• Nie Znaleziono Wyników

Erosive aggressiveness of collapsing cavitating structures

N/A
N/A
Protected

Academic year: 2021

Share "Erosive aggressiveness of collapsing cavitating structures"

Copied!
8
0
0

Pełen tekst

(1)

Delft University of Technology

Erosive aggressiveness of collapsing cavitating structures

Schenke, Sören; van Terwisga, Thomas DOI

10.1115/1.861851_ch69 Publication date

2018

Document Version Final published version Published in

Proceedings of the the 10th International Symposium on Cavitation (CAV2018)

Citation (APA)

Schenke, S., & van Terwisga, T. (2018). Erosive aggressiveness of collapsing cavitating structures. In J. Katz (Ed.), Proceedings of the the 10th International Symposium on Cavitation (CAV2018) (pp. 357-362). [05073] New York, NY, USA: ASME. https://doi.org/10.1115/1.861851_ch69

Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

This work is downloaded from Delft University of Technology.

(2)

Green Open Access added to TU Delft Institutional Repository

‘You share, we take care!’ – Taverne project

https://www.openaccess.nl/en/you-share-we-take-care

Otherwise as indicated in the copyright section: the publisher

is the copyright holder of this work and the author uses the

Dutch legislation to make this work public.

(3)

357 presence of the curve soft boundary the gelatin is in a pre-stress condition and appears bright. Compared

to the previous collapse near a wall, the situation is inverted, the bubble collapses away from the free surface and the quasi-jetting happens on the free surface side and the shear wave ”tail” appears between the interface and the bubble.

Figure 9: Bubble collapse near a free surface in 4− % gelatin using plane-polariscope with θ = 0

Discussion and conclusion

The present study is to our understanding the first report of elastic waves generated from non-spherical bubbles collapsing in a tissue-mimicking material. While spherical bubbles generate a stress field, non-spherical bubbles create stress waves due to center-of-mass translation. There we expect that the bubble moves with a speed much faster than the elastic wave velocity. The resulting wave pattern thus are Mach cones, resembling the wave generation in supersonic shearwave elastography [1]. Besides the four demonstrated cases of spherical, non-spherical, rigid and free boundary collapses, we expect also for the shock wave-gas bubble interaction the formation of stress waves. Although not shown here, the stress wave may not be generated at the location of bubble nucleation but at the location of the gas bubble impacted by the shock thus far from the origin of nucleation. This may have important consequences for medical applications of shock waves. The research presented here is only a starting point and demands for a quantitative analysis and simulations of the wave propagation. Monitoring biological cells at various distance from the bubble may allow to evaluate the importance of stress waves for cell viability or drug delivery. At last we expect that the strength of the elastic waves may be strong function of the gelatin concentration and bubble size and have some optimum at intermediate values. This again needs more experimental and numerical work.

References

[1] J´er´emy Bercoff, Mickael Tanter, and Mathias Fink. Supersonic shear imaging: a new technique for soft tissue elasticity mapping. IEEE transactions on ultrasonics, ferroelectrics, and frequency control, 51(4):396–409, 2004.

[2] Christopher E Brennen. A review of cavitation uses and problems in medicine. In WimRC Forum, volume 70, 2006.

[3] Emil-Alexandru Brujan and Alfred Vogel. Stress wave emission and cavitation bubble dynamics by nanosecond optical breakdown in a tissue phantom. Journal of Fluid Mechanics, 558:281, July 2006. [4] Achu G Byju and Ankur Kulkarni. Mechanics of gelatin and elastin based hydrogels as tissue

engi-neered constructs. In ICF13, 2013.

[5] J.-L. Gennisson, T. Deffieux, M. Fink, and M. Tanter. Ultrasound elastography: Principles and techniques. Diagnostic and Interventional Imaging, 94(5):487–495, May 2013.

[6] Ryota Oguri and Keita Ando. Cloud cavitation induced by shock-bubble interaction in a viscoelastic solid. Journal of Physics: Conference Series, 656:012032, December 2015.

[7] K Ramesh. Digital photoelasticity, 2000. 10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05072

*Corresponding author, Sören Schenke: s.schenke@tudelft.nl

Erosive Aggressiveness of Collapsing Cavitating Structures

1Sören Schenke*; 1,2Tom J.C. van Terwisga

1Delft University of Technology, Delft, the Netherlands; 2Maritime Research Institute Netherlands, Wageningen, the Netherlands

Abstract

The erosive aggressiveness of idealized cavities collapsing on a flat surface is investigated by numerical simulation, employing the cavitation intensity approach by Leclercq et al (2017) as a measure of the local energy impact rate. We propose a more straight forward formulation of the cavitation intensity model and verify that it satisfies energy conservation requirements for the accumulated surface energy. Based on the cavitation intensity model, statistical aggressiveness indicators are proposed. The indicators account for the rapidness and frequency of the collapse events. The aggressiveness indicators are further applied to a NACA0015 hydrofoil surface.

Keywords: cavitation erosion, impact energy, aggressiveness indicator, pressure recovery

Introduction

From an energy point of view, the erosive aggressiveness of a cavity imploding close to a solid surface depends for one part on its ability to focus its impact energy on the surface. The surface distribution of impact energy accumulating throughout a collapse event partially depends on the cavity shape and its orientation relative to the surface. Starting from the potential energy approach by Vogel and Lauterborn (1988), Leclercq et al (2017) have proposed a cavitation intensity model for the instantaneous energy impact rate on a solid surface caused by collapsing cavities. The cavitation intensity model is derived from the solid angle projection of the released power on discrete triangular elements of the impacted surface. We propose a more straight forward and fully continuous form of the cavitation intensity model. This facilitates its application and allows to make analytical predictions on the amount of accumulated surface energy in certain situations, which are employed for numerical validation in this study. The erosive aggressiveness is further thought to depend on the rapidness of the collapse event and, in periodic flow, the impact frequency. Based on the cavitation intensity model, aggressiveness indicators are proposed that account for both effects. A distinct weak point of this modelling approach is that it involves the knowledge of the ambient pressure driving the cavity collapse, which is typically unknown in more complex flow situations. A solution to this problem is suggested by Arabnejad and Bensow (2017), who trace the cavity collapses individually to reconstruct the driving ambient pressure from a set of kinematic parameters employing the Rayleigh-Plesset equation. However, this approach requires any cavitating structure to be simplified to a spherical bubble of equivalent volume. Since we deliberately address the effect of cavity shape and surface orientation, we assume that the driving pressure field of a periodic cavitating flow can be approximated by the time averaged pressure field, thereby accounting for the effect of spatial pressure recovery on statistical average. The effect of pressure recovery and the feasibility of the aggressiveness indicators is demonstrated for the cavitating flow around a NACA0015 hydrofoil.

