Skip to main content
Log in

A time-implicit representation of the lift force for coupled translational–rotational galloping

  • Original paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

The lift force acting on a purely translational galloping oscillator can be well approximated by using the quasi-steady theory, which states that the flow around a galloping body in motion is very similar to the known flow around a fixed body provided a minimum free stream velocity threshold and a similarity principle are satisfied. However, for oscillators undergoing coupled translational–rotational galloping, the rotation of the bluff body breaks the similarity principle, and therewith the quasi-steady assumption. Traditionally, the effects of rotation are accounted for by including explicit time-dependent terms in the lift force. However, we argue that the time-explicit representation of the lift force is unnecessary because the phenomenon of galloping is not time-explicit and acts only as a function of the body motion. Thus, we propose a modified time-implicit lift force representation, which has the same form of the quasi-steady theory, but with an additional dependence on the free stream velocity. The modified lift force is obtained by studying the transient growth of the amplitude of the bluff body oscillation. The proposed approach is used to model the lift force for square and trapezoidal prisms undergoing coupled translational–rotational galloping showing excellent prediction capabilities.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. The reader should bear in mind that perfectly symmetric bodies like a square are prone to galloping. This is because although the production of lift requires asymmetry, the asymmetry does not need to be in the geometry of the bluff body, but can also occur due to the angle of incidence between the moving body and the flow.

  2. The reader can refer to the book by Païdoussis for a detailed review on the galloping instability [6].

References

  1. Den Hartog, J.P.: Mechanical Vibrations. Courier Corporation, North Chelmsford (1985)

    MATH  Google Scholar 

  2. Parkinson, G.: Phenomena and modelling of flow-induced vibrations of bluff bodies. Prog. Aerosp. Sci. 26(2), 169–224 (1989)

    Article  MathSciNet  Google Scholar 

  3. Parkinson, G.V., Smith, J.D.: The square prism as an aeroelastic non-linear oscillator. Q. J. Mech. Appl. Math. 17(2), 225–239 (1964)

    Article  Google Scholar 

  4. Nakamura, Y., Tomonari, Y.: Galloping of rectangular prisms in a smooth and in a turbulent flow. J. Sound Vib. 52(2), 233–241 (1977)

    Article  Google Scholar 

  5. Nakamura, Y., Hirata, K., Urabe, T.: Galloping of rectangular cylinders in the presence of a splitter plate. J. Fluids Struct. 5(5), 521–549 (1991)

    Article  Google Scholar 

  6. Païdoussis, M.P., Price, S.J., De Langre, E.: Fluid–Structure Interactions: Cross-Flow-Induced Instabilities. Cambridge University Press, Cambridge (2010)

    Book  Google Scholar 

  7. Fung, Y.C.: An Introduction to the Theory of Elasticity. Wiley, New York (1955)

    MATH  Google Scholar 

  8. Blevins, R.D.: Flow-Induced Vibration. Van Nostrand Reinhold Co., New York (1977)

    MATH  Google Scholar 

  9. Blevins, R.D.: Flow-Induced Vibration, 2nd edn. Van Nostrand Reinhold Co., New York (1990)

    MATH  Google Scholar 

  10. Bearman, P.W., Gartshore, I.S., Maull, D.J., Parkinson, G.V.: Experiments on flow-induced vibration of a square-section cylinder. J. Fluids Struct. 1(1), 19–34 (1987)

    Article  Google Scholar 

  11. Daqaq, M.F., Bibo, A., Akhtar, I., Alhadidi, A.H., Panyam, M., Caldwell, B., Noel, J.: Micropower generation using cross-flow instabilities: a review of the literature and its implications. J. Vib. Acoust. 141(3), 030801 (2019)

    Article  Google Scholar 

  12. Ryabinin, A.N., Lyusin, V.D.: Galloping of small aspect ratio square cylinder. ARPN J. Eng. Appl. Sci. 10(1), 134–138 (2015)

    Google Scholar 

  13. Ewere, F., Wang, G.: Performance of galloping piezoelectric energy harvesters. J. Intell. Mater. Syst. Struct. 25(14), 1693–1704 (2013)

    Article  Google Scholar 

  14. Nakamura, Y.: On the aerodynamic mechanism of torsional flutter of bluff structures. J. Sound Vib. 67(2), 163–177 (1979)

