You are on page 1of 47

DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER

PARTICLES I-1

X. DEVELOP MENT AND IMPROVEMENT OF EROSION MODELS


FOR SAND AND OTHER PARTICLES

Introduction
Erosion of materials due to the impingement of solid particles is one form of wear
degradation that jeopardizes integrity of the flow boundaries and functionality of moving
components in particle-contained flows. The application includes but not limited to
production, process, and transportation facilities in petroleum, power plant and aerospace
industries and calculation of erosive damage is of great importance from both economical and
safety aspects. Sand production from oil and gas reservoirs may cause rapid erosion and wear
of production and process components and transportation lines. Predicting erosion caused by
the particles of varying sizes and shapes is of great importance and accurate erosion equations
are needed to predict erosion for a desired condition given particle impact velocity and angle.
The particles that may cause erosion are of various sizes, shapes and hardnesses, and
the effects of these parameters are not properly understood. The experimental results obtained
for erosion caused by sand particles and glass beads show that shape and size of the particle
have great influence on the erosion of materials not only as a sharpness or size factor but also
as an important parameter on the form of angle function, especially on the location of
maximum erosion obtained from the direct impingement tests at different impacting angles at
a given particle velocity. The influence of these parameters is characterized to some extent in
the previous work.
In this work, a new mechanistic erosion equation is developed based on the
experimental data from direct impingement testing that accounts for the particle and target
material properties by considering some of the studies in the literature. Particle velocity is
measured by PIV and model predictions are compared to multiple cases to validate the final
erosion equation. Experimental and numerical studies have shown that when particles are
transported by liquid (e.g. slurry transport in the pipe or pump) or dense gas, the particle
impact angles are very The impact angles in these cases are sometimes less than the smallest
value that can be obtained in a direct impingement erosion test in gas. For these cases, the
abrasion erosion equation should be used to calculate erosion. So, the mechanistic erosion
equation developed previously is extended to near zero impact angles for sharp particles. The
abrasion erosion equation is developed by introducing an initial penetration of the sharp
particle in the equation of the motion. This initial penetration may be attributed to the sharp
particle rotation and/or turbulent flow fluctuations of the carrier fluid near the wall.

November 2016
I-2 EROSION/CORROSION RESEARCH CENTER

Background
The parameters that erosion depends on may be classified into three categories, impact
condition (impact velocity and angle), erodent particle characteristics and target material
properties. The most important particle parameters are hardness, shape and size. The effects
of particle size and shape are now currently expressed as explicit sharpness and size functions,
but experimental data revealed that there are some inter-relations between some of these
parameters, and the angle function may depend on impact velocity and shape of the particle.
Properties of the eroding surface that are important in erosion are ductility, hardness and
density. Erosional behavior of ductile materials is different than brittle materials. Figure X-1
shows typical erosional behavior of ductile and brittle materials as a function of particle
impact angle. For ductile materials, erosion increases with the impact angle up to a maximum
point (approximately between 15-30 degrees) and then coasts down to a certain value at 90
degree impact, but for brittle materials, erosion increases with impact angle and the maximum
erosion is obtained at normal impact.

Ductile

Brittle
Erosion Ratio (ER)

30 60 90
Impact Angle, θ (deg)

Figure X-1. Typical erosion behavior of ductile and brittle materials versus impact angle

The effect of particle hardness is studied in a separate work supported by Chevron and
will be discussed briefly. The effect of particle size and shape is now currently expressed as
explicit sharpness and size function but experimental data revealed that there are some inter-
relations between some of these parameters. As we will see, impact speed and shape of the
particle change the angle function. Properties of the target material that are important in
erosion are ductility, hardness and density. Erosional behavior of ductile materials is different
than Brittle materials. Material hardness is another important parameter, and density of the

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-3

material is used when we need to convert removed mass to removed volume. So, the erosion
ratio (ER) which is the ratio of target mass loss to the mass of impinged particle is
Mass of Removed Material
𝐸𝑅 = = 𝑓(𝑉, 𝜃, 𝐻𝑣, 𝜌, 𝐷) (1)
Mass of Erodent
where V and θ are speed and angle of impact, Hv and ρ are target material hardness and
density, D and Fs are particle size and sharpness factor, respectively.
For a given geometry and flow properties, one can use his own particle tracking code
or commercial CFD software to estimate the particle impact speed and angle and put these
values in the erosion equation to find the erosion ratio. The penetration rate is the erosion ratio
multiplied by the impinged particle mass divided by the impacted area and material density.
The erosion prediction models such as CFD based erosion models and previous
versions of Sand Production Pipe Saver (SPPS) program which is developed at the
Erosion/Corrosion Research Center (E/CRC) relied on empirical erosion equations. These
equations do not account for the particle size and shape accurately and they have been
developed for each erodent particle and target material separately. There are many cases in
industry where the particle impact angle is very low (less than 5o). Figure X-2 shows CFD
simulation and particle tracking results for a 4” elbow where the fluid is water and the tracked
particles are 150 µm. The left figure is a contour plot of impact velocity and on the right is a
contour of average impact angle. It is observed that the maximum impact angle is less than
4o, considering a liquid viscosity of 1 cP. So, for a more viscous fluid the, particle impact
angle would be even less and a typical erosion equation may not be valid.

Avg. Impact Velocity (m/s) Avg. Impact Angle (deg)

Flow
Flow

Figure X-2. Average particle impact velocity and angle (VSL= 15 m/s, D = 4”)
There are also some other particles that may cause erosion such as scale products.
Experimental observations imply that the erosion damage caused by sand particles is more

November 2016
I-4 EROSION/CORROSION RESEARCH CENTER

than the erosion caused by scale products as sand particles are harder than the scale products.
But hardness is not the only dominant parameter of erosivity of a particle. So, more work has
been done to characterize the effect of other factors such as particle size and shape on erosion.

Goals and Approach


The main goals of this work are to predict erosion failure in production and
transportation facilities in oil and gas industry and their components due to the impingement
of different particles. Being able to predict erosion resulting from various particles, would
lead to lower costs of erosion inspection and maintenance and also, the risk of component
failure would be reduced by improving geometry and utilizing more appropriate materials.
A comprehensive approach to erosion modeling consists of flow modeling, particle
tracking and erosion equations. So, the first step is to model the flow either by computational
fluid dynamics (CFD) or use approximations for flow near the wall. Then, particles are tracked
as they move toward the wall, and the impact velocity and angle is estimated. The final step
is to apply the erosion equation for the estimated impact velocity and angle and calculate the
erosion ratio which is the ratio of target mass loss to the mass of erodent particle. The erosion
ratio equation plays a major role in this calculation and depends on many parameters including
but not limited to erodent particle characteristics, target material properties and speed and
angle of impact. Thus, in the present work, erosion models are investigated including erosion
resulting from sand and other particles.
The approach of this work is first to search the literature for erosion equations for solid
particles. Then, conduct erosion tests in a direct impingement configuration for different
materials and develop a mechanistic model for erosion prediction which accounts for the
speed and angle of impact, particle characteristics and target material properties. Particle
hardness will be also correlated to erosion based on the experimental data obtained in a
separate experimental facility.
The mechanistic erosion models have been implemented in latest version of SPPS
(Sand Production Pipe Saver program developed at E/CRC) and the effect of particle hardness
on erosion is included as well. The erosion equation is used in form of User-Defined Function
(UDF) in commercial CFD software (ANSYS Fluent) to predict erosion caused by different
particles in the oil and gas industry.

Literature Review
Solid particle erosion has been studied extensively in the literature for aerospace
industry applications and oil and gas production and slurry transport systems. In early studies,

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-5

most of the attention was paid to the material-related aspects of erosion [1]. The researchers
[2]–[9] proposed equations generally in the form of

𝐸𝑅 = 𝐾 𝑉 𝑛 𝑓(𝜃) (2)

where K is the erosion constant, V is the particle velocity and 𝑓(𝜃) is the impact angle
function. The erosion constant, K, velocity exponent, n, and the angle function have been
determined from experimental data or theoretical analysis of material behavior under particle
impacts.
The particle impact velocity and angle in the erosion equation are unknown and their
value depends on the environment surrounding the particle. Laitone [10], Chein et al. [11],
Clark [12] and Nguyen [13] proposed analytical quantification methods to estimate the
particle-wall collision information in particle-laden flows.
Assisted by the advancement of computational resources, Dosanjh [14] and Schuh et
al. [15] accounted for the influence of turbulence in predicting motion of the particle and used
CFD simulations in their studies, but comprehensive CFD simulations along with particle
tracking have been done in more recent studies by Edwards [16], Niu et al. [17], [18], Chen
[19] and Zhang [20].
Specific to oil and gas industry, there are some calculation guidelines and
methodologies proposed in the literature. American Petroleum Institute Recommended
Practice 14E (API RP 14E) [21] proposed a correlation for erosional velocity, Ve (in ft/s) for
gas-liquid mixtures as follows,
𝑐
𝑉𝑒 = (3)
√𝜌 𝑚

where c is an empirical constant and ρm is the gas/liquid mixture density in lb/ft3. The basis
for development API correlation is not clear, but it should not be applied to sand erosion
conditions as it does not account for many parameters in erosion calculation such as particle
and wall properties and geometry specifications. Salama and Venkatesh [22] and Salama [23]
proposed alternate correlations to API RP 14E in which they assumed that the velocity of
particles is similar to fluid velocity. Bourgoyne [24] developed another empirical correlation
and Svedeman and Arnold [25] calculated threshold velocities from Bourgoyne’s correlation.
These empirical correlations highly depend on the experimental conditions and most of them
do not account for the particle size, shape and fluid physical properties.
Shirazi, et al. [26], [27] and McLaury, et al. [28] presented a comprehensive
mechanistic model to predict erosion in elbows and tees in single and multiphase flows using
a stagnation length concept. The stagnation length was determined from experiments or CFD

