Skip to main content

Advertisement

Log in

Fluidic energy harvesting beams in grid turbulence

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

Much of the recent research involving fluidic energy harvesters based on piezoelectricity has focused on excitation through vortex-induced vibration while turbulence-induced excitation has attracted very little attention, and virtually no previous work exists on excitation due to grid-generated turbulence. The present experiments involve placing several piezoelectric cantilever beams of various dimensions and properties in flows where turbulence is generated by passive, active, or semi-passive grids, the latter having a novel design that significantly improves turbulence generation compared to the passive grid and is much less complex than the active grid. We experimentally show for the first time that the average power harvested by a piezoelectric cantilever beam placed in decaying isotropic, homogeneous turbulence depends on mean velocity, velocity and length scales of turbulence as well as the electromechanical properties of the beam. The output power can be modeled as a power law with respect to the distance of the beam from the grid. Furthermore, we show that the rate of decay of this power law closely follows the rate of decay of the turbulent kinetic energy. We also introduce a forcing function used to model approximately the turbulent eddies moving over the cantilever beam and observe that the feedback from the beam motion onto the flow is virtually negligible for most of the cases considered, indicating an effectively one-way interaction for small-velocity fluctuations.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  • Akaydin HD, Elvin N, Andreopoulos Y (2010) Wake of a cylinder: a paradigm for energy harvesting with piezoelectric materials. Exp Fluids 49:291–304. doi:10.1007/s00348-010-0871-7

    Article  Google Scholar 

  • Akaydin HD, Elvin N, Andreopoulos Y (2010) Energy harvesting from highly unsteady fluid flows using piezoelectric materials. J Intell Mater Syst Struct 21:1263–1278. doi:10.1177/1045389X10366317

    Article  Google Scholar 

  • Batchelor GK, Townsend AA (1948) Decay of isotropic turbulence in the initial period. Proc R Soc Lond A Math Phys Eng Sci 193:539–558. doi:10.1098/rspa.1948.0061

    Article  MATH  Google Scholar 

  • Batchelor GK (1959) The theory of homogeneous turbulence. Cambridge University Press, Cambridge

    Google Scholar 

  • Bryant M, Garcia E (2011) Modeling and testing of a novel aeroelastic flutter energy harvester. J Vib Acoust 133:011010

    Article  Google Scholar 

  • Comte-Bellot G, Corrsin S (1966) The use of a contraction to improve the isotropy of grid-generated turbulence. J Fluid Mech 25:657–682

    Article  Google Scholar 

  • Comte-Bellot G, Corrsin S (1971) Simple Eulerian time correlation of full-and narrow-band velocity signals in grid generated, ‘isotropic’ turbulence. J Fluid Mech 48:273–337

    Article  Google Scholar 

  • Dai HL, Abdelkafi A, Wang L (2014) Piezoelectric energy harvesting from concurrent vortex-induced vibrations and base excitations. Nonlinear Dyn 77:1–15. doi:10.1007/s11071-014-1355-8

    Article  Google Scholar 

  • Danesh-Yazdi AH, Elvin N, Andreopoulos Y (2014) Green’s function method for piezoelectric energy harvesting beams. J Sound Vib 333:3092–3108. doi:10.1016/j.jsv.2014.02.023

    Article  Google Scholar 

  • Elvin NG, Elvin AA (2009) A general equivalent circuit model for piezoelectric generators. J Intell Mater Syst Struct 20:3–9

    Article  Google Scholar 

  • Erturk A, Inman DJ (2008) A distributed parameter electromechanical model for cantilevered piezoelectric energy harvesters. J Vib Acoust 130:041002. doi:10.1115/1.2890402

    Article  Google Scholar 

  • Erturk A, Viera WGR, De Marqui C Jr, Inman DJ (2010) On the energy harvesting potential of piezoaeroelastic systems. Appl Phys Lett 96:184103. doi:10.1063/1.3427405

    Article  Google Scholar 

  • Goushcha O, Akaydin HD, Elvin N (2015) Andreopoulos Y energy harvesting prospects in turbulent boundary layers by using piezoelectric transduction. J Fluids Struct 54:823–847

    Article  Google Scholar 

  • Hobeck JD, Inman DJ (2010) Artificial piezoelectric grass for energy harvesting from turbulence-induced vibration. Smart Mater Struct 21:105024. doi:10.1088/0964-1726/21/10/105024

    Article  Google Scholar 

  • Kolmogorov AN (1941) The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. Dokl Akad Nauk SSSR 30:299–303

