You are on page 1of 15

Sixth International Symposium on Cavitation

CAV2006, Wageningen, The Netherlands, September 2006

INCUBATION TIME AND CAVITATION EROSION RATE


OF WORK-HARDENING MATERIALS

Jean-Pierre FRANC and Michel RIONDET


LEGI, BP 53, 38041 Grenoble Cedex 9, France, Jean-Pierre.Franc@hmg.inpg.fr

ABSTRACT Secondary effects may affect this schematic behavior


A phenomenological model for the prediction of and more complicated evolutions are found in the
incubation time and erosion rate of ductile materials is literature depending in particular upon the test facility
proposed. It is based upon a simplified description of the (see e.g. Karimi and Martin 1986). For example, the
erosive potential of the cavitating flow made of three steady-state period may be almost inexistent or the
integral parameters: rate, mean amplitude and mean size maximum may be followed by a deceleration or even
of hydrodynamic impact loads. The limited number of oscillations of the erosion rate. These effects are generally
parameters considered here allows us to derive simple due to an interaction between the cavitating flow and the
equations for the prediction of the incubation time and wall via for instance changes of roughness or wall shape
erosion rate. These equations point out two characteristic induced by the erosion itself. They are ignored in the
scales for cavitation erosion of ductile materials. The present paper which focuses on the three basic stages
relevant time scale of the phenomenon is the covering described above.
time (the time necessary for the impacts to cover the This paper is based upon an analysis proposed in
material surface) whereas the pertinent length scale for 1987 by Karimi and Leo and adapted in 2002 by
ductile materials is the thickness of the hardened layer. Berchiche et al. The method is applicable to ductile
The incubation time is proportional to the covering time materials which undergo work hardening when exposed
with a multiplicative factor which is strongly dependent to cavitation.
upon flow aggressiveness in terms of the mean amplitude The starting point of the predictive technique is made
of impact loads. As for the erosion rate, it is scaled by the of pitting tests from which the erosive potential of the
ratio of the depth of hardened layers to the covering time cavitating flow is characterized. The measured size and
with still a dependence upon flow aggressiveness. The depth of each pit are considered as the signature of the
model is supported by erosion tests conducted in a collapse of a bubble and used to infer the characteristics
cavitation tunnel at a velocity of 65 m/s on stainless steel of the original hydrodynamic impact load. Berchiche et
316 L. Flow aggressiveness is characterized from an al. (2002) characterized the aggressiveness of the
analysis of pitting tests. Long duration tests are performed cavitating flow by a spectrum of impact loads. They
in order to determine the incubation time and the mean reproduced it numerically a large number of times and by
depth of penetration rate MDPR and validate the modeling the material response, they could compute mass
theoretical approach. loss as a function of the exposure time. Because of the
wide spectrum of pits, they used a numerical treatment
1. INTRODUCTION which gives realistic evolutions in time and space of the
The evolution of cavitation erosion with the exposure damage but makes difficult a phenomenological analysis.
time shows different periods. Three main stages are In order to point out the basic parameters of the
generally distinguished (see e.g. Thiruvengadam 1974, model and their influence mainly on the incubation time
Hammitt 1979). At the very beginning, the erosion and the steady state erosion rate, the approach has been
damage is made of isolated pits which progressively re-formulated in a simplified way. The simplification
overlap as the exposure time increases. Usually, consists in representing the whole spectrum by a limited
negligible mass loss occurs during this incubation period. number of integral parameters. Analytical relationships
Then, mass loss actually begins and an acceleration can then be derived, from which it is much easier to draw
period is observed with a damage rate gradually general trends and in particular point out the relevant
increasing up to a maximum. A steady state period with a length and time scales of the erosion mechanism of
constant erosion rate is then expected as long as the flow ductile materials.
pattern remains unchanged.

1
2. PHENOMENOLOGICAL ANALYSIS In the present paper we have chosen to use a Ludwik
type consolidation relationship between stress σ and
2.1 Erosive Potential strain ε (figure 1):
The real cavitating flow comprises a large variety of σ = σY + K ε n (1)
vapor structures. When collapsing, each structure is σ
supposed to generate a load on the wall. A large variety of σU
impact loads of various sizes and amplitudes is then
generated. The simplification which is made here consists
in characterizing the whole impact load spectrum by a σ = σY + K ε n
σY
mean value σ of the amplitude and a mean value S of
the impacted area.
The present model of impact is the simplest one
which can be imagined since the radial distribution of εU ε
stress is considered as uniform: stress is equal to σ over
the surface S and zero outside. Moreover, no indication Figure 1: Typical stress / strain relationship.
is given on the time evolution of the applied stress. This Let us observe that, although equation (1) is strictly
means in particular that the strain rate, which is known to speaking valid only up to the ultimate strength, it will be
be high in cavitating flows, is ignored at this step. extrapolated beyond σ U . This is the basis of the
Another important parameter with respect to flow computation of mass loss rate.
aggressiveness is the number N per unit time and unit All along this paper, we will take the example of
surface area of these impacts. In summary, the stainless-steel 316 L which has been used for the
aggressiveness of the cavitating flow is characterized in experimental part of this work. Typical values for
this paper by a set of three parameters σ , S and N . SS 316 L are (Berchiche et al. 2002):
The averaging procedure to represent a complex ⎧σ Y = 400 MPa
spectrum by a unique set of values ( σ , S , N ) is ⎪σ = 1020 MPa
presented in more detail in section 4.1. The general idea ⎪ U
⎨ (2)
is to preserve the total energy impacting the wall. ⎪ K = 900 MPa
From a purely dimensional viewpoint, a fundamental ⎪⎩n = 0.5
characteristic time can be built on the basis of the last two The erosion damage depends primarily upon the
parameters. It is defined by τ = 1 / NS . From a physical mean amplitude σ of the impact loads relatively to the
viewpoint, this time can be interpreted as the covering two previous thresholds. Three cases can happen.
time i.e. the time necessary for the surface to be fully (i) If σ < σ Y , the impacts are supposed to cause no
covered exactly once by the cavitation impacts. This damage at all, whatever their size and rate may be. The
characteristic time plays a major role as shown later. elastic behavior is dominating.
(ii) If σ lies between σ Y and σ U , successive
2.2 Material thresholds
impacts cause first the progressive hardening of the
To estimate damage caused to the wall, it is initially virgin material and then its damage after the
necessary to compare the mean amplitude of impact loads work hardening process is completed. This is why an
σ to specific material characteristics used as thresholds. incubation time is needed before erosion actually starts.
In the present paper and in accordance with the initial (iii) Finally, if σ > σ U , mass loss appears from the
version of the model (Karimi and Leo, 1987), we very beginning of exposure to cavitation. This can be
systematically refer to tensile tests to define these considered as the limit of cavitating flows of high
thresholds using the ordinary yield strength σ Y and aggressiveness. Beyond σ U , severe cavitation damage is
ultimate tensile strength σ U . expected.
More appropriate material tests could certainly be
used to better account for the very specific type of 2.3 Work-hardening
loading due to bubble collapse. One limitation among For a virgin material, strain is zero everywhere inside
others is that ordinary tensile tests are conducted at a the material. As it is exposed to bubble collapses, the
strain rate several orders of magnitude smaller than that successive hydrodynamic impacts lead to a progressive
of cavitation. However, within the same simplified superficial hardening. The strain profile inside an eroded
approach, the high strain rate encountered in cavitation sample can be determined from microhardness
erosion can be taken into account, to some extent, by measurements on a cross section (Berchiche et al., 2002).
artificially increasing σ Y and σ U with respect to the A classical representation of the strain profile inside
values given by a classical tensile test at much smaller the material is (figure 2):
strain rate. θ
⎡ x⎤
ε ( x ) = ε 0 ⎢1 − ⎥ (3)
⎣ l⎦