November 2016
I-6 EROSION/CORROSION RESEARCH CENTER

simulations. The proposed method predicts a representative particle impact velocity to be used
in the erosion equation.
Zhang et al. [29] implemented an empirical erosion equation which had been obtained
from gas testing into a CFD code to predict the erosion ratio occurring on a flat specimen and
bend for air and water flows. Also, Wong et al. [30] utilized an empirical erosion equation
originally proposed by Chen et al. [19] to predict the erosion ratio in a pipe annular cavity via
CFD simulation. In addition, many other works are conducted to predict the erosion rate in
various geometries by coupling the CFD simulation and an erosion equation (Njobuenwu et
al. [31], Pereira et al. [32], Mansouri et al. [33]).
Mechanistic erosion equations that are available in the literature are developed based
on the calculation of the displaced volume by a single particle or energy dissipation during
particle impact. Finnie et al. [34] developed an erosion equation for ductile materials based
on the material cutting volume by a single particle. Bitter [35], [36] used an energy balance
and proposed that erosion is proportional to the part of the particle kinetic energy that is
absorbed by the target material and caused plastic deformation. Sheldon et al. [37] developed
an equation based on single particle indentations for spherical and angular particles for normal
impacts and at low velocity. Two erosion models developed by Hutching [38], [39] are based
on the deformed volume by spherical particles at normal incidence and cutting action of a
particle at oblique impacts. Sundararajan [40] proposed an erosion model by assuming that
deformation beyond the critical strain and energy dissipation of the particle, caused by friction
force between the particle and the eroding material, are responsible for the erosion at normal
and oblique impacts. Bingley et al. [41] implemented the equations of Hutching and
Sundararajan for nine heat treated steels, and Harsha et al. [42] used Hutching’s equations for
some ferrous and non-ferrous metals. In their work, the constants in the erosion equations
were calculated from experimental data, and a relation was found between the empirical
constants and the mechanical properties of the eroding material including hardness. However,
the effect of impact angle was not properly investigated. A relation was found by Levin et al.
[43] between the mechanical properties of some ductile alloys and the volumetric erosion
obtained from experiments, but the relation was developed for normal impact only. In a recent
study, a mechanistic erosion equation has been developed by Huang et al. [44]using
approximations for removed material by a spherical particle. They calculated the volume
removed by the vertical and tangential components of particle velocity separately. Generally
in these equations, many assumptions have been made to find a closed form solution of the
problem and find the relation between the properties of target materials and erosion. They
compared the results mostly to experimental data for pure materials such as iron, aluminum,
and copper, not alloy metals that are being used in industry.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-7

Feng et al. [45] empirically studied the dependency of erosion of ductile and brittle
materials on impact velocity and particle size for different erodent particles but did not report
an equation to be used for erosion prediction. Oka et al. [46], [47] developed empirical
correlations of erosion for many particles and materials, but his equation only has been
validated at impact velocities more than 50 m/s which are rarely applicable to the oil and gas
industry. Experiments in a slurry pot tester for two ductile materials and three erodent particles
by Desale et al. [48] implied that the material removal mechanism is a function of particle
shape and density, but no equation was proposed.
In E/CRC, Ahlert [49] proposed an empirical correlation for the erosion ratio and
McLaury [50] later used two term angle function to be used in that equation. Zhang [29]
changed the angle function to a polynomial form and Okita [51] used the angle function that
was originally proposed by Oka [46].
Although many authors have examined slurry erosion and abrasive erosion of
materials, very few have developed models that can be used to determine erosion caused by
abrasive wear. Hashish [52] conducted an experimental and theoretical study of an abrasive
jet and developed a model utilizing the theory originally proposed by Finnie [53]. Jain et al.
[54], [55] presented a stochastic methodology to evaluate the interaction between spherical
abrasive grains and a workpiece surface in abrasive flow machining. The models developed
in these works are for a specific geometry and cannot be applied to other configurations. Much
research [56]–[59] has been conducted in the literature to determine slurry erosion, but they
used empirical models based on experiments with gas to determine slurry erosion.

Experimental details
In order to develop erosion equations, direct impingement testing has been performed
on different materials to provide the experimental data base. Figure X-3 shows schematics of
the test apparatus. Pressurized air that is supplied to the nozzle takes in particles from sand
feeder at a constant flow rate. These particles are accelerated by the gas flow to impact the
target material and cause material loss of the coupon. These tests have been performed at
different impact velocities (9, 18 and 28 m/s) and angles (15, 30, 45, 60, 75 and 90 degrees)
to determine the speed and angle dependence of the erosion equation for each material. A
sample of the erosion testing experimental results is shown in Figure X-4. At each impact
angle, the mass loss of the specimen is measured at three intervals after blasting with a specific
amount of sand which is 300 grams in most of the cases considered here. More particles were
required to get measurable mass losses at low impact velocities. A linear trendline is fit
through these three points which are cumulative mass loss versus sand mass throughput. The
slope of this line is the steady-state erosion ratio which is dimensionless and plotted for all

November 2016
I-8 EROSION/CORROSION RESEARCH CENTER

impact angles on the right of Figure X-4. In the right hand side of Figure X-4, vertical axis
is the erosion ratio and corresponding impact angle is shown on the horizontal axis. Erosion
equation is the dashed line that passes through these points. Most of the erosion tests have
been repeated at the same condition to confirm the repeatability of the experiments.
Sand feeder

Compressor

Specimen holder
Flow meter

Nozzle Valve

Figure X-3. Schematics of experimental test facility

Figure X-4. Calculation of erosion ratio from experimental data

In these experiments, it is essential to determine the particle impact velocity which


depends on the gas velocity. The gas velocity has been measured by means of a Pitot tube and
manometer. The manometer measures the difference between the stagnation and static
pressure which is dynamic pressure.

stagnation pressure = static pressure + dynamic pressure (4)


or
1
𝑝𝑡 = 𝑝𝑠 + 𝜌𝑢2 (5)
2

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-9

where 𝑝𝑡 is the total or stagnation pressure, 𝑝𝑠 is the static pressure, 𝜌 is the fluid density and
u is the fluid velocity. Solving for the fluid velocity yields the following equation.
2 (𝑝𝑡 − 𝑝𝑠 )
𝑢=√ (6)
𝜌
Two of the most accurate methods of measuring particle velocity are laser Doppler
velocimetry (LDV) and particle image velocimetry (PIV). The LDV, also called laser Doppler
anemometry (LDA), is the technique of measuring the velocity of a moving object from the
Doppler shift of the light that it scatters. This technique can be used to measure the velocity
of fluid, particles or bubbles. In the case of fluid velocity measurement, seeding
particles/bubbles are required and it is assumed that they are moving at the same velocity with
the fluid. This non-invasive method is used extensively in the literature for the measurement
of turbulent flows, flows around solid objects and in other environments, but there are some
limitations in measuring the velocity near the wall or when the particle is not a good light
reflector.
In this work, the particle velocity is measured using PIV at different gas velocities at
the exit of the nozzle to correlate the particle velocity to the gas velocity. This correlation is
used to estimate the particle impact velocity in the erosion experiments. The PIV system (TSI
Inc.) utilizes a double-pulsed Nd:YAG laser, CCD camera, synchronizer and processor.
Figure X-5 shows a schematic of the PIV system, and Figure X-6 show the whole velocity
measurement system and details of the location of measurement box and the camera. The light
sheet is provided by a double-pulsed laser to illuminate the particles, and the camera is
synchronized with the laser pulses. The camera captures two consecutive images of the field.
The images are then transferred to the processor to correlate between the two images and track
particle movement and finally produce the measured flow/particle velocity field.
A sample of tracked particles and measured particle velocity distribution (for 150 µm
sand) at gas velocity of 65 m/s is shown in Figures X-7 and X-8. The circle size is proportional
to the particle size and the arrow shows the direction of movement and its size and color shows
velocity magnitude. The wide particle velocity distribution may be due to the particle size
distribution or turbulent flow fluctuations. The same procedure has been repeated for other
gas velocity and corresponding particle velocity has been measured. The calibration curve
shows the average particle velocity versus gas velocity (Figure X-9). In direct impingement
tests, there is always slippage between the entrained particles and gas. So, gas velocity is
measured by the Pitot tube and the corresponding particle velocity is extracted from the
calibration curve.

November 2016
I-10 EROSION/CORROSION RESEARCH CENTER

Figure X-5. Schematics of Particle Image Velocimeter

Figure X-6. Particle velocity measurement setup

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-11

Nozzle

Figure X-7. Tracked particle sample (for 150 µm sand) at gas velocity of 65 m/s

35%

30%
Number percent (%)

25%

20%

15%

10%

5%

0%
6.3 12.1 17.9 23.7 29.5 35.2 41.0 46.8 52.6 58.4
Particle velocity (m/s)

Figure X-8. Particle velocity distribution (for 150 µm sand) at gas velocity of 65 m/s

November 2016
I-12 EROSION/CORROSION RESEARCH CENTER

40

35
Sand 150 mic y = 0.31x
30
Particle Velocity (m/s)

Sand 300 mic y = 0.32x


25
Glass Beads 150 mic y = 0.31x
20

15

10

0
0 20 40 60 80 100 120
Gas Velocity (m/s)

Figure X-9. Velocity calibration curve for 150µm, 300µm sand and 150µm glass beads

Seven target materials were selected for testing including two carbon steels (1018 and
4130), stainless steel 2205, 13 chrome duplex, Inconel 625 and aluminum alloy 6061. Some
of these materials are very common in many industries including oil and gas, and their
mechanical properties are distributed over a wide range to demonstrate the capability of the
erosion equation. Table X-1 shows the properties of these materials. The hardness of these
materials is reported as the annealing Vickers hardness (for some materials it is converted
from Brinell hardness) because it will be shown that work or thermal hardening processes
have negligible effect on the erosion characteristics.

