{j}\cdot\boldsymbol{u}}{c_{s}^{4}}\boldsymbol{e}_{j}]\cdot\boldsymbol{f}, (10)

where ωi\omega_{i} are the weights, cs=Δx/(3Δt)c_{s}=\Delta x/(\sqrt{3}\Delta t) is the the sound speed, and 𝒇\boldsymbol{f} is the body force. The kinematic viscosity ν\nu of the fluid is determined by the relaxation time τ\tau, according to ν=(τ0.5)cs2Δt\nu=(\tau-0.5)c_{s}^{2}\Delta t (Qian et al., 1992; Chen & Doolen, 1998; Wang et al., 2022).

The macro variables including the density, pressure and momentum can be directly calculated from the distribution functions as follows,

ρ=i=19gi,p=ρcs2,ρ𝒖=i=19gi𝒆i+12𝒇Δt.\rho=\sum_{i=1}^{9}g_{i},\quad p=\rho c_{s}^{2},\quad\rho\boldsymbol{u}=\sum_{i=1}^{9}g_{i}\boldsymbol{e}_{i}+\frac{1}{2}\boldsymbol{f}\Delta t. (11)

Here, the feedback immersed boundary method proposed (see e.g., Goldstein et al., 1993; Huang & Tian, 2019) is employed to handle the no-slip boundary conditions at the fluid–structure interface,

𝒇(𝒙,t)=Γ𝑭L(s,t)δh(𝑿(s,t)𝒙)𝑑s,\displaystyle\boldsymbol{f}(\boldsymbol{x},t)=-\int_{\Gamma}\boldsymbol{F}_{L}(s,t)\delta_{h}(\boldsymbol{X}(s,t)-\boldsymbol{x})ds, (12)
𝑭L=αib0t(𝑼ib𝑼)𝑑t+βib(𝑼ib𝑼),\displaystyle\boldsymbol{F}_{L}=\alpha_{ib}\int_{0}^{t}(\boldsymbol{U}_{ib}-\boldsymbol{U})dt+\beta_{ib}(\boldsymbol{U}_{ib}-\boldsymbol{U}), (13)
𝑼ib(s,t)=V𝒖(x,t)δh(𝑿(s,t)𝒙)𝑑𝒙,\displaystyle\boldsymbol{U}_{ib}(s,t)=\int_{V}\boldsymbol{u}(x,t)\delta_{h}(\boldsymbol{X}(s,t)-\boldsymbol{x})d\boldsymbol{x}, (14)

where 𝑭L\boldsymbol{F}_{L} is the Lagrangian force, 𝑼ib\boldsymbol{U}_{ib} is the boundary velocity obtained by the interpolation from the fluid domain, α\alpha and β\beta are positive feedback constants, 𝒖\boldsymbol{u} is velocity of fluid nodes, 𝑿\boldsymbol{X} and 𝒙\boldsymbol{x} are respectively the coordinates of structural nodes and fluid nodes, ss is arc coordinate, VV denotes the fluid domain, Γ\Gamma denotes the structure domain, and δh\delta_{h} is the smoothed Delta function. In this work, the delta function proposed by Peskin (Peskin, 2002) is used. In addition, αib=0\alpha_{ib}=0 and βib=2.0\beta_{ib}=2.0 are used. More details can be found in Refs. (Wang & Tian, 2018; Tian et al., 2011; Xu et al., 2018; Huang et al., 2022).

Refer to caption

Figure 2: A schematic of the multi-block for the numerical simulation: (a) the computational domain, (b) the multi-block mesh and (c) the fluid and structure nodes in the IBM.

The computational domain shown in Fig. 2(a) extends respectively from 150c¯-150\bar{c} to 150c¯150\bar{c} in both directions for 2D and from 50c¯-50\bar{c} to 50c¯50\bar{c} in all three directions for 3D. A sponge layer is applied respectively at a distance of 100c¯100\bar{c} (2D) and 45c¯45\bar{c} (3D) for the outer domain to damp the solution to the far-field quiescent state, and a characteristic non-reflective boundary condition is also applied at the external boundaries to guarantee the reflections from the external boundaries are negligible. Seven (2D) and five (3D) levels of mesh are used in present study, and the mesh is refined with a factor of two with the finest mesh spacing of 0.005c¯0.005\bar{c} (2D) and 0.02c¯0.02\bar{c}, respectively. Three mesh blocks close to the wing is shown in Fig. 2(b) to illustrate the mesh strategy. The finest mesh covers the sweeping range of the wing for the near-field flow, and the acoustics are far away from the wing is resolved with mesh resolution of 0.16c¯0.16\bar{c} namely the acoustics region in Fig. 2(a). The Lagrangian mesh spacing is about 0.5Δx0.5\Delta x, as demonstrated in Fig. 2(c). In the current simulation, a reference density ρ=1.0\rho=1.0 and a sound velocity cs=1/3c_{s}=1/\sqrt{3} are used. The sound pressure is nondimensionalised by ρcs2\rho c_{s}^{2}. The validation of the numerical solver is given in Appendix A, and the mesh convergence study is provided in Appendix B. It should be noted that the force and sound power are transformed into the physical space assuming a wing mean chord length of 2mm2mm in earth atmosphere environment, to make it intuitive for readers.