2
where ε 0 is the strain on the surface, x the distance from Equation (6) shows that the thickness l of the
the surface, l the thickness of the hardened layers and θ hardened layer increases as ε 1 / θ . In particular, l is
a shape factor for the strain profile. connected to the maximum value L by:
θ
ε0 ε0 ⎡ l ⎤
material surface ε = (7)
εU ⎢⎣ L ⎥⎦

2.4 Single impact


θ
⎡ x⎤ The present approach is based upon energy
ε ( x ) = ε 0 ⎢1 − ⎥
⎣ l⎦ considerations. A key concept is the energy absorbed by
l the material when submitted to a single hydrodynamic
x impact ( σ , S ). This energy is assumed to be independent
of the degree of hardening of the material and is
Figure 2: Typical strain profile inside a material subject evaluated from the particular case of a virgin material.
to work-hardening. Let us denote by ε 1 the strain on the material surface
The thickness l of the hardened layer progressively resulting from an impact of amplitude σ on the virgin
increases with the exposure time and reaches a maximum material (figure 4a). There is a unique correspondence
(denoted L ) when the work-hardening process is between σ and ε 1 via the material stress-strain
completed. The surface strain is then εU corresponding relationship (1). The ε 1 -variable can then be used instead
to σ U (cf. figure 1): material is ready to rupture at its of σ . In case σ > σ U , the approach remains valid
surface. provided the stress/strain relationship is extrapolated
Typical values of L and θ for stainless-steel 316 L beyond σ U .
are (Berchiche et al., 2002): The material surface behaves as indicated
⎧ L = 200 µm schematically on figure 4a. The surface strain is ε 1 and
⎨ (4) corresponds to the impact load σ . The energy absorbed
⎩θ = 5 per unit volume by the most superficial layer is the area
To predict the evolution of l with ε 0 , let us observe below the stress/strain curve. As for internal layers, strain
that the strain profile can be considered as simply ε ( x ) decreases with distance x from the surface
translated inside the material as work hardening proceeds. according to equation (3) and the energy absorbed (which
This is a property of the power-law chosen to describe the is still the area below the curve) is smaller for inner layers
strain profile. In other words, if ε 0 increases up to ε '0 than for the most superficial ones (figure 4b).
whereas l increases up to l' (figure 3), the new strain The total energy absorbed by the material can be
profile is: computed as:
θ l ⎡ ε( x )
⎡ x⎤ ⎤
ε ( x ) = ε '0 ⎢1 − ⎥
⎣ l' ⎦
and we have ε( x ) = ε 0 for x = l' −l , i.e.
(5)
∫ ∫
W ( ε1 ) = ⎢
x =0 ⎢⎣ ε =0
σ dε ⎥ S dx
⎥⎦
(8)

It is given by:
θ
ε0 ⎡ l ⎤ 1/θ
= (6) ⎛ε ⎞ σ Y + β Kε 1n
ε '0 ⎢⎣ l' ⎥⎦ W ( ε 1 ) = ε 1 S L ⎜⎜ 1 ⎟⎟ (9)
⎝ εU ⎠ 1+θ
material surface ε 0 ε '0 ε with
1+θ
β= (10)
( 1 + n )( 1 + θ + nθ )
σ σ
σU σU
σ
l
σY σY

l'
ε1 εU ε ε( x ) εU ε
x (a) (b)
Figure 3: Work hardening progress. Figure 4: Behavior of a virgin material after being loaded
by a hydrodynamic impact of amplitude σ . (a) case of
the superficial layer. (b) case of a layer at depth x.

3
L is the maximum thickness of the hardened layers when thickness L of the hardened layers) and upon the flow
work hardening is complete whereas the quantity aggressiveness σ , S and N . Concerning the last two
1/θ parameters, the incubation period appears to be simply
⎛ε ⎞
L ⎜⎜ 1 ⎟⎟ (11) proportional to the covering time τ = 1 / NS .
⎝ εU ⎠ The dependence with respect to the amplitude of
is the thickness of the hardened layers for a partial impact loads σ is more complicated. When σ
hardening due to the only considered impact σ (cf. approaches σ Y , the incubation time tends to infinity.
equation 7). This is consistent with the assumption that no damage
will occur if the amplitude of impacts is smaller than the
2.5 Incubation period yield strength.
On the other hand, for σ = σ U , one has T = τ . This
As exposure time increases, the material is hardened
means that mass loss is expected to happen after the
and the surface strain progressively increases. After the
material surface has been covered only once. This is quite
first covering of the initially virgin material (i.e. after
understandable from a physical viewpoint. Beyond σ U ,
time τ ), the surface strain is uniformly equal to ε 1 .
the incubation time takes no sense since erosion occurs
After the second covering (time 2τ ), the surface
from the very beginning of exposure.
strain is increased from ε 1 up to a value ε 2 which is
Between σ Y and σ U , the incubation period
deduced from the conservation of energy. Since there has
progressively decreases as flow aggressiveness (i.e. σ )
been two coverings, the flow has discharged into the
increases. As an example, the evolution of the incubation
material the energy 2 W ( ε 1 ) . Moreover, if the surface
period with the mean amplitude of impact loads is
strain is ε 2 , the total energy absorbed by the material is
presented in figure 6 for stainless steel 316 L using the
actually W ( ε 2 ) . The energy balance writes
values given before in (2) and (4). In case the mean
W ( ε 2 ) = 2 W ( ε1 ) . This equation allows the
amplitude of impact loads is only a little bit higher than
determination of the ε 2 -value.
the yield strength i.e. for flows of small aggressiveness,
σ the number of coverings necessary before mass loss
σU happens may be quite large and the incubation time may
be several orders of magnitude larger than the covering
time.
σ
10000
SS 316 L
σY
incubation time T/τ
Non-dimensional
g