Table X-1. Target materials properties


Material Density (kg/m3) Hardness (VHN)
Carbon steel 1018 7870 131
Carbon steel 4130 7850 162
Stainless steel 316 8000 224
Stainless steel 2205 7820 305
13 chrome duplex 7720 268
Inconel 625 8440 252
Aluminum alloy 6061 2700 31

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-13

Erosion modeling
There are lots of studies in the literature conducted on erosion to develop an erosion
equation theoretically or empirically. The application of empirical correlations is limited to
the experimented materials with specific particles and impact conditions and theoretical
formulations may not be in agreement with experimental data as they have been developed
with many simplifying assumptions. The approach in this study is to combine the mechanistic
and empirical methods to come up with an erosion equation that can capture the erosion
mechanisms while providing agreement with experimental data.
In this work, 150 µm semi-rounded sand is used as the main erodent particle, but the
effect of particle size and shape is studied with other particles including 300 µm sharp sand
and 150 µm glass beads. Figure X-10 shows SEM micrographs of these particles.

150 µm semi-rounded sand 300 µm sharp sand 150 µm glass beads

Figure X-10. SEM micrographs of three erodent particles

Power [60] classified particles into six categories based on their roundness: very
angular, angular, sub-angular, sub-rounded, rounded and well rounded. Experimental results
at the Erosion/Corrosion Research Center (E/CRC) showed that well rounded particles (glass
beads) cause four to five times less erosion than very angular particles (sharp sand), and this
phenomenon has been observed by other researchers [48], [61]. So, angularity numbers have
been assigned to each category from 0.2 for well rounded up to 1.0 for very angular particles
and the other categories take numbers in between. Figure X-11 shows the particles and
angularity numbers assigned to each particle type. The angularity factor that has been assigned
to the particles in this study ranges from 0.5 to1.0. Angularity factors less than 0.5 is not
applicable to these particles because none of these particles are rounded or well-rounded as
observed in the SEM micrographs. So the possible error associated with this method will be
confined in the specified range.

November 2016
I-14 EROSION/CORROSION RESEARCH CENTER

(Sharp) (Semi-rounded) (Glass Beads)

Figure X-11. Erodent particle angularity [60]

Meng et al. [62] divided wear mechanisms into three categories: mechanical, chemical
and thermal action. Chemical reaction of the material is classified as corrosion, and removal
of the material due to melting occurs at relatively high impact velocities which are out the
scope of this context. The mechanical removal of the material may be due to two mechanisms:
cutting and deformation [35]. When a particle impact a surface at grazing angles, shearing
tension is applied to the material at the contact area and plastic deformation is caused when
kinetic energy of the particle is high enough. This process is repeated for consequent particle
impacts until a piece of material is removed from the surface. At normal impacts, the repeated
collisions of many particles cause plastic deformation on the surface if they exceed the elastic
limit and form platelets on the surface that may lead to material failure. The erosion scar and
SEM micrographs of eroded areas of the stainless steel 316 samples with 150 µm sand
particles with impact angles of 30o and 90o are shown in Figure X-12. The mechanism of
erosion varies from scouring erosion at grazing impact angles where long crates are formed
to platelet formation erosion at normal or near normal impact angles. So, the erosive damage
by solid particles is caused by two mechanisms namely cutting and deformation. Cutting wear
is the process of displacing a piece of material by particle which will be removed totally or
partially by consequent impacts. Deformation wear is removal of the material by repeated
impact of the particle in normal direction which result in material plastic deformation,
hardening, sub-surface cracking and finally a piece of material will break off. The total wear
is the summation of two terms, cutting plus deformation:

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-15

𝐸𝑅 = 𝐸𝑅𝐶 + 𝐸𝑅𝐷 (7)


The new model will consider both of the erosion mechanisms for ductile materials
based on these experimental observations. The volumetric loss of the material due to particle
can be estimated from the volume of the materials that is swept by single rigid particle as it
cut into a ductile surface. The forces that resist against motion of the particle are shown in
Figure X-13. The equations of motion of the particle in horizontal and vertical directions
originally developed by Finnie et al. [34] are as follows,

(a) (b)
Figure X-12. SEM micrographs of SS-316 surface eroded with 150 µm sand at two impact
angles: a) 30o and b) 90o

= =

= = ( )

Figure X-13. Force balance of the particle cutting into the surface

𝑑2𝑦
𝑚 +𝑃𝑛𝑅𝑦 = 0 (8)
𝑑𝑡 2

𝑑2𝑥 𝑃 𝑛 𝑅 𝑦
𝑚 2+ =0 (9)
𝑑𝑡 𝐾

November 2016
I-16 EROSION/CORROSION RESEARCH CENTER

in which m is mass of the particle, P is the flow pressure for annealed material which is
assumed to be the Vickers hardness of the material, n is the ratio of contact area to the removed
area and R is the particle size so that (according to the Hertz contact stress [63])

𝐴𝑦 = 𝑅 𝑦 (10)

It is assumed that contact area in x direction is a fraction of contact area in y direction


as particle is in arbitrary shape and K depends on shape of the particle and material
deformation behavior. In contrast to the constant value that is assigned by Finnie et al. [34],
the value of K will be determined from experimental data for each material. Integration of
Eqs. (8) and (9) and using initial velocity and location of the particle yields,

𝑈 𝑠𝑖𝑛(𝜃)
𝑦= 𝑠𝑖𝑛(𝛽𝑡) (11)
𝛽

𝑈 𝑠𝑖𝑛(𝜃)[𝑡𝛽 − 𝑠𝑖𝑛(𝛽𝑡)]
𝑥 = 𝑡𝑈 𝑐𝑜𝑠(𝜃) − (12)
𝐾𝛽

where
𝑛𝜋𝑃𝑅
𝛽=√ (13)
𝑚

The swept volume is

𝑚 𝑈 2 𝑠𝑖𝑛(𝜃)[2𝐾 𝑐𝑜𝑠(𝜃) − 𝑠𝑖𝑛(𝜃)]


𝑓𝑜𝑟 𝜃 ≤ 𝑡𝑎𝑛−1 𝐾
𝑉𝑜𝑙𝐶 = ∫ 𝐴𝑥 𝑑𝑥 = 2𝐾 2 𝑃 (14)
𝑚 𝑈 2 𝑐𝑜𝑠(𝜃)2 −1
𝑓𝑜𝑟 𝜃 ≥ 𝑡𝑎𝑛 𝐾
{ 2𝑃

If the impact angle is greater than 𝑡𝑎𝑛−1 𝐾, particle will have velocity component in x direction
when it leaves the surface and the first equation applies. Otherwise, the x component of
velocity will be zero earlier and the second equation is obtained.
Bitter [36] proposed the following equation for deformation erosion,

1 𝑚 (𝑈 𝑠𝑖𝑛 𝜃 − 𝑈𝑡𝑠ℎ )2
𝑉𝑜𝑙𝐷 = (15)
2 𝜀

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-17

in which U is the particle initial velocity, Utsh is the threshold velocity below which the
deformation erosion is negligible and ε is the deformation wear factor.
The final form of erosion equation with incorporated empirical factors is

𝑘𝑔 𝐶 𝑉𝑜𝑙𝐶 + 𝑉𝑜𝑙𝐷
𝐸𝑅 [ ] = 𝐹𝑠 𝜌 (16)
𝑘𝑔 𝑚

where Fs is the sharpness factor of the particle, ρ is the material density to convert volumetric
loss to mass loss and C is the cutting erosion coefficient that is multiplied by the calculated
displaced volume above as every impact is not as ideal as what is assumed here and multiple
particle impact is required to remove a piece of material. Experimental data implied that

𝑉𝑜𝑙𝐶 ∝ 𝑈 2 41 (17)

This may be due to the efficiency of cutting during impact or the effect of velocity on the
material resistant forces and velocity exponent in cutting erosion will updated according to
the experimental data.
Many studies in the literature [61], [64]–[66] showed that the effect of particle size on
erosion is not considerable for particles larger than 100 µm. This is observed in the derivation
of the cutting erosion equation (Eq. 14) which is not a function of particle size, although the
particle size, R, is considered in the equations of motion (Eqs. 8, 9). But in the deformation
erosion equation (Eq. 15) which is based on the kinetic energy of the particle, the value of
threshold velocity needs to be a function of particle size. The ratio of the kinetic energy of a
particle with average size of R2 to that of the reference particle with average size of R1 with
the same velocity is
𝐾𝐸2 𝑚2 𝑅2 3
= =( ) (18)
𝐾𝐸1 𝑚1 𝑅1
So, the ratio of the threshold velocity that the particle with size of R2 may cause deformation
erosion to the corresponding value for the reference particle (R1) is
(𝑈𝑡𝑠ℎ )2 𝐾𝐸1 𝑅1 3
=√ = √( ) (19)
(𝑈𝑡𝑠ℎ )1 𝐾𝐸2 𝑅2
Particle shape has an influence on two things: first, the dependency of erosion on the
impact angle and second, on particle erosion effectiveness. Rickerby and Macmillan [67]
derived an equation for the volume of the crater caused by a spherical particle, but their
equation needs to be solved numerically. The experimental study by Hutchings [68] on the