4 Results and discussions

4.1 Purely translating wings

The purely translating wing is first considered here for its simplicity. For a steady wing at a given angle of attack (see Fig. 1 a), the flow separates near the leading edge of the wing and forms a leading edge vortex (LEV). The strength of the LEV is related to the shear flow on the wing, and the aerodynamic force generated by the wing can be estimated from the strength of the LEV by using the potential flow theory as F=ρΓc¯F=\rho\Gamma\bar{c} with Γ\Gamma being the strength of the circulation (Dickinson, 1996; Wu et al., 2007). For a translating wing, the aerodynamic force can be estimated alternatively as (Lee et al., 2015)

F=ΔI/T,F=\Delta I/T, (15)

where ΔI\Delta I is the momentum induced by the vortical structures and TT is the translating period of the wing. ΔIρΓS\Delta I\sim\rho\Gamma S and S=c¯S=\bar{c} in 2D. Therefore, we obtain FρΓfc¯F\sim\rho\Gamma f\bar{c} by using Eq. 15. According to the potential flow theory, the circulation can be expressed as ΓUc¯cos(αm)\Gamma\sim U\bar{c}\cos(\alpha_{m}) for a wing translating a given angle of attack, where UU is the translating velocity of the wing and αm\alpha_{m} is the angle of attack which is a constant here. Therefore, the aerodynamic force generated by the translating wing can be written as

FρUfc¯2cos(αm),F\sim\rho Uf\bar{c}^{2}\cos({\alpha_{m}}), (16)

and the lift (LL) is the projection of FF in the vertical direction, i.e.,

LρUfc¯2sin(2αm).L\sim\rho Uf\bar{c}^{2}\sin({2\alpha_{m}}). (17)

For the purely translating wing, the maximum velocity of the wing UmAmfU_{m}\sim A_{m}f according to the kinematics described in Eq. 1. Therefore, the amplitude of the lift follows

LmρAmf2c¯2sin(2αm).L_{m}\sim\rho A_{m}f^{2}\bar{c}^{2}\sin({2\alpha_{m}}). (18)

It should be noted that LmL_{m} is the amplitude of the lift rather than the mean lift. The mean lift is expected to be zero for the translating wing due to the symmetric lift during left and right translation stages.

According to the FW-H equation, the far-field sound generated by a moving objects has three components with its solution consisting of a surface integral of the monopole source, a surface integral of the dipole source and a volume integral of the quadrupole source. If the sources are compact and the observation location is far away from the wing, the solution of the FW-H equation becomes (Tian, 2020; Seo et al., 2019),

Δpd𝑭dt𝒓^,\Delta p\sim\frac{d\boldsymbol{F}}{dt}\cdot\hat{\boldsymbol{r}}, (19)

where 𝑭\boldsymbol{F} is the aerodynamic force which is mainly determined by the pressure, and 𝒓^\hat{\boldsymbol{r}} is the unit vector pointing from the wing to the far-field observer. Using the scaling law for the amplitude of lift in Eq. 18, the scaling law for the amplitude of sound pressure generated in the vertical direction is

ΔpLmfρAmf3c¯2sin(2αm).\Delta p\sim L_{m}f\sim\rho A_{m}f^{3}\bar{c}^{2}\sin({2\alpha_{m}}). (20)

To confirm the validity of the constructed scaling laws for the amplitude of lift and sound generated in the vertical direction, direct numerical simulation is conducted by using an IB-LBM method as mentioned in § 3 with the parameters shown in Tab. 1 being considered which gives 35 simulations in total. It should be noted that we randomly chose parameters shown in Tab.  1 instead of simulating all parameters to save computational resource and cover the listed parameter range.

Refer to caption


Figure 3: Scaling laws of a translating wing: (a) the lift, and (b) the sound pressure θ=90o\theta=90^{o} and r=85c¯r=85\bar{c} scaled by using the fluid density and sound speed at quiescent state.

Fig. 3 shows the amplitudes of lift and the root-mean-square of sound pressure in the vertical direction generated by the translating wing from 40 numerical simulations according to scaling laws in Eqs. 18 and 20. It shows that the lift collapses to a straight line confirming the validity of the derived scaling laws. The sound pressure also collapses to a line as expected in the scaling law, although it shows slightly larger discrepancies compared with the lift which could be attributed to the non-perfect sinusoidal sound pressure in time domain due to the viscous effects.

Refer to caption

Figure 4: Sound generation by a translating wing at M=0.05M=0.05, Re=400Re=400, αm=30o\alpha_{m}=30^{o} and Am=c¯A_{m}=\bar{c}: (a) time histories of CDC_{D} (horizontal direction); (a) time histories of CLC_{L} (vertical direction); (c) time histories of sound pressure measured at a distance of 100c¯100\bar{c}, θ=90o\theta=90^{o} and 270o270^{o}; (d) decay of sound pressure with distance measured at θ=90o\theta=90^{o}; (e) sound pressure directivity measured at a distance of 100c¯100\bar{c}.

Refer to caption

Figure 5: Instantaneous acoustics and flow fields generated by a translating wing at t=0T/4t=0T/4 with M=0.04M=0.04, Re=400Re=400, αm=45o\alpha_{m}=45^{o} and Am=0.8c¯A_{m}=0.8\bar{c}: (a) sound pressure contours; (b) pressure coefficient contours near the wing; (c) vorticity contours near the wing.