Erosion Intensity Model

Following Vogel and Lauterborn (1988), the potential energy of a cavity with volume 𝑉𝑉 and liquid volume fraction 𝛾𝛾 = (𝜌𝜌 − 𝜌𝜌𝑣𝑣)/(𝜌𝜌𝑙𝑙− 𝜌𝜌𝑣𝑣) is given by 𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝= (1 − 𝛾𝛾)(𝑝𝑝∞−𝑝𝑝𝑣𝑣)𝑉𝑉, where 𝑝𝑝𝑣𝑣 denotes the vapor pressure and 𝑝𝑝∞ the ambient pressure driving the cavity collapse. As the cavity collapses, the instantaneous rate of energy release is given by 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝/𝜕𝜕𝜕𝜕. Due to the linearity of these expressions, the energy release rate of the entire cavity is equal to the volume integral of the differential point source release rate 𝜕𝜕𝜕𝜕𝑝𝑝𝑝𝑝𝑝𝑝/𝜕𝜕𝜕𝜕. As indicated by Figure 1, this allows to reconstruct both the local surface impact rate 𝜕𝜕𝜕𝜕𝑆𝑆/𝜕𝜕𝜕𝜕 as well as the overall surface integrated impact rate 𝜕𝜕𝐸𝐸𝑆𝑆/𝜕𝜕𝜕𝜕 from either the local change of volume fraction 𝛾𝛾 or the volume change of the entire cavity, resulting in the notations

𝜕𝜕𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕 |𝛾𝛾̇= −X 𝜕𝜕𝜕𝜕𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 , 𝜕𝜕𝐸𝐸𝑆𝑆 𝜕𝜕𝜕𝜕 |𝛾𝛾̇ = ∫ 𝜕𝜕𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕 |𝛾𝛾̇ 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 𝑑𝑑𝑑𝑑, 𝜕𝜕𝜕𝜕𝑆𝑆𝜕𝜕𝜕𝜕 |𝑉𝑉̇= − ∫ X𝜕𝜕𝜕𝜕𝑝𝑝𝑝𝑝𝑝𝑝𝜕𝜕𝜕𝜕 𝑑𝑑𝑉𝑉 and 𝜕𝜕𝐸𝐸𝑆𝑆𝜕𝜕𝜕𝜕 |𝑉𝑉̇ 𝑣𝑣𝑝𝑝𝑙𝑙 = ∫ 𝜕𝜕𝜕𝜕𝑆𝑆𝜕𝜕𝜕𝜕 |𝑉𝑉̇ 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 𝑑𝑑𝑑𝑑, (1) where X is the projection operator that converts locally released power to local surface impact power, depending on the distance from the emission source and the surface orientation relative to the source. Flageul et al (2012) determine

10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05073

(4)

358

the local potential energy release 𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝/𝜕𝜕𝜕𝜕 from the Lagrangian time derivative of mixture density. Alternatively, the energy release can be calculated from the Lagrangian time derivative of liquid volume fraction, which gives

𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 = − ( 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕)𝑑𝑑𝑑𝑑𝑑𝑑 + (𝑝𝑝∞−𝑝𝑝𝑑𝑑), where (𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕)𝑑𝑑𝑑𝑑𝑑𝑑+ = max [−∇ ⋅ 𝑢𝑢 (𝜕𝜕 +𝜌𝜌𝑙𝑙𝜌𝜌− 𝜌𝜌𝑑𝑑𝑑𝑑 ) , 0]. (2) The ‘+’ index indicates that only condensation is taken into account. Assuming that each point source emits its potential energy as a radial wave of infinitely large propagation speed, Leclercq et al (2017) define the potential power impact per discrete surface element Δ𝑆𝑆 from the solid angle projection of the radial source on a planar triangle as presented in the work by Van Oosterom and Strackee (1983). A continuous form of the energy impact rate on a surface location with local normal vector 𝑛𝑛⃗ is derived by employing the projection operator

X =4𝜋𝜋 (1 𝑥𝑥 ⋅ 𝑛𝑛⃗ |𝑥𝑥 |3), (3)

where 𝑥𝑥 denotes the vector from the impacted surface location to the center of the emission source as shown in Figure 1. Integration of the bracket term in Equation (3) over an arbitrary convex surface gives the solid angle Ω as used in the work by Leclercq et al (2017). From the continuous form of the cavitation intensity model, it can be shown analytically that the integrated impact rate is equal to the emission rate of a volume source on any closed convex surface around the source. It is equal to half of the emission rate on a flat surface stretched to infinity, hence

∮ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕 |𝑉𝑉̇𝑑𝑑𝑆𝑆 = 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝑐𝑐𝑝𝑝𝑐𝑐𝑑𝑑𝑐𝑐𝑐𝑐 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑐𝑐𝑐𝑐 and ∫ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕 |𝑉𝑉̇𝑑𝑑𝑆𝑆 ∞ 𝑝𝑝𝑙𝑙𝑠𝑠𝑑𝑑𝑐𝑐 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑐𝑐𝑐𝑐 =1 2 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 . (4)

For simplicity, 𝑒𝑒̇𝑆𝑆 and 𝐸𝐸̇𝑆𝑆 are further on referred to as the local and the surface integrated impact rate caused by volume sources. A statistical measure for the aggressiveness distribution is given by the aggressiveness indicators