November 2016
I-18 EROSION/CORROSION RESEARCH CENTER

deformation caused by square plates did not result in an equation. Explanation of erosion
mechanisms of material removal by an angular particle is presented by Papini et al. [69] and
Dhar et al. [70]. According to Figure X-13 and Eqs. (8, 9), K is the ratio of the contact area
in the y-direction to the contact area in the x-direction. Simple geometrical relations for a
sharp particle represented by a square impacting the surface by one of its vertices and a
rounded particle represented by a circle imply that
𝐾𝑠𝑝ℎ𝑒𝑟𝑒
≈25 (20)
𝐾𝑠𝑞𝑢𝑎𝑟𝑒
So, the value of K (which is 0.4 for most of the materials eroded with sand) needs to
be multiplied by the factor obtained above for round particles, and this increase in the value
of K will change the angle dependency in Eq. (14). Particle sharpness changes the erosion
effectiveness through stress concentration and changes the contribution of ploughing and
microcutting contributions in erosion [61]. It is observed from the experiments that angular
particles can cause four to five times more erosion than spherical particles. So, the sharpness
factor (Fs) introduced above (Eq. 16) varies between 0.2 for fully rounded particles up to 1.0
for fully sharp particles.
The target materials introduced earlier have been tested at different impact velocities
and angles with 150 µm sand (for which the sharpness factor is 0.5) to find the empirical
constants in the erosion equation. The empirical constants are provided in Figure X-14 and
X-16. The cutting erosion constant, C is observed to be function of material annealing
hardness for these materials and it follows a trend for these metals. This factor correlates with
the square root of the target material hardness. So, the cutting erosion will be proportional to
the inverse square root of Vickers hardness. Similar trend has been observed in another study
in the literature [53]. Finnie et al. conducted erosion experiments on some pure materials as
well as some alloys with silicon carbide at 20 degree and impact velocity of 250 ft/s (76 m/s).
The results are shown in Figure X-15. The specimens in annealed form are represented by
open circle markers, work hardened represented by filled circle markers and square markers
are data points of thermally hardened materials. Vertical axis is the erosion resistance which
is defined as one over volumetric erosion in g/mm3 and horizontal axis is the Vickers hardness
of the target material. It is observed that erosion resistance correlate mainly with annealed
hardness of the specimen and work and thermal hardening processes have negligible effect.
For some pure materials, the erosion resistance is proportional to the annealing Vickers
hardness and for some others including iron and three alloy steels 1213, 1045 and tool steel,
it is proportional to the square root of annealing Vickers hardness.
For deformation erosion and its empirical constant, finding the correlation is more
difficult. In Figure X-9, the threshold velocity on the left axis and deformation wear factor

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-19

on the right axis are plotted versus target material hardness. Open markers represents
threshold velocity and filled markers are deformation wear factors for aluminum, two carbon
steels and four stainless steels. But an approximate correlation may be found for each group
of materials between the empirical constants and material hardness and we can estimate
erosion behavior of other steels with these correlations. The threshold velocity is decreasing
with the material hardness for all of the samples. The deformation wear factor increased with
materials hardness for aluminum and stainless steel but showed dissimilar behavior for carbon
steels.

0.018
0.016 2205
13 Cr
0.014 SS 316
Empirical Constant, C

0.012 1018 CS Inconel 625


0.010
4130 CS
0.008
0.006 1
∝ 𝐻𝑣 → 𝐸𝑅𝐶 ∝
0.004 𝐻𝑣
Al 6061
0.002
0.000
0 50 100 150 200 250 300 350
Vickers hardness (Hv)
Figure X-14. Cutting erosion empirical constants for tested materials

In order to verify if the developed erosion equations are valid at impact angles less
than 15o, additional testing has been done for erosion at low impact angle (5o) for stainless
steel 316 and aluminum alloy 6061 and the results are compared with experimental data in
Figures X-17 and X-18.

Figures X-19 to X-23 show experimental data for carbon steel 1018, carbon steel
4130, stainless steel 2205, 13 Cr duplex and Inconel 625 and corresponding values from the
erosion equation. Fair agreement is observed between the model predictions and experimental
values.

November 2016
I-20 EROSION/CORROSION RESEARCH CENTER

Figure X-15. Erosion resistance versus material hardness

Stainless steel (U) Carbon steel (U) Aluminum (U)


Stainless steel (ε) Carbon steel (ε) Aluminum (ε)
8 6.E+11

Deformation wear factor, ε (kg/m s2)


7
5.E+11
Threshold Velocity, U (m/s)

6
4.E+11
5

4 3.E+11

3
2.E+11
2
1.E+11
1

0 0.E+00
0 50 100 150 200 250 300 350
Vickers hardness (Hv)

Figure X-16. Deformation erosion empirical constants for tested materials

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-21
3.5E-05
U=28 m/s (Exp data)
U=18 m/s (Exp data)
3.0E-05
U=9 m/s (Exp data)
U=28 m/s (Model)
2.5E-05
U=18 m/s (Model)
Erosion Ratio (kg/kg)

U=9 m/s (Model)


2.0E-05

1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-17. Erosion ratio of stainless steel 316 at different impact velocities and angles

3.0E-05
U=28 m/s (Exp data)
U=18 m/s (Exp data)
2.5E-05 U=9 m/s (Exp data)
U=28 m/s (Model)
U=18 m/s (Model)
U=9 m/s (Model)
2.0E-05
Erosion Ratio (kg/kg)

1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-18. Erosion ratio of aluminum alloy 6061 at different impact velocities and angles

November 2016
I-22 EROSION/CORROSION RESEARCH CENTER

4.0E-05
U=28 m/s (Exp data)
3.5E-05 U=18 m/s (Exp data)
U=9 m/s (Exp data)
3.0E-05 U=28 m/s (Model)
U=18 m/s (Model)
Erosion Ratio (kg/kg)

2.5E-05 U=9 m/s (Model)

2.0E-05

1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-19. Erosion ratio of carbon steel 1018 at different impact velocities and angles

3.5E-05
U=28 m/s (Exp data)
U=18 m/s (Exp data)
3.0E-05
U=9 m/s (Exp data)
U=28 m/s (Model)
2.5E-05
Erosion Ratio (kg/kg)

U=18 m/s (Model)


U=9 m/s (Model)
2.0E-05

1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-20. Erosion ratio of carbon steel 4130 at different impact velocities and angles

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-23

3.0E-05
U=28 m/s (Exp data)

U=18 m/s (Exp data)


2.5E-05
Up=9 m/s (Exp data)

U=28 m/s (Model)


Erosion Ratio (kg/kg)

2.0E-05
U=18 m/s (Model)

U=9 m/s (Model)


1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-21. Erosion ratio of stainless steel 2205 at different impact velocities and angles

3.0E-05
U=28 m/s (Exp data)
U=18 m/s (Exp data)
2.5E-05
U=9 m/s (Exp data)
U=28 m/s (Model)
Erosion Ratio (kg/kg)

2.0E-05
U=18 m/s (Model)
U=9 m/s (Model)
1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-22. Erosion ratio of 13 Cr duplex at different impact velocities and angles

November 2016
I-24 EROSION/CORROSION RESEARCH CENTER

3.5E-05
U=28 m/s (Exp data)
3.0E-05 U=18 m/s (Exp data)
U=9 m/s (Exp data)

2.5E-05 U=28 m/s (Model)


Erosion Ratio (kg/kg)

U=18 m/s (Model)


U=9 m/s (Model)
2.0E-05

1.5E-05

1.0E-05

5.0E-06

0.0E+00
0 15 30 45 60 75 90
Impact Angle (Degrees)

Figure X-23. Erosion ratio of Inconel 625 at different impact velocities and angles

Figure X-24 shows the erosion scars on SS-316 specimens after tests with 150 µm
spherical glass beads, 150 µm semi-round sand and 300 µm sharp sand. The erosion ratio of
these three samples is 5.3 ×10-7, 1.0 ×10-5, 2.3 ×10-5 kg/kg. The erosion caused by glass beads
is about 20 times less than that observed for semi-round sand which indicates that the effect
of particle sharpness is significant at low impact angles. The erosion caused by sharp sand is
approximately twice the value for semi-round sand, and the difference can be addressed by
the sharpness factor for these two particles. As described in detail in [71], the sharpness factor
(Fs) is 1 and 0.5 for sharp and semi-round particles, respectively. The sharpness effect may
be attributed to the efficiency of cutting by the particle or stress concentration on the target
material.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-25

(a)

(b)

(c)

Figure X-24. Erosion scars on SS-316 created by a) 150 µm glass beads, b) 150 µm sand,
and c) 300 µm sand

In this erosion equation the angle dependency of erosion varies with velocity. At
impact velocity lower than threshold velocity the deformation erosion is negligible. By
increasing the particle impact velocity, the deformation erosion term become appreciable
especially at normal impact and change the angle dependence of the erosion ratio. Figure X-
25 shows normalized erosion ratio with respect to the maximum value for each impact
velocity. It is observed that the angle function is not the same at all impact velocities and
feature of the new model is that it captures angle function variation with impact velocity. This
phenomenon is not important at impact velocities relatively higher than the threshold velocity.
[46]
In another comparison, model predictions are compared to experimental data of
stainless steel 316 eroded with 300 μm sand particles at different velocities and impact angles
(Figure X-26). It is important to note that the model is developed with experimental data for
150 μm sand and only two parameters have been changed. Threshold velocity is decreased
from 5.8 to 2 m/s based on the concept that kinetic energy of a 300 μm particle is about 8
times that of a 150 μm particle with the same velocity and the sharpness factor is also
increased to 1 because 300 μm sand particles are very angular.

November 2016
I-26 EROSION/CORROSION RESEARCH CENTER

Figure X-25. Normalized ER of Inconel 625 at different impact velocities and angles

Figure X-26. Erosion ratio of stainless steel 316 at different impact velocities and angles

For the cases where the impact angle is very low (≈1o), the abrasive erosion equation
should be used to calculate erosion. Assuming the same condition for the particle motion in
the development of mechanistic erosion equation, forces in x and y (Figure X-27) direction
resist against the motion of particle with respect to the surface but for sharp particle there is
an initial penetration to the surface that is caused by sharp particle rotation or turbulent flow
fluctuations.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-27

𝜃
𝑈 𝐴𝑦
𝐾=
𝐴𝑥
𝑥
𝑦0
𝑛𝑅𝑦
𝑦 𝐹𝑥 = 𝑃 𝐴𝑥 = 𝑃
𝐾

𝐹𝑦 = 𝑃 𝐴𝑦 = 𝑃 𝑛 𝑅 𝑦

Figure X-27. Force balance of the particle cutting into the surface

The equations of the motion for the particle in low impact angle are the same as the equations
for high impact angles but an initial penetration is introduced in the initial condition

𝑈 𝑠𝑖𝑛(𝜃)
𝑦 = 𝑦0 cos(𝛽𝑡) + 𝑠𝑖𝑛(𝛽𝑡) (21)
𝛽

𝑦𝑜 𝑐𝑜𝑠(𝛽𝑡) 𝑈 𝑠𝑖𝑛(𝜃)[𝑡𝛽 − 𝑠𝑖𝑛(𝛽𝑡)]


𝑥 = 𝑡𝑈 𝑐𝑜𝑠(𝜃) − (22)
𝐾𝛽
The swept volume is

𝑈𝛽𝑦0 𝛽 2 𝑦0 2
𝑙𝑖𝑚 𝑉𝑜𝑙𝐶 = 𝑚 − (23)
𝜃→0 𝐾𝑃 2𝐾 2 𝑃

and the resulting abrasion erosion ratio equation will be

𝜌𝛽𝑦0 𝛽𝑦0
𝐸𝑅𝐴 = (𝑈 − ) (24)
𝐾𝑃 2𝐾

The initial indentation (y0) needs to be quantified based on experimental observations.