To illustrate the major features of both force and sound generations by a translating wing, Fig. 4 shows the force histories, sound pressure time histories, decay of sound pressure with distance and sound pressure directivity; and Fig. 5 shows the instantaneous acoustics and flow fields at t=0T/4t=0T/4 with M=0.04M=0.04, Re=400Re=400, αm=45o\alpha_{m}=45^{o} and Am=0.8c¯A_{m}=0.8\bar{c}. Both drag and lift forces are periodic and symmetric (see Fig. 4 (a) and (b)). This is because that when the wing translates from the left to the right, a suction region and two vortices are generated at bottom side of the wing as shown in Fig. 5 (b) and (c), which are “rotational symmetric” to those generated when the wing moves from the right to the left (not shown for conciseness). Due to the compact sound sources, the sound field is quasi-symmetric about the wing, as shown in Fig. 4 (a). A quantitative illustration is shown in Fig. 4 (c), which measures the sound pressure at a distance of 100c¯100\bar{c} and θ=90o\theta=90^{o} and 270o270^{o}. It shows that the measured sound pressure signals are close to a sinusoidal shape which aligns with the prescribed kinematics, and the same amplitude and out-of-phase pressure at the two measure points confirms the symmetric sound fields, as shown in the acoustics field in Fig. 5 (a) and the polar plot of the sound diretivity in Fig. 4 (e). In addition, the decay of sound pressure measured at θ=90o\theta=90^{o} with the distance is shown in Fig. 4 (d), which agrees well with the linear relation from theoretical analysis.

4.2 Translating and pitching wings

Refer to caption

Figure 6: A schematic of the velocity of a translating and pitching wing.

In the real world, insects and birds normally flap theirs wings with a pitching motion to get high aerodynamic performance. Here, the wing with such a combined flapping and pitching motion is further considered. Fig. 6 shows a schematic of the wing velocity when the combined kinematics is used. In addition to the circulation generated due to the leading edge vortex induced by the translation, the pitching motion induce extra circulation which could enhance the lift generation and should be considered. According to the potential flow theory (see e.g., Dickinson, 1996; Wu et al., 2007), the force is determined by

F=FT+FRρ(ΓT+ΓR)fc¯,F=F_{T}+F_{R}\sim\rho(\Gamma_{T}+\Gamma_{R})f\bar{c}, (21)

where FTF_{T} and FRF_{R} are respectively the force caused by the translating and pitching components, and ΓT\Gamma_{T} and ΓR\Gamma_{R} are the circulation induced by the translation and pitching, respectively. According to Section 4.1, FTρΓTfc¯ρUTfc¯2sin(0.5παm)=ρAmf2c¯2sin(0.5παm)F_{T}\sim\rho\Gamma_{T}f\bar{c}\sim\rho U_{T}f\bar{c}^{2}\sin{(0.5\pi-\alpha_{m})}=\rho A_{m}f^{2}\bar{c}^{2}\sin(0.5\pi-\alpha_{m}). The force generated by the pitching can be written as FRρΓRfc¯ρURfc¯2F_{R}\sim\rho\Gamma_{R}f\bar{c}\sim\rho U_{R}f\bar{c}^{2}, where URU_{R} is the averaged rotating velocity along the chord of the wing, i.e., the rotating velocity measured at the middle of the wing chord. According to the kinematics described in Eq. 2, UR0.5αmfc¯U_{R}\sim 0.5\alpha_{m}f\bar{c}, and thus the amplitude of the aerodynamic force generated by a wing with combined translation and pitching motion follows

FmρAmf2c¯2sin(0.5παm)+0.5ραmf2c¯3.F_{m}\sim\rho A_{m}f^{2}\bar{c}^{2}\sin(0.5\pi-\alpha_{m})+0.5\rho\alpha_{m}f^{2}\bar{c}^{3}. (22)

By projecting the force on to the vertical direction, then the amplitude of the lift has the following scaling law

LmλρAmf2c¯2sin(2αm),L_{m}\sim\lambda\rho A_{m}f^{2}\bar{c}^{2}\sin(2\alpha_{m}), (23)

with

λ=1+αmc¯2Amsinαm.\lambda=1+\frac{\alpha_{m}\bar{c}}{2A_{m}\sin\alpha_{m}}. (24)

By using Eq. 19, a scaling law for the amplitude of the sound generated by a wing with combined translation and pitching motion in the vertical direction is written as

ΔpdLdtλρAmf3c¯2sin(2αm).\Delta p\sim\frac{dL}{dt}\sim\lambda\rho A_{m}f^{3}\bar{c}^{2}\sin(2\alpha_{m}). (25)

Refer to caption

Figure 7: Scaling laws of 2D wings undergoing translating only, combined translating and pitching motion: (a) the amplitude of lift, and (b) sound pressure at θ=90o\theta=90^{o} and r=85c¯r=85\bar{c} scaled by using the fluid density and sound speed at quiescent state. λ\lambda is unit for the translating wing and determined by Eq. 24 for the translating and pitching wing.

It is noted that the scaling laws described by Eqs. 23 and 25 for the wing undergoing combined translating and pitching motion are consistent with those described by Eqs. 18 and 20 for the wing undergoing purely translating motion. The coefficient λ\lambda is derived for the modification due to the pitching motion. To confirm the validity of the scaling laws for a wing with combined translation and pitching kinematics, another 30 simulations are conducted with the parameters selected from the Tab. 1. Fig. 7 shows the lift and sound pressure from all numerical simulations. The results show that the numerical data collapse well to a line for both the lift and sound pressure. It is also noted that the wings with pitching motion generate smaller lift amplitudes and thus smaller acoustic outputs compared with those wings translating only, as the pitching motion considered in this work relieves the relative velocity near the wing tip during the start of each stroke to the flow as discussed in Wang & Tian (2019).