    Article  Google Scholar 

  15. Van Oudheusden, B.W.: Rotational one-degree-of-freedom galloping in the presence of viscous and frictional damping. J. Fluids Struct. 10(7), 673–689 (1996)

    Article  Google Scholar 

  16. Van Oudheusden, B.W.: Aerodynamic stiffness and damping effects in the rotational galloping of a rectangular cross-section. J. Fluids Struct. 14(8), 1119–1144 (2000)

    Article  Google Scholar 

  17. Van Oudheusden, B.W.: Aerodynamic stiffness effects in rotational galloping at high wind speeds. J. Wind Eng. Ind. Aerodyn. 64(1), 31–46 (1996)

    Article  Google Scholar 

  18. Kluger, J.M., Moon, F.C., Rand, R.H.: Shape optimization of a blunt body vibro-wind galloping oscillator. J. Fluids Struct. 40, 185–200 (2013)

    Article  Google Scholar 

  19. Scanlan, R.H.: Problematics in formulation of wind-force models for bridge decks. J. Eng. Mech. 119(7), 1353–1375 (1993)

    Article  Google Scholar 

  20. Dickinson, M.: The effects of wing rotation on unsteady aerodynamic performance at low Reynolds numbers. J. Exp. Biol. 192(1), 179–206 (1994)

    Google Scholar 

  21. Dickinson, M.H., Lehmann, F.-O., Sane, S.P.: Wing rotation and the aerodynamic basis of insect flight. Science 284(5422), 1954–1960 (1999)

    Article  Google Scholar 

  22. Sane, S.P., Dickinson, M.H.: The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight. J. Exp. Biol. 205(8), 1087–1096 (2002)

    Google Scholar 

  23. Sane, S.P.: The aerodynamics of insect flight. J. Exp. Biol. 206(23), 4191–4208 (2003)

    Article  Google Scholar 

  24. Nayfeh, A.H., Mook, D.T.: Nonlinear Oscillations. Wiley, New York (2008)

    MATH  Google Scholar 

  25. Nayfeh, A.H.: Introduction to Perturbation Techniques. Wiley, New York (2011)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mohammed F. Daqaq.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Equation of motion

Following linear Euler–Bernoulli beam theory for which the thickness, \(t_b\), and deflection w(xt) of the beam are assumed to be much smaller than its length L, we can write the equation governing the motion of the beam as

$$\begin{aligned} M_b \ddot{w}+c_v {\dot{w}} +E_b I_b w^{\prime \prime \prime \prime }=F_N \delta (x-L), \end{aligned}$$
(A.1)

where t represents the temporal variable, and x represents the spatial variable along the length of the beam, the dot represents a derivative with respect to t and the prime represents a derivative with respect to x, \(M_b\) is the mass of the beam, \(c_v\) is the linear damping coefficient, \(E_b\) and \(I_b\) are, respectively, the modulus of elasticity, and mass moment of inertia of the beam, \(F_N\) is the vertical component of the aerodynamic force, and \(\delta \) is the Dirac delta function. Note that the aerodynamic force is assumed to be concentrated at the tip, which implicitly assumes that the width of the bluff body, D, is much smaller than the length of the beam, L.

The proper boundary conditions pertinent to Fig. 2 are given by

$$\begin{aligned} \begin{aligned}&w(0,t)=0, \quad w^{\prime }(0,t)=0, \\&E_bI_b w^{\prime \prime }(L,t)=I_t \ddot{w}^{\prime }(L,t), \\&E_bI_b w^{\prime \prime \prime }(L,t)=M_t \ddot{w}^{\prime }(L,t), \end{aligned} \end{aligned}$$
(A.2)

where \(M_t\) and \(I_t\) are, respectively, the mass and mass moment of inertia of the bluff body.

As shown in Fig. 2, the aerodynamic force acting on the moving tip body is produced by a relative wind velocity \(V_{\mathrm{rel}}\) at an effective angle of attack \(\alpha \), which can be written as

$$\begin{aligned} V_{\mathrm{rel}}=\sqrt{V^2+{\dot{w}}^2(L,t)},\quad \alpha =\alpha _t+w^{\prime }(L,t), \end{aligned}$$
(A.3)

where \(\alpha _t=\tan ^{-1}\left[ \frac{{\dot{w}}(L,t)}{V}\right] \). The corresponding drag and lift force components are given by

$$\begin{aligned} F_D= & {} \frac{1}{2}\rho _aC_{F_D}(\alpha )D H V^2_{\mathrm{rel}},\nonumber \\ F_L= & {} \frac{1}{2}\rho _aC_{F_L}(\alpha )D H V^2_{\mathrm{rel}}, \end{aligned}$$
(A.4)

where \(\rho _a\) is the air density, \(C_{F_D}\) and \(C_{F_L}\) are, respectively, the drag and lift coefficients. The normal force acting in the y-direction can be, therefore, written as

$$\begin{aligned} F_N=F_D \sin \alpha _t + F_L \cos \alpha _t. \end{aligned}$$
(A.5)

