Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-24T08:16:28.764Z Has data issue: false hasContentIssue false

The influence of fluid–structure interaction on cloud cavitation about a flexible hydrofoil. Part 2.

Published online by Cambridge University Press:  17 June 2020

Samuel M. Smith*
Affiliation:
Australian Maritime College, University of Tasmania, Launceston, TAS 7250, Australia
James A. Venning
Affiliation:
Australian Maritime College, University of Tasmania, Launceston, TAS 7250, Australia
Bryce W. Pearce
Affiliation:
Australian Maritime College, University of Tasmania, Launceston, TAS 7250, Australia
Yin Lu Young
Affiliation:
Department of Naval Architecture and Marine Engineering, University of Michigan, Ann Arbour, MI 48109, USA
Paul A. Brandner
Affiliation:
Australian Maritime College, University of Tasmania, Launceston, TAS 7250, Australia
*
Email address for correspondence: ssmith18@utas.edu.au

Abstract

The influence of fluid–structure interaction on cloud cavitation about a hydrofoil is investigated by comparing results from a relatively stiff reference hydrofoil, presented in Part 1, with those obtained on a geometrically identical flexible hydrofoil. Measurements were conducted with a chord-based Reynolds number $Re=0.8\times 10^{6}$ for cavitation numbers, $\unicode[STIX]{x1D70E}$, ranging from 0.2 to 1.2 while the hydrofoil was mounted at an incidence, $\unicode[STIX]{x1D6FC}$, of $6^{\circ }$ to the oncoming flow. Tip deformations and cavitation behaviour were recorded with synchronised force measurements utilising two high-speed cameras. The flexible composite hydrofoil was manufactured as a carbon/glass-epoxy hybrid structure with a lay-up sequence selected principally to consider spanwise bending deformations with no material-induced bend–twist coupling. Hydrodynamic bend–twist coupling is seen to result in nose-up twist deformations causing frequency modulation from the increase in cavity length. The lock-in phenomenon driven by re-entrant jet shedding observed on the stiff hydrofoil is also evident on the flexible hydrofoil at $0.70\leqslant \unicode[STIX]{x1D70E}\leqslant 0.75$, but occurs between different modes. Flexibility is observed to accelerate cavitation regime transition with reducing $\unicode[STIX]{x1D70E}$. This is seen with the rapid growth and influence the shockwave instability has on the forces, deflections and cavitation behaviour on the flexible hydrofoil, suggesting structural behaviour plays a significant role in modifying cavity physics. The reduced stiffness causes secondary lock-in of the flexible hydrofoil’s one-quarter sub-harmonic, $f_{n}/4$, at $\unicode[STIX]{x1D70E}$ = 0.4. This leads to the most severe deflections observed in the conditions tested along with a shift in phase between normal force and tip deflection.