g
g

1000
ver in

verin

verin
verin

2 nd co

3 rd co

4 th co
1 st co

100

10
ε1 ε2 ε3 ε4 εU ε
Figure 5: Material response during the incubation 1
0 0.2 0.4 0.6 0.8 1
period.
yield strength ultimate strength
The work hardening process is continued between
the second covering (instant 2τ ) and the third one Non-dimensional amplitude of impact loads
(instant 3τ ). The surface strain increases up to a ε 3 - Figure 6: Typical example of the variation of the
value given by W ( ε 3 ) = 3 W ( ε 1 ) and so on (cf. figure incubation period with the amplitude of impact loads σ .
5). Case of stainless steel 316 L. The values of the material
No mass loss is expected until the surface strain constants are given in (2) and (4). The abscissa
reaches the ultimate strain εU corresponding to σ U . The represents the non-dimensional amplitude of impact loads
number of coverings necessary to reach this critical point defined by ( σ − σ Y ) /( σ U − σ Y ) whereas the ordinate is
is n = W ( εU ) / W ( ε 1 ) and the corresponding time is the incubation time T made non-dimensional using the
nτ . This time is precisely the incubation period T covering time τ.
beyond which mass loss is expected. Using equation (9)
2.6 Steady-state mass loss rate
together with the stress-strain relationship (1), the
incubation time is given in a non-dimensional form by: After the incubation period, surface strain is
1+θ maximum and equal to εU . Any further impact will
T ⎡ σ U − σ Y ⎤ nθ
σY + β ( σU − σY ) cause material removal.
=⎢ ⎥ (12)
τ ⎣ σ − σY ⎦ σY + β ( σ − σY ) After a further complete covering of the material
The incubation period depends upon the material surface, strain at surface reaches a value ε ' greater than
characteristics σ Y , σ U , n and θ (but not upon the εU (cf. figure 7). This ε ' -value is extrapolated beyond
the ultimate tensile strength on the stress-strain

4
relationship. It can be considered as virtual in so far as all layers (which is a feature of the material only) whereas
the layers of the material with a strain greater than εU the characteristic time to be considered is the covering
are indeed assumed to be ruptured and eroded. This ε ' - time τ (which is a feature of the fluid flow). The term in
value is computed, as previously, from an energy balance brackets is a multiplicative factor which depends
which writes: primarily upon flow aggressiveness. It is estimated below
W ( ε' ) = W ( εU ) + W ( ε 1 ) (13) in the case of stainless steel 316 L.
A practical way to compute MDPR is as follows.
σ
From the mean value of the impact load σ , strain ε 1 is
σU computed using the stress/strain relationship and the
corresponding energy W ( ε 1 ) is computed using equation
σ
(9). The energy conservation equation (13) is then solved
in order to determine ε' . Since this equation cannot be
σY solved analytically, we cannot get a fully analytical
equation for MDPR. The mean depth of penetration rate
can then be computed using equation (15).
A typical evolution of the erosion rate with flow
aggressiveness is presented on figure 9, still in the case of
ε1 εU ε' ε stainless steel 316 L. In the limit case σ = σ Y , one has
Figure 7: Material behavior after the incubation period W ( ε 1 ) = 0 , ε ' = εU and then MDPR=0 as expected.
during steady-state erosion. Figure 9 shows that MDPR increases first slowly with σ
around the yield strength and tends to be more or less
Once the new surface strain ε ' is known, the new linear with σ around the ultimate strength.
strain profile inside the material can be computed. Both For σ < σ U , damage is due to fatigue associated to
profiles, just after the incubation period and after a further the accumulation of energy by the material. The present
covering, are compared in figure (8). They are simply approach remains valid for severe cavitation erosion
translated of a length ∆L given by: ( σ > σ U ).
1/θ
∆L⎡ ε' ⎤ 0.5
Non-dimensional erosion rate

=⎢ ⎥ −1 (14) SS 316 L
L ⎣ εU ⎦ 0.4
as explained in section 2.3. Previous equation is a direct
MDPR/(L/τ)

consequence of equation (7). 0.3

initial surface εU ε' ε 0.2

0.1
∆L
0
surface 0 0.5 1 1.5 2 2.5
after erosion
yield strength ultimate strength
L Non-dimensional amplitude of impact loads
Figure 9: Typical example of the variation of the mean
depth of penetration rate (MDPR) with the mean
L + ∆L
amplitude of impact loads. Case of stainless steel 316 L
(see (2) and (4) for values of constants). MDPR is made
x
non dimensional using the characteristic erosion rate L/τ
Figure 8: Principle of the computation of mass loss. whereas the non-dimensional amplitude of impact loads
Mass loss is computed by assuming that the material is defined as in figure 6.
cannot sustain a strain greater than the ultimate strain εU
so that thickness ∆L is actually eroded and removed 3. EXPERIMENTS
during time τ. The mean depth of penetration rate is then:
1/θ 3.1 Experimental facility
∆L L ⎡⎢⎛ ε ' ⎞ ⎤

MDPR = = ⎜ ⎟ − 1 (15)
τ τ ⎢⎜⎝ εU ⎟⎠ ⎥ To support the theoretical approach, erosion tests
⎣ ⎦ were conducted in a cavitation tunnel (figure 10). The
A major outcome of this equation is that the steady facility is designed for a maximum pressure of 40 bars.
state erosion rate MDPR is proportional to L / τ . In other This relatively high pressure provides with high velocities
words, from a dimensional viewpoint, it appears that the and consequently high erosive potential for the cavitating
characteristic length which is pertinent for the flow. This is an essential condition to obtain significant
computation of MDPR is the thickness L of the hardened mass loss within reasonable exposure times and make

5
possible the investigation of the most advanced stages of changing the downstream pressure pd . Cavitation
erosion. number is defined by:
The pump is a KSB centrifugal pump driven by an p − pv
electric motor of 80 kW in power. Maximum flowrate is σ= d (16)
pu − pd
11 l/s. The facility includes a downstream tank of about
The σ -value is chosen equal to 0.9 which leads to a
1 m3. Pressurisation is achieved by means of a bottle
cavity of about 25 mm in mean length. This ensures that
connected to the tank by a pipe of small diameter in order
the zone of maximum erosion which corresponds to
to limit diffusion of nitrogen used for pressurisation.
cavity closure is in an appropriate region of the target.
Liquid is tap water without any special control of
The tests presented in this paper are conducted at a
dissolved gas.
flowrate of 6.25 l/s. The upstream pressure is 21.3 bar
Water temperature is kept constant by means of a
and the pressure drop through the test section given by:
heat exchanger. It is a countercurrent exchanger made of
85 tubes of 11 mm in diameter with a nominal power of 1
pu − pd ≅ pu (17)
80 kW. Tests are conducted at ambient temperature. The 1+σ
rise in temperature after a run of 5 hours is typically of is 11.2 bar at σ = 0.9 . The mean velocity in the 16 mm
the order of 0.5°C. nozzle is 31 m/s. The reference velocity on the cavity
The tunnel is equipped with several transducers to where pressure is assumed to be equal to the vapour
determine the operating conditions: an electromagnetic pressure can be estimated from Bernoulli equation in
flowmeter, a pressure transducer to measure the absolute which the vapour pressure is neglected in comparison to
pressure upstream of the test section pu , a differential the upstream pressure:
pressure transducer to measure the pressure drop through 2 pu
Vc ≅ (18)
the test section and a temperature probe. ρ
For present tests, the reference velocity on the cavity
is Vc = 65.3 m / s .
cavity

