Вы находитесь на странице: 1из 9

Wear 302 (2013) 10261034

Contents lists available at SciVerse ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Comparison of some laboratory wear tests and eld wear in slurry pumps
C.I. Walker n, P. Robbie
Weir Minerals Australia Ltd., 1 Marden Street, Artarmon, NSW, Australia

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 23 August 2012
Received in revised form
19 November 2012
Accepted 22 November 2012
Available online 7 December 2012

A number of different laboratory wear tests have been undertaken to measure the wear resistance of a
natural rubber and a eutectic and hypereutectic white iron under abrasion and erosion conditions. Laboratory
work included two different slurry jet erosion tests, a Coriolis test and an ASTM dry sand rubber wheel test.
The laboratory results were compared with wear of the same materials in a centrifugal slurry pump
application in a mineral processing plant. The pump application has been monitored for over 2 years and over
40 parts run to destruction. Analysis of the wear data shows a factor of almost 3 difference in wear rate
between the rubber and the best white iron. Coefcient of variance of the data was in line with typical wear
results from the eld.
The laboratory wear tests were conducted with a silica sand slurry and average particle size range of 300
500 mm to match the eld conditions. The Coriolis and one of the jet erosion tests showed order of magnitude
similarity with the eld test results for the metals, but the other tests gave very different trends. The jet and
Coriolis erosion tests on the rubber showed a much lower wear rate than seen in the eld, while the DSRW
test found that the eutectic white iron wear rate was lower than that of the hypereutectic iron (all opposite of
the eld test).
Explanation for the different wear rates between the laboratory and eld tests was postulated to be nonrepresentative wear mechanisms. This is compounded by the lack of understanding of specic wear
conditions in the pump (local velocity, concentration, particle size, size distribution and particle shape) as
well as microstructure of the samples.
& 2012 Elsevier B.V. All rights reserved.

Keywords:
Wear
Slurry
Erosion
Pump

1. Introduction
The wear of centrifugal slurry pumps in mill circuit applications in mineral processing plant is generally quite severe. Typical
life of pumps is in the range 15004000 h and material wear
rates can be over 2 mm/day. Part section thickness in large pumps
of over 100 mm is not uncommon. A typical Warmans MC mill
circuit slurry pump application is shown in Fig. 1.
The mill circuit pump shown above has a number of key
internal parts subject to wear. These include the rotating impeller
that imparts energy to the uid, the casing liner, the frame liner
(or back liner) and the throatbush or inlet side-liner. The impeller
and throatbush orientation is shown in Fig. 2.
Walker [1] shows that the throatbush in a mill circuit slurry
pump often wears faster than the other components due to a
combination of the sharp particle shape, the coarse particle size
distribution and the high velocity recirculating ow. Given both
the aggressive nature of the mill circuit duty and the preferential
wear of the throatbush, it is this part that is used to compare the
3 different materials in the current study.

The type of wear that occurs on the throatbush is not well


understood. It is hypothesised to be a combination of 2 or 3 body
abrasion and mostly erosion. During operation with the above pump,
the throatbush is adjusted regularly (often weekly) up to touch with
the rotating impeller causing both metalmetal contact and wedging
of any particles in the gap. There is however only a relatively short
touching period and as both parts subsequently wear there is
increased ow in the gap and erosion wear. As the gap increases
there is less abrasion wear and more erosion wear. This cycle repeats
with 810 adjustments not uncommon over the life of the part.
The objective of the current research is to compare the wear life
seen in the eld application of the throatbush material with that of
similar material in laboratory simulated wear tests. The laboratory
tests included two tests in Australia (using the authors erosion jet
tester and a commercial dry sand rubber wheel (DSRW) abrasion test
at a technical institute) and two tests (slurry jet erosion (SJE) and
Coriolis erosion) undertaken at research establishments in Canada.

2. Field wear data


2.1. Wear rate measurement method

Corresponding author. Tel.: 61 2 99345100.


E-mail address: craig.walker@weirminerals.com (C.I. Walker).