〈𝑒𝑒̇𝑆𝑆〉𝑐𝑐𝑆𝑆= 1 𝑒𝑒𝑆𝑆∫ 𝑒𝑒̇𝑆𝑆𝑑𝑑𝑒𝑒𝑆𝑆 and 〈𝑒𝑒̇𝑆𝑆〉𝑠𝑠= ( 1 𝑇𝑇 ∫ 𝑒𝑒̇𝑆𝑆𝑑𝑑𝑒𝑒𝑆𝑆 𝑐𝑐𝑆𝑆 0 ) 1 2⁄ , 𝑐𝑐𝑆𝑆 0 where 𝑒𝑒𝑆𝑆(𝜕𝜕) = ∫ 𝑒𝑒̇𝑆𝑆(𝜕𝜕)𝑑𝑑𝜕𝜕. 𝑝𝑝 0 (5) The indicator 〈𝑒𝑒̇𝑆𝑆〉𝑐𝑐𝑆𝑆, being the local energy impact rate 𝑒𝑒̇𝑆𝑆(𝜕𝜕) averaged over the accumulated energy 𝑒𝑒𝑆𝑆, has a tendency

to amplify local extreme events, because for the same amount of accumulated energy, it is proportional to the rate at which the surface is impacted. In case of a periodic impact signal with uniform amplitude, it is independent from the impact frequency due to the normalization by the accumulated energy, whereas the indicator 〈𝑒𝑒̇𝑆𝑆〉𝑠𝑠 is also proportional to the impact frequency. Normalization by the sample time 𝑇𝑇 causes it to converge as 𝑇𝑇 → ∞.

Figure 1: Terminology of energy time derivatives

Numerical Testcases

The cavitation intensity model is implemented in the open source CFD environment OpenFOAM (2017). Viscous forces and surface tension are neglected in the momentum equation of the flow model. Phase transition is modelled by the mass transfer approach, where the model by Merkle et al (1998) is employed in a slightly modified form as presented in the work by Schenke and Van Terwisga (2017). In the pure liquid and vapor limits, the flow is assumed to be incompressible and a pressure equation is solved with subsequent momentum correction. The erosive

𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕|𝛾𝛾̇ 𝜕𝜕𝐸𝐸𝑆𝑆 𝜕𝜕𝜕𝜕|𝛾𝛾̇ 𝜕𝜕𝐸𝐸𝑆𝑆 𝜕𝜕𝜕𝜕|𝑉𝑉̇ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕|𝑉𝑉̇ 𝑛𝑛⃗ 𝑛𝑛⃗ 𝑥𝑥

Baltimore, Maryland, USA, May 14 – 16, 2018

aggressiveness of three different cavity types is investigated. Those are, as shown in Figure 2, a horseshoe cavity, a parallel ring cavity and a bubble, all collapsing on a flat surface which is large enough to verify Equation (4). The initial bubble volume is equal to the horseshoe volume and the ring volume is twice the horseshoe/bubble volume. The torus diameter is 𝑅𝑅𝑡𝑡 = 6 mm and the tube radius 𝑟𝑟 = 2 mm. Both the parallel ring and the bubble are initialized at ℎ = 0.5 mm above the surface. The inner part of the computational domain is indicated in Figure 3. It consists of a cubic block structured grid of 20 mm edge length, uniformly subdivided by 55 cells in all directions. The overall volume of the domain is 32 m3. A time step size of 𝛥𝛥𝛥𝛥 = 1.0e-7 s turned out to be sufficiently small to obtain a

converged cavity collapse time. Additionally, a free bubble collapse is simulated and the collapse time is compared to the analytical solution of the Rayleigh-Plesset equation. As indicated in Figure 3, the solid bottom plane, treated as a slip wall, is replaced by a symmetry plane in this case. With 𝐶𝐶𝑐𝑐 and 𝐶𝐶𝑣𝑣 being the condensation and evaporation mass transfer coefficients of the modified Merkle model (Schenke and Van Terwisga 2017), the simulation is conducted for 𝐶𝐶𝑐𝑐 = 𝐶𝐶𝑣𝑣 = 1000 kgs/m5, which is sufficiently large for the collapse to be inertia driven. The pressure field is initialized with 𝑝𝑝𝑣𝑣= 2340 Pa inside the cavities and 𝑝𝑝∞= 1 bar outside the cavities. To avoid unrealistic behavior due to the initial pressure jump across the interface, the pressure equation is solved first, thereby providing a solution of the Laplace equation as the flow is at rest initially. Figure 3 exemplarily shows the pressure field close to the beginning of the horseshoe cavity collapse. It is further noted that the reconstruction of the positive Lagrangian time derivative (𝜕𝜕𝜕𝜕 𝜕𝜕𝛥𝛥⁄ )𝑑𝑑𝑑𝑑𝑣𝑣+ in Equation (2) involves numerical errors that may eventually violate the energy balance. To verify the cavitation intensity model itself, this effect is eliminated for the cavity collapse study by assuming that the change of vapor volume in the entire domain predicted from the divergence field differs from the actual vapor volume change predicted by the 𝜕𝜕-time derivative by a correction factor. This factor is applied to Equation (2) at each cell individually.

Figure 2: Idealized cavities investigated in this study Figure 3: Initial horseshoe cavity (iso-surface for 𝜕𝜕0= 0.5) and

bottom plane with initial 𝜕𝜕-field and pressure field at 2e-05 s

Figure 4: Grid refinement levels and observation point on the foil

surface at 20% cord length Figure 5: Pressure signal and evolution of the corresponding time averaged pressure 〈𝑝𝑝〉𝑡𝑡 at the observation point in Figure 4. The analysis is further carried out for the cavitating flow around a NACA0015 hydrofoil in a tunnel section. The flow conditions are in line with an experiment by Van Rijsbergen et al (2012), where the angle of attack is 8∘ at an inflow speed of 17.3 m/s and an ambient pressure of 302295 Pa. Cordlength and span are 6 cm and 4 cm, respectively. The grid, as shown in Figure 4, corresponds to what is referred to as the fine grid in previous work by Schenke and Van Terwisga (2017), where further details on grid properties and solver settings are found. The simulation is conducted

(5)

