Вы находитесь на странице: 1из 11

Erosive wear rate

Erosive wear (Mechanical wear prediction and prevention. Bayer pp.393)

It is possible to approximate erosive wear, e, as the product of:

𝒆(𝒕) = 𝒌(𝒕)𝑨𝑰 𝑬𝒒𝒖𝒂𝒕𝒊𝒐𝒏 𝒈𝒆𝒏𝒆𝒓𝒂𝒍 𝒕𝒐 𝒆𝒓𝒐𝒔𝒊𝒗𝒆 𝒘𝒆𝒂𝒓 (𝟏)

Where

- k(t) is a wear coefficient that depends on the material. (erodent’s composition,


size and shape) and environmental conditions (temperature). It can be
determinate by simulative erosion test and may be a function of time.
- A is a function of the impingement angle.
- I is a factor expressing the intensity of wear situation.

For determinate both k(t) and the forms of A and I depend on erosive wear situation.
For cavitation, data indicates that long-term values of k can be as low as a third of the
peak value. For liquid and particle erosion however, the effect appears to be less
severe (e.g., typically less than a 20% reduction).

k(t) as a function of time. Principles reasons:

1. Fatigue wear process in erosion (Incubation period in fatigue wear process


before material loss occurs).

Since erosion situations generally involve a distribution of impact conditions, this


characteristics would be exhibited as a gradual increase to stable erosion rate rather
than as a sharp transition.

2. Modifications of the surface caused by erosion process.

These are used to explain differences between short and long-term behaviour.

Once these changes occur and a stable conditions is established, k becomes


independent of time. Prior to the development of this stable condition, erosion rates
may be higher or lower, depending on the effects that these changes have in erosion
behaviour.
Particle erosion

For the case of particle erosion, we have to models to consider, one is ductile mode
and the other one is brittle mode. Actual erosion behaviour is considered to be the
sum of these components.

For ductile behaviour: For brittle behaviour:

𝝅∝ 𝑨 = 𝑺𝒆𝒏𝒎 ∝
𝑨 = 𝑪𝒐𝒔𝒏 ∝ 𝑺𝒆𝒏 ( ) ∝≤ 𝜷𝟎
𝟐𝜷𝟎 𝒊 = 𝒗𝒎
𝑨 = 𝑪𝒐𝒔𝒏 ∝ ∝≥ 𝜷𝟎 𝒌 = 𝒌𝒃
𝒊 = 𝒗𝒏
𝒌 = 𝒌𝒅

Therefore:

𝝅∝ 𝒏
𝒆(𝒕) = [𝒌𝒅 𝑪𝒐𝒔𝒏 ∝ 𝑺𝒆𝒏 ( ) 𝒗 + 𝒌𝒃 𝑺𝒆𝒏𝒎 ∝ 𝒗𝒎 ]𝑴 ∝≤ 𝜷𝟎 (𝟐)
𝟐𝜷𝟎

𝒆(𝒕) = (𝒌𝒅 𝒗𝒏 𝑪𝒐𝒔𝒏 ∝ +𝒌𝒃 𝒗𝒎 𝑺𝒆𝒏𝒎 ∝)𝑴 ∝≥ 𝜷𝟎 (𝟑)

where

𝑘𝑑 , 𝑘𝑏 Wear coeffients in ductile and brittle behaviour, respectly


𝑣 Velocity of erosive particle
𝒎, 𝒏 Experimental exponents function of material
∝ Impact angle of erosive particle respect of tarjet
𝑀 The rate of supply of the erodent

Erosive wear rate (experimentally)

Wear measured (Mechanical wear fundamentals. Bayer pp. 270)

Wight loss is the method used for determining the amount of wear that occurs.
However, the resistance to erosion is measured in terms of the wear volume per gram
of abrasive. This is obtained through the use of a wear curve that is generated by to
determinate an average wear rate. The slope of this curve is then used to determine
an average wear rate.

The mass loss rate is converted to a volume loss rate by dividing by the density of the
specimen. This value wear rate is then normalized to the abrasive flow rate to provide
the erosion value, which is defining as wear volume per gram of abrasive. The smaller
erosion value, the more wear resistant the material is.
Guidelines for test duration are provided with the intention that the measurements
are made in a period of stable wear behaviour. Since 2 min or less is typically required
for stabilization. It is specified that the firs measurement be taken after 2 min.

Changes in the macro-and-micro geometry of a wearing surface (Solid particle


erosion, Kleis pp. 51) 13/03/17

In the process of erosion, changes take place both in the micro-and-macro geometry
or a wearing part. These changes are also accompanied by changes in the parameters
of wear (especially the impact angle).