    Google Scholar 

  • Makita H (1991) Realization of a large-scale turbulence field in a small wind tunnel. Fluid Dyn Res 8:1–4

    Article  Google Scholar 

  • Mehmood A, Abdelkafi A, Hajj MR, Nayfeh AH, Akhtar I, Nuhait AO (2013) Piezoelectric energy harvesting from vortex-induced vibrations of circular cylinder. J Sound Vib 332:4656–4667

    Article  Google Scholar 

  • Mohamed MS (1989) Experimental investigation of the near and far fields of the turbulent velocity and thermal mixing layer in nearly homogeneous turbulence. Dissertation, University of California, Irvine

  • Mohamed MS, Larue JC (1990) The decay power law in grid-generated turbulence. J Fluid Mech 219:195–214

    Article  Google Scholar 

  • Mydlarski L, Warhaft Z (1996) On the onset of high-Reynolds-number grid-generated wind tunnel turbulence. J Fluid Mech 320:331–368

    Article  Google Scholar 

  • Myers R, Vickers M, Kim H, Priya S (2007) Small scale windmill. Appl Phys Lett 90:054106. doi:10.1063/1.2435346

    Article  Google Scholar 

  • Pope SB (2000) Turbulent flows. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Sodano HA, Inman DJ, Park G (2004) Estimation of electric charge output for piezoelectric energy harvesting. Strain 40:49–58

    Article  Google Scholar 

  • Tan-Atichat J, Nagib HM, Loehrke RI (1982) Interaction of free-stream turbulence with screens and grids: a balance between turbulence scales. J Fluid Mech 114:501–528

    Article  Google Scholar 

  • Vonlanthen R, Monkewitz PA (2011) A novel tethered-sphere add-on to enhance grid turbulence. Exp Fluids 51:579–585

    Article  Google Scholar 

Download references

Acknowledgments

The authors wish to thank Professor John Larue and his student Alejandro Puga for allowing access to the active grid wind tunnel at the Turbulence Research Laboratory at the University of California, Irvine. We are also grateful to Mr. Loïc Boiton-Margant, Mr. Hugo Kuntz, and Mr. Nicolas Fornes for their help with the construction of the passive and semi-passive grids. The present work is supported by the National Science Foundation under Grant No. CBET 1033117.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Y. Andreopoulos.

Additional information

This article belongs to a Topical Collection of articles entitled Extreme Flow Workshop 2014. Guest editors: I. Marusic and B. J. McKeon.

Appendix: Average piezoelectric power decay

Appendix: Average piezoelectric power decay

To relate the average power of the piezoelectric cantilever beam to the distance from the grid, the coupled electromechanical equations (Elvin and Elvin 2009; Sodano et al. 2004) are utilized:

$$\begin{aligned} {{\mathbf {M}}}{\ddot{\mathbf {r}}}+{\mathbf {D}}{\dot{\mathbf {r}}}+ {\mathbf {Kr}}-{\varvec{\Theta }}v_p= \,& {} {\mathbf {g}} \end{aligned}$$
(12)
$$\begin{aligned} C_p{\dot{v}}_p+\frac{v_p}{R_l}+{\varvec{\Theta }}^T {\dot{\mathbf {r}}}=0 \end{aligned}$$
(13)

where \({{\mathbf {M}}}\), \({\mathbf {D}}\) and \({{\mathbf {K}}}\) are the mass, damping and stiffness matrices, respectively, and \({\varvec{\Theta }}\) is the piezoelectric coupling vector, \({\mathbf {g}}\) is the effective forcing vector, \(C_p\) is the capacitance of the piezoelectric layer, \(v_p\) is the voltage generated by the piezoelectric layer. The piezoelectric harvester is assumed to be attached to an external resistance \(R_l\). The local displacement along the beam \(y(X,x/M;t)\) can be expressed as:

$$\begin{aligned} y(X,x/M;t) =\sum _{j=1}^{N} \phi _j(X)r_j(x/M;t) = \varvec{\Phi }^T(X){\mathbf {r}}(x/M;t) \end{aligned}$$
(14)

where \({\varvec{\Phi }}(X)\) is a vector of assumed mode shapes for the beam, \({\mathbf {r}}(x/M;t)\) is the time-dependent component of the displacement vector, and N is the number of modes under consideration.