359 the local potential energy release 𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝/𝜕𝜕𝜕𝜕 from the Lagrangian time derivative of mixture density. Alternatively, the

energy release can be calculated from the Lagrangian time derivative of liquid volume fraction, which gives 𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 = − ( 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕)𝑑𝑑𝑑𝑑𝑑𝑑 + (𝑝𝑝∞−𝑝𝑝𝑑𝑑), where (𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕)𝑑𝑑𝑑𝑑𝑑𝑑+ = max [−∇ ⋅ 𝑢𝑢 (𝜕𝜕 +𝜌𝜌𝑙𝑙𝜌𝜌− 𝜌𝜌𝑑𝑑𝑑𝑑 ) , 0]. (2) The ‘+’ index indicates that only condensation is taken into account. Assuming that each point source emits its potential energy as a radial wave of infinitely large propagation speed, Leclercq et al (2017) define the potential power impact per discrete surface element Δ𝑆𝑆 from the solid angle projection of the radial source on a planar triangle as presented in the work by Van Oosterom and Strackee (1983). A continuous form of the energy impact rate on a surface location with local normal vector 𝑛𝑛⃗ is derived by employing the projection operator

X =4𝜋𝜋 (1 𝑥𝑥 ⋅ 𝑛𝑛⃗ |𝑥𝑥 |3), (3)

where 𝑥𝑥 denotes the vector from the impacted surface location to the center of the emission source as shown in Figure 1. Integration of the bracket term in Equation (3) over an arbitrary convex surface gives the solid angle Ω as used in the work by Leclercq et al (2017). From the continuous form of the cavitation intensity model, it can be shown analytically that the integrated impact rate is equal to the emission rate of a volume source on any closed convex surface around the source. It is equal to half of the emission rate on a flat surface stretched to infinity, hence

∮ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕 |𝑉𝑉̇𝑑𝑑𝑆𝑆 = 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝑐𝑐𝑝𝑝𝑐𝑐𝑑𝑑𝑐𝑐𝑐𝑐 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑐𝑐𝑐𝑐 and ∫ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕 |𝑉𝑉̇𝑑𝑑𝑆𝑆 ∞ 𝑝𝑝𝑙𝑙𝑠𝑠𝑑𝑑𝑐𝑐 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑐𝑐𝑐𝑐 =1 2 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 . (4)

For simplicity, 𝑒𝑒̇𝑆𝑆 and 𝐸𝐸̇𝑆𝑆 are further on referred to as the local and the surface integrated impact rate caused by volume sources. A statistical measure for the aggressiveness distribution is given by the aggressiveness indicators

〈𝑒𝑒̇𝑆𝑆〉𝑐𝑐𝑆𝑆= 1 𝑒𝑒𝑆𝑆∫ 𝑒𝑒̇𝑆𝑆𝑑𝑑𝑒𝑒𝑆𝑆 and 〈𝑒𝑒̇𝑆𝑆〉𝑠𝑠= ( 1 𝑇𝑇 ∫ 𝑒𝑒̇𝑆𝑆𝑑𝑑𝑒𝑒𝑆𝑆 𝑐𝑐𝑆𝑆 0 ) 1 2⁄ , 𝑐𝑐𝑆𝑆 0 where 𝑒𝑒𝑆𝑆(𝜕𝜕) = ∫ 𝑒𝑒̇𝑆𝑆(𝜕𝜕)𝑑𝑑𝜕𝜕. 𝑝𝑝 0 (5) The indicator 〈𝑒𝑒̇𝑆𝑆〉𝑐𝑐𝑆𝑆, being the local energy impact rate 𝑒𝑒̇𝑆𝑆(𝜕𝜕) averaged over the accumulated energy 𝑒𝑒𝑆𝑆, has a tendency

to amplify local extreme events, because for the same amount of accumulated energy, it is proportional to the rate at which the surface is impacted. In case of a periodic impact signal with uniform amplitude, it is independent from the impact frequency due to the normalization by the accumulated energy, whereas the indicator 〈𝑒𝑒̇𝑆𝑆〉𝑠𝑠 is also proportional to the impact frequency. Normalization by the sample time 𝑇𝑇 causes it to converge as 𝑇𝑇 → ∞.

Figure 1: Terminology of energy time derivatives

Numerical Testcases

The cavitation intensity model is implemented in the open source CFD environment OpenFOAM (2017). Viscous forces and surface tension are neglected in the momentum equation of the flow model. Phase transition is modelled by the mass transfer approach, where the model by Merkle et al (1998) is employed in a slightly modified form as presented in the work by Schenke and Van Terwisga (2017). In the pure liquid and vapor limits, the flow is assumed to be incompressible and a pressure equation is solved with subsequent momentum correction. The erosive

𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝑒𝑒𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝐸𝐸𝑝𝑝𝑝𝑝𝑝𝑝 𝜕𝜕𝜕𝜕 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕|𝛾𝛾̇ 𝜕𝜕𝐸𝐸𝑆𝑆 𝜕𝜕𝜕𝜕|𝛾𝛾̇ 𝜕𝜕𝐸𝐸𝑆𝑆 𝜕𝜕𝜕𝜕|𝑉𝑉̇ 𝜕𝜕𝑒𝑒𝑆𝑆 𝜕𝜕𝜕𝜕|𝑉𝑉̇ 𝑛𝑛⃗ 𝑛𝑛⃗ 𝑥𝑥

10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05073

aggressiveness of three different cavity types is investigated. Those are, as shown in Figure 2, a horseshoe cavity, a parallel ring cavity and a bubble, all collapsing on a flat surface which is large enough to verify Equation (4). The initial bubble volume is equal to the horseshoe volume and the ring volume is twice the horseshoe/bubble volume. The torus diameter is 𝑅𝑅𝑡𝑡 = 6 mm and the tube radius 𝑟𝑟 = 2 mm. Both the parallel ring and the bubble are initialized at ℎ = 0.5 mm above the surface. The inner part of the computational domain is indicated in Figure 3. It consists of a cubic block structured grid of 20 mm edge length, uniformly subdivided by 55 cells in all directions. The overall volume of the domain is 32 m3. A time step size of 𝛥𝛥𝛥𝛥 = 1.0e-7 s turned out to be sufficiently small to obtain a