When \({\dot{w}}(L,t)/V \ll 1\), the previous equation becomes

$$\begin{aligned} F_N=F_D \frac{{\dot{w}}(L,t)}{V} + F_L. \end{aligned}$$
(A.6)

Furthermore, under the assumption that \({\dot{w}}(L,t)/V \ll 1\), \(V_{\mathrm{rel}}\) can be approximated by V, which yields

$$\begin{aligned} \begin{aligned}&F_N=\frac{1}{2}\rho _aC_{N}(\alpha )D H V^2, \\&\text {where}\quad C_{N}=\left[ C_{F_D}\frac{{\dot{w}}(L,t)}{V}+C_{F_L}\right] \end{aligned} \end{aligned}$$
(A.7)

Assuming that the beam is free of internal resonances and that the fluid forces are incapable of directly exciting the higher vibration modes, a single-mode representation of the beam dynamics can be written as

$$\begin{aligned} w(x,t)= \phi (x) {\bar{y}}(t) \end{aligned}$$
(A.8)

where \(\phi (x)\) is a spatial representation of the beam shape at a given time, chosen here to be the first mode shape of a cantilever beam with a tip mass. The function \({\bar{y}}(t)\) is a temporal function representing the time variation of the mode shape. The function \(\phi (x)\) can be written as

$$\begin{aligned} \phi (x)= & {} A\left[ \cos \frac{\lambda }{L}x-\cosh \frac{\lambda }{L}x \right. \nonumber \\&\left. +\xi \left( \sin \frac{\lambda }{L}x-\sinh \frac{\lambda }{L}x\right) \right] , \end{aligned}$$
(A.9)

where

$$\begin{aligned}&\xi =\frac{\left( \sin {\lambda }-\sinh {\lambda }\right) +R_1\lambda \left[ \cos {\lambda }-\cosh {\lambda }\right] }{\left( \cos {\lambda }-\cosh {\lambda }\right) -R_1\lambda \left[ \sin {\lambda }-\sinh {\lambda }\right] },\nonumber \\&R_1= \frac{M_t}{m_bL}. \end{aligned}$$

Substituting Eq. (A.8) into Eq. (A.1), multiplying by \(\phi (x)\) and integrating over the length of the beam yields

$$\begin{aligned} m\ddot{{\bar{y}}}+c\dot{{\bar{y}}}+k {\bar{y}} =f_0 F_N, \end{aligned}$$
(A.10)

where

$$\begin{aligned} m= & {} \int _0^{L}M_b \phi ^2 \mathrm{d}x, \quad c= \int _0^L c_v \phi ^2 \mathrm{d}x,\nonumber \\ k= & {} 7\int _0^{L}E_b I_b \phi ^{\prime \prime \prime \prime } \phi \mathrm{d}x, \quad f_0= \phi (L). \end{aligned}$$
(A.11)

Normalizing the mode shape, \(\phi (x)\), by choosing A such that \(\phi (L)=1\), yields \(f_0=1\) and \(\alpha = \dot{{\bar{y}}}/V+ \phi ^{\prime }(L) {\bar{y}}\). Furthermore, for small deflections \({\dot{w}}(L,t)/L < 0.3\), the function \(\phi ^{\prime }(L)\) can be further approximated by \(\phi ^{\prime }(L) \approx -\frac{3}{2 L} \phi (L)\) . Thus, Eq. (A.10) can be further simplified as

$$\begin{aligned} \begin{aligned}&m\ddot{{\bar{y}}}+c\dot{{\bar{y}}}+k {\bar{y}} = \frac{1}{2}\rho _aC_{N}(\alpha )D H V^2, \\&\quad \alpha =\frac{\dot{{\bar{y}}}}{V}-\frac{3}{2L}{\bar{y}}, \end{aligned} \end{aligned}$$
(A.12)

which is the same as Eq. (2) used to study the dynamics of the galloping oscillator.