0043-1648/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2012.11.053

In the current work, the wear rate in mm/day is used as the


base measure for comparison. This assumes a relatively constant

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

1027

Gouge

Baseline

Fig. 3. Part wear measurement using a template.

18

frequency of failure

16
14
12
10
8
6
4
2
0
Fig. 1. Warmans MC mill circuit (cyclone feed) slurry pump.

96

336

576

816
1056
age at failure (hrs)

1296

1536

Fig. 4. Wear life distribution at functional failure [2].

Impeller

Throatbush
(or inlet side-liner)

Fig. 2. Sectional view showing relationship between impeller and throatbush.

mean and standard deviation so as to meaningfully compare the


different material performance. In eld slurry applications (operating plants) this is generally far easier said than done, as many
variables are not controllable and equipment often starts and
stops to meet production or other equipment maintenance
requirements.
The current application is almost unique in the authors 35
years experience. With co-operative plant personnel, a very
consistent duty and multiple operating pumps on the same slurry
it has been possible to accumulate a signicant data set over a
2 year period.
2.2. Slurry particle size distribution

mass ow rate through the pump for all data points. Wear depth
on the throatbush was measured using a simple template as
shown in Fig. 3.
A point to note in Fig. 3 is that the wear is not uniform, but
rather there is a baseline wear which is the overall surface wear
and a gouging wear that occurs locally and is generally much
deeper. The gouging wear is used in the current work as in most
mill circuit pump applications it is this wear that is life limiting
for the part.
In examining data from different pump applications (including
some mill circuit), Walker [2] found signicant variability in
measured wear life, with the coefcient of varianceCOV (standard deviation/mean) on the order of 0.20.3. A typical wear life
distribution curve is shown in Fig. 4.
Given this sort of variability, it is essential that there is
sufcient data to be able to statistically determine a wear rate,

The slurry particle size analysis is shown in Fig. 5. As illustrated


in Walker [3], mill circuit applications have a d50 particle size
typically250500 mm and a d85 size of 6003000 mm. The current
application has a relatively coarse large fraction (d85 of 3000 mm),
but is otherwise reasonably typical with a d50 of 300 mm.
2.3. Velocity and contact conditions
As mentioned in the introduction, an exact determination of
the ow in the gap between the impeller and throatbush (and
which is the primary inuence on the wear of the throatbush) is
difcult to determine. Further, the gap is changing all the time at
a rate of up to 2 mm/day. To give some perspective to the likely
particle velocity seen in the gap a number of assumptions can be
made. Given that a typical gap for the application is around 6 mm,

1028

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

recirculation ow of around 5% of the net ow through the pump


and tangential velocity of half the impeller velocity (say13 m/s),
in the area around the inlet the slurry velocity could be on the
order of 16 m/s.
The other complicating factor for ow in the gap is determining a typical particle size distribution. Because of the strong
centrifugal eld due to the rotating impeller (at the impeller tip,
tangential velocity is 425 m/s), it is hypothesised that larger
particles entering the gap will be expelled and that only the
smaller particles would return to the eye of the impeller.
Typically, the drag force on particles smaller than 5070 mm
would be sufciently high relative to the centrifugal force to
ensure their return to the impeller eye, but for any larger particles
their presence at the impeller eye is somewhat moot.

35% Cr hypereutectic (650 HBW) white iron to ISO21988/JW/


HBW600XCr35. The rubber part was compression moulded as a
monolithic part over a mild steel reinforcing, while the metal
parts were inserts in the rubber that extended to about 50% of the
diameter of the rubber part (as illustrated in Fig. 6).
Table 1 shows the wear rates for the 3 different materials. The
results show that the wear rate for the 27% Cr white iron was 30%
less than that of the rubber, while that of 35% Cr white iron was
some 60% less. The variability (COV) for the rubber was about
double that for the metal parts.
The very good performance of the 35% Cr iron is a function of
the microstructure and increased volume of hard carbide phase
relative to the 27% Cr iron. The 35% Cr is part of a unique family of
hypereutectic white irons (tradename Hyperchromes) that are
inoculated to rene the microstructure during casting [5].