Depending on the impact angle, properties of the material and particle shape, impact
scars appearing in the surface were found to vary in shape as well.

Figure 2.2a–e. Impact craters


produced by particles hitting the
metal surface at α = 90°: a –
impact of particles of sand of
0.4–0.6 mm on a cobalt target,
v0 = 80 m/s; b – the same as “a”
but on a tungsten surface; c –
impact of 0.6–0.8 mm particle
on a WC-6Co hardmetal
surface, v0 = 225 m/s; d –
impact crater produced by a 0.9
mm spherical cast iron pellet on
the 0.2% C steel target surface,
v0 = 50 m/s; e – the same as “d”,
v0 = 225 m/s

Figure 2.2a,c shows the emergence of block-shaped formations around the impact
crater brought about by shearing deformations of cobalt, whereas the trace of
impact on the surface of brittle tungsten is surrounding by radical cracks (Figure
2.2b). At a moderate impact velocity (𝑣0 = 50 𝑚/𝑠), the trace in the surface of
plastic steel – with its clear-cut outline – resembles the identation typical for Brinell
hardness test (Figure 2.2d); the impact crater obtained at high velocity (Figure
2.2e), however, is surrounded by a ridge of sparse metal which consists of material
squeezed out as a result of shearing strain.
Figure 2.3a–f. Impact craters
produced by particles hitting
metal surface at α<90°, with the
velocity vector directed from
left to right: a – craters of 0.4–
0.6 mm particles of sand from
the Männiku quarry on the
surface of 0.2% C steel, α = 3°,
v0 = 100 m/s; b – craters of 0.3–
0.4 mm particles of sand from
the Männiku quarry on 0.2% C
steel, α = 30°, v0 = 150 m/s; c –
the same as “b”; d – impact scar
left by 0.3–0.4 mm particle
from the Männiku quarry on the
hardened steel 710 HV, α = 45°,
v0 = 80 m/s; e,f – scars of 0.3–
0.4 mm particle of sand from
the Männiku quarry on the
surface of 0.2% C steel, α = 40°,
v0 = 150 m/s; e – focused on the
bottom of the crater; f – focused
on the squeezed-out lip

Figure 2.3 shows impact traces if 𝛼 < 90° . In this case, the material excluded from
the impact trace is submitted to directed shear strain. Depending on the value of 𝛼,
particle size and its position at the moment of hitting the surface, the shape of the
trace can vary. Thus, the material squeezed out from the crater may become
pressed to the front and the sides of the impact scars (Figure 2.3b) or form a lip the
front of it (Figure 2.3d,f), without any material removal from the surface.
Alternatively, the particle may remove the whole volume of material out of the
crater at the first impact already (Figure 2.3a,c).

As far as ductile materials are concerned, wear rate in the initial stage of the
process is generally lower than in permanent conditions, because only a little piece
of crater volume material is removed.

(14/03/17 pp.55)

Availible data about ratio between volumen impact craters and volume wear
material removed at low impact angles is present in table 2.1
Where

Could these results indicate for removing material from surface (i.e. the objetive of DIB
process) is better work with low angles?

Analogous test on hardened steel indicates that the values of ratio 𝑘𝑔 were
considerably higher than those of ductile steel. (Figure 2.5)

Then, Could ductile materials be more resistant to loss volume wear than brittle
materials?
Erosive wear (Fricction and wear, Giovanni Straffelini. pp.149) 14/03/17

Erosive wear

In the case of solid particle erosion (SPE), wear damage is essentially due to
mechanism of abrasive wear. In case of liquid droplet erosion (LDE) and the liquid
jets erosion (also called, cavition erosion), wear is mainly due to contact fatigue.

Solid particle erosion

The main tribological parameters affecting this wear process are the impact
velocity and the impact (or attack) angle.

The abrasive interaction with ductile materials can occur by microcutting or


microploughing. If all the plastically deformed material is removed, wear is by
microcutting and it is maximum. If all the plastically deformed material flows to the
sides of the groove, wear is by microploughing and it is zero, even if the surface
can be severely damaged (4.3.1 Abrasive wear of ductile material pp.96).

There is a contradiction with Bayer wear definition (Mechanical wear prediction and
prevention pp.1)

(15/03/17)
To understand this, it is necessary check the following topics:

Contribution of abrasion to friction (pp.44)

Abrasion takes place when a hard body exerts a ploughing or grooving action on
the surface of a softer body.

There are two types of abrasion indeed:


- Two-body abrasion takes place when hard particles are firmly fixed in one
body and are allowed to plastically penetrate the counterface and plough
groove on it. Such hard particles may be part of the material microstructure
(such as in ceramic-reinforced composites), or may come from the
surrounding environment (typical examples are the sand particles that
contaminate tribological systems). Also, two-body abrasion body process
may be exerted by the hard asperities of a surface sliding on a softer
counterface.