4.3 Three-dimensional flapping wings

In this section, we further examine three-dimensional flapping wings in hovering flight. According to the potential flow theory discussed in Section. 4.1, the aerodynamic force generated by the flapping wing FρΓSfF\sim\rho\Gamma Sf, where the area S=ARc¯2S=AR\bar{c}^{2}, circulation ΓUc¯sin(0.5παm)\Gamma\sim U\bar{c}\sin(0.5\pi-\alpha_{m}) and the maximum wing tip velocity UmϕmfARc¯U_{m}\sim\phi_{m}fAR\bar{c}. Take the aspect ratio effects ARAR+2\frac{AR}{AR+2} into account, a scaling law of the amplitude of lift and sound pressure in the vertical direction for the 3D flapping wing can be obtained as

Lm=Fcos(0.5παm)ρλϕmf2sin(2αm)AR2c¯4ARAR+2,\displaystyle L_{m}=F\cos(0.5\pi-\alpha_{m})\sim\rho\lambda\phi_{m}f^{2}\sin(2\alpha_{m})AR^{2}\bar{c}^{4}\frac{AR}{AR+2}, (26)
ΔpdLdtλρϕmf3AR2c¯4sin(2αm)ARAR+2,\displaystyle\Delta p\sim\frac{dL}{dt}\sim\lambda\rho\phi_{m}f^{3}AR^{2}\bar{c}^{4}\sin(2\alpha_{m})\frac{AR}{AR+2}, (27)

where λ\lambda is the same as that in Eq. 24.

Refer to caption


Figure 8: Scaling laws of a 3D wing undergoing combined stroke and pitching motion: (a) the lift, and (b) sound pressure in the vertical direction at r=40c¯r=40\bar{c} scaled by using the fluid density and sound speed at quiescent state.

Refer to caption

Figure 9: Sound generation by a 3D flapping wing at M=0.05M=0.05, Re=200Re=200, ϕm=45o\phi_{m}=45^{o}, αm=45o\alpha_{m}=45^{o} and AR=2.0AR=2.0: (a) time histories of CLC_{L}; (b) time histories of sound pressure measured at a distance of 30c¯30\bar{c} on the top of the wing; (c) decay of sound pressure with distance measured on the top of the wing; (d) sound pressure directivity measured at a distance of 40c¯40\bar{c} on the stroke plane and vertical plane; (e) sound pressure directivity measured over a spherical surface of a radius of 40c¯40\bar{c}.

To confirm the validity of the scaling laws for the 3D flapping wing in hovering flight, 16 direct numerical simulations are conducted with the parameters selected from Tab. 1 covering a range of Reynolds numbers, Mach numbers, wing kinematics and aspect ratios. According to Ellington et al. (1996), the insect wing can be characterised by r^\hat{r} which is normally in a range between 0.4 and 0.6. Here r^=0.53\hat{r}=0.53 in the middle range is selected to reduce the parameters in the simulation. Fig. 8 shows the constructed scaling laws with the data from DNS which shows that the numerical results are well positioned around the scaling laws confirming the validity of the scaling laws for the aerodynamic force and the sound pressure.

To demonstrate the properties of aerodynamic force and sound generated by 3D flapping wing, a specific case at M=0.05M=0.05, Re=200Re=200, ϕm=45o\phi_{m}=45^{o}, αm=45o\alpha_{m}=45^{o} and AR=2.0AR=2.0 are selected for analysis, as shown in Fig. 9. It is found that the lift coefficient is quasi-sinusoidal (see Fig. 9a), while the sound pressure measured at a distance of 40c¯40\bar{c} on the top of the wing reaches a periodic but non-sinusoidal state (see Fig. 9b). Although complex acoustic features in time space are observed in Fig.  9 (b), the second harmonic is the dominating frequency of the sound, and is consistent with the frequency of the lift force (as shown by the insert which is the Fast Fourier Transform of the sound pressure). This observation agrees well with that observed by Sueur et al. (2005) considering real flies with more complex wing kinematics. The coincidence of the frequencies of the lift and the sound pressure explains why a simple scaling law can be constructed by relating the aeroacoutics to the aerodynamic force directly without losing the major physics, as shown in Fig. 8 (a) and (b). Fig. 9 (c) shows the decay of the pressure measured on the top of the wing versus the distance, which agrees well with the linear decay law in the far-field region, as indicated by the solid line confirming the accuracy of the numerical simulation. For the near field with r/c¯<25r/\bar{c}<25, it also shows that the decay of sound pressure does not follow the linear trend due to the near-field effects. This is the reason that the sound pressure at a distance of 40c¯40\bar{c} is used in Fig. 8 to validate the scaling law eliminating the effects from the near-field unsteady flow. Fig. 9 (d) shows the directivity of the sound pressure on the stroke plane (xyxy) and vertical plane (yzyz), indicating a nearly symmetric profile about the 9090^{\circ} line due to the symmetric kinematics. A more comprehensive directivity in 3D space is shown in Fig. 9 (e), which shows that the maximum sound is observed on the stroke plane, indicating that the scaling law for the maximum sound pressure in the vertical direction corresponding to the lift does not necessarily reflect the overall acoustic outputs.