2.4. Materials tested


3. Laboratory wear tests
The throatbush conguration is shown in Fig. 6. The 3
materials tested were a carbon black lled natural rubber of
50 Duro hardness, a heat treated 27% Cr eutectic (600HBW)
white iron to ISO21988/JN/HBW555XCr27 [4] and a nominal

120

Cum. % Passing

100

Given hypothesis that the throatbush wear is a mix of abrasion


and erosion, the laboratory wear test results compared were a
similar mix. Two different jet testers in different laboratories
were used for erosion and a dry sand rubber wheel test (DSRW)
for abrasion. Previously published results from a Coriolis test [7,8]
for similar materials were also compared.
3.1. Jet tester (eductor)

80
60
40
20
0
1

10

1000
100
Size (microns)

100000

10000

Fig. 5. Slurry particle size distribution from mill circuit pump.

template

metal

The jet test (eductor) is described by Walker and Bodkin [6]


and consists of municipal (tap) water with solid particles
entrained being blasted onto the surface of the sample. The
sample holder can rotate to give a variety of impingement angles
between 151 and 901. The water ow rate can be metered and the
resultant slurry jet velocity can be varied between 10 and 20 m/s.
Both the solid particle sizing and the liquid velocity determine the
concentration of the slurry in the nozzle. Particles are passed
through the rig only once to ensure no degradation of shape. The
test conditions used which most closely represent the slurry
conditions of the eld trial are shown in Table 2.
Wear results from the original work [6] are shown in Table 3.
As can be seen the wear rate for the rubber is less than that of the
metal in all cases. The wear rate of the 35% Cr iron is greater than
that of the 27% Cr for the sharp particles, but less for the rounded
particles.

worn
material
rubber

Fig. 6. Throatbush showing typical wear pattern with 40 mm of material


removed.

Table 2
Jet test (eductor) test conditions.
Impingement angle (deg.)
Jet velocity (m/s)
Duration (min)
Sand ow (g/min)
Nozzle dia. (mm)
Stand-off distance (mm)
Sharp particle (mm)
Round particle (mm)

30
20
10
1400
4.5
20
Alumina 300600
Silica sand 250500

Table 1
Comparative wear rates of different throatbush materials in mill circuit application.
Material description

Mean gouging wear rate


(mm/day)

Coefcient of
variance

No. of data points

Wear rate relative


to 27% Cr

50 Duro natural rubber


Composite mix of natural rubber and 27% Cr white iron insert
Composite mix of natural rubber and 35% Cr white iron insert

2.26
1.56
0.888

0.660
0.306
0.347

25
9
10

1.45
1.00
0.57

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

Table 3
Comparative wear rates of different materials [6].
Wear rate
Material
Sharp particle Round particle Wear rate
wear rate
relative to 27% relative to
description wear rate
27% Cr
(mm3/min)
Cr (sharp)
(mm3/min)
(round)
3.53
40 Duro
natural
rubber
6.05
27% Cr
white
iron
7.44
35% Cr
white
iron, heat
treated

0.605

0.58

0.31

1.95

1.0

1.0

1029

steel) reference. Additional variants of each composition were


testedwith and without heat treatment and with microstructural
renement for the 35% Cr material. Results are shown in Table 5.
As can be seen the 35% Cr material that performed so well in
the eld was the same as or worse in this test than the 27% Cr
material. Heat treatment (and higher hardness) seemed to
improve wear resistance, while the ner carbide 35% Cr iron
performed better than the coarser carbide material (when both
were not heat treated).
3.3. Slurry jet erosion (SJE)

1.49

1.2

0.76

Table 4
DSRW test conditions.
Test load (N)
Wheel speed (r/min)
Surface speed (m/s)
Duration (min)
Sand ow (g/min)
Particle (mm)
Rubber wheel