Φ 16 nozzle
2.5 mm

eroded target
zone of 19 mm
maximum
erosion 32 mm

0 1 2 3 4 5 6 7 8 9 10 11 12 mm
0
0.2
0.4
0.6
0.8
1
1.2
1.4

mm

3.05 µm

13 mm
1.5 mm

Figure 11: Schematic view of the test section and of an


eroded target. The size of the eroded surface shown is
13 mm x 1.5 mm. The operating conditions are
Vc = 65.3 m / s , pu = 21.3 bar , pd = 10.1 bar , σ = 0.9 .
Figure 10: View of the cavitation erosion tunnel. Test duration: 30 mn.
The test section is about 6 m above the pump in order The target to be eroded is facing the nozzle exit.
to avoid any cavitation of the pump. It is axisymmetric Erosion appears in the form of a ring centred on cavity
and made of a nozzle of 16 mm in diameter followed by a closure where erosive potential is maximum. Both sides
radial divergent of 2.5 mm in thickness (figure 11). Under of the channel are expected to be eroded, but the nozzle is
cavitating conditions, a cavity is attached to the nozzle made of a highly resistant material so that no significant
exit whose length can be adjusted by changing the erosion is observed on the nozzle itself.
cavitation number which is obtained in practice by Erosion tests are conducted on stainless steel 316 L
free of any corrosion effect in water. Prior to exposure to

6
cavitation, a metallurgical polishing of the specimen major importance to limit overlapping in order to obtain
surface is carried out with successive diamond pastes an unbiased estimate of erosive potential.
down to 0.25 µm which ensures a mirror-type polishing. The degree of overlapping is controlled by the
The effect of roughness on cavitation erosion damage is exposure time and, statistically speaking, it can be
not considered in the present investigation. reasonably assumed that it is negligible if the exposure
time is small enough. In practice, it has to be compared to
3.2 Pitting tests the covering time introduced in section 2.1. In the case of
figure 12, the covering time is estimated to 88 mn
Pitting tests are used here to characterize the erosive
whereas the exposure time is only 5 mn. It can then be
potential of the cavitating flow. The idea of using the
concluded that overlapping is negligible. Photograph in
material itself as a kind of transducer to reveal the flow
figure 12 confirms that, at least qualitatively, overlapping
aggressiveness is classical (see e.g. Knapp et al., 1970).
is only occasional. Provided the exposure time is much
The procedure consists in considering that each
smaller than the covering time, it can be expected that the
indentation is the signature of a bubble collapse. In a first
analysis of a pitting test to evaluate flow aggressiveness is
approach, it can reasonably be assumed that pit depth is
independent of exposure time.
representative of maximum load and pit size of the extent
The smaller the exposure time, the smaller the pit
of the loaded area. This is the assumption made in section
number and the less accurate the estimate. Conversely,
4.1 below.
the larger the exposure time, the larger the degree of
An alternative for characterizing flow aggressiveness
overlapping. It is then necessary to find a compromise. A
is to use pressure transducers. Li et al. (1986) have shown
possibility can be to make several pitting tests of rather
that the temporal pressure fluctuations given by a
short duration and cumulate the data to get a stable
transducer flush-mounted in the cavitating region is made
statistics (cf. section 4.1).
of several components. In addition to basic flow noise
Furthermore, it is essential that the exposure time be
and low-frequency fluctuations associated to the global
sufficiently short so that large portions of the original
behaviour of the two-phase region, Li et al. observed high
surface remains easily detectable between indentations.
frequency pulses due to cavitation bubble collapse.
This is important for the analysis technique since the
Several investigators (e.g. Nguyen The et al., 1987, Fry,
original surface is used as a reference for the
1989, Okada et al., 1995) have shown that a good
determination of pit depth.
correlation exists between erosion damage and the
number of pulses above a suitable threshold.
However, the quantitative evaluation of pressure load
due to bubble collapse by pressure transducers raises a
few basic difficulties. The very high characteristic
frequency of cavitation pulses requires transducers of
small rise time in order to follow reliably the fast pressure
rise. Very significant progress has been made in this field
by the development of special transducers (Okada et al.,
1995, Momma & Lichtarowicz, 1995). But a major
difficulty remains concerning the size of the sensitive
surface which, although miniaturized, is much bigger than Figure 12: Typical view of the surface after a pitting test.
produced indentations. Even though realistic estimates Operating conditions: Vc = 65.3 m / s , pu = 21.3 bar ,
can be obtained by dividing the measured load in terms of pd = 10.1 bar , σ = 0.9 , exposure time: 5 mn. The size of
force by the area of the indentation (cf. e.g. Momma & the surface is 2 mm x 4 mm.
Lichtarowicz, 1995), pressure transducers cannot directly Several techniques are available to analyze a pitted
supply with pressure load. This limitation lead us to surface. Belahadji et al. (1991) used an interferometric
prefer pitting tests to evaluate flow aggressiveness. technique and Fortes Patella et al. (2000) a laser
Nevertheless, the estimate of pressure load from profilometer. In the present work, we used a Taylor-
indentation characteristics is not straightforward and a Hobson contact-type profilometer with a stylus of tip
special procedure presented in section 4.1 is needed. radius 2 µm and a vertical maximum resolution of
Figure 12 presents a typical view of a pitted surface. 3.2 nm. The surface is scanned along parallel lines distant
A pitting test can be considered as appropriate if the of 1 µm with a resolution of 0.5 µm between two
degree of pit overlapping is low enough. It is clear that successive points. For most tests, a surface of 2 mm x
the response of a ductile material is not the same whether 4 mm is scanned.
the impact falls on a virgin part of its surface or an Treatment of the data starts by subtracting a
already pitted area. This is because of the increase in polynomial obtained by the least mean square technique.
superficial hardness due to pitting. Since the analysis of After several tests, a polynomial of degree 5 was
the pitting test presented in section 4.1 is based upon the considered. It proved to satisfactorily remove the mean
assumption that each pit falls on a virgin area, it is of shape of the surface often due to large scale machining or

7
polishing defects without altering the shape of pits at both parameters can be considered as independent. In
much smaller scale. other words, large pits are not necessarily deeper and pit
Next step consists in applying a threshold to the profiles are not geometrically similar.
surface in order to define the pit border. For all the
analyses presented here, the threshold was fixed at 5
0.5 µm below the reference surface corresponding to the
original material surface. All pits whose depth is smaller
4
than 0.5 µm are then ignored. This threshold is constant
and independent of the indentation. Its influence on the

Pit depth (µm)


subsequent prediction of incubation time and mass loss 3
rate is discussed in section 4.2. The chosen value of
0.5 µm appears to give a correct description of the pits
(cf. figure 28). Other investigators (Belahadji et al., 1991, 2
Fortes Patella et al., 2000, for instance) consider a pit
dependent threshold often taken as a fraction (typically 1
10%) of maximum pit depth. The influence of the
analysis technique on the estimate of the erosive potential
of the flow and the further estimate of long-term damage 0
has not been investigated. 0 50 100 150 200 250
Pit diameter (µm)
Figure 15: Pit depth versus pit diameter. Each point
represents a pit. The total number of pits is 797.