The mean lift and the sound power are also important in measuring the flapping performance. For the hovering flapping wing considered in this work, the mean lift is proportional to its amplitude, as only small negative lift is generated as shown in Fig. 9 (a). This limited negative lift is actually common in hovering flight as observed by Nagai et al. (2010) in experiment and Young et al. (2009) in simulation with a pair of flexible wings. Therefore, the mean lift (L¯\bar{L}) should follow the same scaling law as the amplitude of the lift, which can be given as

L¯\displaystyle\bar{L} \displaystyle\sim λρϕmf2sin(2αm)AR2c¯4ARAR+2.\displaystyle\lambda\rho\phi_{m}f^{2}\sin(2\alpha_{m})AR^{2}\bar{c}^{4}\frac{AR}{AR+2}. (28)

It should be noted that this scaling law is consistent with that obtained by Lee et al. (2015), who ignored the effects from pitching (as reflected by λ\lambda) so that their scaling law covers a large range of lift (up to 2nd-order difference) where the pitching effects are not significant. The consistent scaling law is constructed based on the fact that the mean lift scales with the lift amplitude in hovering flight which was not explained in Lee et al. (2015). The current DNS data shown in Fig. 10 (a) also confirms the validity of the scaling law in Eq. 28. The sound power (SPSP) which measures the overall acoustic outputs is further obtained as

SP\displaystyle SP \displaystyle\sim (Δp)2(dFdt)2λ2ρϕm2f6AR4c¯8cos2(αm)(ARAR+2)2,\displaystyle(\Delta p)^{2}\sim\left(\frac{dF}{dt}\right)^{2}\sim\lambda^{2}\rho\phi_{m}^{2}f^{6}AR^{4}\bar{c}^{8}\cos^{2}(\alpha_{m})(\frac{AR}{AR+2})^{2}, (29)

which is confirmed to capture the major physics well by the data from DNS, as shown in Fig. 10(b). It is noted that the derived scaling law is consistent with Curl’s theory indicating that the dipole sound power is proportional to the sixth power of the flow velocity (Curle, 1955). This also indicates that the sound generated by the flapping wing is dominated by the dipole sound and/or the combination of multiple dipole sounds, as shown in Fig. 9 (d) and (e).

Refer to caption


Figure 10: Scaling laws for the mean lift (a), and sound power at r=40c¯r=40\bar{c} (b) of the 3D hovering flapping wing.

5 Conclusions

This work presents the scaling laws for the sound generation of bio-inspired flapping wings during hovering flight, which are derived by using the classical quasi-steady flow theory and the FW-H acoustic equation. Three groups of wings are consider: 2D purely translating wings, 2D translating and pitching wings, and 3D stroking and pitching wings. Direct numerical simulations considering a range of parameters including the Reynolds number, Mach number and wing kinematics confirm that the constructed scaling laws capture the major physics and agree well with the direct numerical simulation results.

It is noted that the simple scaling laws directly relate the kinematics of bio-inspired flapping wings to their noise emission. The scaling laws are able to explain the properties of sound generated by most insects with different wing sizes and kinematics. The success of this work is due to the fact that the major sound by the bio-inspired flapping wings considered here is the Gutin sound. The scaling analysis of other sound sources should be considered for complicated wings and high-speed flight, which will be our future effort.

Funding

This work was supported by the Australian Research Council Discovery Project (grant No. DP200101500) and conducted with the assistance of computational resources from the National Computational Infrastructure (NCI), which is supported by the Australian Government.

Declaration of interests

The authors report no conflict of interest.

Appendix A Validation of the numerical solver

This section presents a validation of the numerical solver considering the aerodynamics and acoustics of flow around a cylinder at Re=ρUD/μ=150Re=\rho UD/\mu=150 and M=U/cs=0.2M=U/c_{s}=0.2 (where ρ\rho and μ\mu are respectively the density and viscosity of the fluid, DD is the diameter of the cylinder, UU is the velocity of the uniform incoming flow and csc_{s} is the sound speed of the fluid flow at quiescent state), as which is a benchmark and has been extensively validated by Inoue & Hatakeyama (2002); Wang et al. (2020).

Refer to caption

Figure 11: Flow around a circular cylinder at Re=150Re=150 and M=0.2M=0.2: (a) the time histories of CLC_{L} and CDC_{D}; (b) the decay of sound pressure measured at θ=90o\theta=90^{o} versus distance.

In this validation, the fluid domain has a dimension of 150D×150D150D\times 150D, and six mesh blocks are used with the finest mesh spacing of D/100D/100. The quantitative comparison of the lift and drag coefficients, as well as the sound pressure measured at the top of the cylinder are shown in Fig. 11. As we can see, the lift and drag coefficients predicted by the current method agree well with the data from Inoue & Hatakeyama (2002). In addition, the sound pressure with the distance from the cylinder centre is proportional to 1/sqrtr1/sqrt{r}, which is consistent with the result reported in the reference. Therefore, the IB-LBM solver for the direct simulation of acoustics at low Reynolds numbers is reliable.

