Vous êtes sur la page 1sur 13

http://cav.safl.umn.edu/tutorial/introduction.htm Introduction to Cavitation Cavitation is defined as the formation of the vapor phase in a liquid.

The term cavitation can imply anything from the initial formation of bubbles (inception) to large-scale, attached cavities (supercavitation). The formation of individual bubbles and subsequent development of attached cavities, bubble clouds, etc., is directly related to reductions in pressure to some critical value, which in turn is associated with dynamical effects, either in a flowing liquid or in an acoustical field. Any device handling liquids is subject to cavitation. Cavitation can affect the performance of turbomachinery resulting in a drop in head and efficiency of pumps and decreased power output and efficiency of hydroturbines. The thrust of propulsion systems can be cavitation limited and the accuracy of fluid meters can be degraded by the process. Noise and vibration occur in many applications. In addition to the deleterious effects of reduced performance, noise and vibration, there is the possibility of cavitation damage. The extent of the damage can range from a relatively minor amount of pitting after years of service to catastrophic failure in a relatively short period of time. Cavitation Scaling The fundamental parameter in the description of cavitation is the cavitation index that is a special form of the Euler Number:

wherein p0 and U0 are a characteristic pressure and velocity, respectively, is the density, and pv is the vapor pressure of the liquid. Various hydrodynamic parameters such as lift and drag coefficient, torque coefficient, and efficiency, are assumed to be unique functions of when there is correct geometric similitude between the model and prototype. Generally speaking, these parameters are independent of above a certain critical value of . This critical value is often referred to as the incipient cavitation number, i. It should be emphasized, however, that the point where there is a measurable difference in performance is not the same value of where cavitation can be first detected visually or acoustically. Thoma's Sigma is another version of the cavitation number that is used in turbomachinery tests. It is defined by the pump or turbine head:

where Hsv is the net positive suction head and H is the total head under which a given machine is operating. T and are qualitatively equivalent. We can think of
y y
i

as a performance boundary such that

> i, no cavitation effects < i, cavitation effects such as performance degradation, noise, vibration, and damage
TC,

Other definitions of the Cavitation Inception

is also used in pump and turbine testing.

Cavitation is normally assumed to occur when the minimum pressure, pm in a flow is equal to the vapor pressure. For calculations of steady flow over a streamlined body, the incipient value of is normally determined by setting pvm:

where Cpm is the minimum pressure coefficient defined in the normal manner. This is generally an over-simplification since cavitation inception is governed by both the single phase flow characteristics (including turbulence) and the critical pressure, pc. Hence a more general form of the inception cavitation index is given by

where the second and third terms on the right-hand side of Equation 4 incorporate the effects of unsteadiness and bubble dynamics respectively. The second term, which is proportional to the intensity of pressure fluctuations due to turbulence, is very important in free shear flows and boundary layers adjacent to roughened walls [Arndt, 1981, 1995]. As shown in Figure 1, experimental data suggest that the rms pressure fluctuations are proportional to the shear stress coefficient:

Shear

stress

coefficient

is

defined

as

or

Figure 1. Cavitation inception in turbulent shear flows. (Arndt, 1981) With permission, form the annual Review of Fluid Mechanics, Volume 13, copyright 1981, by Annual Reviews Inc. T is defined as the tensile strength of the liquid, (pv - pc) that can be an important factor in cavitation testing. It is generally accepted that cavitation inception occurs as a consequence of the rapid or explosive growth of small bubbles or nuclei that have become unstable due to a change in ambient pressure. These nuclei can be either imbedded in the flow or find their origins in small cracks or crevices at the bounding surfaces of the flow. The tension that a liquid can sustain before cavitating depends on the size of nuclei in the negative, i.e., the flow is locally in tension. Measured nuclei size distributions vary greatly in various facilities. This leads to significant discrepancies in the measured value of i as shown in Figure 2. The amount of tension that can be supported in a given flow is sometimes referred to as the water quality. This factor is carefully monitored in our testing.