130
200
2.3
30
465
Silica sand 150300
223 mm, 60 Shore A

The SJE is a recirculating slurry loop with an air operated


diaphragm pump that is controlled to ensure constant velocity of
ow through the nozzle onto the samples. Test conditions are
shown in Table 6 below.
Sample volume loss is measured indirectly using a mass/loss/
density calculation or directly using a non-contacting laser
prolometer. A photo of the rig and typical sample prole are
shown in Figs. 8 and 9, respectively.
Results for the SJE test are shown in Table 7. Although very
similar test rigs, the wear rates for the SJE are much lower than
those for the jet tester. Overall, the 35% Cr materials did not wear
as well as the 27% Cr material. Wear rates for the rubber were
very low compared to the white irons.
3.4. Coriolis test
The Coriolis test results are taken from the work of Llewellyn
et al. [7] and Jones [8]. In the Coriolis test, sand (2.2 kg) is preweighed and dispensed continuously into a plastic funnel where
it is mixed with metered municipal water to produce a 10 wt%
slurry that ows at a constant rate of 60 ml/s through a feed tube
containing a controlling constriction onto and across the sample.
The rotor spins at 5000 rpm. The slurry is accelerated by centrifugal force at 17 mm distance from the rotor centre, through
two radial 6.3 mm high and 1 mm wide channels in each arm of
the rotor, until it contacts exposed specimen surfaces at a
distance of 39.5 mm.The slurry subsequently exits at the end of
the specimen at a radial distance of 68.5 mm. At least two tests
were carried out on both faces of pairs of samples for each
material. Each surface was exposed to 1.1 kg of material.
The erosive particles used are AFS 50/70 silica sand grains with
a size range of 212300 mm (similar to the SJE test) and a SiC
Table 5
DSRW laboratory wear test results on white irons.

Fig. 7. DSRW tester.

3.2. Dry sand rubber wheel (DSRW)


The two metal samples were also tested using an ASTM G65
DSRW test under the conditions outlined in Table 4. A photo of
the rig is shown in Fig. 7.
Each sample was tested twice, and the 35% Cr samples 4 times.
Sample mass loss was measured relative to a K110 (heat treated

Test material

Mass loss relative


to K110 steel

Wear rate relative


to 27% Cr

27%
35%
35%
35%

0.94
0.89
1.61
0.91

1.00
0.95
1.71
0.97

Cr
Cr
Cr
Cr

(heat treated)
(heat treated)
coarse carbide (no H/T)
ne carbide (no H/T)

Table 6
SJE test conditions.
Impingement angle (deg.)
Jet velocity (m/s)
Duration (min)
Sand concentration (w/w %)
Particle detail (mm)
Nozzle dia. (mm)
Stand-off distance (mm)

20
16
120
10
Silica sand 212300
5
100

1030

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

angular particle with a size range of 205365 mm. Test conditions


are summarised in Table 8 and a schematic of the test rig is
shown in Fig. 10.
Volume loss determinations for the wear scars were carried
out with a non-contacting optical 3D surface imaging system.
The vertical resolution of this instrument is 3 nm, and the spatial

resolution is 26 mm for all acquired maps. A countor representation of a wear scar with X and Y line proles is illustrated in
Fig. 11.
The wear test results are shown in Table 9. In all cases the
wear rates of the Hyperchromes materials were less than that of
the 27% Cr iron, wear rate decreasing with increasing carbide
volume. The impact of carbide size was very apparent, with very
ne carbide structure giving wear rates some 3 times lower than
those of similar composition materials with coarser structure,
regardless of heat treatment. The sharper SiC particles had less
relative effect on 35% Cr, showing some 20% lower wear rate than
that of the 27% Cr iron.
For the rubber tests, the wear with the rounded sand particles
was negligible. For the sharper particles the wear of the rubber
was less than that of 27% Cr, with the softer 40 Duro material
performing better than the 60 Duro material.

Table 7
SJE laboratory wear test results on white irons.

Fig. 8. Slurry jet erosion tester.