Figure X-28 shows a 3D scanning microscopic image of SS-316 after the 5o impact test with
150 µm sand. Individual indentations found on the surface of the specimen and the depth of
each crater is measured. The results of average depth of penetration versus the normal
component of impact velocities are summarized in Table X-1.

November 2016
I-28 EROSION/CORROSION RESEARCH CENTER

Figure X-28. 3D scanning microscopic image of SS-316 specimen after the test

Table X-1. Average particle depth of penetration of semi-rounded 150 µm particles


impacting SS-16

Impact vel. mag. Particle normal vel. Depth of penetration


(m/s) (m/s) (µm)
57 5 1.2
29 14 1.4
32 16 1.6
29 29 2.4
32 32 2.8

It can be observed that the average depth of penetration is proportional to the normal
component of impact velocity and not the impact velocity magnitude. These numbers are
obtained for 150 µm particles and may be scaled based on the momentum of the particle to
estimate the initial penetration for other particle sizes.
The final form of the erosion equation is

𝑈 2 41 𝑠𝑖𝑛(𝜃)[2𝐾 𝑐𝑜𝑠(𝜃) − 𝑠𝑖𝑛(𝜃)]


𝐶1 2
𝜃 < 𝑡𝑎𝑛−1 (𝐾)
𝐸𝑅𝐶 = 2𝐾
(25)
𝑈 2 41 𝑐𝑜𝑠 2 (𝜃)
𝐶1 𝜃 > 𝑡𝑎𝑛−1 (𝐾)
{ 2
𝐸𝑅𝐷 = 𝐶2 (𝑈 𝑠𝑖𝑛 𝜃 − 𝑈𝑡𝑠ℎ )2 (26)
𝐸𝑅 = 𝐹𝑆 (𝐸𝑅𝐶 + 𝐸𝑅𝐷 + 𝐸𝑅𝐴 ) (27)

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-29

where C1, C2, K and Utsh are empirical constants, FS is particle sharpness factor that is 1 for
sharp, 0.5 for semi-round and 0.25 for fully rounded particles. The empirical constants for 7
materials tested are provided in Table X-2.
Table X-2. Empirical constants for the erosion equation
Material C1 C2 K Utsh (m/s)
Carbon steel 1018 5.90E-08 4.25E-08 0.5 5.5
Carbon steel 4130 4.94E-08 3.02E-08 0.4 3.0
Stainless steel 316 4.58E-08 5.56E-08 0.4 5.8
Stainless steel 2205 3.92E-08 2.30E-08 0.4 2.3
13 chrome steel 4.11E-08 3.09E-08 0.5 5.1
Inconel 625 4.58E-08 4.22E-08 0.4 5.5
Aluminum alloy 6061 3.96E-08 3.38E-08 0.4 7.3

CFD simulation of submerged slurry jet


Models describing particle-laden flows can be roughly divided into two groups,
Eulerian-Lagrangian models and Eulerian-Eulerian models. In the Eulerian-Lagrangian
approach, the fluid phase is treated as a continuum by solving Navier-Stokes equations while
the dispersed phase is solved by tracking a large number of particles. The Eulerian-Eulerian
model is the most complex multiphase model in ANSYS FLUENT as it solves a set of
equations for each phase. This model treats different phases as interpenetrating continua by
introducing phasic volume fraction. It solves momentum and continuity equations for each
phase and tracks volume fractions which are continuous functions of space and time. A single
pressure field is shared for all phases and interphase drag coefficient is used for modelling
momentum transfer between phases. The manner in which this coupling is handled depends
upon the type of phases involved. For granular flows, the properties are obtained from
application of kinetic theory. An overview of these two approaches can be found in [72].
In this study, the geometry and mesh is created by ANSYS Design Modeler and
Meshing tool, respectively and ANSYS FLUENT 15.0 is used to solve the Time-Averaged
Navier-Stokes equations to obtain the fluid dynamics solution. Figure X-29-a shows the
computational domain and boundary conditions for the submerged impinging jet, and the
nozzle dimensions and stand-off distance are provided in Figure X-29-b. Although two
symmetry planes can be found in the geometry, the full body is modeled as the results will be
used for particle tracking in liquid. The Discrete Particle Model (DPM) performance is poor

November 2016
I-30 EROSION/CORROSION RESEARCH CENTER

in geometries with symmetry planes. Moreover, it is sensitive to the surface mesh on the wall
and will not work well if the surface mesh on the wall is not uniform. So, the mesh that is
created in this study is structured in the nozzle and unstructured for the rest of geometry.

ressure outlet
Inlet

o le

Impingement plate
(b)

(a)

Figure X-29. a) Computational domain and b) mesh configuration of submerged jet

ANSYS FLUENT calculates the trajectory of a discrete phase particle by integrating


the force balance on the particle, which is done in a Lagrangian frame. This force balance
equates the particle inertia with forces acting on the particles such as drag, pressure, buoyancy
and added mas or virtual forces. This can be written as [73]:
𝑑𝑢⃗𝑝 𝑔 𝜌𝑝 − 𝜌
= 𝐹𝐷 𝑢 ⃗ −𝑢⃗𝑝 + +𝐹 (28)
𝑑𝑡 𝜌𝑝
where 𝑢 ⃗ is the fluid velocity, 𝑢
⃗ 𝑝 is the particle velocity, 𝜌 is the fluid density, 𝜌𝑝 is the
density of particle, 𝐹𝐷 𝑢 ⃗ −𝑢⃗ 𝑝 is the drag force per unit particle mass and 𝐹 is an additional
acceleration term that accounts for virtual mass force and Saffman’s lift force.
A large number of particles are tracked in the field after obtaining a converged flow
solution. In this approach, the particles are modeled as a point mass that does not affect the
flow. The particle-particle interaction is also neglected in the standard formulation of DPM.
It is though possible to incorporate the effect of the discrete phase on the continuum and
achieve a two-way coupling between the phases. This is accomplished by alternately solving

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-31

the discrete and continuous phase equations until the solutions in both phases have stopped
changing.
The individual particle-particle interaction may be resolved through Discrete Element
Model (DEM), but it is very computationally expensive. Dense Discrete Phase Model
(DDPM) is a special form of DPM that can model the particle-particle interaction statistically
instead of resolving the individual interactions. It also incorporates the effect of particles to
the primary phase flow by modeling the discrete phase as granular phase which is sharing the
pressure with the primary Eulerian phase. The Eulerian model makes it possible to model
multiple separate but interacting phases. Simulations with almost any combination of solid,
liquid and gas phases can be performed where an Eulerian approach is used for each phase.
The particle-particle interaction is modelled through solid stresses acting on the particles in a
dense flow by an additional acceleration in the particle force balance. This addition results in
the following expression [73]

𝑑𝑢⃗𝑝 𝑔 𝜌𝑝 − 𝜌
= 𝐹𝐷 𝑢
⃗ −𝑢
⃗𝑝 + + 𝐹 + 𝐹𝑖𝑛𝑡𝑒𝑟𝑎𝑐𝑡𝑖𝑜𝑛 (29)
𝑑𝑡 𝜌𝑝

The additional term accounts for interparticle interaction and is computed from the
stress tensor given by the kinetic theory of granular flows as [73]:
1
𝐹𝑖𝑛𝑡𝑒𝑟𝑎𝑐𝑡𝑖𝑜𝑛 = − ∇𝜏̅𝑠 (30)
𝜌𝑝
Mansouri [74] has conducted submerged erosion experiments and measured the
erosion pattern with a 3D profilometer. The comparison of the CFD predictions and erosion
measurements for the two particle sizes are presented in Figs. 9 and 10. The value of integral
erosion ratio (in kg/kg) is also compared with experimental data with and without considering
abrasive erosion in Table X-3. The results in Figure X-30 are for semi-rounded particles
where abrasive erosion effects are less than those shown in Figure X-31 with 300 µm particles
where both impact and abrasive erosion effects are more severe.