converged cavity collapse time. Additionally, a free bubble collapse is simulated and the collapse time is compared to the analytical solution of the Rayleigh-Plesset equation. As indicated in Figure 3, the solid bottom plane, treated as a slip wall, is replaced by a symmetry plane in this case. With 𝐶𝐶𝑐𝑐 and 𝐶𝐶𝑣𝑣 being the condensation and evaporation mass transfer coefficients of the modified Merkle model (Schenke and Van Terwisga 2017), the simulation is conducted for 𝐶𝐶𝑐𝑐 = 𝐶𝐶𝑣𝑣 = 1000 kgs/m5, which is sufficiently large for the collapse to be inertia driven. The pressure field is initialized with 𝑝𝑝𝑣𝑣= 2340 Pa inside the cavities and 𝑝𝑝∞= 1 bar outside the cavities. To avoid unrealistic behavior due to the initial pressure jump across the interface, the pressure equation is solved first, thereby providing a solution of the Laplace equation as the flow is at rest initially. Figure 3 exemplarily shows the pressure field close to the beginning of the horseshoe cavity collapse. It is further noted that the reconstruction of the positive Lagrangian time derivative (𝜕𝜕𝜕𝜕 𝜕𝜕𝛥𝛥⁄ )𝑑𝑑𝑑𝑑𝑣𝑣+ in Equation (2) involves numerical errors that may eventually violate the energy balance. To verify the cavitation intensity model itself, this effect is eliminated for the cavity collapse study by assuming that the change of vapor volume in the entire domain predicted from the divergence field differs from the actual vapor volume change predicted by the 𝜕𝜕-time derivative by a correction factor. This factor is applied to Equation (2) at each cell individually.

Figure 2: Idealized cavities investigated in this study Figure 3: Initial horseshoe cavity (iso-surface for 𝜕𝜕0= 0.5) and

bottom plane with initial 𝜕𝜕-field and pressure field at 2e-05 s

Figure 4: Grid refinement levels and observation point on the foil

surface at 20% cord length Figure 5: Pressure signal and evolution of the corresponding time averaged pressure 〈𝑝𝑝〉𝑡𝑡 at the observation point in Figure 4. The analysis is further carried out for the cavitating flow around a NACA0015 hydrofoil in a tunnel section. The flow conditions are in line with an experiment by Van Rijsbergen et al (2012), where the angle of attack is 8∘ at an inflow speed of 17.3 m/s and an ambient pressure of 302295 Pa. Cordlength and span are 6 cm and 4 cm, respectively. The grid, as shown in Figure 4, corresponds to what is referred to as the fine grid in previous work by Schenke and Van Terwisga (2017), where further details on grid properties and solver settings are found. The simulation is conducted

10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05073

(6)

360

for 𝐶𝐶𝑐𝑐/𝐶𝐶𝑣𝑣 = 2, 𝐶𝐶𝑐𝑐 = 5000 kgs/m5 and 𝛥𝛥𝛥𝛥 = 7.5e-07 s. For these settings, a converged pressure impact frequency of 193

Hz has been found by FFT analysis, being in good agreement with the shedding frequency of 188 Hz found by Van Rijsbergen et al (2012) in their experiment. In addition to the previous work, the time averaged pressure field 〈𝑝𝑝〉𝑡𝑡 under cavitating flow conditions is computed. We assume the time averaged pressure field to be the steady field driving the cavity collapses such that 〈𝑝𝑝〉𝑡𝑡 locally represents the driving pressure 𝑝𝑝∞ in Equation (2). Figure 5 exemplarily depicts the pressure signal and the evolution of the corresponding time averaged pressure at an observation on the foil surface at mid-span and 20% cord length (see Figure 4).

Results & Discussion

Figure 6 depicts the evolution of vapor volume 𝑉𝑉𝑣𝑣 and accumulated surface energy 𝐸𝐸𝑆𝑆 over time for the cavities shown in Figure 2. Towards the end of the collapse, the accumulated surface energy 𝐸𝐸𝑆𝑆 converges to 50% of the initial potential energy 𝐸𝐸𝑝𝑝𝑝𝑝𝑡𝑡, thereby verifying Equation (4). The analytical solution obtained from the Rayleigh-Plesset equation without viscous and surface tension forces is included as a reference solution for the free bubble collapse. Satisfactory agreement of the predicted collapse time is achieved. Since the presence of the wall weakens the lower half of the pressure field driving the close wall bubble collapse, the free bubble collapses faster. For the same reason the horseshoe cavity collapses faster than the parallel ring because it is not as strongly subjected to wall interaction.

Figure 6: Evolution of vapor volume and accumulated surface energy for the cavity types depicted in Figure 2 collapsing on a flat surface

Figure 7: Accumulated surface energy 𝑒𝑒𝑆𝑆 (left) and distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 (right) for the horseshoe cavity (a), the parallel ring cavity (b) and the spherical bubble (c) with the green line indicating the outline of the corresponding initial cavity shape

To gain more insight into the energy focusing abilities of the different cavity types, Figure 7 depicts the corresponding distributions of accumulated surface energy 𝑒𝑒𝑆𝑆 and the distributions of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆. In this

particular case, all three cavities exhibit similar abilities to focus their potential energy towards the surface. The different collapse times, however, clearly reflect in the distributions of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆, indicating

that the horseshoe cavity is most aggressive, followed by the bubble of equivalent volume. The aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 also predicts a more focused impact distribution than the accumulated surface energy 𝑒𝑒𝑆𝑆. It is further

observed for the horseshoe and the ring cavity that especially the peak value of the indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 exhibits a slight

eccentricity from the torus center line towards the torus center, which is motivated by the circumstance that a Laplacian pressure field exhibits a larger driving force outside from the torus center line than from the inside. This is a distinct shape effect and shows that it is not entirely correct to assume the initial potential cavity energy to be proportional to the ambient pressure 𝑝𝑝∞. Even in the absence of wall interaction, the driving pressure may vary over the cavity surface

(a)

(b)

(c)

Baltimore, Maryland, USA, May 14 – 16, 2018

because, depending on its shape, the cavity also interacts with itself to some extent. The same applies to cavity-cavity interaction. This makes the determination of the driving ambient pressure 𝑝𝑝∞ a problem of distinct difficulty.