Test material

Wear rate
(mm3/min)

Wear rate relative


to 27% Cr

40 Duro natural rubber


50 Duro natural rubber
27% Cr (heat treated)
35% Cr coarse carbide
35% Cr ne carbide

0.0017
0.00275
0.0561
0.0813
0.105

0.03
0.05
1.00
1.86
1.45

Fig. 9. Typical SJE white iron worn sample.

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

4. Discussion of results
The relative wear rates for all the materials on different testers
are shown in Table 10. In general the DSRW and the SJE test have
Table 8
Coriolis test conditions [7,8].
Impingement angle (deg.)
Slurry velocityapprox. (m/s)
Duration (min)
Solidsconcentration (w/w %)
Round particle detail (mm)
Sharp particle detail (mm)
Wear scar width (mm)
Stand-off distance (mm)

08
1424
6
10
Silica sand 212300
Silicon carbide 205365
1
40

1031

the 35% Cr iron with a higher wear rate than the 27% Cr iron
which is the opposite of what was seen in the eld test. The SJE,
jet eductor and Coriolis test have the rubber performing better
than the 27% Cr iron which again is the opposite of what was seen
in the eld test. The Coriolis test and the jet test have the 35% Cr
iron with a similar lower wear rate compared to the eld test;
however predictive accuracy was relatively poor.
4.1. Comparison of test conditions
Mean particle sizes are similar for all laboratory tests
(150500 mm) and also similar to the mean particle size of the eld
test (300 m). The major difference is with particle size distribution,
where the eld slurry had a much broader particle distribution (and
much larger d85 particle size) than the laboratory tests.
Whilst particle shape was not measured, it is obviously a
major factor in the different results seen both between the
laboratory tests and the laboratary and eld tests. In the eld
test, the particles are angular, freshly crushed from the grinding
and milling process. In the laboratory tests the SiC and Al2O3 sand
particles were also relatively sharp while for the silica sand the
particles are considerably more rounded. In the SJE test the
particles would have been most rounded due to attrition as the
slurry is recirculated over 120 min test period.
The effect of particle shape and size has a marked impact on the
relative rubber wear as seen with the jet and Coriolis laboratory
tests. The rounded particles exhibited negligible wear, while the
sharp particles of similar size had a much greater wear rate.
Extrapolating this effect might go some way to explain the much
lower rubber wear rates seen in the laboratory wear tests relative to
the eld test. The quantity of larger particles seen in the eld test
would certainly contribute to a higher relative wear rates for the
rubber also.
Particle hardness would have been similar for the siliceous ore
of the eld test and the sand laboratory tests, but the sharp SiC
Table 9
Coriolis laboratory wear test results on white irons.

Fig. 10. Coriolis wear tester arrangement [7].

Test material

Wear rate relative Wear rate relative


to 27% Cr (sharp) to 27% Cr (round)

40 Shore A natural rubber


60 Shore A natural rubber
27% Cr (heat treated)
27% Crvery ne carbide (H/T)
35% Cr coarse carbide (heat treated)
35% Cr coarse carbide (no H/T)
35% Crvery ne carbide (no H/T)

0.22
0.86
1.00

0.54

Fig. 11. Coriolis test typical wear scar prole [7].

1.00
0.3
0.64
0.43
0.11

1032

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

Table 10
Relative wear rates of different materialscompared to 27% Cr white iron for both silica sand (round) and SiC (sharp) 150500 mm particles.
Material description
40 Duro natural rubber
50 Duro natural rubber
60 Duro natural rubber
35% Cr coarse carbide (heat treated)
35% Cr coarse carbide (no H/T)
35% Cr ne carbide (no H/T)

Field test

DSRW test

Coriolis test (round)

0.03
0.05

1.45

0.57

SJE test

0.95
1.71
0.97

1.86
1.45

0.64
0.43
0.11

Coriolis test (sharp)

Jet test (round)

Jet test (sharp)

0.22

0.31

0.58

0.86
0.54

0.76

1.2

Fig. 12. Effect of chill cast microstructure on wear of Hyperchromes [7].