November 2016
I-32 EROSION/CORROSION RESEARCH CENTER

-20

-40
 m)
Thickness Loss (

Exp data
-60 CFD w/o abrasive
CFD w/ abrasive

-80

-100

-120
-2 0 2 4 6 8 10 12 14 16 18
r (mm)

Figure X-30. Comparison of CFD prediction with experimental data for 150 µm sand
and c = 1% kg/kg

-20

-40

-60
 m)

-80
Thickness Loss (

Exp data
-100 CFD w/o abrasive
CFD w/ abrasive
-120

-140

-160

-180

-200
-2 0 2 4 6 8 10 12 14 16
r (mm)

Figure X-31. Comparison of CFD prediction with experimental data for 300 µm sand
and c = 1% kg/kg

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-33

Table X-3. Comparison of Integral Erosion Ratio (ER)


Particle size Particle conc. Experimental ER CFD w/o abrasive CFD w/ abrasive
(µm) (% kg/kg) (kg/kg) (kg/kg) (kg/kg)
150 1 2.94 ×10-6 1.15 ×10-6 1.78×10-6
300 1 4.35×10-6 2.61×10-6 3.92×10-6

Erosion study of particles other than sand


There are some cases where erosion is not caused by sand particles and other particles
with lower or higher hardness even with smaller sizes may cause erosion. The most important
particles are scale products such as calcite, magnetite, hematite and barite. Calcite is one of
the main scale products which is formed due to changes in conditions of the fluid in the
reservoir. Black powder which is mainly composed of magnetite is another scale product
which forms in the presence of water, oxygen, hydrogen sulfide and carbon dioxide from
corrosion of ferrous steel pipe, and hematite is another corrosion scale of ferrous steel pipe.
Barite is a scale product that forms when sea water which is rich in sulfate ions comes in
contact with the barium ion in the brine. Moreover, some of these particles are being used as
a densifying agent in particulate well kill fluids. So, the erosional behavior of particles other
than sand is studied in a separate work supported by Chevron TUCoRE.
In this study, two experimental apparatuses have been used, (1) particles entrained in
submerged liquid jet in a slurry tank and (2) gas testing with liquid droplets containing
particles. In the submerged apparatus, particles are suspended in the slurry tank by means of
a stirrer. As sketched in Figure X-32-a, the slurry mixture is deducted from the bottom of the
tank and pumped through a nozzle to impact the specimen. It should be noted that the particle
impact velocities are not the same as the liquid velocities as significant drag is expected as
particles interact with the flowing submerged jet impacting a target. This is similar to what
happens in the slurry mix flow through an elbow. The second experimental apparatus consists
of a slurry mixer container, stirrer and nozzle (Figure X-32-b). Pressurized air is supplied to
the nozzle to deduct the slurry mix from the container which leads to formation of droplets
containing particles. The flow condition in this experimental apparatus resembles gas-liquid
mixture flow in the pipe with low liquid loading where droplets are entrained in the gas core.
Water with density of 1000 kg/m3 and viscosity of 1 cP was used in these tests and the liquid
velocity of the submerged jet and gas velocity in the mist flow tests were kept constant at 16.8
and 45.7 m/s, respectively. The orientation angle between the nozzle and specimen (which is

November 2016
I-34 EROSION/CORROSION RESEARCH CENTER

SS-316) was 90o and the distance from nozzle to specimen was 0.5 in (12.7 mm) in both tests.
Table X-4 shows flow parameters and testing conditions of these loops. Particle concentration
in the slurry mix flowing through the nozzle was assumed to be consistent with the particle
concentration in the slurry tank as particles are so small (2 – 38 µm) that they will be easily
transported by the liquid. A stirrer was also used to keep homogeneity of the slurry mix in the
tank.

(a) (b)
Figure X-32. Experimental apparatus, a) submerged, b) air/water mist

Table X-4. Summary of testing conditions in two experimental apparatuses

Parameter submerged air/water mist


Jet velocity (m/s) 16.8 (liquid) 45.7 (gas)
Particle concentration (kg/kg) 1% 1%
Liquid flow rate (lit/s) 0.715 0.013

The small erodent particles that have been selected have a wide range of properties
(see Table 2). Iron powder is the softest particle with a hardness of 65 using the Vickers’s
scale, and silicon carbide is the hardest (3000 kgf mm-2). The average size varies from 2 µm
(for magnetite) to 38 µm (for barite), and the density range is from 2650 to 7860 for silica
flour and iron powder, respectively. Scanning electron microscope images of some of these
particles are shown in Figure X-33.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-35

Table X-5. Erodent particle properties


Particle Density (kg/m3) Average size (µm) Hardness (VHN)
Iron powder (Fe) 7860 32 65
Calcite (CaCO3) 2710 6 145
Barite (BaSO4) 4300 38 173
Hematite (Fe2O3) 5260 30 600
Magnetite (Fe3O4) 5170 2 680
Silica Flour (SiO2) 2650 24 1000
Alumina (Al2O3) 3950 20 2000
Silicon Carbide (SiC) 3210 20 3000
Density and size of particles affect their impact velocity as they move in the liquid
layers formed on the target wall by the jet or droplet impact. It is required to estimate the
impact velocities to find a correlation between the particle hardness and induced erosion. CFD
simulations and particle tracking were used to estimate the average impact velocity for each
case, but droplet velocity is not the same as gas velocity in mist flow testing and was measured
by Particle Image Velocimeter (PIV).

30 µm 30 µm 30 µm
Alumina Silicon carbide Iron powder

30 µm 30 µm 5 µm

Silica flour Barite Magnetite


Figure X-33. SEM images of erodent particles

Mass loss of the specimens after 72 hours eroded with eight types of particles in the
two experimental configurations are shown in Figure X-34. Squares are data points for
submerged tests, and diamonds represents mist flow data. Mass losses of the specimens in the

November 2016
I-36 EROSION/CORROSION RESEARCH CENTER

submerged test were higher than what was observed in the mist flow tests because the liquid
flow rate and particle impingement rate were higher, but the erosion ratio will be relatively
consistent in the two experiments for most of the particles as it is normalized by the amount
of particles throughput. It should be noted that particles that are impacting the target are
impacting at different impact velocities due to their densities and size as they slip through the
liquid protecting the target, and thereby their velocities are much smaller than the liquid jet
and gas velocities.
Table X-5 shows angularity numbers assigned to each particle based on the visual
observation under SEM (Figure X-33). The angularity factor that has been assigned to the
particles in this study ranges from 0.5 to 1.

Figure X-34. Erosion ratio of the SS-316 specimens for different particles

Figure X-35 shows surfaces of SS-316 specimens eroded by silica flour, iron powder,
barite and magnetite. The top image is in the center of impacting area and the lower
photograph shows the eroded surfaces at 4 mm from the center. At the center, particles impact
the specimen normally and platelets formed on the surface which may lead to surface failure
after repetitive impacts. At locations far from the this point, particles impact the specimen at
grazing angles and craters are formed on the surface in so called scouring erosion phenomena.
The SEM images did not show any fragments of particles embedded as the impact velocity of
these particles with the SS-316 material are fairly low as shown below. This was confirmed
by using Energy Dispersive X-ray Spectroscopy (EDX) in the locations that were suspicious
to have embedded particles after the test. It is noted that craters created by silica sand are

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-37

much deeper than others, and even tiny magnetite particles create craters that are similar to
larger silica sand.

Silica flour Iron powder Barite Magnetite

Figure X-35. SEM images of different locations on SS-316 specimens


eroded with different particles

It is well known that the major contributor to erosion is impact velocity of particles.
Thus, significant work was dedicated to estimate the representative impact velocities of these
particles. To estimate impact velocity of particles that were entrained in the liquid droplets,
the velocities of droplets were first measured by PIV. Then for both submerged and mist flow
tests, CFD simulations and particle tracking inside the liquid droplet or jet [15,16] were
conducted to estimate the impact velocity of particles that penetrate into the liquid layer
formed by droplets or continuous jet impact.
ANSYS Fluent [73] is used to study relative motion of the particles with respect to
liquid when it is spreading out on the wall. The simulation for submerged slurry jet impact
was done with turbulence Reynolds stress model and after obtaining the steady state flow
solution; particles were injected at the nozzle exit with modeling particles as a discrete phase
(DPM) and traced until they leave the simulation domain. Figures X-36-a and X-36-b shows
velocity contours and particle traces in the submerged configuration. For droplets containing
small particles the simulation is more complex. Transient multiphase flow with volume of
fluid (VOF) model was used to simulate a droplet moving with an initial velocity measured
by PIV toward the wall. The initial particle location, size and velocity was set through an
injection file in Fluent and particles traced until they leave the simulation domain. It is
assumed that the droplet and particles that are inside the droplets have the same initial velocity

November 2016
I-38 EROSION/CORROSION RESEARCH CENTER

and particles are distributed uniformly in the droplet. Figure X-36-c shows sequences of a
droplet with particle inside during impact. The simulation results for both cases have been
summarized in Table X-6.
Figure X-37 shows test results for submerged and droplet testing and their comparison
to Levy’s data in the literature that was gathered using an experimental facility, procedure and
velocities that were much different than the current results and were obtained for much larger
particles entrained in gas streams [75]. The vertical axis is erosion ratio divided by the normal
impact velocity squared to consider the effect of particle size and density that affect
deceleration of a particle when it enters the viscous layer near the wall and divided by the
angularity factor obtained from visual observation of the particles under SEM. Considering
that these data were gathered with three experimental facilities that were considerably
different with particle impact velocities ranging from 2 to 80 m/s, the normalized data appear
to line up well as a function of Vicker’s hardness of the particles.

Nozzle Nozzle
Exit
Particle
Traces

Specimen

a) b)
Droplet Entrained
Particles

c)
Figure X-36. CFD simulation and particle tracking results,
a) velocity contours and b) particle traces in submerged jet flow
and c) sequences of droplet impact with particles

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-39

Table X-6. Average particle impact velocity and angularity

Representative normal Avg. normal component


Particle Angularity component of impact vel. of impact vel. in mist
in submerged test (m/s) flow test (m/s)
Iron powder (Fe) 0.50 3.76 6.86
Calcite (CaCO3) 0.75 1.67 2.66
Barite (BaSO4) 0.70 3.45 6.17
Hematite (Fe2O3) 0.75 2.99 3.70
Magnetite (Fe3O4) 0.75 1.62 2.10
Silica Flour (SiO2) 1.00 2.75 3.10
Alumina (Al2O3) 1.00 2.11 3.33
Silicon Carbide (SiC) 1.00 1.97 2.89

Figure X-37. Correlation between normalized erosion and particle hardness

Summary and conclusions


A unified erosion equation has been developed based on some of the studies in the
literature to calculate the erosion of various metallic samples. The equation is composed of
two parts, cutting erosion that is related to cutting into the surface target material by striking
particles at grazing impact angles and deformation erosion which is caused by platelet