Type
JFM Papers
Copyright
© The Author(s), 2020. Published by Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Akcabay, D. T., Chae, E. J., Young, Y. L., Ducoin, A. & Astolfi, J. A. 2014 Cavity induced vibration of flexible hydrofoils. J. Fluids Struct. 49 (Supplement C), 463484.CrossRefGoogle Scholar
Akcabay, D. T. & Young, Y. L. 2014 Influence of cavitation on the hydroelastic stability of hydrofoils. J. Fluids Struct. 49, 170185.CrossRefGoogle Scholar
Akcabay, D. T. & Young, Y. L. 2015 Parametric excitations and lock-in of flexible hydrofoils in two-phase flows. J. Fluids Struct. 57, 344356.CrossRefGoogle Scholar
Ashkenazi, Y., Golfman, I., Rezhkov, L. & Sidorov, N. 1974 Glass-Fiber-Reinforced Plastic Parts in Ship Machinery. Sudostroyenniye Publishing House.Google Scholar
Ausoni, P., Farhat, M., Escaler, X., Egusquiza, E. & Avellan, F. 2007 Cavitation influence on von Karman vortex shedding and induced hydrofoil vibrations. Trans. ASME J. Fluids Engng 129 (8), 966973.CrossRefGoogle Scholar
Brandner, P. A., Lecoffre, Y. & Walker, G. J. 2007 Design considerations in the development of a modern cavitation tunnel. In 16th Australasian Fluid Mechanics Conference, pp. 630637. School of Engineering, University of Queensland.Google Scholar
Clarke, D. B., Butler, D., Crowley, B. & Brandner, P. A. 2014 High-speed full-field deflection measurements on a hydrofoil using digital image correlation. In 30th Symposium on Naval Hydrodynamics, pp. 113. Office of Naval Research.Google Scholar
Ducoin, A., Astolfi, J. A. & Sigrist, J. 2012 An experimental analysis of fluid structure interaction on a flexible hydrofoil in various flow regimes including cavitating flow. Eur. J. Mech. (B/Fluids) 36, 6374.CrossRefGoogle Scholar
Ducoin, A. & Young, Y. L. 2013 Hydroelastic response and stability of a hydrofoil in viscous flow. J. Fluids Struct. 38, 4057.CrossRefGoogle Scholar
Harwood, C., Felli, M., Falchi, M., Garg, N., Ceccio, S. L. & Young, Y. L. 2019 The hydroelastic response of a surface-piercing hydrofoil in multiphase flows. Part 1. Passive hydroelasticity. J. Fluid Mech. 881, 313364.CrossRefGoogle Scholar
Harwood, C., Felli, M., Falchi, M., Garg, N., Ceccio, S. L. & Young, Y. L. 2020 The hydroelastic response of a surface-piercing hydrofoil in multiphase flows. Part 2. Modal parameters and generalized fluid forces. J. Fluid Mech. 884, A3.CrossRefGoogle Scholar
Kato, K., Dan, H. & Matsudaira, Y. 2006 Lock-in phenomenon of pitching hydrofoil with cavitation breakdown. JSME Intl J. Ser. (B/Fluids) 49 (3), 797805.CrossRefGoogle Scholar
Le, Q., Franc, J. & Michel, J. 1993 Partial cavities: global behaviour and mean pressure distribution. Trans. ASME J. Fluids Engng 115, 243243.CrossRefGoogle Scholar
Leroux, J., Astolfi, J. A. & Billard, J. Y. 2004 An experimental study of unsteady partial cavitation. Trans. ASME J. Fluids Engng 126 (1), 94101.CrossRefGoogle Scholar
Leroux, J., Coutier-Delgosha, O. & Astolfi, J. 2005 A joint experimental and numerical study of mechanisms associated to instability of partial cavitation on two-dimensional hydrofoil. Phys. Fluids 17 (5), 052101.CrossRefGoogle Scholar
Liao, Y., Martins, J. & Young, Y. L. 2019 Sweep and anisotropy effects on the viscous hydroelastic response of composite hydrofoils. Compos. Struct. 230, 111471.CrossRefGoogle Scholar
Motley, M. R., Liu, Z. & Young, Y. L. 2009 Utilizing fluid–structure interactions to improve energy efficiency of composite marine propellers in spatially varying wake. Compos. Struct. 90 (3), 304313.CrossRefGoogle Scholar
Mouritz, A. P., Gellert, E., Burchill, P. & Challis, K. 2001 Review of advanced composite structures for naval ships and submarines. Compos. Struct. 53 (1), 2142.CrossRefGoogle Scholar