and Al2O3 particles are obviously much harder (42000 HV) and
should have worn the white iron carbides (14001900 HV)
relatively more than that of the silica sand (1000 HV).
Relative slurry velocities in the laboratory erosion tests (1020
m/s) were similar to those of the eld test (16 m/s estimated) while
the DSRW test rubbing speed was considerably lower at 2.3 m/s.
While the slurry concentrations of the laboratory tests (10 wt%)
were lower than the eld condition (55 wt%) it is thought not to be a
signicant inuence on the relative performance of various materials.
4.2. Effect of microstructure on white iron results
It is well known that the ner the microstructure, the smaller
the inter-carbide distance, and the more effective the carbide
volume can be in preventing wear [7]. In all of the tests undertaken and reviewed here (Coriolis, jet, SJE and DSRW) the ner
carbide material performed better.
As mentioned previously, Llewellyn et al. [7] found that the
chill cast surface wear in the Coriolis test had mean wear rate
values some 34  greater than those for coarser structures of the
same casting composition (see Fig. 12).
4.3. Wear mechanisms
The worn 35% Cr iron throatbush surface from the eld trial is
shown magnied in Fig. 13. The wear mechanisms seem to be a
combination of ploughing and fracturing/spalling of the carbides.
In some areas it looks like the carbides have been plucked from
the matrix. The plough marks in the surface of the carbides are
mostly in the same direction and could potentially be a result of
the 3 body abrasion (particles jammed between rotating impeller
and throatbush) or directional erosion due to some of the larger
particles. The matrix has been worn at a large rate, leaving the
carbides standing proud of the surface.
The worn surface morphology of the DSRW and jet eductor
test samples are quite different from the eld test surface as can

Fig. 13. Photo of worn surface of 35% Cr iron following eld test.

be seen in Fig. 14. In both laboratory test cases the carbides and
matrix of the material seem to be worn at the same rate,
indicating greater particle impact forces (i.e. higher hardness of
the carbides has not helped). A further indicator of the high
stresses seen in the laboratory test wear is that the carbide
trailing edges are cracked at right angles to the wear direction.
Wear direction is that of the rubber wheel rotation in the DSRW
and the particle direction in the jet eductor test.
In the jet eductor test (Fig. 14b) signicant ploughing is evident
along with carbide spalling and breakout from the matrix surface.
Fig. 15 shows the worn 35% Cr white iron sample from the
Coriolis test work of Jones [8]. As can be seen in Fig. 15(a) for the
rounded particle wear, the bulk of the material removal occurs in
the matrix due to micro-ploughing while the carbides have edge
rounding and stand out from the matrix. In Fig. 15(b) with the
hard and sharp SiC particles, the wear is considerably different,
with ploughing and micro-machining occurring uniformly across
the surface, regardless of the carbide structure.
There is no carbide fracturing evident with the Coriolis test
like that seen from the eld sample (viz. Fig. 13) and this may be
due to the much smaller top size of the particles in the laboratory
test (300 mm) compared to the eld test (10,000 mm).

4.4. Predictive ability of different tests


Llewellyn et al. [7] believe that the Coriolis test provides
discriminative power that is almost two orders of magnitude
greater than those of traditional slurry jet tests on white iron
materials. In jet tests, the damage is mainly caused by high
normal impacts; thus, the wear rate mainly reects the materials
resistance to plastic deformation and carbide cracking. In Coriolis

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

1033

Fig. 14. Photo of worn surface of 35% Cr iron heat treated. (a) DSRW test and (b) Jet eductor test.

Fig. 15. Photo of worn surface of 35% Cr white iron following Coriolis test [8]. (a) Rounded sand particles and (b) Sharp SiC particles.