1,5

1,3
Pit depth (µm)

1,1

Figure 13: 3D view of the pitted surface presented in 0,9


figure 12. The size of the volume shown is 2 mm x 4 mm x
2.8 µm. Same operating conditions as for figure 12. A
0,7
polynomial of degree 5 was removed to eliminate large
scale defects.
0,5
Each pit is then identified, and its main
20 30 40 50
characteristics (surface, volume, maximum depth and Pit diameter (µm)
mean diameter) are computed. Pit size is defined as the
equivalent diameter of the section of the pit by the plane Figure 16: Same representation as figure 15 but in a
x = 0.5 µm . Pit depth is counted from the original limited range of depth and diameter. No clear correlation
material surface x = 0 and volume is defined as shown in between pit diameter and pit depth is visible.
figure 14.
The distribution of pits as a function of diameter is
x =0 shown in figure 17. The smaller the pits, the higher the
pitting rate.
depth
As for the contribution of each class of size to the
0.5 µm
damage, it can be estimated by considering the eroded
diameter volume surface defined here as the total surface of pits for the
considered class of size. The corresponding cumulative
x and probability density functions are shown in figure 18.
Figure 14: Sketch showing the definition of pit depth, PDF exhibits a maximum for pit diameters around
diameter and volume. 100 µm. Small pits are many but do not contribute
significantly to the eroded surface because of their small
Figures 15 and 16 give a representation of the sample size. The contribution of large pits is also negligible but
of almost 800 pits analyzed here in a diagram whose because of their small probability. The existence of a
coordinates are pit depth and pit diameter. No correlation maximum is then quite understandable. It has also been
is noticeable between pit depth and pit diameter so that

8
observed by Belahadji et al. (1991) for pitting tests in a 1 0,01
Venturi.

Cumul. fraction of erod. vol


The conclusion is the same when considering the

PDF of fraction of erod. vol.


eroded volume defined as the total volume of pits for a 0,8 0,008
given class of size. In particular, the PDF of the eroded
volume exhibits a maximum in the same range of size,
0,6 0,006
around 100 µm, as shown in figure 19.
For the prediction of incubation time and mass loss
rate, it is essential that the technique of analysis of pitting 0,4 0,004
tests focuses on the range of pit diameter which
contributes mostly to the damage. In the present case, the
0,2 0,002
contribution of pits smaller than 20 µm was ignored
whereas the largest observed pit was 220 µm in diameter.
0 0
5 0,1 0 50 100 150 200 250
Pit diameter (µm)
Cumul. pit dens. (pits/cm²/s)

4 0,08 Figure 19: Cumulative distribution and probability


PDF of pit density

density functions of the fraction of eroded volume.


3 0,06
Operating conditions: Vc = 65.3 m / s , pu = 21.3 bar ,
pd = 10.1 bar , σ = 0.9 , total number of pits: 797,
analyzed surface: 59.9 mm²).
2 0,04
3.3 Mass loss tests

1 0,02 A series of mass loss tests with increasing exposure


times have been conducted on a SS 316 L target for the
same operating conditions as the pitting tests presented in
0 0
section 3.2.
0 50 100 150 200 250
Figure 20 presents a view of the material surface
Pit diameter (µm)
after the maximum exposure time of 104 h. Damage is
Figure 17: Cumulative distribution and probability concentrated on a ring corresponding to the closure
density functions of pit density per unit time and unit region of the cavity with almost no damage inside the
surface area. Operating conditions: Vc = 65.3 m / s , cavity. On the whole, damage appears well enough
pu = 21.3 bar , pd = 10.1 bar , σ = 0.9 , total number of axisymmetric despite a few undulations in the upper left
pits: 797, analyzed surface: 59.9 mm²). quarter. The measurements were then concentrated on the
three other quarters free of defects.
1 0,01
PDF of fraction of erod. surf.
Cumul. fraction of erod. surf

0,8 0,008

0,6 0,006

0,4 0,004

0,2 0,002

0 0
0 50 100 150 200 250
Pit diameter (µm)
Figure 20: Photograph of an eroded sample after an
Figure 18: Cumulative distribution and probability exposure time of 104 h. The external diameter of the
density functions of the fraction of eroded surface. sample is 100 mm. Operating conditions: Vc = 65.3 m / s ,
Operating conditions: Vc = 65.3 m / s , pu = 21.3 bar , pu = 21.3 bar , pd = 10.1 bar , σ = 0.9 .
pd = 10.1 bar , σ = 0.9 , total number of pits: 797,
analyzed surface: 59.9 mm²). The mass loss test was bracketed by two pitting tests
in order to detect any possible alteration in flow

9
aggressiveness during the long duration test. No sample. Strictly speaking, present estimate is then a
significant variation was observed between both pitting volume loss and not a mass loss measurement.
tests so that it can be considered that flow aggressiveness The comparison of three different radial profiles
has remained invariable all along the mass loss test. This shown in figure 22 confirms a satisfactory axisymmetry
results from (i) a negligible change in geometry since the of the erosion pattern as already concluded from the
maximum depth of penetration was 70 µm after 104 h, photograph of figure 20. The data on MDPR presented
which is negligible in comparison with the 2.5 mm gap below are relative to the average of these three smoothed
(figure 11) and (ii) of a precise control of the operating profiles. On the basis of these mean profiles, two
conditions. Let us observe however that the influence of quantities are considered : (i) the maximum depth of
target roughness on cavitation was not investigated. penetration and (ii) a mean depth of penetration defined
Radial profiles of erosion are presented in figure 21 over a radial distance of 2 mm centred on the point of
for an increasing exposure time. During about the first maximum erosion. This last procedure presents the
30 hours of exposure to cavitation, there is no measurable advantage of smoothing the data and reducing the
penetration of damage which is essentially characterized dispersion since the damage is defined on a more global
by an increase in roughness. After this incubation period, basis. Both estimates of depth of penetration are shown in
material is removed and the depth of penetration regularly figure 23.
increases. Raw profiles shown in figure 21 exhibit a
strong roughness and require an averaging procedure for 0 34h
smoothing. A moving averaging technique on 1 mm was 54h
systematically applied before estimating the depth of 64h
penetration. -50 74h