Appendix B: Approximate analytical solution of the equation of motion

An approximate solution of Eq. (6) can be obtained using the method of multiple scales; a method widely utilized to analyze the response of weakly nonlinear systems containing mechanisms that act on distinct time scales [25]. In the case of a galloping body, a “true” time scale describes the instantaneous position of the body, while a slow time scale tracks the weakly nonlinear evolution of oscillation amplitude.

The scaling is accomplished using an arbitrarily small bookkeeping parameter \(\epsilon \) such that “true” time is \(T_0 = \tau \) and “slow” time is \(T_1 = \epsilon \tau \). For the present problem, higher orders of \(\epsilon \) (even slower time scales) are neglected. The derivative operator is adapted such that

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d} \tau }= & {} D_0 + \epsilon D_1 + {\mathcal {O}}(\epsilon ^2), \nonumber \\ \frac{\mathrm{d}^2 }{\mathrm{d}\tau ^2}= & {} D_0^2 + 2\epsilon D_0 D_1 + {\mathcal {O}}(\epsilon ^2) . \end{aligned}$$
(B.1)

where

$$\begin{aligned} D_0 = \frac{\mathrm{d}}{\mathrm{d} T_0},\quad D_1 = \frac{\mathrm{d}}{\mathrm{d} T_1} . \end{aligned}$$
(B.2)

Additionally, the response \(y(\tau )\) is described into multiple scales such that

$$\begin{aligned} y(\tau ) = y_0 (T_0,T_1) + \epsilon y_1 (T_0,T_1) + {\mathcal {O}}(\epsilon ^2). \end{aligned}$$
(B.3)

It can be reasoned that the mechanism of amplitude growth occurs slowly because the energy input to the system is on the order of the energy dissipated in the context of underdamped vibrations, energy dissipation must be small (controlled by \(\zeta _m\)). As such, a small net energy input requires small input flow energy (controlled by \(\mu \)). Therefore, the appropriate parameters are scaled by \(\epsilon \) as

$$\begin{aligned} \mu = \epsilon \mu ,\quad \zeta _m=\epsilon \zeta _m. \end{aligned}$$
(B.4)

When Eqs. (B.1), (B.3), and (B.4) are substituted into Eq. (6), the resulting system can be divided into subsystems by collecting the zeroth and first power of \(\epsilon \) as following:

For \(\epsilon ^0\):

$$\begin{aligned} D_0^2 y_0 + y_0 = 0 \end{aligned}$$
(B.5)

admits a zeroth-order solution

$$\begin{aligned} y_0 = a(T_1) \cos (T_0 + \gamma (T_1)), \end{aligned}$$
(B.6)

where a is the amplitude of oscillation and \(\gamma \) is a phase shift function.

For \(\epsilon ^1\):

$$\begin{aligned} D_1^2 y_1 + y_1 = -2D_0D_1 y_0 - 2\zeta _m D_0y_0 + 2\mu U^2 C_N(\alpha ). \end{aligned}$$
(B.7)

Letting \(\phi = T_0 + \gamma \) and substituting the zeroth-order solutions, as well as the expression for \(C_N\) from Eq. (3) into Eq. (B.7) yields

$$\begin{aligned} \begin{aligned}&D_1^2 y_1 + y_1 = 2D_1a\sin \phi +2aD_1\beta \cos \phi + 2\zeta _m a\sin \phi \\&\quad {+2\mu U^2} \sum _{\begin{array}{c} j=1\\ \text {odd} \end{array}}^{n} A_j \left( \alpha \right) ^j + {2\mu U^2} \sum _{\begin{array}{c} j=1\\ \text {even} \end{array}}^{n} A_j \left| \alpha _0\right| ^j\mathrm {sgn}(\alpha ). \end{aligned} \end{aligned}$$
(B.8)

Because the physical system does not have oscillations which grow to infinity, the “forcing” terms on the right-hand side of Eq. (B.8) must not resonate with \(y_1\). This condition is enforced by ensuring the right-hand side of Eq. (B.8) is orthogonal to its homogeneous solution throughout one cycle of oscillation. That is, the right-hand side of the equation should be orthogonal to both \(\sin \phi \) and \(\cos \phi \).