- Three-body abrasion takes place when hard particles trapped between two
contacting surfaces are quite free to roll. This typically occurs when both
bodies in contact are comparatively hard and the particles are not able to
remain fixed in neither of two. The hard particles may even come from a
damaging (wear) action of one of the two bodies.

Hard abrasive particles are angular in shape. A simplified model to quantify the
contribution of two-body abrasion to friction has been provided by Rabinowicz. In
this model, an angular particle is represented by a cone having a given attack
angle, 𝜃 (Fig. 2.25). Following the application of a normal load, 𝐹𝑁 , the cone
penetrates the softer material to a certain depth, determined by its hardness (or the
yield pressure). The tangential force, 𝐹𝑇 , necessary to horizontally move the cone
can be expressed by the product of the section of the groove (𝐴𝑝 ) to the average
2𝑟(𝑟 tan 𝜃)
stress for plastic deformation (𝑃𝑌 ). 𝐴𝑝 is given by: 𝐴𝑝 = = 𝑟 2 tan 𝜃, where 𝑟
2
is here the radius of the base of the cone. Therefore:

𝐹𝑇 = 𝑃𝑦 𝑟 2 tan 𝜃 (4)

𝜋𝑟 2
Since 𝐹𝑁 is given by: (𝑃𝑌 ) (note that, due to the movement of translation, only
2
the frontal projected area sustains the normal load), the following relation for the
abrasive contribution to friction, 𝜇𝑎𝑏𝑟 , is obtained:

𝐹𝑇 2 tan 𝜃
𝐹𝑇 = 𝐹𝑁 𝜇𝑎𝑏𝑟 (5) ∴ 𝜇𝑎𝑏𝑟 = = (6)
𝐹𝑁 𝜋

In general, the friction coefficient is then given by the sum of two contributions:

𝜇 = 𝜇𝑎𝑏𝑟 + 𝜇𝑎𝑑 (7)

Where the subscripts ad and abr indicate the contributions of adhesion and two-
body abrasion. Such contributions have clearly different weights, depending on the
specific characteristics of the tribological system.
(Abrasive wear of ductile material pp.96)

In general, microcutting and microploughing work together, then the wear rate can
be determinate by:

𝐹𝑁
𝑊 = 𝐾𝑎𝑏𝑟 (8)
𝐻

Where 𝐾𝑎𝑏𝑟 is the wear coefficient of abrasive wear, given by:

2 tan 𝜃
𝐾𝑎𝑏𝑟 = Φ (9)
𝜋

Where 𝜙 varies between 1 and 0. When Φ = 1, wear is ideally by microcutting


only. In this case 𝐾𝑎𝑏𝑟 = 𝜇𝑎𝑏𝑟 and the wear coefficient is very high. On the other
hand, when Φ = 0, wear is ideally by microploughing and 𝐾𝑎𝑏𝑟 = 0.

Back to solid particle erosion (16/03/17)


The wear volume, V, can be evaluated using eq. 8 which can be rewritten in the
2 tan 𝜃
following way (considering that 𝜋 = 𝜇𝑎𝑏𝑟 ):

𝐹𝑁 𝑠 𝐹𝑇 𝑠 𝐿
𝑉 = Φ𝜇𝑎𝑏𝑟 =Φ =Φ (10)
𝐻 𝐻 𝐻

Where 𝐿 is the work done by the hard particles on the abraded surface. In the case
of solid particle erosion, such a work is roughly given 1⁄2 𝑚𝑣 2 , where 𝑚 is the total
mass of erosive particles and 𝑣 their average impact velocity. Then the eq.10 is
obtained:

𝑚𝑣 2
𝑉 = Φ𝑆𝑃𝐸 (11)
2𝐻

Where Φ𝑆𝑃𝐸 is a constant that is representative of the efficiency of the erosive


process. It is also called wear coefficient for erosive wear, and varies between 0
and 1. It is theoretically 0 when the interaction is by microploughing and involves
just a displacement of material, and it is 1 when all the material deformed by the
particle is removed by microcutting. Therefore, the main factors that influence Φ𝑆𝑃𝐸
are:

1. The impact angle 𝜃. The erosive wear in ductile materials is expected to be


maximum at small impact angles, in correspondence of which the shear
stress exerted on the surface is maximum. Such a stress typically induces a
considerable plastic deformation, which is incremented by repeated
collisions up to the fulfilment of the critical conditions for fracture, either by
low cycle fatigue or by ratcheting.
2. Morphology of the particles. Angular particles are more effective than
rounded particles to produce a microcutting interaction.
3. The hardness of the target material. Microcutting is favoured in materials
with low ductility and high hardness.