November 2016
I-40 EROSION/CORROSION RESEARCH CENTER

formation and surface failure as a result of multiple collisions of particles in normal direction.
The distinction of erosion mechanisms for normal and tilted angle impacts is supported by
SEM images of the sample surface as rough indented surface is observed at locations where
particle impact the surface normally and long craters found at locations where particles hit the
target at grazing angles. In order to develop the mechanistic erosion equation experimental
data have been collected for different materials using direct impingement testing in gas. The
tests were conducted originally at six impact angles (15, 30, 45, 60, 75 and 90 deg) and three
particle velocities (27.6, 18.4 and 9.2 m/s) with 150 μm sand. Particle velocity has been
measured with Particle Image Velocimetry (PIV) at different gas velocities. The model is
developed by assuming that erosion caused by particle impact is due to two mechanisms,
cutting and deformation. This hypothesis is supported by the SEM images of the surface of
the eroded specimens eroded at different angles. Rough indented surface is observed at normal
impact (90o) and long craters found at tilted impacts (30o). Empirical constants have been
obtained for seven target materials (Stainless Steel 2205, 13 Chrome Duplex, 1018 and 4130
Carbon Steel, Inconel 625, Stainless Steel 316 and 6061 Aluminum Alloy). In contrast to the
angle functions that are introduced in the literature for all particles and impact velocities, angle
dependence in the new model changes with the particle shape and velocity and showed fair
agreement with experimental data. The model accounts for the particle shape and size and has
been validated with experimental data from direct impingement testing. The particle velocity
in gas has been measured using Particle Image Velocimeter (PIV) and empirical constants
have incorporated into the final erosion equation based on the experimental data. It was
concluded from experimental data that velocity exponent may be increased from 2 that is
obtained from the formulation to 2.41. Fair agreement has been observed between the
experimental data and model predictions for various samples at different impact conditions.
It was also found that most of the empirical constants follow a trend and can be correlated to
the mechanical properties of the materials indicating that the equations are capturing the
physics of the problem. Also the effects of particle size on the deformation erosion threshold
velocity and particle sharpness on erosion were justified physically.
Experimental and numerical studies have shown that when particles are transported
by liquid (e.g. slurry transport in the pipe or pump) or dense gas, the particle impact angles
are very low. The impact angles in these cases are sometimes less than the smallest value that
can be obtained in a direct impingement erosion test in gas. For these cases, the abrasion
erosion equation should be used to calculate erosion. In this work, the mechanistic erosion
equation developed previously is extended to near zero impact angles for sharp particles. The
abrasive erosion equation is developed by introducing an initial penetration of the sharp
particle in the equation of the motion. This initial penetration may be attributed to the sharp

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-41

particle rotation and/or turbulent flow fluctuations of the carrier fluid near the wall. 3D
scanning microscope is used to find a correlation for initial penetration and the empirical
constants are obtained from submerged erosion experiments in liquid. The equation has been
implemented in a commercially available Computational Fluid Dynamics (CFD) code
(ANSYS FLUENT) to calculate erosion for a submerged impingement jet geometry, and the
result is compared with the experiment. It is shown that by neglecting the abrasive term in the
erosion equation, the specimen mass loss does not match the experimental measured mass
loss. While by including the abrasive term, the total mass loss of the specimen agrees well
with the experimental data. Moreover, the erosion pattern becomes closer to the
experimentally measured pattern especially farther from the center of the highly eroded area
where abrasive wear is expected to dominate.
Erosive behavior of very fine particles (iron powder, calcite, barite, hematite,
magnetite, silica flour, alumina and silicon carbide) entrained in liquid have been studied in
two experimental configurations, submerged and mist flow. Particle concentration was 1% by
mass and particles were scanned by SEM to characterize representative size distribution,
shape and angularity. PIV was used to measure the velocity of liquid droplets containing
particles, and CFD simulations were performed to estimate the impact velocity of particles
that are entrained by droplets in air/water mist flow or by the liquid jet spreading over the wall
in the submerged case.
Mass loss of a specimen is measured after 72 hours of nonstop testing and converted
to erosion ratio which is defined as the ratio of mass loss of the specimen to mass of erodent
throughput. The results indicate that the induced erosion (mass loss) of target specimen is not
only a function of particle hardness, but other parameters such as particle impact kinetic
energy and angularity also contribute. So, the erosion ratio values obtained from the two test
configurations were divided by the estimated particle impact velocity squared and particle
angularity to find the erodent hardness dependency of the erosion. The hardness effect found
to be remarkable when the hardness of particle is less than the hardness of target material and
erosion ratio found to not increase significantly when the particle hardness is relatively higher
than the material hardness and the particle keeps its integrity during impact.

Future Work

It is proposed to continue the development of abrasion erosion equation for other


particle sizes and shapes. 3D profilometry microscope would be used to extract the erosion
profile and advance the abrasion erosion equation. The unified equation will be applied to

November 2016
I-42 EROSION/CORROSION RESEARCH CENTER

other geometries and flow conditions such as high concentration cases and the results will be
compared with experimental data.

References
[1] J. A. . Humphrey, “Fundamentals of fluid motion in erosion by solid particle impact,”
Int. J. Heat Fluid Flow, vol. 11, no. 3, pp. 170–195, Sep. 1990.

[2] I. Finnie, “Erosion of surfaces by solid particles,” Wear, vol. 3, no. 2, pp. 87–103,
Mar. 1960.

[3] G. L. Sheldon and I. Finnie, “The Mechanism of Material Removal in the Erosive
Cutting of Brittle Materials,” J. Eng. Ind., vol. 88, no. 4, p. 393, 1966.

[4] J. E. Goodwin, W. Sage, and G. P. Tilly, “Study of erosion by solid particles,” Arch.
Proc. Inst. Mech. Eng. 1847-1982 (vols 1-196), vol. 184, no. 1969, pp. 279–292, Jun.
1969.

[5] W. J. Head and M. E. Harr, “The development of a model to predict the erosion of
materials by natural contaminants,” Wear, vol. 15, no. 1, pp. 1–46, Jan. 1970.

[6] G. L. Sheldon, “Similarities and Differences in the Erosion Behavior of Materials,” J.


Basic Eng., vol. 92, no. 3, p. 619, 1970.

[7] W. Grant, G., Tabakoff, An Experimental Investigation of the Erosive Characteristics


of 2024 Aluminum Alloy. 1973.

[8] J. H. Williams and E. K. Lau, “Solid particle erosion of graphite-epoxy composites,”


Wear, vol. 29, no. 2, pp. 219–230, Aug. 1974.

[9] G. Sundararajan and P. G. Shewmon, “A new model for the erosion of metals at
normal incidence,” Wear, vol. 84, no. 2, pp. 237–258, Jan. 1983.

[10] J. a. Laitone, “Aerodynamic effects in the erosion process,” Wear, vol. 56, pp. 239–
246, 1979.

[11] R. Chein and J. N. Chung, “Particle dynamics in a gas-particle flow over normal and
inclined plates,” Chem. Eng. Sci., vol. 43, no. 7, pp. 1621–1636, 1988.

[12] H. M. Clark, “The influence of the flow field in slurry erosion,” Wear, vol. 152, no. 2,
pp. 223–240, 1992.

[13] a. V. Nguyen and C. a. J. Fletcher, “Particle interaction with the wall surface in two-
phase gas–solid particle flow,” Int. J. Multiph. Flow, vol. 25, no. 1, pp. 139–154,
1999.

[14] S. Dosanjh and J. a. C. Humphrey, “The influence of turbulence on erosion by a


particle-laden fluid jet,” Wear, vol. 102, no. 4, pp. 309–330, 1985.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-43

[15] M. J. Schuh, C. A. Schuler, and J. A. C. Humphrey, “Numerical calculation of


particle-laden gas flows past tubes,” AIChE J., vol. 35, no. 3, pp. 466–480, Mar.
1989.

[16] J. Edwards, “Development, validation, and application of a three-dimensional, CFD-


based erosion prediction procedure,” University of Tulsa, 2000.

[17] Y.-Y. Niu, “Evaluation of erosion in a two-way coupled fluid-particle system,” Int. J.
Numer. Methods Fluids, vol. 36, no. 6, pp. 711–742, Jul. 2001.

[18] C.-S. Tsai, Yang-Yao Niu, “SIMULATION OF EROSION BY THE DILUTE


PARTICULATE FLOW IMPACT,” Numer. Heat Transf. Part A Appl., vol. 37, no. 2,
pp. 167–187, Feb. 2000.

[19] X. Chen, “Application of computational fluid dynamics (CFD) to singlephase and


multiphase flow simulation and erosion prediction,” The University of Tulsa, 2004.

[20] Y. Zhang, “Application and improvement of computational fluid dynamics(CFD) in


solid particle erosion modeling,” 2006.

[21] F. Edition, “Recommended Practice for Design and Installation of Offshore


Production Platform Piping Systems,” 1981.

[22] M. M. M. Salama and E. S. S. Venkatesh, “Evaluation of API RP 14E Erosional


Velocity Limitations for Offshore Gas Wells,” Offshore Technol. Conf., Apr. 1983.

[23] M. M. Salama, “An Alternative to API 14E Erosional Velocity Limits for Sand-
Laden Fluids,” J. Energy Resour. Technol., vol. 122, no. 2, p. 71, 2000.

[24] A. T. Bourgoyne, “Experimental Study of Erosion in Diverter Systems Due to Sand


Production,” in SPE/IADC Drilling Conference, 1989.

[25] S. J. Svedeman, “Criteria for Sizing Multiphase Flowlines for Erosive/Corrosive


Service,” SPE Prod. Facil., vol. 9, no. 1, pp. 74–80, Feb. 1994.

[26] S. A. Shirazi, B. S. McLaury, J. R. Shadley, and E. F. Rybicki, “Generalization of the


API RP 14E Guideline for Erosive Services,” J. Pet. Technol., vol. 47, no. 8, pp. 693–
698, Aug. 1995.

[27] S. A. Shirazi, J. R. Shadley, B. S. McLaury, and E. F. Rybicki, “A Procedure to


Predict Solid Particle Erosion in Elbows and Tees,” J. Press. Vessel Technol., vol.
117, no. 1, p. 45, Feb. 1995.