Figure 8: Instantaneous vapor structures for iso-surfaces of 𝛾𝛾 = 0.5 and energy impact rate for uniform driving pressure 𝑝𝑝∞ (a) and variable

driving pressure 〈𝑝𝑝〉𝑡𝑡 (b)

Figure 9: Accumulated surface energy on the NACA0015 hydrofoil surface for uniform 𝑝𝑝∞ (a) and variable driving pressure 〈𝑝𝑝〉𝑡𝑡 (b) after 0.27 s

Figure 10: Distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 on the NACA0015 hydrofoil surface, accounting for pressure recovery

Figure 11: Distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑓𝑓 on the NACA0015 hydrofoil

surface, accounting for pressure recovery

Figure 12: Damage pattern on the NACA0015 hydrofoil surface identified from paint test experiments by Van Rijsbergen et al (2012) In case of a periodic flow, we can at least account for the effect of spatial pressure recovery on statistical average by assuming that the driving pressure field is given by the time averaged pressure field under cavitating flow conditions. Figure 8 (a) shows the distribution of the instantaneous energy impact rate 𝑒𝑒̇𝑆𝑆 for uniform driving pressure 𝑝𝑝∞. The same flow situation is depicted in Figure 8 (b), however with the driving pressure 𝑝𝑝∞ being equal to the time averaged pressure 〈𝑝𝑝〉𝑡𝑡 as depicted on the tunnel side wall. This inevitably results in a larger predicted impact rates towards the leading edge, which is also clearly reflected by the distribution of accumulated surface energy 𝑒𝑒𝑆𝑆, shown in Figure 9 (a) for uniform driving pressure 𝑝𝑝∞ and in Figure 9 (b) for variable driving pressure 〈𝑝𝑝〉𝑡𝑡. The rapidness of the impacts is assessed by the aggressiveness indicators 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 and 〈𝑒𝑒̇𝑆𝑆〉𝑓𝑓 given by Equation (5). Both result in similar impact

distributions, indicating regions of distinct periodic impacts caused by the re-entrant jet mechanism rather than regions of scattered impacts. The indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 further amplifies isolated events at the trailing edge because its magnitude

is independent from the impact frequency. The so identified regions of high erosion risk qualitatively agree with the damage pattern obtained by Van Rijsbergen et al (2012) from experimental paint tests (Figure 12), although the main impact region identified from the simulation is somewhat closer to the leading edge and less focused towards the mid-span. The latter observation may partially be explained by the neglect of viscous forces which, if present at the tunnel wall, may cause a deflection of the re-entrant jet towards the mid-span (Dular and Petkovšek, 2015). It should further

(a)

(b)

lea di ng e dg e lea di ng e dg e lea di ng e dg e lea di ng e dg e

(a)

(b)

(7)

361 for 𝐶𝐶𝑐𝑐/𝐶𝐶𝑣𝑣 = 2, 𝐶𝐶𝑐𝑐 = 5000 kgs/m5 and 𝛥𝛥𝛥𝛥 = 7.5e-07 s. For these settings, a converged pressure impact frequency of 193

Hz has been found by FFT analysis, being in good agreement with the shedding frequency of 188 Hz found by Van Rijsbergen et al (2012) in their experiment. In addition to the previous work, the time averaged pressure field 〈𝑝𝑝〉𝑡𝑡 under cavitating flow conditions is computed. We assume the time averaged pressure field to be the steady field driving the cavity collapses such that 〈𝑝𝑝〉𝑡𝑡 locally represents the driving pressure 𝑝𝑝∞ in Equation (2). Figure 5 exemplarily depicts the pressure signal and the evolution of the corresponding time averaged pressure at an observation on the foil surface at mid-span and 20% cord length (see Figure 4).

Results & Discussion

Figure 6 depicts the evolution of vapor volume 𝑉𝑉𝑣𝑣 and accumulated surface energy 𝐸𝐸𝑆𝑆 over time for the cavities shown in Figure 2. Towards the end of the collapse, the accumulated surface energy 𝐸𝐸𝑆𝑆 converges to 50% of the initial potential energy 𝐸𝐸𝑝𝑝𝑝𝑝𝑡𝑡, thereby verifying Equation (4). The analytical solution obtained from the Rayleigh-Plesset equation without viscous and surface tension forces is included as a reference solution for the free bubble collapse. Satisfactory agreement of the predicted collapse time is achieved. Since the presence of the wall weakens the lower half of the pressure field driving the close wall bubble collapse, the free bubble collapses faster. For the same reason the horseshoe cavity collapses faster than the parallel ring because it is not as strongly subjected to wall interaction.

Figure 6: Evolution of vapor volume and accumulated surface energy for the cavity types depicted in Figure 2 collapsing on a flat surface

Figure 7: Accumulated surface energy 𝑒𝑒𝑆𝑆 (left) and distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 (right) for the horseshoe cavity (a), the parallel ring cavity (b) and the spherical bubble (c) with the green line indicating the outline of the corresponding initial cavity shape

To gain more insight into the energy focusing abilities of the different cavity types, Figure 7 depicts the corresponding distributions of accumulated surface energy 𝑒𝑒𝑆𝑆 and the distributions of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆. In this

particular case, all three cavities exhibit similar abilities to focus their potential energy towards the surface. The different collapse times, however, clearly reflect in the distributions of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆, indicating

that the horseshoe cavity is most aggressive, followed by the bubble of equivalent volume. The aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 also predicts a more focused impact distribution than the accumulated surface energy 𝑒𝑒𝑆𝑆. It is further

observed for the horseshoe and the ring cavity that especially the peak value of the indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 exhibits a slight

eccentricity from the torus center line towards the torus center, which is motivated by the circumstance that a Laplacian pressure field exhibits a larger driving force outside from the torus center line than from the inside. This is a distinct shape effect and shows that it is not entirely correct to assume the initial potential cavity energy to be proportional to the ambient pressure 𝑝𝑝∞. Even in the absence of wall interaction, the driving pressure may vary over the cavity surface

(a)

(b)

(c)

10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05073

because, depending on its shape, the cavity also interacts with itself to some extent. The same applies to cavity-cavity interaction. This makes the determination of the driving ambient pressure 𝑝𝑝∞ a problem of distinct difficulty.