0 84h
Depth of penetration (µm)

-100
0h
54h 3h
-50
64h 8h
22h -150 94h
34h
-100 74h
Depth of penetration (µm)

-200
-150 84h

-200 94h -250 104h

-250 -300

-300 104h
-350
15 20 25 30
-350
Radial distance (mm)

-400 Figure 22: Effect of direction of analysis on radial profile


of erosion for the sample presented in figure 20. For each
exposure time, four profiles are superposed (i) light gray:
-450
three mean profiles along three different radial directions
15 20 25 30
(ii) thick black: average of the three profiles. (The origins
Radial distance (mm) of horizontal and vertical scales are arbitrary).
Figure 21: Influence of the exposure time on radial The evolution of the mean depth of penetration with
profile of erosion for the sample presented in figure 20. exposure time presented in figure 23 is quite
For each time, two profiles are presented: (i) light gray: conventional. After an incubation period during which the
raw data with a step of 10 µm (ii) thick black: moving depth of penetration remains zero, an acceleration period
average data on 100 points or 1 mm. (The origins of is observed followed by a quasi linear increase of the
horizontal and vertical scales are arbitrary). depth of penetration with exposure time. The mean depth
Let us observe that, because of the massive feature of penetration rate (MDPR) is obtained by derivation with
and large weight of the sample, mass loss could not be respect to time. For further validation, special attention is
measured with a sufficient accuracy by weighting the paid to the constant value of MDPR during the steady-
state regime of erosion. The existence of a steady-state
regime characterized by a linear increase of depth of

10
penetration with time is a good indicator of the invariance This equation is used to estimate the impact load σ
of the erosive potential of the cavitating flow, confirmed from the measured depth H of the indentation. It is valid
by pitting tests. The duration of this steady-state period for flows of moderate aggressiveness, i.e. for σ < σ U .
and the further evolution of mass-loss were not In the limit case of an impact load equal to the
investigated. material ultimate strength σ = σ U , the depth of the
indentation is:
80 2 ε L
Maximum depth H= U (22)
70 Mean depth (over 2 mm) 1+θ
Depth of penetration (µm)

MDPR ≅
In the case of high erosive potential defined by
60
1.16 µm/h
1,5 σ > σ U , material is removed and the depth of removed

MDPR (µm/h)
50 material given by equation (14) has to be added to the
previous deformation (22) to obtain the actual pit depth:
40 1
Incubation time ⎡ 1 ⎤
≅ 30 h εU L ⎢⎛ σ − σ Y ⎞ nθ ⎥
30 H= + L ⎢⎜⎜ ⎟⎟ − 1⎥ (23)
1+θ σ −σY ⎠
⎢⎝ U ⎥
20 0,5 ⎣ ⎦
10
Previous equation is valid only for σ > σ U .
Equations (21) and (23) are represented on figure 24.
0 0 For stainless steel 316 L, the limit value (22) estimated
0 20 40 60 80 100 120 using experimental data (2) and (3) is 15.8 µm. Since the
Exposure time (h) maximum measured pit depth on the sample of 800 pits
considered here is 4.8 µm (cf. figure 15), it can be
Figure 23: Mean depth of penetration (in black, left concluded that all impact loads are below the material
scale) and mean depth of penetration rate (in gray, right ultimate strength. The maximum load corresponding to
scale) versus exposure time. The gray curve is the time the maximum pit depth is 777 MPa. In other words, the
derivative of the thick black curve. Data are relative to present cavitating flow at velocity 65.3 m/s can be
the average of the three profiles presented in figure 22. considered of moderate aggressiveness. The influence of
flow velocity was not investigated here.
4. ANALYSIS 40
SS 316 L
4.1 Estimation of erosive potential

The first step for the computation of incubation time 30


Pit depth (µm)

ultimate strength: 1020 MPa


and erosion rate is the estimation of the erosive potential
of the cavitating flow in terms of the three integral
20
parameters considered here: mean impact load σ , mean
yield strength: 400 MPa
area of impact load S (or equivalent mean diameter D )
and pitting rate N . 10
For each pit, the impact load responsible for it is
deduced from the measurement of its maximum depth.
Let us first consider the case of an impact load σ smaller 0
400 600 800 1000 1200
than the ultimate strength σ U of the material. Plastic
deformation occurs without any material removal and the Amplitude of impact load (MPa)
depth of the resulting indentation is: Figure 24: Pit depth versus amplitude of impact load for
l stainless steel 316 L. The two curves correspond to
H=
∫ x =0
ε ( x ) dx (19) equations (21) if σ < σ U and (23) if σ > σ U .
The previous correspondence between pit depth and
Using equation (3) for the strain profile ε ( x ) inside
impact load is systematically applied to each pit. A
the material, we obtain:
distribution of hydrodynamic impact loads is then
ε l
H= 0 (20) obtained. In order to define a mean value representative
1+θ of the whole spectrum, an averaging procedure is used
and finally, by means of equation (7): based upon energy considerations.
1+θ
From a dimensional viewpoint, the product of pit
ε L ⎛ σ −σY ⎞ nθ
volume V and impact load σ has the units of energy.
H = U ⎜⎜ ⎟⎟ (21)
1 + θ ⎝ σU − σ Y ⎠ Hence, it was decided to define the mean impact load σ
by the following averaging procedure:

11
σ =
∑Viσ i (24)
For the present cavitating flow, the mean diameter of
impact loads is about 72 µm. This value is near the
∑Vi characteristic value of pit diameter leading to maximum
where summation is relative to all identified pits. damage as observed in figures 18 and 19.
Previous definition of mean impact load is based upon the
principle of conservation of energy discharged in the 6,5
material. In the present case, the mean impact load

N (impacts/cm²/s)
6
determined by equation (24) on the sample of 797
indentations is equal to 621 MPa. This value lies between 5,5
yield strength and ultimate strength which confirms that
5
the erosive potential of the present cavitating flow is
moderate and that fatigue and work-hardening processes 4,5
are prevailing.
The mean surface of impact load S is defined by a 4 ±6%
classical averaging procedure: 3,5
1
S= Si
ν
∑ (25) 0 200 400 600
Pit number
800

where ν is the total number of pits analyzed. As for the Figure 27: Influence of pit number on impact rate per
pitting rate N , it is simply the number of pits per unit unit time and unit surface area.
time and unit surface area.
4.2 Incubation time and MDPR
650
Once the flow aggressiveness defined by the
640 quantities σ , S and N has been quantified, it is
possible to apply the predictive method developed in
630 section 2.5 and 2.6 and particularly equations (12) and
σ (Mpa)

± 0,5%
(15) to predict the incubation period and mass loss rate.
620
Results are presented in the table below. The flow
aggressiveness has been estimated from eight different
610
surfaces similar to the one presented in figure 12 relative
to the same pitting test but to eight different angular
600
0 200 400 600 800
positions on the target at the same radius corresponding to
Pit number maximum damage.