Appendix B Mesh convergence study

This section presents a mesh convergence study of the numerical solver considering the aerodynamics and acoustics of a 3D flapping wing in hovering flight. All the parameters considered here are same as those used in Fig. 9. Three mesh sizes in the near-field region are considered, i.e., Δx=0.08c¯,0.04c¯\Delta x=0.08\bar{c},0.04\bar{c} and 0.02c¯0.02\bar{c}. Fig. 12 shows the time histories of lift coefficient CLC_{L} and the sound pressure Δp\Delta p measured above the wing at a distance of 30c¯30\bar{c}. Both of CLC_{L} and Δp\Delta p show a good convergence with the refinement of the mesh. The finest mesh of 0.02c¯0.02\bar{c} is used for the current study to provide a better sound visualisation.

Refer to caption


Figure 12: Mesh convergence study of 3D flapping wing at M=0.05M=0.05, Re=200Re=200, ϕm=45o\phi_{m}=45^{o}, αm=45o\alpha_{m}=45^{o} and AR=2.0AR=2.0: (a) time histories of CLC_{L}, and (b) time histories of sound pressure measured above the wing at a distance of 30c¯30\bar{c}.

References

  • Ayancik et al. (2019) Ayancik, Fatma, Zhong, Qiang, Quinn, Daniel B., Brandes, Aaron, Bart-Smith, Hilary & Moored, Keith W. 2019 Scaling laws for the propulsive performance of three-dimensional pitching propulsors. Journal of Fluid Mechanics 871, 1117–1138.
  • Bae & Moon (2008) Bae, Youngmin & Moon, Young J 2008 Aerodynamic sound generation of flapping wing. The Journal of the Acoustical Society of America 124 (1), 72–81.
  • Bennett (1970) Bennett, Leon 1970 Insect flight: lift and rate of change of incidence. Science 167 (3915), 177–179.
  • Bomphrey et al. (2017) Bomphrey, Richard J, Nakata, Toshiyuki, Phillips, Nathan & Walker, Simon M 2017 Smart wing rotation and trailing-edge vortices enable high frequency mosquito flight. Nature 544 (7648), 92–95.
  • Chen & Doolen (1998) Chen, Shiyi & Doolen, Gary D 1998 Lattice Boltzmann method for fluid flows. Annual Review of Fluid Mechanics 30 (1), 329–364.
  • Clark (2021) Clark, Christopher J 2021 Ways that animal wings produce sound. Integrative and Comparative Biology 61, 696–709.
  • Coveney et al. (2002) Coveney, P. V., Succi, S., d’Humières, Dominique, Ginzburg, Irina, Krafczyk, Manfred, Lallemand, Pierre & Luo, Li-Shi 2002 Multiple–relaxation–time lattice Boltzmann models in three dimensions. Philosophical Transactions of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences 360 (1792), 437–451.
  • Curle (1955) Curle, Newby 1955 The influence of solid boundaries upon aerodynamic sound. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 231 (1187), 505–514.
  • Dai et al. (2012) Dai, Hu, Luo, Haoxiang & Doyle, James F 2012 Dynamic pitching of an elastic rectangular wing in hovering motion. Journal of Fluid Mechanics 693, 473–499.
  • Dewey et al. (2013) Dewey, P. A., Boschitsch, B. M., Moored, K. W., Stone, H. A. & Smits, A. J. 2013 Scaling laws for the thrust production of flexible pitching panels. Journal of Fluid Mechanics 732, 29–46.
  • Dickinson (1996) Dickinson, Michael H 1996 Unsteady mechanisms of force generation in aquatic and aerial locomotion. American zoologist 36 (6), 537–554.
  • Dickinson et al. (1999) Dickinson, Michael H, Lehmann, Fritz-Olaf & Sane, Sanjay P 1999 Wing rotation and the aerodynamic basis of insect flight. Science 284 (5422), 1954–1960.
  • Eldredge & Jones (2019) Eldredge, Jeff D & Jones, Anya R 2019 Leading-edge vortices: mechanics and modeling. Annual Review of Fluid Mechanics 51, 75–104.
  • Ellington et al. (1996) Ellington, Charles P, Van Den Berg, Coen, Willmott, Alexander P & Thomas, Adrian LR 1996 Leading-edge vortices in insect flight. Nature 384 (6610), 626–630.
  • Ewing & Bennet-Clark (1968) Ewing, Arthur W & Bennet-Clark, HC 1968 The courtship songs of drosophila. Behaviour 31 (3-4), 288–301.
  • Floryan et al. (2017) Floryan, D., Van Buren, T., Rowley, C. W. & Smits, A. J. 2017 Scaling the propulsive performance of heaving and pitching foils. Journal of Fluid Mechanics 822, 386–397.
  • Geng et al. (2017) Geng, Biao, Xue, Qian, Zheng, Xudong, Liu, Geng, Ren, Yan & Dong, Haibo 2017 The effect of wing flexibility on sound generation of flapping wings. Bioinspiration & Biomimetics 13 (1), 016010.
  • Goldstein et al. (1993) Goldstein, D., Handler, R. & Sirovich, L. 1993 Modeling a no-slip flow boundary with an external force field. Journal of Computational Physics 105 (2), 354–366.