Figure 8: Instantaneous vapor structures for iso-surfaces of 𝛾𝛾 = 0.5 and energy impact rate for uniform driving pressure 𝑝𝑝∞ (a) and variable

driving pressure 〈𝑝𝑝〉𝑡𝑡 (b)

Figure 9: Accumulated surface energy on the NACA0015 hydrofoil surface for uniform 𝑝𝑝∞ (a) and variable driving pressure 〈𝑝𝑝〉𝑡𝑡 (b) after 0.27 s

Figure 10: Distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 on the NACA0015 hydrofoil surface, accounting for pressure recovery

Figure 11: Distribution of the aggressiveness indicator 〈𝑒𝑒̇𝑆𝑆〉𝑓𝑓 on the NACA0015 hydrofoil

surface, accounting for pressure recovery

Figure 12: Damage pattern on the NACA0015 hydrofoil surface identified from paint test experiments by Van Rijsbergen et al (2012) In case of a periodic flow, we can at least account for the effect of spatial pressure recovery on statistical average by assuming that the driving pressure field is given by the time averaged pressure field under cavitating flow conditions. Figure 8 (a) shows the distribution of the instantaneous energy impact rate 𝑒𝑒̇𝑆𝑆 for uniform driving pressure 𝑝𝑝∞. The same flow situation is depicted in Figure 8 (b), however with the driving pressure 𝑝𝑝∞ being equal to the time averaged pressure 〈𝑝𝑝〉𝑡𝑡 as depicted on the tunnel side wall. This inevitably results in a larger predicted impact rates towards the leading edge, which is also clearly reflected by the distribution of accumulated surface energy 𝑒𝑒𝑆𝑆, shown in Figure 9 (a) for uniform driving pressure 𝑝𝑝∞ and in Figure 9 (b) for variable driving pressure 〈𝑝𝑝〉𝑡𝑡. The rapidness of the impacts is assessed by the aggressiveness indicators 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 and 〈𝑒𝑒̇𝑆𝑆〉𝑓𝑓 given by Equation (5). Both result in similar impact

distributions, indicating regions of distinct periodic impacts caused by the re-entrant jet mechanism rather than regions of scattered impacts. The indicator 〈𝑒𝑒̇𝑆𝑆〉𝑒𝑒𝑆𝑆 further amplifies isolated events at the trailing edge because its magnitude

is independent from the impact frequency. The so identified regions of high erosion risk qualitatively agree with the damage pattern obtained by Van Rijsbergen et al (2012) from experimental paint tests (Figure 12), although the main impact region identified from the simulation is somewhat closer to the leading edge and less focused towards the mid-span. The latter observation may partially be explained by the neglect of viscous forces which, if present at the tunnel wall, may cause a deflection of the re-entrant jet towards the mid-span (Dular and Petkovšek, 2015). It should further

(a)

(b)

lea di ng e dg e lea di ng e dg e lea di ng e dg e lea di ng e dg e

(a)

(b)

10th International Symposium on Cavitation - CAV2018

Baltimore, Maryland, USA, May 14 – 16, 2018 CAV18-05073

(8)

362

be noted that the paint test by Van Rijsbergen et al (2012) was conducted for one hour, whereas the simulation time is only 0.27 s, not yet enough to obtain a fully converged impact distribution.

Conclusion

A continuous form of the cavitation intensity model by Leclercq et al (2017) is presented in this study, providing a measure for the instantaneous impact power on a surface subjected to cavity collapses. The distribution of accumulated surface energy shows that the idealized horseshoe cavity, the parallel ring cavity and the bubble exhibit similar efficiencies in focusing their initial potential energy towards a confined space on the surface, given that the bubble and the parallel ring collapse very close to the surface. More pronounced differences have been found regarding the rapidness of the collapse event, which is strongly affected by the cavity shape and wall interaction effects. An aggressiveness indicator derived from the cavitation intensity approach indicates that the horseshoe cavity is more aggressive than a bubble of equivalent volume, which again turns out to be more aggressive than a parallel ring cavity of same torus and tube diameter as the horseshoe. Further taking into account that the impact power of the bubble and the parallel ring decreases considerably with increasing distance from the impacted surface, leads to the conclusion that the horseshoe cavity attached to the surface is most likely to result in damaging collapse events. Next to the rapidness of the collapse event, the effect of spatial pressure recovery has been identified as a second crucial effect since it strongly affects the distribution of initial potential cavity energy eventually impacting the solid surface. On time average, the flow around the NACA0015 hydrofoil investigated in this study exhibits a pronounced pressure recovery gradient, thereby increasing the flow aggressiveness towards the trailing edge. However, the exact determination of the pressure effectively driving the individual cavity collapses remains an issue of distinct difficulty and requires further research.

Acknowledgements

This research is funded by the CaFE ITN initiative and the MARIN academy. We thank both institutions for their collaboration and we highly appreciate the inspiring discussions on cavitation erosion mechanisms.

References

[1] Arabnejad, M.H. and Bensow, R. (2017). A Methodology to Identify Erosive Collapse Events in Incompressible Simulation of

Cavitating Flows. Proceedings of the 20th Numerical Towing Tank Symposium, Wageningen, the Netherlands.

[2] Dular, M. and Petkovšek, M. (2015). On the Mechanisms of Cavitation Erosion – Coupling High Speed Videos to Damage Patterns. Experimental Thermal and Fluid Science. 68(2015), pp. 359-370.

[3] Flageul, C., Fortes-Patella, R. and Archer, A. (2012). Cavitation Erosion Prediction by Numerical Simulations. Proceedings of the 14th International Symposium on Transport Phenomena and Dynamics of Rotating Machinery, Honolulu, HI, USA.

[4] Leclercq, C., Archer, A., Fortes-Patella, R. and Cerru, F. (2017). Numerical Cavitation Intensity on a Hydrofoil for 3D

Homogeneous Unsteady Viscous Flows. International Journal of Fluid Machinery and Systems. 10(3), pp. 254-263.

[5] Merkle, C.L., Feng, J.Z., and Buelow, P.E.O. (1998). Computational Modelling of the Dynamics of Sheet Cavitation. Proceedings of the 3rd International Symposium on Cavitation, Grenoble, France.

[6] OpenFOAM (2017). OpenFOAM web site, 2017.

[7] Schenke, S. and van Terwisga, T.J.C. (2017). Numerical Prediction of Vortex Dynamics in Inviscid Sheet Cavitation. Proceedings of the 20th Numerical Towing Tank Symposium, Wageningen, the Netherlands.