Figure 2. Cavitation inception index measured for the same body in different facilities. Influence of Dissolved Gas Non-condensible gas in solution can also play a role in vaporous cavitation, since the size and number of nuclei in the flow are related to the concentration of dissolved gas. Under certain circumstances, cavitation can also occur when the lowest pressure in the flow is substantially higher than vapor pressure. In this case bubble growth is due to diffusion of dissolved gas across the bubble wall. This can occur when nuclei are subjected to pressures below the saturation pressure for a relatively long period of time. Thus, for gaseous cavitation an upper limit on i is given by (Holl, 1960):

is Henry's constant and Cg is the concentration of dissolved gas. Henry's constant is a function of the type of gas in solution and the water temperature. As a rule of thumb is about 6700 Pa/ppm for air, when concentration is expressed in a mole/mole basis. In other words, water is saturated at one atmosphere when the concentration is approximately 15 ppm. This is another factor that is carefully monitored in our research. Effects of Cavitation Once cavitation occurs, a given flow field is significantly modified because the lowest pressure in the flow is vapor pressure. Thus

Since the lift coefficient of hydrofoils scales with -Cpm, this parameter will decrease with

decreasing as shown in Figure 3. The effect of cavitation on lift is directly related to the observed degradation of performance of turbomachinery due to cavitation.

Figure 3. Variation in lift coefficient on an NACA 0015 hydrofoil. Also shown is variation in unsteady lift and noise over the same range of cavitation number tested. Cavitation can also influence the vortex dynamics of a flow in subtle ways. An example is shown in Figure 4, which indicates the dependence of the frequency of vortex shedding behind wedges on the cavitation index [Young and Holl 1966]. Here the Strouhal number is normalized with respect to its non-cavitating value St, while the cavitation number is normalized with respect to its incipient value (denoted as i). The values of St and i are a function of wedge angle. Because cavitation modifies the forcing frequency due to flow over a body, there is a possibility of hydroelastic vibration if a closer match between forcing frequency and a structural mode of vibration occurs.

Figure 4. Strouhal number for vortex shedding as a function of normalized cavitation number. (Young and Holl, 1966) With permission, from the Annual Review of Fluid Mechanics, Volume 13, copyright 1981, by Annual Reviews Inc. Even in "steady' flow cavitation is basically a non-steady phenomenon. As is lowered below i, a sheet of cavitation develops on submerged bodies as shown in Figure 5 which is an example of cavitation on a hydrofoil. To the naked eye the extent of this cavitation over the surface of the body appears to be stable. However, high speed photography reveals a more complex process. Typically, a cavity forms, fills with water, detaches.

Figure 5. View of sheet and cloud cavitation on a NACA 4215 M hydrofoil. Flow is from right to left. Note the cavitating vortices at the trailing edge of cloud. Photo made in the Obernach, Germany water tunnel, courtesy of Prof. R. E. A. Arndt and Dr. A. P. Keller In many situations this process is periodic. Under these conditions, the frequency can be calculated to be approximately [Arndt et al, 1995]:

where LC is the length of the cavity. Cavitation Damage The physics of cavitation damage is a complex problem. At the heart of the problem is the impulsive pressures created by collapsing bubbles [Rayleigh, 1917]. Recent numerical techniques permit detailed examination of the collapse of individual bubbles [Blake and Gibson 1987]. This work has been complemented by a wide variety of experimental studies [Lauterborn and Bolle 1975, Tomita and Shima 1986, Vogel et al 1989]. All of these studies indicate that the final stages of collapse result in the formation of a microjet that can be highly erosive. The collapse pressure is estimated to be greater than 1500 atmospheres. In practical problems, the collective collapse of a cloud of bubbles is an important mechanism. Hansson and Mrch [1980] suggested an energy-transfer model of concerted collapse of clusters of cavities. Because of mathematical difficulties this problem has not been studied in detail until recently [Prosperetti et al 1993]. Earlier work [Wijngaarden, 1964] had already indicated the damage potential of a collapsing cloud of bubbles. Recent work supports this contention [Wang

and Brennen, 1994]. Very little is known about the correlation between cavitation damage and the properties of a given flow field. However, it is important to have in mind that cavitation erosion will scale with a high power of velocity at a given cavitation number and that cavitation erosion does not necessarily increase with a decrease in the cavitation index. It has also been observed that the cavitation pitting rate is measurably reduced with an increased concentration of gas [Stinebring et al, 1977]. An important factor is that the pitting rate scales with a very high power of velocity (typically in the neighborhood of 6). Since the velocity in turbomachinery passages is proportional to the square root of head, this also implies that the magnitude of the erosion problem is more severe in high-head machinery. Thiruvengadam [1971] has analyzed a great deal of erosion data and has concluded that, for engineering purposes, the erosive intensity of a given flow field can be quantified in terms of depth of penetration per unit time and the strength Se of the material being eroded,