  • Guo & Zheng (2008) Guo, Zhaoli & Zheng, Chuguang 2008 Analysis of lattice Boltzmann equation for microscale gas flows: relaxation times, boundary conditions and the Knudsen layer. International Journal of Computational Fluid Dynamics 22 (7), 465–473.
  • Huang et al. (2023) Huang, Qiuxiang, Bhat, Shantanu S, Yeo, Eng Chow, Young, John, Lai, Joseph CS, Tian, Fang-Bao & Ravi, Sridhar 2023 Power synchronisations determine the hovering flight efficiency of passively pitching flapping wings. Journal of Fluid Mechanics 974, A41.
  • Huang et al. (2022) Huang, Qiuxiang, Liu, Zhengliang, Wang, Li, Ravi, Sridhar, Young, John, Lai, Joseph C. S. & Tian, Fang-Bao 2022 Streamline penetration, velocity error, and consequences of the feedback immersed boundary method. Physics of Fluids 34 (9), 097101.
  • Huang & Tian (2019) Huang, Wei-Xi & Tian, Fang-Bao 2019 Recent trends and progress in the immersed boundary method. Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science 233 (23-24), 7617–7636.
  • Inada et al. (2009) Inada, Yoshinobu, Aono, Hikaru, Liu, Hao & Aoyama, Takashi 2009 Numerical analysis of sound generation of insect flapping wings. Theoretical and Applied Mechanics Japan 57, 437–447.
  • Inoue & Hatakeyama (2002) Inoue, Osamu & Hatakeyama, Nozomu 2002 Sound generation by a two-dimensional circular cylinder in a uniform flow. Journal of Fluid Mechanics 471, 285–314.
  • Jardin & David (2015) Jardin, Thierry & David, Laurent 2015 Coriolis effects enhance lift on revolving wings. Physical Review E 91 (3), 031001.
  • Jaworski & Peake (2020) Jaworski, Justin W & Peake, Nigel 2020 Aeroacoustics of silent owl flight. Annual Review of Fluid Mechanics 52, 395–420.
  • Kang & Shyy (2013) Kang, Chang-kwon & Shyy, Wei 2013 Scaling law and enhancement of lift generation of an insect-size hovering flexible wing. Journal of The Royal Society Interface 10 (85), 20130361.
  • Kweon & Choi (2010) Kweon, Jihoon & Choi, Haecheon 2010 Sectional lift coefficient of a flapping wing in hovering motion. Physics of Fluids 22 (7), 071703.
  • Lee et al. (2015) Lee, Jeongsu, Choi, Haecheon & Kim, Ho-Young 2015 A scaling law for the lift of hovering insects. Journal of Fluid Mechanics 782, 479–490.
  • Lighthill (1975) Lighthill, James 1975 Aerodynamic aspects of animal flight. In Swimming and Flying in Nature: Volume 2, pp. 423–491. Springer.
  • Lighthill (1973) Lighthill, M J 1973 On the Weis-Fogh mechanism of lift generation. Journal of Fluid Mechanics 60 (1), 1–17.
  • Liu & Aono (2009) Liu, Hao & Aono, Hikaru 2009 Size effects on insect hovering aerodynamics: an integrated computational study. Bioinspiration & Biomimetics 4 (1), 015002.
  • Liu et al. (1998) Liu, Hao, Ellington, Charles P, Kawachi, Keiji, Van Den Berg, Coen & Willmott, Alexander P 1998 A computational fluid dynamic study of hawkmoth hovering. The journal of Experimental Biology 201 (4), 461–477.
  • Liu et al. (2024) Liu, Hao, Wang, Shizhao & Liu, Tianshu 2024 Vortices and forces in biological flight: Insects, birds, and bats. Annual Review of Fluid Mechanics 56.
  • Nagai et al. (2010) Nagai, Hiroto, Isogai, Koji & Fujimoto, Tatsumi 2010 Experimental study on flow interaction between fore-and hindwings of dragonfly in hovering and forward flight. In 27th Congress of International Council of the Aeronautical Sciences, Nice, France, pp. 1–10.
  • Nakata & Liu (2012a) Nakata, Toshiyuki & Liu, Hao 2012a Aerodynamic performance of a hovering hawkmoth with flexible wings: a computational approach. Proceedings of the Royal Society B: Biological Sciences 279 (1729), 722–731.
  • Nakata & Liu (2012b) Nakata, Toshiyuki & Liu, Hao 2012b A fluid–structure interaction model of insect flight with flexible wings. Journal of Computational Physics 231 (4), 1822–1847.
  • Nedunchezian et al. (2019) Nedunchezian, Kabilan, Kang, Chang-kwon & Aono, Hikaru 2019 Effects of flapping wing kinematics on the aeroacoustics of hovering flight. Journal of Sound and Vibration 442, 366–383.
  • Ozen & Rockwell (2012) Ozen, Cem A & Rockwell, D 2012 Three-dimensional vortex structure on a rotating wing. Journal of Fluid Mechanics 707, 541–550.
  • Peskin (2002) Peskin, Charles S 2002 The immersed boundary method. Acta Numerica 11, 479–517.
  • Poelma et al. (2006) Poelma, C, Dickson, WB & Dickinson, MH 2006 Time-resolved reconstruction of the full velocity field around a dynamically-scaled flapping wing. Experiments in Fluids 41, 213–225.
  • Qian et al. (1992) Qian, Yue-Hong, d’Humières, Dominique & Lallemand, Pierre 1992 Lattice BGK models for Navier-Stokes equation. Europhysics Letters 17 (6), 479.