Figure 25:Influence of pit number on mean amplitude of Surface MDPR Incubation


impact load. Pit number
N° (µm/h) time (h)
2 67 1.61 17.2
85
9 105 1.20 22.9
44 98 0.85 32.3
80 47 125 1.19 23.2
D (µm)

49 110 1.08 25.5


75 59 103 1.18 23.4
91 72 0.88 31.4
70 ±1% 93 117 1.02 27.3

Total mean
65 number of mean MDPR incubation
0 200 400 600 800 pits time
Pit number 797 pits 1.1 µm/h 25.4 h
Figure 26: Influence of pit number on mean diameter of Table 1: Estimation of incubation time and erosion rate
impact load. under steady state conditions from 8 different pitting test
samples
Figures 25 to 27 present the influence of the number
of pits used for the determination of flow aggressiveness There is obviously a non negligible dependence of
on each of the three variables σ , D and N . It appears the prediction on the pitted surface considered to evaluate
that flow aggressiveness can be considered as correctly the erosive potential. Following previous discussion on
defined by a sample of typically a few hundreds of pits. the influence of the size of the sample on the evaluation
of the erosive potential of the cavitating flow, it can be

12
expected that the mean values corresponding to the total ultimate strain and subsequent mass loss are uniformly
number of pits analyzed here (797 pits) should be reached on the whole material surface. There is no
representative of the long term damage. difference between the end of the incubation period and
Table 1 shows that MDPR and incubation time are the beginning of the steady-state erosion period so that a
strongly correlated. A small MDPR is associated to a long step-like behaviour for MDPR versus exposure time is
exposure time, both being the consequence of a relatively found. A more realistic variation with an acceleration
small aggressiveness. This has been observed by several period would be obtained by considering the whole
investigators. Soyama and Futakawa (2004) for instance spectrum of impact loads (and not only average values)
have recently shown that the incubation time can be and assuming that impact loads are randomly distributed
estimated by evaluating the amount of erosion at a post- in space as shown by Berchiche et al. (2002). If so,
incubation point. ultimate strain is not instantaneously but progressively
The mean value of the predicted MDPR based on reached on the material surface depending upon the
797 pits (1.1 µm/h) is in pretty good agreement with the amplitude of successive impact loads and their possible
experimental measurement given in figure 23 partial overlapping. This results in a transitional regime
(1.16 µm/h). The predicted incubation time is 25.4 h between incubation and steady-state erosion characterized
which appears in reasonable agreement with the by a gradual increase of the erosion rate. However, in the
experimental value (about 30 h as shown in figure 23). present paper, we have deliberately chosen to ignore the
random distribution in space, size and amplitude of
impact loads in order to be able to derive analytical
relationships and more easily point out the key
parameters of the damage process.
A final remark concerns the evaluation of the
threshold

influence of the threshold chosen for the analysis of


pitting tests on the prediction of the incubation period and
erosion rate. An excessive value of the threshold tends to
underestimate the number and size of pits whereas a too
small value will overestimate them. It is difficult to derive
an objective criterion for the choice of this threshold.
Here, it is done by a qualitative comparison of the binary
0.8µm

image obtained after applying the threshold to the original


one. From figure 28, it is clear that the two extreme
thresholds 0.8 µm and 0.3 µm are not acceptable whereas
the threshold 0.5 µm is considered to give a relatively
satisfactory description of the pitted surface.

80 4,8
chosen threshold: 0.5µm

70 4,2
e
tim
0.5µm

Incubation time (h)

60 3,6
on
MD

at i

MDPR (µm/h)
ub
PR

in c

50 3,0

40 2,4
exp.
30 incubation 1,8
time
20 experimental MDPR 1,2

10 0,6
0.3µm

0 0,0
0,20 0,30 0,40 0,50 0,60 0,70 0,80 0,90
Depth threshold (µm)
Figure 29: Influence of depth threshold on predicted
incubation period and erosion rate.
Figure 28: Influence of depth threshold on the treatment
of a pitting test (same surface as in figure 12). The influence of this threshold on the predicted
incubation time and erosion rate is presented in figure 29
Let us observe that the present simplified model fails together with the experimental measurements. The higher
at predicting the acceleration period clearly visible on the threshold, the higher the incubation time and the
figure 23. This is due to the consideration of a unique smaller the erosion rate. Although the predicted values
mean value σ for the impact load. As a consequence,

13
depend upon the precise value of the threshold, figure 29 values obtained from mass loss tests. The tested material
shows that in the range 0.4 - 0.6 which is considered here is stainless steel 316 L which is known to exhibit high
as acceptable, they remain of the same order of work hardening. On the whole, the agreement between
magnitude as in experiments. predicted and measured values of the incubation time and
MDPR are satisfactory.
5. CONCLUDING REMARKS The present approach is based upon an elementary
The present paper focuses on the prediction of modeling of the flow aggressiveness. It consists in
incubation time and erosion rate of ductile materials approximating the whole impact load spectrum by mean
exposed to a cavitating flow. From a dimensional values of both the impact loads and the impacted areas.
viewpoint, the erosion rate measured in terms of the mean This results in a step-like variation of the erosion rate
depth of penetration rate (MDPR) has the dimension of a versus the exposure time with no mass loss within the
velocity i.e. of a characteristic length divided by a incubation time and a constant steady state erosion
characteristic time. The present work suggests that : beyond. The acceleration period which actually exists
(i) The relevant time scale is the covering time τ, i.e. between the incubation and the steady-state periods is in
the time necessary for the surface to be exactly covered fact due to the wide distribution of impact loads. This
by the hydrodynamic impacts. This time depends upon makes that the ultimate strength is not uniformly reached
the impact rate and the mean size of the loaded areas. It is on the material surface so that highly loaded areas
a characteristic of the cavitating flow in so far as it undergo rupture before less loaded areas. The spatially
depends mainly upon hydrodynamic features as the random nature of the hydrodynamic impacts is another
bubble density, size, distance to the wall… reason responsible for the non uniformity of the surface
(ii) The relevant length scale for ductile materials is strain. Both result in a progressive increase of the erosion
the thickness of the hardened layer L. It can be rate which is typical of the acceleration period.
determined from microhardness measurements on cross The objective of the present work was essentially to
sections of eroded specimens. This is essentially a contribute to elucidate the erosion mechanism of ductile
material characteristic, although it might integrate to materials and point out the major parameters which
some extend the very specific type of loading due to control the damaging process. Although it is still under
bubble collapses since it is determined on specimens development, the present technique of prediction of
actually eroded by cavitation. cavitation erosion can be considered as a future
MDPR can then be estimated by the ratio L/τ with a alternative to more conventional techniques based on
multiplicative factor which depends principally upon the correlations with various material properties as Vickers
flow aggressiveness in terms of the mean amplitude of the hardness (Hattori et al., 2004), fatigue tests (Bedkowski et
hydrodynamic impact loads. It is zero in the limit case of al., 1999) or others.
impact loads equal to the yield strength and progressively
increases with the flow aggressiveness. A quasi-analytical ACKNOWLEDGEMENTS
formulation is proposed to estimate this factor. The present research has been conducted in the
As for the incubation time, it is shown that it is framework of the European Project "Prevero" under
proportional to the covering time and that the ratio is contract No. ENK6-CT-2002-00605. The authors are
strongly dependent upon the flow aggressiveness. For particularly grateful to AVL, coordinator of the project,
impact loads close to the ultimate strength of the material, and all other partners of the consortium for fruitful
the incubation time matches with the covering time. discussions. The authors would also like to acknowledge
When the impact load approaches the yield strength of the the help provided by J.C. Jay, Y. Lecoffre and
material i.e. for flows of relatively small aggressiveness, J.M. Michel for the design of the test facility.
the incubation time increases considerably and can be
several orders of magnitude larger than the covering time. NOMENCLATURE
A simple model is proposed here to account for this H Maximum pit depth (m)
variation. K Constant in stress/strain relationship (1) (Pa)
The model is supported by an experimental l Thickness of hardened layer for partial
investigation in which erosion is produced by a cavitating hardening (m)
flow in a radial divergent. Pitting tests are used to L Thickness of hardened layer for complete
quantify the erosive potential of the flow. Hydrodynamic hardening (m)
impact loads are deduced from the measurement of MDPR Mean depth of penetration rate (m/s)
maximum pit depth by means of a model of response of n Exponent in stress/strain relationship (1) (-)
the material. As for the rate and surface area of the N Impact rate per unit surface area (impacts/m2/s)
impacts, they are deduced from pitting rate and pit size. pd Downstream pressure in test section (Pa)
The estimated aggressiveness of the cavitating flow is pu Upstream pressure in test section (Pa)
then introduced in the model in order to predict the pv Vapour pressure (Pa)
incubation period and mass loss rate under steady state S Mean size of impacted area (m2)
conditions. Predictions are compared to experimental T Incubation time (s)