The intensity I is a function of a given flow field. Many different forms of Se have been tried. The most used value appears to be ultimate strength, which is basically a weighted value of the area under a stress-strain curve [Arndt, 1990]. Although various materials have different rates of weight loss when subjected to the same cavitating flow, a normalized erosion rate versus time characteristic is often similar for a wide range of materials [Thiruvengadam, 1971]. Hence, a simplified theory allows for rapid determination of I for a given flow by measuring penetration per unit time for a soft material in the laboratory. Service life for a harder material can then be predicted from the ratio of the strengths of the hard and soft material. Although the basic physics of the damage process in turbomachinery are complex, the essential features can be simulated by experiments with partially cavitating hydrofoils in a water tunnel [Avellan et al 1988,1991, Bourdon et al 1990, Abbot et al 1993, Le et al 1993a, 1993b]. This is the focus of our current hydrofoil studies. Partial Cavitation and its Relation to Erosion A particularly important form of cavitation from a technical point of view is attached cavitation on lifting surfaces. At typical angles of attack, this takes the form of a sheet terminated at the trailing edge by a highly dynamic form of cloud cavitation. Vortex cavitation is often observed in the cloud that is caused by vorticity shed into the flow field. These cavitating micro-structures are highly energetic and are responsible for significant levels of noise and erosion, [Arndt et al, 1995]. (See Figure 6) Laboratory experimentation indicates that a variety of cavitating flow patterns are possible within the sigma-angle of attack ( - ) plane. This is illustrated in Fig.6, which is adapted from our current hydrofoil studies. In this particular example, the variety in the cavitation patterns is dominated by the interrelationship between the cavitation and the boundary layer characteristics at various angles of attack.

In spite of many excellent studies, the actual structure of this type of cavitation is still not understood. From a design point of view, cavitating flows must be modeled over a given performance envelope in the - plane in order to accurately predict performance at off- design conditions and to assess the potential for noise and erosion. This requirement is still far from being realized at the present time. For example, it is well known that the modeling of partial, steady cavities is not simple, due to the inverse character of the flow representation in the vicinity of the cavity and its wake [Furuya, 1980; Yamaguchi and Kato, 1983; Ito, 1986; Lemmonier and Rowe, 1988]. In addition, as pointed out by Kubota et al, 1992, partial cavity models cannot explain the break-up of sheet cavitation at the trailing edge into detached cavitation clouds. The process is inherently unsteady, even for steady free stream conditions. Within a certain envelope of and the process is also periodic [Franc and Michel, 1985; Le et al , 1993a,b]. This creates a modulation of the trailing cloud cavitation that is highly erosive [Arndt et al, 1995; Abbot et al, 1993, Avellan et al , 1988]. Although these details cannot be modeled with current numerical codes, Professor Song and his colleagues at the SAFL are making good progress in this direction. Hydrofoil experiments, also currently underway at SAFL, are providing insight for the development of robust numerical models. Developed Cavitation When a vapor or gas filled cavity is very long in comparison to the body dimensions, it is classified as a supercavity. Generally speaking, the shape and dimensions of vapor filled and ventilated cavities (sustained by air injection) are the same when correlated with the cavitation number based on cavity pressure. The engineering importance of supercavitation relates especially to the design of very high speed hydrofoil vessels as well as supercavitating propellers for very high speed watercraft and low-head pumps for use as supercavitating inducers for rocket pumps and other applications that require the pumping of highly volatile liquids. An example of supercavitation behind a sharp edged disk is shown in Figure 7. A detailed discussion of supercavitation can be found in Knapp et al [1979].