Pearce, B. W., Brandner, P. A., Garg, N., Young, Y. L., Phillips, A. W. & Clarke, D. B. 2017 The influence of bend-twist coupling on the dynamic response of cavitating composite hydrofoils. In 5th International Symposium on Marine Propulsors (SMP17), pp. 803813. VTT Technical Research Center of Finland Ltd.Google Scholar
Phillips, A. W., Cairns, R., Davis, C., Norman, P., Brandner, P. A., Pearce, B. W. & Young, Y. 2017 Effect of material design parameters on the forced vibration response of composite hydrofoils in air and in water. In 5th International Symposium on Marine Propulsors (SMP17), pp. 813822. VTT Technical Research Center of Finland Ltd.Google Scholar
Shamsborhan, H., Coutier-Delgosha, O., Caignaert, G. & Nour, F. A. 2010 Experimental determination of the speed of sound in cavitating flows. Exp. Fluids 49 (6), 13591373.CrossRefGoogle Scholar
Smith, S. M., Venning, J. A., Pearce, B. W., Young, Y. L. & Brandner, P. A. 2020 The influence of fluid-structure interaction on cloud cavitation about a stiff hydrofoil. Part 1. J. Fluid Mech. 896, A1.CrossRefGoogle Scholar
Smith, S. M., Venning, J. A., Brandner, P. A., Pearce, B. W., Giosio, D. R. & Young, Y. L. 2018 The influence of fluid–structure interaction on cloud cavitation about a hydrofoil. In Proceedings of the 10th International Symposium on Cavitation (CAV2018). ASME Press.Google Scholar
Smith, S. M., Venning, J. A., Giosio, D. R., Brandner, P. A., Pearce, B. W. & Young, Y. L. 2019b Cloud cavitation behavior on a hydrofoil due to fluid–structure interaction. Trans. ASME J. Fluids Engng 141 (4), 041105.CrossRefGoogle Scholar
Towne, A., Schmidt, O. T. & Colonius, T. 2018 Spectral proper orthogonal decomposition and its relationship to dynamic mode decomposition and resolvent analysis. J. Fluid Mech. 847, 821867.CrossRefGoogle Scholar
Turnock, S. R. & Wright, A. M. 2000 Directly coupled fluid structural model of a ship rudder behind a propeller. Mar. Struct. 13 (1), 5372.CrossRefGoogle Scholar
Wu, Q., Huang, B., Wang, G. & Gao, Y. 2015 Experimental and numerical investigation of hydroelastic response of a flexible hydrofoil in cavitating flow. Intl J. Multiphase Flow 74, 1933.CrossRefGoogle Scholar
Young, Y. L. 2007 Time-dependent hydroelastic analysis of cavitating propulsors. J. Fluids Struct. 23 (2), 269295.CrossRefGoogle Scholar
Young, Y. L. 2008 Fluid-structure interaction analysis of flexible composite marine propellers. J. Fluids Struct. 24 (6), 799818.CrossRefGoogle Scholar
Young, Y. L., Garg, N., Brandner, P. A., Pearce, B. W., Butler, D., Clarke, D. & Phillips, A. W. 2018a Load-dependent bend-twist coupling effects on the steady-state hydroelastic response of composite hydrofoils. Compos. Struct. 189, 398418.CrossRefGoogle Scholar
Young, Y. L., Garg, N., Brandner, P. A., Pearce, B. W., Butler, D., Clarke, D. & Phillips, A. W. 2018b Material bend-twist coupling effects on cavitating response of composite hydrofoils. In 10th International Cavitation Symposium (CAV2018). ASME.Google Scholar
Young, Y. L., Harwood, C. M., Montero, F. M., Ward, J. C. & Ceccio, S. L. 2017 Ventilation of lifting bodies: review of the physics and discussion of scaling effects. Appl. Mech. Rev. 69 (1), 010801.Google Scholar
Young, Y. L., Motley, M. R., Barber, R., Chae, E. J. & Garg, N. G. 2016 Adaptive composite marine propulsors and turbines: progress and challenges. Appl. Mech. Rev. 68 (6), 060803.Google Scholar
Zarruk, G. A., Brandner, P. A., Pearce, B. W. & Phillips, A. W. 2014 Experimental study of the steady fluid–structure interaction of flexible hydrofoils. J. Fluids Struct. 51, 326343.CrossRefGoogle Scholar