14
V Pit volume (m3) Hattori S., Ishikura R. & Zhang Q. (2004) "Construction
Vc Reference velocity on cavity (equation 18) (m/s) of (s)
database on cavitation erosion and analyses of
W Energy absorbed by the material (equation 8) (J) carbon steel data" Wear, Vol.257, Issues 9-10,
x Distance inside the material from surface (m) November 2004, pp.1022-1029
β Parameter defined by equation (10) (-) Hattori S., Maeda K. & Zhang Q. (2004) "Formulation of
εU Ultimate strain (-) cavitation erosion behavior based on logistic analysis",
ε Strain (-) Wear, Vol. 257, Issues 9-10, November 2004, pp.1064-
ε' Virtual surface strain during steady-state erosion 1070
period (cf. section 2.6) (-) Karimi A. and Leo W.R. (1987) "Phenomenological
θ Metallurgical shape factor (equation 3) (-) model for cavitation rate computation" Mater. Sci.,
σ Cavitation number (equation 16) (-) Eng, 95, pp.1-14
or Stress (Pa) Karimi A. and Martin J.L. (1986) "Cavitation erosion of
σ Mean amplitude of impact loads (Pa) materials", Int. Met. Rev., 31, N°1, pp.1-26
σU Material ultimate tensile strength (Pa)
Knapp R.T., Daily J.W. & Hammitt F.G. (1970)
σY Material yield strength (Pa) "Cavitation", McGraw-Hill
τ Covering time 1 / NS (s)
Li S., Zhang Y. & Hammitt F.G. (1986) "Characteristics
of cavitation bubble collapse pulses, associated
REFERENCES
pressure fluctuations, and flow noise" Journal of
Bedkowski W., Gasiak G., Lachowicz C., Lichtarowicz Hydraulic Research, Vol.24, N°2, pp.109-122
A., Lagoda T. & Macha E. (1999) "Relations between
Momma T. & Lichtarowicz A. (1995) A study of
cavitation erosion resistance of materials and their
pressures and erosion produced by collapsing
fatigue strength under random loading" Wear, Vol.230,
cavitation" Wear, Vol.186-187, pp.425-436
pp.201-209
Nguyen The M., Franc J.P. & Michel J.M. (1987) "On
Belahadji B., Franc J.P. & Michel J.M. (1991) "A
correlating pitting rate and pressure peak measurements
statistical analysis of cavitation erosion pits" Journal of
in cavitating flows" Holl J.W. & Billet M.L. (eds),
Fluids Engineering, Vol.113, pp.700-706
Proc. of ASME Int. Symp. on Cavitation Research
Berchiche N., Franc J.P. and Michel J.M. (2002) "A Facilities and Techniques, FED-Vol.57, ASME, Boston,
cavitation erosion model for ductile materials" ASME December 13-18, 1987, pp.207-216
J. Fluids Eng., 124, pp.601-606
Okada T., Iwai Y., Hattori S. and Tanimura N. (1995)
Fortes Patella R., Reboud J.L. & Archer A. (2000) "Relation between impact load and the damage
"Cavitation damage measurement by 3D laser produced by cavitation bubble collapse" Wear, Vol.184,
profilometry" Wear, Vol.246, pp.59-67 pp.231-239
Franc J.P. & Michel J.M. (1997) "Cavitation erosion Soyama H., Lichtarowicz A., Momma T. & Williams E.J.
research in France: the state of the art" Journal of (1998) "A new calibration method for dynamically
Marine Science and Technology, Vol.2, pp.233-244 loaded transducers and its application to cavitation
Franc J.P., Michel J.M., Nguyen Trong H. & Karimi impact measurement" Journal of Fluids Engineering,
(1994) "From pressure pulses measurements to mass Vol.120, December 1998, pp.712-717
loss prediction: the analysis of a method" Kato H. (ed), Soyama Y. & Futakawa M. (2004) "Estimation of
Proc. 2nd Int. Symp. on Cavitation, April 1994, Tokyo, incubation time of cavitation erosion for various
Japan, pp.231-236 cavitating conditions" Tribology letters, Vol.17, N°1,
Fry S.A. (1989) "The damage capacity of cavitating flow July 2004, pp.27-30
from pulse height analysis" Journal of Fluids Thiruvengadam A. (1974) "Handbook of cavitation
Engineering, Vol.111, December 1989, pp.502-509 erosion" Hydronautics Inc., Technical Report 7301-1
Hammitt F.G. (June 1979) "Cavitation erosion: the state
of the art and predicting capability" Applied Mechanics
Reviews, Vol.32, N°6, pp.665-675

⎯⎯⎯⎯

15

You might also like