Figure 7. Supercavitation on a sharped edged disk. Flow is from right to left. Facilities and Techniques

Most cavitation observations and measurements are made in the laboratory. The exception to this is the recent development of cavitation monitoring techniques for hydroturbines(1). Typical laboratory facilities include: 1. 2. 3. 4. 5. Water Tunnels De-pressurized flumes De-pressurized towing tanks Pump and Turbine Test Loops Cavitation Erosion Test Apparatus

Important features that are necessary for cavitation tests include accurate, stable, and independent control of pressure and velocity, measurement equipment for velocity, pressure, temperature, dissolved gas content and nuclei content and control, and photographic and video equipment. Because of the unsteady nature of cavitation and the extremely rapid physical processes that occur during bubble collapse and erosion, many laboratories are equipped with highly specialized high speed video and photographic cameras that are capable of very high framing rates. The latest in video equipment is capable of framing rates as high as 40,500 fps. Pump and Turbine Test Loops Pump and turbine test loops are similar in concept to water tunnels. Model testing is an important element in the design and development phases of turbine manufacture. For this reason most laboratories equipped with model turbine test stands are owned by manufacturers. However, there are some independent laboratories, such as SAFL (see Figure 8), where test loops are available for cavitation testing and relative performance evaluations between competing manufacturers.

Figure 8. SAFL water tunnel

All test loops perform basically the same function. A model turbine is driven by high-pressure water from a head tank and discharges into a tail tank. The flow is recirculated by a pump, usually positioned well below the elevation of the model to ensure cavitation-free performance of the pump while performing cavitation testing with the turbine model. One important advantage of a recirculating turbine test loop is that cavitation testing can be done over a wide range of cavitation indices at constant head and flow, which is difficult, if not impossible, to accomplish in the field. Cavitation Erosion Test Facilities In many cases the service life of equipment and hydraulic structures subject to cavitation erosion can range from months to years. Because of the relatively lengthy periods required to observe measurable erosion in the field, many different techniques have been developed in the laboratory to achieve significant time compression. The time compression factor achieved in accelerated erosion tests is as high as 105 [Durrer, 1986]. Many of the devices used have little relationship to actual field conditions. For this reason they have been typically used for screening tests of different types of materials. Recent research at SAFL is aimed at relating screening tests to predictions of service life in various applications [Arndt et al, 1995] The most commonly used device is the ASTM vibratory apparatus. An oscillating horn produces a periodic pressure field that induces the periodic growth and collapse of a cloud of cavitation bubbles. A sample placed at the tip of the horn or immediately below it is easily eroded. The standard frequency of operation is 20 KHz which produces a very high erosion rate due to the rapid recycling of the cavitation process. This is the technique used at SAFL. As already mention new methods are being developed for measuring erosion rate in the field. Usually erosion rate is inferred from the measurement of noise or vibration. References P.A. Abbot, R.E.A. Arndt, and T.B. Shanahan, "Modulation Noise Analysis of Cavitating Hydrofoils" Proc. Symp. on Bubble Noise and Cavitation Erosion in Fluid Systems, ASME FED Vol 176, 1993. R.E.A. Arndt, "Cavitation in Fluid Machinery and Hydraulic Structures", Ann. Rev. Fluid Mech., Vol. 13, 1981. R.E.A. Arndt, "Hydraulic Turbines" Chapt. 4 in Hydropower Engineering Handbook (J. S. Gulliver and R. E. A. Arndt, eds.), McGraw-Hill Inc., 1990. R.E.A. Arndt, C.R. Ellis, and S. Paul, "Preliminary Investigation of the Use of Air Injection to Mitigate Cavitation Erosion" J. Fluids Engineering, Sept. 1995, (see also Proc. Symp. on Bubble Noise and Cavitation Erosion in Fluid Systems, ASME FED Vol 176, 1993).