The \(j{{\rm th}}\) entry in the effective forcing vector, \(g_j(x/M;t)\), is given as:

$$\begin{aligned} g_j(x/M;t)= \frac{1}{L} \int _0^L {\phi _j(X)f(X,x/M;t)}\, {{\rm d}}X \end{aligned}$$
(15)

where \(\phi _j(X)\) is the \(j{{\rm th}}\) entry of the \(\varvec{\Phi }(X)\) vector and is defined in Erturk and Inman (2008).

Thus, from (12) and (13), the voltage in the frequency domain, \(V_p(x/M;\omega )\), is found to be:

$$\begin{aligned} V_p(x/M;\omega )={\mathbf {H}}_v(X;\omega ){\mathbf {G}}(x/M;\omega ) \end{aligned}$$
(16)

where

$$\begin{aligned} {\mathbf {H}}_v(\omega )= & {} \frac{1}{C_p}\left[ \frac{i\omega R_lC_p}{ 1+i\omega R_lC_p}\right] \varvec{\Theta }^T\left[ {\mathbf {K}}- \omega ^2{\mathbf {M}} +i\omega \mathbf {D} \right. \nonumber \\&\left. +\,\frac{1}{C_p}\left[ \frac{i\omega R_lC_p}{1+i\omega R_lC_p} \right] \varvec{\Theta }\varvec{\Theta }^T\right] ^{-1} \end{aligned}$$
(17)

and

$$\begin{aligned} G_j(x/M;\omega )= \frac{1}{L} \int _0^L {\phi _j(X) F(X,x/M;\omega )}\, {{\rm d}}X \end{aligned}$$
(18)

We will limit the analysis to the first mode only, since the beams typically vibrate at or near the first mode. Thus, the expected power \({\mathscr {E}}\left\{ P(x/M)\right\} \) can be expressed as:

$$\begin{aligned} {\mathscr {E}}\left\{ P(x/M)\right\}= & {} \int _{-\infty }^\infty { P(x/M;\omega )}\, {{\rm d}}\omega = \int _{-\infty }^\infty \frac{|V_p(x/M;\omega )|^2}{R_l} {{\rm d}}\omega \nonumber \\= & {} \frac{1}{R_l}\int _{-\infty }^\infty {H_v{\widehat{H_v}} G_1{\widehat{G_1}}}\, {{\rm d}}\omega \end{aligned}$$
(19)

where \({\widehat{H_v}}(\omega )\) and \({\widehat{G_1}}(x/M;\omega )\) are complex conjugates of \(H_v(\omega )\) and \(G_1(x/M;\omega )\), respectively.

Equation (19) shows that the expected power generated in the beam primarily depends on x/M. The x/M dependence of the expected power is manifested through the term \(G_1{\widehat{G_1}}=|G_1|^2\). Parseval’s theorem can be applied to the \(|G_1|^2\) term to give:

$$\begin{aligned} \int _{-\infty }^\infty {|g_1(x/M;t)|^2} {{\rm d}}t = \frac{1}{2\pi } \int _{-\infty }^\infty {|G_1(x/M;\omega )|^2} {{\rm d}}\omega \end{aligned}$$
(20)

From (20), the x/M dependence of \(|G_1(x/M;\omega )|^2\) is equivalent to that of \(|g_1(x/M;t)|^2\). Thus, using the definition of average power, one can conclude that, with respect to x/M:

$$\begin{aligned} {\overline{P}}(x/M)&\propto {\mathscr {E}}\left\{ P(x/M)\right\} \propto |G_1(x/M;\omega )|^2 \nonumber \\&\propto \sigma ^2_{g_1}(x/M) \end{aligned}$$
(21)

where \(\sigma ^2_{g_1}(x/M)=\overline{|g_1(x/M)|^2}\) is the variance of \(g_1(x/M;t)\).

The effect of the turbulence length scale does not explicitly enter into the result in (21), even though we expect it to be a factor since the beam vibrates because of the presence of turbulence in the flow. The turbulence length scale can be included in (21) by expanding on the relation between the average power and \(\sigma ^2_{g_1}(x/M)\):

$$\begin{aligned} {\overline{P}} \propto \sigma ^2_{g_1}(x/M) = \frac{1}{L^2} \int _0^L \int _0^L {\phi _1(X_1)\phi _1(X_2) \overline{f(X_1)f(X_2)}}\, {{\rm d}}X_1 {{\rm d}}X_2 \end{aligned}$$
(22)