Smith et al. supplementary movie 1

At σ = 1.20, the flexible hydrofoils increased effective incidence due to twist deformations causes cavity break-up to interchange between interfacial instabilities and re-entrant jet formation over time on the flexible hydrofoil. Small scale vapour structures can be seen shedding from positions along the span due to re-entrant jet formation which is only observed on the stiff hydrofoil once σ is reduced to 1.0. The cavity is also seen to still be broken up to due to inter-facial instabilities such as Kelvin-Helmholtz driven spanwise vorticity lines and turbulent transition.

Download Smith et al. supplementary movie 1(Video)
Video 6.4 MB

Smith et al. supplementary movie 2

At σ = 1.00, the flexible hydrofoil is experiencing Type IIa shedding where the re-entrant jet mechanism is confined to around mid-span due to three-dimensional flows effects as mentioned in part 1. Shedding is seen to occur at approximately St = 0.61 which is lower compared to the stiff hydrofoil which experiences shedding at St =74.

Download Smith et al. supplementary movie 2(Video)
Video 6.3 MB

Smith et al. supplementary movie 3

Reduction in σ down to 0.8 sees the cavity along with the re-entrant jet thickness grow, giving it enough momentum to overcome spanwise flow components and reach the leading edge for majority of the span. This results in the formation of two shedding sites with the Type IIa and IIb modes both active.

Download Smith et al. supplementary movie 3(Video)
Video 7.9 MB

Smith et al. supplementary movie 4

Further reduction in σ to 0.7 sees the hydrofoil enter lock-in where the Type IIb tip shedding frequency matches the first mode of the hydrofoil leading to amplified structural excitations. The Type IIa shedding is still evident in the upper portion of the span while the Type IIb shedding events in the lower span are more defined to that occurring on the stiff hydrofoil due to FSI.

Download Smith et al. supplementary movie 4(Video)
Video 7.9 MB

Smith et al. supplementary movie 5

As σ reaches 0.65, the attached cavity now extends to the trailing edge causing the cloud cavitation to no longer have spanwise compatibility with the hydrofoil, leading to irregular shedding along the span. The extended cavity lengths reaching the trailing edge leads to the formation of shockwaves with both the re-entrant jet and shockwave mechanisms influencing shedding. Small scale break-up of the cavity from surface perturbations forming as the re-entrant jet moves upstream preconditions the flow for condensation shockwaves to form. Similar to the stiff hydrofoil, the shockwave causes the shedding of cloud cavitation where the re-entrant jet instability drives the frequency. However, due in increased FSI, the emergence of the shockwave instability is accelerated on the flexible hydrofoil with similar behaviour seen between the stiff hydrofoil at σ = 0.60 and the flexible at σ = 0.65.

Download Smith et al. supplementary movie 5(Video)
Video 7.9 MB

Smith et al. supplementary movie 6

With σ reducing to 0.55, the cavitation behavior becomes highly irregular with both the re-entrant jet and shockwave instability active as well as a lack of spanwise compatibility between the attached cavity and the hydrofoil. Strong interactions between the deforming hydrofoil, attached cavity and multiple shedding mechanisms results in highly complex cavitation behavior.

Download Smith et al. supplementary movie 6(Video)
Video 7.9 MB

Smith et al. supplementary movie 7

At σ = 0.5 the attached cavity appears more perturbed on the flexible hydrofoil than that of the stiff. This results in small-scale cavity break-up increasing the bubble population in the flow and hence, activating the shockwave instability at a higher σ as well as increasing its severity. The re-entrant jet is still evident in the flow, however, the shockwave instability dominates the cavity dynamics.

Download Smith et al. supplementary movie 7(Video)
Video 11 MB

Smith et al. supplementary movie 8

At σ = 0.4, due to FSI, the flexible hydrofoil experiences solely shockwave driven shedding before the stiff hydrofoil. A re-entrant jet still forms but lacks sufficient time and momentum to reach the upstream extent of the cavity. This results in the shockwave driving the shedding frequency at a rate of St = 0.11 where unlike on the stiff hydrofoil with multiple shedding events in a cycle, the flexible hydrofoil experience full span coherent shedding.

Download Smith et al. supplementary movie 8(Video)
Video 6.4 MB