[28] B. . McLaury, S. A. Shirazi, J. R. Shadley, and E. F. Rybicki, “Parameters affecting


flow accelerated erosion and erosion-corrosion,” in CORROSION 95, 1995.

[29] Y. Zhang, E. P. Reuterfors, B. S. McLaury, S. a. Shirazi, and E. F. Rybicki,


“Comparison of computed and measured particle velocities and erosion in water and

November 2016
I-44 EROSION/CORROSION RESEARCH CENTER

air flows,” Wear, vol. 263, pp. 330–338, 2007.

[30] C. Y. Wong, C. Solnordal, A. Swallow, and J. Wu, “Experimental and computational


modelling of solid particle erosion in a pipe annular cavity,” Wear, vol. 303, no. 1–2,
pp. 109–129, Jun. 2013.

[31] D. O. Njobuenwu and M. Fairweather, “Modelling of pipe bend erosion by dilute


particle suspensions,” Comput. Chem. Eng., vol. 42, pp. 235–247, Jul. 2012.

[32] G. C. Pereira, F. J. de Souza, and D. A. de Moro Martins, “Numerical prediction of


the erosion due to particles in elbows,” Powder Technol., vol. 261, pp. 105–117, Jul.
2014.

[33] A. Mansouri, S. A. Shirazi, and B. S. McLaury, “Experimental and Numerical


Investigation of the Effect of Viscosity and Particle Size on the Erosion Damage
Caused by Solid Particles,” in ASME Fluids Engineering Division Summer
Meeting,Volume 1D, 2014.

[34] I. Finnie and D. H. McFadden, “On the velocity dependence of the erosion of ductile
metals by solid particles at low angles of incidence,” Wear, vol. 48, no. 1, pp. 181–
190, May 1978.

[35] J. G. A. Bitter, “A study of erosion phenomena : Part I,” Wear, vol. 6, no. 1, pp. 5–21,
Jan. 1963.

[36] J. G. A. Bitter, “A study of erosion phenomena: Part 2,” Wear, vol. 6, no. 3, pp. 169–
190, May 1963.

[37] G. L. Sheldon and A. Kanhere, “An investigation of impingement erosion using


single particles,” Wear, vol. 21, no. 1, pp. 195–209, Aug. 1972.

[38] I. M. Hutchings, “A model for the erosion of metals by spherical particles at normal
incidence,” Wear, vol. 70, no. 3, pp. 269–281, Aug. 1981.

[39] I. M. Hutchings, “Mechanisms of wear in powder technology: A review,” Powder


Technol., vol. 76, no. 1, pp. 3–13, Jul. 1993.

[40] P. S. Babu, B. Basu, and G. Sundararajan, “The influence of erodent hardness on the
erosion behavior of detonation sprayed WC-12Co coatings,” Wear, vol. 270, no. 11–
12, pp. 903–913, 2011.

[41] M. S. Bingley and D. J. O’Flynn, “Examination and comparison of various erosive


wear models,” Wear, vol. 258, no. 1–4, pp. 511–525, Jan. 2005.

[42] A. P. Harsha and D. K. Bhaskar, “Solid particle erosion behaviour of ferrous and non-
ferrous materials and correlation of erosion data with erosion models,” Mater. Des.,
vol. 29, no. 9, pp. 1745–1754, Oct. 2008.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-45

[43] B. F. Levin, K. S. Vecchio, J. N. DuPont, and A. R. Marder, “Modeling solid-particle


erosion of ductile alloys,” Metall. Mater. Trans. A, vol. 30, no. 7, pp. 1763–1774, Jul.
1999.

[44] C. Huang, S. Chiovelli, P. Minev, J. Luo, and K. Nandakumar, “A comprehensive


phenomenological model for erosion of materials in jet flow,” Powder Technol., vol.
187, no. 3, pp. 273–279, Nov. 2008.

[45] Z. Feng and A. Ball, “The erosion of four materials using seven erodents — towards
an understanding,” Wear, vol. 233–235, pp. 674–684, Dec. 1999.

[46] Y. I. Oka, K. Okamura, and T. Yoshida, “Practical estimation of erosion damage


caused by solid particle impact,” Wear, vol. 259, no. 1–6, pp. 95–101, Jul. 2005.

[47] Y. I. Oka and T. Yoshida, “Practical estimation of erosion damage caused by solid
particle impact,” Wear, vol. 259, no. 1–6, pp. 102–109, Jul. 2005.

[48] G. R. Desale, B. K. Gandhi, and S. C. Jain, “Effect of erodent properties on erosion


wear of ductile type materials,” Wear, vol. 261, no. 7–8, pp. 914–921, Oct. 2006.

[49] K. R. Ahlert, “Effects of particle impingement angle and surface wetting on solid
particle erosion of AISI 1018 Steel,” University of Tulsa, 1994.

[50] B. S. McLaury, “Predicting solid particle erosion resulting from turbulent fluctuations
in oilfield geometries,” University of Tulsa, 1996.

[51] R. Okita, “Effects of Viscosity and Particle Size on Erosion Measurements and
Predictions,” University of Tulsa, 2010.

[52] M. Hashish, “A Modeling Study of Metal Cutting With Abrasive Waterjets,” J. Eng.
Mater. Technol., vol. 106, no. 1, p. 88, 1984.

[53] I. Finnie, J. Wolak, and Y. Kabil, “Erosion of Metals by Solid Particles,” J. Mater.,
vol. 2, no. 3, pp. 682–700, 1967.

[54] V. K. Jain and S. G. Adsul, “Experimental investigations into abrasive flow


machining (AFM),” Int. J. Mach. Tools Manuf., vol. 40, no. 7, pp. 1003–1021, 2000.

[55] R. K. Jain and V. K. Jain, “Stochastic simulation of active grain density in abrasive
flow machining,” J. Mater. Process. Technol., vol. 152, no. 1, pp. 17–22, 2004.

[56] R. J. K. Wood, T. F. Jones, J. Ganeshalingam, and N. J. Miles, “Comparison of


predicted and experimental erosion estimates in slurry ducts,” Wear, vol. 256, no. 9,
pp. 937–947, 2004.

[57] A. Gnanavelu, N. Kapur, A. Neville, J. F. Flores, and N. Ghorbani, “A numerical


investigation of a geometry independent integrated method to predict erosion rates in
slurry erosion,” Wear, vol. 271, no. 5, pp. 712–719, 2011.

November 2016
I-46 EROSION/CORROSION RESEARCH CENTER

[58] S. N. Shah and S. Jain, “Coiled tubing erosion during hydraulic fracturing slurry
flow,” Wear, vol. 264, no. 3, pp. 279–290, 2008.

[59] K. Pagalthivarthi, P. Gupta, V. Tyagi, and M. Ravi, “CFD Prediction of Erosion Wear
in Centrifugal Slurry Pumps for Dilute Slurry Flows,” J. Comput. Multiph. Flows,
vol. 3, no. 4, pp. 225–246, Dec. 2011.

[60] M. C. Powers, “A New Roundness Scale for Sedimentary Particles,” SEPM J.


Sediment. Res., vol. Vol. 23, no. 2, 1953.

[61] S. Bahadur and R. Badruddin, “Erodent particle characterization and the effect of
particle size and shape on erosion,” Wear, vol. 138, no. 1–2, pp. 189–208, Jun. 1990.

[62] H. C. Meng and K. C. Ludema, “Wear models and predictive equations: their form
and content,” Wear, vol. 181–183, pp. 443–457, Mar. 1995.

[63] J. P. Andrews, “LVI. Theory of collision of spheres of soft metals,” London,


Edinburgh, Dublin Philos. Mag. J. Sci., vol. 9, no. 58, pp. 593–610, Apr. 1930.

[64] G. P. Tilly, “A two stage mechanism of ductile erosion,” Wear, vol. 23, no. 1, pp. 87–
96, Jan. 1973.

[65] A. Misra and I. Finnie, “On the size effect in abrasive and erosive wear,” Wear, vol.
65, no. 3, pp. 359–373, Jan. 1981.

[66] M. Liebhard and A. Levy, “The effect of erodent particle characteristics on the
erosion of metals,” Wear, vol. 151, no. 2, pp. 381–390, Dec. 1991.

[67] D. G. Rickerby and N. H. Macmillan, “On the oblique impact of a rigid sphere
against a rigid-plastic solid,” Int. J. Mech. Sci., vol. 22, no. 8, pp. 491–494, Jan. 1980.

[68] I. M. Hutchings, “Deformation of metal surfaces by the oblique impact of square


plates,” Int. J. Mech. Sci., vol. 19, no. 1, pp. 45–52, Jan. 1977.

[69] M. Papini and S. Dhar, “Experimental verification of a model of erosion due to the
impact of rigid single angular particles on fully plastic targets,” Int. J. Mech. Sci., vol.
48, no. 5, pp. 469–482, May 2006.

[70] S. Dhar, T. Krajac, D. Ciampini, and M. Papini, “Erosion mechanisms due to impact
of single angular particles,” Wear, vol. 258, no. 1–4, pp. 567–579, Jan. 2005.

[71] H. Arabnejad, A. Mansouri, S. A. Shirazi, and B. S. McLaury, “Development of


mechanistic erosion equation for solid particles,” Wear, vol. 332–333, pp. 1044–
1050, May 2015.

[72] B. G. M. van Wachem and A. E. Almstedt, “Methods for multiphase computational


fluid dynamics,” Chem. Eng. J., vol. 96, no. 1–3, pp. 81–98, 2003.

November 2016
DEVELOPMENT AND IMPROVEMENT OF EROSION MODELS FOR SAND AND OTHER
PARTICLES I-47

[73] ANSYS, “ANSYS FLUENT User’s Guide.” ANSYS, Inc., Southpointe, PA, 2015.

[74] A. Mansouri, “A combined CFD-experimental method for developing an erosion


equation for both gas-sand and liquid-sand flows,” The University of Tulsa, 2016.

[75] A. Levy, Solid particle erosion and erosion-corrosion of materials. ASM


International, 1995.

November 2016

You might also like