[8] Van Oosterom, A. and Strackee, J. (1983). The Solid Angle of a Plane Triangle. IEEE transactions on Biomedical Engineering, No. 2, pp. 125–126.

[9] Van Rijsbergen, M., Foeth, E.-J., Fitzsimmons, P. and Boorsma, A. (2012). High-Speed Video Observations and

Acoustic-Impact Measurements on a NACA 0015 Foil. Proceedings of the 8th International Symposium on Cavitation, Singapore.

[10] Vogel, A. and Lauterborn, W. (1988). Acoustic Transient Generation by Laser-Produced Cavitation Bubbles near Solid Boundaries. The Journal of the Acoustical Society of America. 84(2), pp. 719-731.

Baltimore, Maryland, USA, May 14 – 16, 2018

*Corresponding Author, Wei Wang: wangw@dlut.edu.cn

Experiment Research on Cavitation Control by Active Injection

1Shengpeng Lu; 1Wei Wang*; 1Tengfei Hou; 2Mindi Zhang; 1Jianxiong Jiao; 1Qingdian Zhang; 1Xiaofang Wang;

1Key Laboratory of Ocean Energy Utilization and Energy Conservation of Ministry of Education, Dalian University of

Technology, Dalian 116024, China; 2School of Mechanical Engineering, Beijing Institute of Technology, Haidian district,

Beijing, 100081, China Abstract

Based on the active control strategy of the flow, this paper proposed a method of suppressing cavitation by arranging water jet holes on the suction surface of the hydrofoil. The idea of cavitation control was tested in the experiments. The NACA0066 hydrofoil was placed in the test channel at the attack angle of 8°. The cavitation conditions were distinguished by cavitation number and the mass flow coefficient of the jet flow. High-speed flow field display technology was used to study the characteristic of cavitation. The influence of the jet flow mass flow on cavitation suppression was examined. The results show that the jet can suppress or weaken cavitation. With the decreasing of the cavitation number, the effect of the cavitation suppression was weakened, the mass flow of the jet needed for cavitation suppressing is increased. And the optimum mass flow coefficient of the jet flow was get under different cavitation number.

Keywords: active control; cavitation; jet hole; jet flow; re-entrant jet

Introduction

It has been proved in the results of experiments [1-3] and numerical simulations [4-6] that the re-entrant jet and the lateral jet are the main reason for sheet cavitation changing into cloud cavitation. So, it can be started with suppressing the re-entrant jets or the lateral jets to suppress the transformation of sheet cavitation to cloud cavitation. At present, there are two main methods to suppress the re-entrant jets, which are based on different flow control strategies respectively. According to the different control methods, the flow control technology can be divided into active control technology and passive control technology. The active control technology is based on the input of external energy, and the control of flow is realized by coupling appropriate disturbance model with the internal flow model, such as injecting or inhaling a liquid [7-11], gas [12] or polymer solution [13] [14] through surface of the test object. Passive control doesn’t require external energy input. It only changes the energy distribution of the fluid by passive devices to control the flow. It is worth noting that, for hydraulic machinery, its working conditions are often not a single working condition. The passive control method [15] [16] (for example, the leading edge flap, vortex generator, hydrofoil surface modification, etc.) has a great influence on the hydrodynamic performance of the hydrofoil. And what's more important, once a certain passive device is applied to the surface of the test object, it is difficult to realize the interactive adjustment under different cavitation conditions. So, this paper pays more attention to the active flow control technology. Based on the active control strategy of the flow, a method of suppressing cavitation by arranging water jet holes on the surface of the hydrofoil is proposed. Through the design of the microstructures on the surface of the hydrofoil, the jet is ejected along the set channel to interfere with the re-entrants, and then the aim of suppressing cavitation is achieved. According to the degree of cavitation, by changing the flow rate of the jet, the momentum and velocity of the jet is adjusted, and thus the interactive adjustment under different cavitation conditions is realized. In the way of achieving the jet ejection by active flow control technology, the pump is used to pump the fluid into the cavity (which is set up in advance) of the hydrofoil, and then the fluid is ejected along the set flow channel. In the process, the velocity of the jet ejection is regulated by adjusting the speed of the pump. The study of this paper is based on the analysis of the visual observations of cavities on three angle of views at attack angle of 8° gathered by high-speed imaging.

Experiment Set up

All experimental study of this paper is carried out in cavitation tunnel of Beijing institute of technology. Its basic structure includes contraction, diffuser, experimental section, curved section, inlet pipe and return pipe, and the upper and lower side are equipped with three high strength transparent organic glass window to observe the flow field of the situation of water tunnel experiment section. The basic sizes of cavitation tunnel as listed in table 1. In the process of experiment, the dynamic system of tunnel is used to control and drive the water circulation in tunnel. Power system

Baltimore, Maryland, USA, May 14 – 16, 2018

Cytaty

Powiązane dokumenty

• Main Points – A novel solver available for wave overtopping over sloping and vertical structures; – Necessary to consider the effects of surface tension and to use an

So to simplify the steps towards the use of a generic Energy Matching Model for combined renewable power supplies, each of the individual energy chains in the

We have analytically derived that for an energy harvester, comprising a piezoelectric capacitor with mechanical load, the maximum output energy per unit volume, under a

1 rozporządzenia w sprawie prowadzenia listy stałych mediatorów nie po- woduje zatem, iż sam fakt przedłożenia któregokolwiek z nich stanowi potwierdzenie posiadanej wiedzy i

Joanna Mrzyk, Martyna Paszek Bibliografia przekładów literatury czeskiej w Polsce w 2013 roku.. Przekłady Literatur Słowiańskich

Een klein deel van de linkervleugel staat vrij dicht op de tegenoverliggende school en van de kamers die daar liggen zijn de ramen zoveel mo- gelijk naar buiten geplaatst, zodat

Do gorącej debaty prasowej wo- kół książki Clarka przyczynił się nie- wątpliwie sam temat − często dysku- towane, a mimo to trudne do zrozu- mienia przyczyny apokalipsy, która

W przypadku zgonu na terenie Związku R adzieckiego obyw atela polskiego, prawo spadkobierców p olskich do przyjęcia pozostaw ionych przez zm arłego nieruchom ości