For orthogonality with \(\cos \phi \), we obtain

$$\begin{aligned} \begin{aligned}&\int _0^{2\pi } \Bigg ( 2D_1a\sin \phi +2aD_1\beta \cos \phi - 2\zeta _m a\sin \phi \\&\quad + {2\mu U^2} \sum _{\begin{array}{c} j=1\\ \text {odd} \end{array}}^{n} A_j \left( \alpha \right) ^j \\&\quad + \sum _{\begin{array}{c} j=1\\ \text {even} \end{array}}^{n} A_j \left| \alpha \right| ^j\mathrm {sgn}(\alpha ) \Bigg ) \cos \phi \text {d}\phi =0. \end{aligned} \end{aligned}$$
(B.9)

Noting that \(\alpha =\frac{y^{\prime }}{U}-\xi U\), and substituting Eq. (B.6) into the expression for \(\alpha \) we obtain

$$\begin{aligned} \alpha = -\frac{a}{U}\sqrt{\xi ^2 U^2 + 1\;\;}\;\sin \left( \phi + \tan ^{-1}\left( \xi U \right) \right) , \end{aligned}$$
(B.10)

which implies that the angle of attack changes sign when

$$\begin{aligned} \phi = k\pi - \tan ^{-1}\left( \xi U\right) , \quad k \equiv \text {integer}. \end{aligned}$$
(B.11)

Thus, Eq. (B.9) becomes

$$\begin{aligned} \begin{aligned} 0&=\int _0^{2\pi } \left[ 2D_1a\sin \phi +2aD_1\beta \cos \phi \right. \\&\quad \left. - 2\zeta _m a\sin \phi \right] \cos \phi \text {d}\phi \\&\quad + 2\mu U^2 \sum _{j=1}^{n} A_j \left( \frac{a}{U}\sqrt{\left( \xi U\right) ^2 + 1}\right) ^j \\&\quad \times \Bigg \{ -\int _0^{\pi -\theta _0}\sin \left( \phi + \theta _0\right) ^j\cos \phi \text {d}\phi \\&\quad +(-1)^j \int _{\pi -\theta _0}^{2\pi -\theta _0}\sin \left( \phi + \theta _0\right) ^j\cos \phi \text {d}\phi \\&\quad - \int _{2\pi -\theta _0}^{2\pi }\sin \left( \phi + \theta _0\right) ^j\cos \phi \text {d}\phi \Bigg \}, \end{aligned} \end{aligned}$$
(B.12)

where \(\theta _0=\tan ^{-1}\left( \xi U\right) \). Upon carrying the integration in Eq. (B.12), we obtain

$$\begin{aligned} a \frac{\mathrm{d} \gamma }{\mathrm{d} \tau } = - 2\mu U^2 \sum _{j=1}^{n} A_j \left( \frac{a}{U}\right) ^j \left( \sqrt{\xi ^2 U^2 + 1}\right) ^{j-1} C_j, \end{aligned}$$
(B.13)

where the \(C_j\) are given by

$$\begin{aligned} \begin{aligned} C_j&= \frac{1-(-1)^{j}}{2^{j+2}}\left( \frac{(j+1)!}{\left( \left( \frac{1}{2}(j+1)\right) !\right) ^2}\right) \\&\quad + \frac{1+(-1)^{j}}{2^j\pi }\sum _{k=0}^{j/2}\frac{(j+1)!\,(-1)^{\frac{1}{2}j-k}}{(j+1-k)!\,k!\,(j+1-2k)}. \end{aligned} \end{aligned}$$
(B.14)

Following a similar procedure to enforce the orthogonality with \(\sin \phi \) yields

$$\begin{aligned} \frac{\text {d} a}{\text {d} \tau } = - \zeta _ma + 2\mu U^2 \sum _{j=1}^{n} A_j \left( \frac{a}{U}\right) ^j \left( \sqrt{\xi ^2 U^2 + 1}\right) ^{j-1} C_j. \end{aligned}$$
(B.15)

At steady state, \(\frac{\text {d} \gamma }{\text {d} \tau } = \frac{\text {d} a}{\text {d} \tau } = 0\). Root-solving algorithms can then be used to solve Eqs. (B.13) and (B.15) for the steady-state phase shift and amplitude of the response, respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Noel, J., Alhadidi, A.H., Alhussien, H. et al. A time-implicit representation of the lift force for coupled translational–rotational galloping. Nonlinear Dyn 103, 2183–2196 (2021). https://doi.org/10.1007/s11071-021-06232-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-021-06232-6

Keywords

Navigation