  • Quinn et al. (2014) Quinn, Daniel B., Lauder, George V. & Smits, Alexander J. 2014 Scaling the propulsive performance of heaving flexible panels. Journal of Fluid Mechanics 738, 250–267.
  • Sane (2003) Sane, Sanjay P 2003 The aerodynamics of insect flight. Journal of Rxperimental Biology 206 (23), 4191–4208.
  • Sane & Dickinson (2002) Sane, Sanjay P & Dickinson, Michael H 2002 The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight. Journal of Experimental Biology 205 (8), 1087–1096.
  • Senturk & Smits (2019) Senturk, Utku & Smits, Alexander J. 2019 Reynolds number scaling of the propulsive performance of a pitching airfoil. AIAA Journal 57 (7), 2663–2669.
  • Seo et al. (2019) Seo, Jung-Hee, Hedrick, Tyson L & Mittal, Rajat 2019 Mechanism and scaling of wing tone generation in mosquitoes. Bioinspiration & Biomimetics 15 (1), 016008.
  • Shyy et al. (2010) Shyy, Wei, Aono, Hikaru, Chimakurthi, Satish Kumar, Trizila, Pat, Kang, C-K, Cesnik, Carlos ES & Liu, Hao 2010 Recent progress in flapping wing aerodynamics and aeroelasticity. Progress in Aerospace Sciences 46 (7), 284–327.
  • Shyy et al. (2013) Shyy, Wei, Aono, Hikaru, Kang, Chang-kwon & Liu, Hao 2013 An introduction to flapping wing aerodynamics, , vol. 37. Cambridge University Press.
  • Sueur et al. (2005) Sueur, Jérôme, Tuck, Elizabeth J & Robert, Daniel 2005 Sound radiation around a flying fly. The Journal of the Acoustical Society of America 118 (1), 530–538.
  • Sum Wu et al. (2019) Sum Wu, Kit, Nowak, Jerome & Breuer, Kenneth S 2019 Scaling of the performance of insect-inspired passive-pitching flapping wings. Journal of the Royal Society Interface 16 (161), 20190609.
  • Tian (2020) Tian, Fang-Bao 2020 Hydrodynamic effects of mucus on swimming performance of an undulatory foil by using the DSD/SST method. Computational Mechanics 65, 751–761.
  • Tian et al. (2013) Tian, Fang-Bao, Luo, Haoxiang, Song, Jialei & Lu, Xi-Yun 2013 Force production and asymmetric deformation of a flexible flapping wing in forward flight. Journal of Fluids and Structures 36, 149–161.
  • Tian et al. (2011) Tian, Fang-Bao, Luo, Haoxiang, Zhu, Luoding, Liao, James C & Lu, Xi-Yun 2011 An efficient immersed boundary-lattice Boltzmann method for the hydrodynamic interaction of elastic filaments. Journal of Computational Physics 230 (19), 7266–7283.
  • Wang et al. (2022) Wang, Li, Liu, Zhengliang & Rajamuni, Methma 2022 Recent progress of lattice boltzmann method and its applications in fluid-structure interaction. Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science p. 09544062221077583.
  • Wang & Tian (2018) Wang, Li & Tian, Fang-Bao 2018 Heat transfer in non-newtonian flows by a hybrid immersed boundary–lattice Boltzmann and finite difference method. Applied Sciences 8 (4), 559.
  • Wang & Tian (2019) Wang, Li & Tian, Fang-Bao 2019 Numerical study of flexible flapping wings with an immersed boundary method: Fluid–structure–acoustics interaction. Journal of Fluids and Structures 90, 396–409.
  • Wang & Tian (2020) Wang, Li & Tian, Fang-Bao 2020 Numerical study of sound generation by three-dimensional flexible flapping wings during hovering flight. Journal of Fluids and Structures 99, 103165.
  • Wang et al. (2020) Wang, Li, Tian, Fang-Bao & Lai, Joseph CS 2020 An immersed boundary method for fluid–structure–acoustics interactions involving large deformations and complex geometries. Journal of Fluids and Structures 95, 102993.
  • Wang (2005) Wang, Z Jane 2005 Dissecting insect flight. Annual Review of Fluid Mechanics 37, 183–210.
  • Weis-Fogh (1973) Weis-Fogh, Torkel 1973 Quick estimates of flight fitness in hovering animals, including novel mechanisms for lift production. Journal of Experimental Biology 59 (1), 169–230.
  • Wu et al. (2007) Wu, Jie-Zhi, Ma, Hui-Yang & Zhou, M-D 2007 Vorticity and vortex dynamics. Springer Science & Business Media.
  • Wu (2011) Wu, Theodore Yaotsu 2011 Fish swimming and bird/insect flight. Annual Review of Fluid Mechanics 43, 25–58.
  • Xu et al. (2018) Xu, Lincheng, Tian, Fang-Bao, Young, John & Lai, Joseph CS 2018 A novel geometry-adaptive cartesian grid based immersed boundary–lattice Boltzmann method for fluid–structure interactions at moderate and high Reynolds numbers. Journal of Computational Physics 375, 22–56.
  • Young et al. (2009) Young, John, Walker, Simon M, Bomphrey, Richard J, Taylor, Graham K & Thomas, Adrian LR 2009 Details of insect wing design and deformation enhance aerodynamic function and flight efficiency. Science 325 (5947), 1549–1552.