R.E.A. Arndt, "Vortex Cavitation,", Chapt. 17, in Fluid Vortices, S. Green, ed., Kluwer, 1995. F. Avellan, P. Dupont, and I. Ryhming, " Generation Mechanism and Dynamics of Cavitation Vortices Downstream of a Fixed Leading Edge Cavity" Proc. 17th ONR Symp. on Naval Hydrodynamics, The Hague, The Netherlands, 1988. F. Avellan, P. Dupont, and M. Farhat, "Cavitation Erosion Power", Proc. ASME-JSME Cavitation '91 Symp., FED-Vol. 116, 1991. J.R. Blake, and D.C. Gibson, "Cavitation Bubbles Near Boundaries", Ann. Rev. Fluid Mech., Vol. 19, 1987. P. Bourdon, R. Simoneau, F. Avellan, and M. Farhat, "Vibratory Characteristics of Erosive Cavitation Vortices Downstream of a Fixed Leading Edge Cavity," Proc. 15th IAHR Symp., Belgrade, Yugoslavia, September, 1990. H. Durrer, "Cavitation Erosion and Fluid Mechanics" Sulzer Technical Review, 3, pp 55-61, 1986. Franc, J.P. and Michel, J. M., "Attached Cavitation and the Boundary Layer: Experimental and Numerical Treatment" J. Fluid Mech., 154, 63-90, 1985. O. Furuya, "Non-linear Theory for Partially Cavitating Cascade Flows" Proc. IAHR 10th Symp. Tokyo, Japan, 1980. I. Hansson, and K.A. Mrch, "The Dynamics of Cavity Clusters in Ultrasonic (Vibratory) Cavitation Erosion", J. Appl. Phys., 51(19), September, 1980. J.W. Holl, "An Effect of Air Content on the Occurrence of Cavitation," J. Basic Eng. 82:941946, 1960. J. Ito, "Calculation of Partially Cavitating Thick Hydrofoil and Examination of a Flow Model at Cavity Termination" Proc. Intl. Symp. on Cavitation, Sendai, Japan, 1986. Knapp, R.T., Daily, J. W., and Hammitt, F.G., Cavitation, University of Iowa Press A. Kubota, H. Kato, and H. Yamaguchi, "A New Modeling of Cavitating Flows: A Numerical Study of Unsteady Cavitation on a hydrofoil section" J. Fluid Mech., 240, pp 59-96, 1992.

W. Lauterborn, and H. Bolle, "Experimental Investigations of Cavitation-Bubble Collapse in the Neighborhood of a Solid Boundary," J. Fluid Mech., Vol. 72, Part 2, 1975. Q. Le, J.P. Franc, and J. M. Michel, "Partial Cavities: Global Behavior and Mean Pressure Distribution", J. Fluids Eng., Vol. 115, No. 2, June, 1993a. Q. Le, J.P. Franc, and J. M. Michel, "Partial Cavities: Pressure Pulse Distribution around Cavity Closure", J. Fluids Eng., Vol. 115, No. 2, June, 1993b. H. Lemonnier, and A. Rowe, "Another Approach in Modeling Cavitating Flows" J. Fluid Mech., 195, pp 557-580, 1988. A. Prosperetti, N.Q. Lu, and H.S. Kim, "Active and Passive Acoustic Behavior of Bubble Clouds at the Ocean's Surface," J. Acoust. Soc. Am., 93, No. 6, June, 1993. Lord Rayleigh, "On the Pressure Developed in a Liquid During the Collapse of a Spherical Cavity", Phil. Mag., 34, August, 1917. D.R. Stinebring, R.E.A. Arndt, and J.W. Holl, "Scaling of Cavitation Damage", J. Hydronautics, Vol. 11, No. 3, July, 1977. A. Thiruvengadam, Hydronautics, Inc., Laurel, Maryland, Tech. Rep. 233-15, 1971. Y. Tomita, and A. Shima, "Mechanisms of Impulsive Pressure Generation and Damage Pit Formation by Bubble Collapse", J. Fluid Mech., Vol. 169, 1986. A. Vogel, W. Lauterborn, and R. Timm, "Optical and Acoustic Investigations of the Dynamics of Laser-Produced Cavitation Bubbles Near a Solid Boundary", J. Fluid Mech., Vol. 206, 1989. Y.C. Wang, and C.E. Brennen, "Shock Development in the Collapse of a Cloud of Bubbles" ASME Cavitation and Multiphase Flow Forum, Vol. 153, 1994. L. van Wijngaarden, "On the Collective Collapse of a Large Number of Cavitation Bubbles in Water," Proc. 11th Int'l Cong. of Appl. Mech., Munich, Springer Verlag, 1964. H. Yamaguchi, and H. Kato, "On Application of Nonlinear Cavity Flow Theory to Thick Foil Sections" Proc. Conf. on Cavitation, Inst. of Mech. Eng., Edinburgh, Scotland, 1983. J.O. Young and J.W. Holl, "Effects of Cavitation on Periodic Wakes Behind Symmetric Wedges," J. Basic Engr. 88:163-176, 1966.

Vous aimerez peut-être aussi