The time-averaged term in the integrand of (22) can be expressed in terms of a two-point force (pressure) correlation:

$$\begin{aligned} J(X_1,X_2,x/M)=\overline{f(X_1,x/M)f(X_2,x/M)} \end{aligned}$$
(23)

From Comte-Bellot and Corrsin (1971) and similar to the approach in Goushcha et al. (2015) for the flow in the turbulent boundary layer, the two-point force correlation in isotropic turbulent flow is approximated as an exponential decay, such that:

$$\begin{aligned} J=\gamma (X,x/M){{\rm e}}^{-a(x/M)\left| X_1-X_2\right| } \end{aligned}$$
(24)

where \(\gamma (X,x/M)\) and a(x/M) are parameters of the exponential decay. If \(X_1=X_2\), equating (23) and (24) leads to:

$$\begin{aligned} \gamma (X,x/M)=\sigma ^2_f(X,x/M) \end{aligned}$$
(25)

Note that in (25), as in Sect. 3.4, the variance of the forcing function \(\sigma ^2_f\) is assumed to be weakly dependent on the beam coordinate X. Thus, \(\gamma (X,x/M)=\gamma (x/M)\). We posit that a(x/M) is inversely proportional to the integral length-scale \(L_I(x/M)\), which characterizes the largest energy-carrying eddies. Thus,

$$\begin{aligned} a(x/M)=\frac{q}{L_I(x/M)} \end{aligned}$$
(26)

where q is a proportionality constant. The reasoning behind the definition in (26) is that we assume that the two-point correlation in (24) is effectively defined by the energy-carrying eddies, i.e., two points that are separated by a distance larger than \(L_I\) are virtually uncorrelated. The proportionality constant q in (26) is defined in a manner such that when \(|X_1-X_2|=L_I\), the correlation function is a small percentage of the force variance, i.e., \(J=0.2\sigma ^2_f\). In this case, \(q=-\ln (0.2)=1.61\). Substituting (24)–(26) into (22) and evaluating the integral results in:

$$\begin{aligned} {\overline{P}}(x/M) \propto \xi (\ell ) \sigma ^2_f(x/M) \end{aligned}$$
(27)

where \(\xi (\ell )\) is referred to as the local variation parameter, which is a known function of the length ratio \(\ell \) that is shown in Fig. 16, and

$$\begin{aligned} \ell =\frac{L}{L_I(x/M)} \end{aligned}$$
(28)

It is important to note that, from Fig. 16, the local variation parameter \(\xi (\ell )\) has a maximum value of \(\xi _{{\rm max}}=\xi (0)=0.153\). This rather unusual number, which is independent of q, arises from the integrand of (22), where the first-mode shape function \(\phi _1(X)\), as defined by Elvin and Elvin (2009), is included.

The variance of the forcing function in (8) is found using Eqs. (1)–(2) to be:

$$\begin{aligned} \sigma ^2_{f}(x/M)=\rho _f^{2} A_f^{2} C_F^{2}{\overline{U}}^{2} {\overline{v^2}}=\frac{2}{3}\rho _f^{2} A_f^{2} C_F^{2} {\overline{U}}^{2} \alpha _k\left( \frac{x}{M}-\frac{x_{0,k}}{M}\right) ^{\beta _{k}} \end{aligned}$$
(29)

From (27) and (29), it is clear that the average power harvested by the piezoelectric cantilever beam displays a power-law behavior in grid turbulence with respect to the distance from the grid. Therefore, we hypothesize that:

$$\begin{aligned} {\overline{P}}(x/M)&= \kappa \left( \rho _f,A_f,C_F,{\overline{U}}, \alpha _k,\ell \right) \left( \frac{x}{M}-\frac{x_{0,k}}{M}\right) ^{\beta _{k}} \nonumber \\&= \alpha _P\left( \frac{x}{M}-\frac{x_{0,P}}{M}\right) ^{\beta _{P}} \end{aligned}$$
(30)

where \(\kappa \) is the proportionality constant of the power law. In this theoretical approach, the exponent of the mean TKE power-law expression (2) is identical to the exponent of the average piezoelectric power-law equation (30), provided that \(C_F\) is independent of x/M.

Fig. 16
figure 16

Function \(\xi (\ell )\) with respect to \(\ell \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Danesh-Yazdi, A.H., Goushcha, O., Elvin, N. et al. Fluidic energy harvesting beams in grid turbulence. Exp Fluids 56, 161 (2015). https://doi.org/10.1007/s00348-015-2027-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-015-2027-2

Keywords

Navigation