tests, the normal impacts are low; therefore, elasticity and other
material properties also play major roles in the extent of damage.
The challenge in using the jet test and the Coriolis test is to
ensure that the particle size, shape and velocity are representative, so that ensuring the wear mechanisms with both white iron
and rubber is similar to what is happening in the eld. This seems
to be particularly the case with rubber, where exceeding a
threshold particle shape and size can radically alter the wear rate
as contact conditions change from purely elastic collisions to
cutting or tearing.
The DSRW abrasion test does not appear to be that helpful in
predicting predominantly erosion wear in a throatbush.
The SJE recirculating slurry jet test is problematic due to the
attrition of the particles over the test period. This results in much
lower wear rates than would be seen in a eld situation,
particularly with rubber materials that have extremely low wear
with rounded particles. The larger stand-off distance of the
sample (100 mm) in the SJE test also allows more diffusion of
the slurry stream than in the jet eductor tester (20 mm stand-off).
This will further reduce the wear rate in the SJE test relative to the
jet test.
A couple of nal comments:

 Cast white iron microstructure is critical in determining wear


rate and any laboratory sample must be the same as the eld
part to ensure reasonable accuracy of the laboratory tests
(reducing carbide size and increasing carbide volume in
hypereutectic white iron can reduce wear rate by up to 10 
over that of standard eutectic white iron). Producing laboratory (small) scale samples of white iron automatically leads to

ner microstructure than would be seen with large thickness


cast parts.
One of the major logistics problems with a laboratory wear
test is trying to get a result in a reasonable period of time,
particularly with elastomeric materials. Inevitably speeding up
the test (by increasing velocity and using sharper, larger
particles) changes the wear mechanisms, invalidating the
result.
The Coriolis test looks most promising in simulating the wear
conditions but suffers because of the limit on maximum
particle size ( o1 mm). A modied tester design that could
handle a 36 mm particle top size may give more accurate
wear results for mill circuit slurries.

5. Conclusions
1 The eld test of a slurry pump throatbush found that the wear
rate on natural rubber was 45% greater and the wear rate of a
hypereutectic white iron was 43% less than that for a standard
high chrome eutectic white iron (ISO21988/JN/HBW555XCr);
2 coriolis and jet eductor lab wear tests gave same order of
magnitude relative results as did the eld test for the hypereutectic white irons; however accuracy was poor;
3 the DSRW and SJE lab test found that 35% Cr hypereutectic
white iron wear rate was higher than that of the standard 27%
Cr eutectic white iron (i.e. opposite of the eld trial);
4 the SJE, Coriolis and jet eductor lab test found that the natural
rubber wear rates were much lower than that of the standard
27% Cr eutectic white iron (i.e. opposite of the eld trial);

1034

C.I. Walker, P. Robbie / Wear 302 (2013) 10261034

5 the differences in the laboratory wear test are largely attributed to the wear mechanisms not being representative of
those seen in the eld situation (due to particle impact energy
and/or impingement angle and particle shape and/or size).

Acknowledgements
The permission of The Weir Group PLC to publish this paper
is gratefully acknowledged. Warmans and Hyperchromes are
registered trademarks of Weir Minerals Australia Ltd.
References
[1] C.I. Walker, Slurry pump side-liner wear: comparison of some laboratory and
eld results, Wear 250 (2001) 8187.

[2] C.I. Walker, Slurry pump wear life uncertainty analysis, in: Proceedings of
Hydrotransport 14, BHR Fluid Engineering, Maastricht, Holland, 1999,
pp. 663679.
[3] C.I. Walker, Aspect of wear in mill circuit pumps, in: Proceedings of SAG 2011,
Vancouver, Canada, 2012.
[4] International Standard ISO21988: 2006 (E), Abrasion resistant cast irons
classication.
[5] K.F. Dolman, C.I. Walker, C.P. Harris, A.W. Thomson, Microstructurally rened
multi-phase castings, US Patent 5,803,152, 1998.
[6] C.I. Walker, G.C. Bodkin, Erosive wear characteristics of various materials, in:
Proceeding of Hydrotransport 12, BHR Fluid Engineering, Brugge, Belgium,
1993, pp. 191210.
[7] R.J. Llewellyn, S.K. Yick, K.F. Dolman, Scouring erosion resistance of metallic
materials used in slurry pump service, Wear 256 (2004) 592599.
[8] L.C. Jones, Low angle scouring erosion behaviour of elastomeric materials,
Wear 271 (2011) 14111417.

Вам также может понравиться