To evaluate solid particle erosion the results are in general referred to the erosion
rate, E, given by the ratio between the mass of erodent material and the mass of
erosive particles striking the surface.

𝜌𝑣 2
𝐸 = Φ𝑆𝑃𝐸
2𝐻

Back to solid particle erosion (20/03/17)


Where 𝜌 is the density of the erodent material.

Velocity influence
From a large number of experimental investigations, it has been obtained that Φ𝑆𝑃𝐸
is quite low, and typically varies between 5𝑥10−3 𝑎𝑛𝑑 10−1 . Let us consider first the
role of impact velocity. If it is too low, deformation of the target material is elastic,
and wear may occur just by contact fatigue. If 𝑣 is high (generally speaking, larger
than 10 m/s), a very extensive and localized plastic deformation occurs, under a
quite large strain rate (in the range 104 − 107 𝑠 −1 ). Such a high strain rate
noticeably increases yield stress, and thus hardness, of the material, even if this
effect may be partially counterbalanced by the softening induced by local thermal
heating. Such a change in hardness modifies Φ𝑆𝑃𝐸 and also renders rather
unreliable the use of room temperature (static) hardness, H in equation 5.10. When
considering the effect of 𝑣 , 𝐸 is then expressed by 𝐸 = 𝐸𝑜 𝑣 𝑛 , where 𝐸𝑜 is a
constant. It has been experimentally obtained that 𝑛-values would range between 2
and 3 in metals and polymers with a ductile behaviour, and between 3 and 5 in
brittle materials.

Particle feed influence


Another parameter that may play a role and is not considered in Eq. 5.10, is
particle feed rate, which is mass of eroding particles per unit area and time. As the
feed rate increases, the erosive wear also increases since the probability of having
local repeated contacts (and therefore plastic strain accumulations) increases too.
But over a specific limit, the wear rate decreases since the arriving particles may
interfere with the rebounding ones. Such a threshold was found to be around
10,000 kg/m2s in metals, and just 1000 kg/m2s for elastomers.

Temperature influence (21/03/17)


Solid particle erosion is also influenced by ambient temperature. In general, the
erosion rate increases with increasing temperature, because of the corresponding
decrease in hardness. In hardened steels, however, erosion rate may decrease if
temperature rise is limited to 200–300 °C, since the hardness decrease induces an
overwhelming decrease in Φ𝑆𝑃𝐸 . In stainless steels and high temperature alloys,
such as Ni- or Co- alloys, the erosion rate increases continuously as temperature is
increased up to 800 °C or more. In oxygen-rich environments, the metal surface
may undergo direct oxidation, and erosion may thus induce a tribo-oxidative form
of wear that syn- ergically increases the overall wear rate.

Angle influence
In the figure 5.23 shows the variation of the erosion rate with the impact angle for
an Inconel 718 Ni-base superalloy tested at 538 ℃ using chromite particles at un
impact velocity of 305 m/s. The maximum wear rate is at about 20°, indicating the
ductile behaviour of the material. In order to reduce the high erosion rates a
ceramic CVD titanium carbide coating was deposited on the superalloy. The results
in figure 5.23 show that the improvement was really significant, with a reduction in
the erosion rate of nearly two orders of magnitude. In addition, it is shown that the
erosion rate increases with the impact angle and reaches a maximum for 𝜃 = 90°.
This behaviour is typical of brittle materials.

Fig. 5.23 Erosion rate as a


function of impact angle for
uncoated and TiC CVD
coated Inconel 718 (538 °C,
305 m/s)

To minimize the erosive wear it is first necessary to reduce the amount of erosive
particles (by filtration for example) and also reduce, if possible, their impact velocity
(which strictly depends on the speed of the fluid that carries the particles). If the
impact angle is low, hard materials should be preferred, whereas when it is close to
90°, high-strength ductile materials would perform better. If possible, changing the
equipment design in order to optimize the impact angle can provide a wear
reduction without changing the wear resistant material.
Erosive wear (Mechanical wear prediction and prevention. Bayer pp.9)
24/03/17

In erosion by solid particles, the particles entrained in the fluid stream typically
induce wear by abrasion. Under cavitation conditions, fatigue is likely mechanism.
Since in both cases the fluids involved may be chemically active, oxidative or
corrosive wear may be a part of the overall wear behaviour in those cases as well.
Similarly impact wear been described in terms of abrasion and fatigue type
mechanisms.

Вам также может понравиться