You are on page 1of 9

Wear 302 (2013) 1026–1034

Contents lists available at SciVerse ScienceDirect

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

Comparison of some laboratory wear tests and field 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: A number of different laboratory wear tests have been undertaken to measure the wear resistance of a
Received 23 August 2012 natural rubber and a eutectic and hypereutectic white iron under abrasion and erosion conditions. Laboratory
Received in revised form work included two different slurry jet erosion tests, a Coriolis test and an ASTM dry sand rubber wheel test.
19 November 2012
The laboratory results were compared with wear of the same materials in a centrifugal slurry pump
Accepted 22 November 2012
Available online 7 December 2012
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
Keywords: between the rubber and the best white iron. Coefficient of variance of the data was in line with typical wear
Wear results from the field.
Slurry
The laboratory wear tests were conducted with a silica sand slurry and average particle size range of 300–
Erosion
500 mm to match the field conditions. The Coriolis and one of the jet erosion tests showed order of magnitude
Pump
similarity with the field 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 field, while the DSRW
test found that the eutectic white iron wear rate was lower than that of the hypereutectic iron (all opposite of
the field test).
Explanation for the different wear rates between the laboratory and field tests was postulated to be non-
representative wear mechanisms. This is compounded by the lack of understanding of specific 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.

1. Introduction 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
The wear of centrifugal slurry pumps in mill circuit applica- abrasion and mostly erosion. During operation with the above pump,
tions in mineral processing plant is generally quite severe. Typical the throatbush is adjusted regularly (often weekly) up to touch with
life of pumps is in the range 1500–4000 h and material wear the rotating impeller causing both metal–metal contact and wedging
rates can be over 2 mm/day. Part section thickness in large pumps of any particles in the gap. There is however only a relatively short
of over 100 mm is not uncommon. A typical Warmans MC mill touching period and as both parts subsequently wear there is
circuit slurry pump application is shown in Fig. 1. increased flow in the gap and erosion wear. As the gap increases
The mill circuit pump shown above has a number of key there is less abrasion wear and more erosion wear. This cycle repeats
internal parts subject to wear. These include the rotating impeller with 8–10 adjustments not uncommon over the life of the part.
that imparts energy to the fluid, the casing liner, the frame liner The objective of the current research is to compare the wear life
(or back liner) and the throatbush or inlet side-liner. The impeller seen in the field application of the throatbush material with that of
and throatbush orientation is shown in Fig. 2. similar material in laboratory simulated wear tests. The laboratory
Walker [1] shows that the throatbush in a mill circuit slurry tests included two tests in Australia (using the author’s erosion jet
pump often wears faster than the other components due to a tester and a commercial dry sand rubber wheel (DSRW) abrasion test
combination of the sharp particle shape, the coarse particle size at a technical institute) and two tests (slurry jet erosion (SJE) and
distribution and the high velocity recirculating flow. Given both Coriolis erosion) undertaken at research establishments in Canada.
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. 2. Field wear data

2.1. Wear rate measurement method

n
Corresponding author. Tel.: þ61 2 99345100. In the current work, the wear rate in mm/day is used as the
E-mail address: craig.walker@weirminerals.com (C.I. Walker). base measure for comparison. This assumes a relatively constant

0043-1648/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2012.11.053
C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034 1027

Gouge
Baseline

Fig. 3. Part wear measurement using a template.

18
16
14

frequency of failure
12
10
8
6
4
2
0
Fig. 1. Warmans MC mill circuit (cyclone feed) slurry pump.
96 336 576 816 1056 1296 1536
age at failure (hrs)

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

Impeller mean and standard deviation so as to meaningfully compare the


different material performance. In field slurry applications (oper-
ating 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
Throatbush requirements.
(or inlet side-liner) 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 significant data set over a
Fig. 2. Sectional view showing relationship between impeller and throatbush. 2 year period.

2.2. Slurry particle size distribution


mass flow rate through the pump for all data points. Wear depth
on the throatbush was measured using a simple template as The slurry particle size analysis is shown in Fig. 5. As illustrated
shown in Fig. 3. in Walker [3], mill circuit applications have a d50 particle size
A point to note in Fig. 3 is that the wear is not uniform, but typically250–500 mm and a d85 size of 600–3000 mm. The current
rather there is a ‘‘baseline wear’’ which is the overall surface wear application has a relatively coarse large fraction (d85 of 3000 mm),
and a ‘‘gouging wear’’ that occurs locally and is generally much but is otherwise reasonably typical with a d50 of 300 mm.
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 2.3. Velocity and contact conditions
for the part.
In examining data from different pump applications (including As mentioned in the introduction, an exact determination of
some mill circuit), Walker [2] found significant variability in the flow in the gap between the impeller and throatbush (and
measured wear life, with the coefficient of variance—COV (stan- which is the primary influence on the wear of the throatbush) is
dard deviation/mean) on the order of 0.2–0.3. A typical wear life difficult to determine. Further, the gap is changing all the time at
distribution curve is shown in Fig. 4. a rate of up to 2 mm/day. To give some perspective to the likely
Given this sort of variability, it is essential that there is particle velocity seen in the gap a number of assumptions can be
sufficient data to be able to statistically determine a wear rate, made. Given that a typical gap for the application is around 6 mm,
1028 C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034

recirculation flow of around 5% of the net flow through the pump 35% Cr hypereutectic (650 þHBW) white iron to ISO21988/JW/
and tangential velocity of half the impeller velocity (say13 m/s), HBW600XCr35. The rubber part was compression moulded as a
in the area around the inlet the slurry velocity could be on the monolithic part over a mild steel reinforcing, while the metal
order of 16 m/s. parts were inserts in the rubber that extended to about 50% of the
The other complicating factor for flow in the gap is determin- diameter of the rubber part (as illustrated in Fig. 6).
ing a typical particle size distribution. Because of the strong Table 1 shows the wear rates for the 3 different materials. The
centrifugal field due to the rotating impeller (at the impeller tip, results show that the wear rate for the 27% Cr white iron was 30%
tangential velocity is 425 m/s), it is hypothesised that larger less than that of the rubber, while that of 35% Cr white iron was
particles entering the gap will be expelled and that only the some 60% less. The variability (COV) for the rubber was about
smaller particles would return to the eye of the impeller. double that for the metal parts.
Typically, the drag force on particles smaller than 50–70 mm The very good performance of the 35% Cr iron is a function of
would be sufficiently high relative to the centrifugal force to the microstructure and increased volume of hard carbide phase
ensure their return to the impeller eye, but for any larger particles relative to the 27% Cr iron. The 35% Cr is part of a unique family of
their presence at the impeller eye is somewhat moot. hypereutectic white irons (tradename Hyperchromes) that are
inoculated to refine the microstructure during casting [5].

2.4. Materials tested


3. Laboratory wear tests
The throatbush configuration is shown in Fig. 6. The 3
materials tested were a carbon black filled natural rubber of Given hypothesis that the throatbush wear is a mix of abrasion
50 Duro hardness, a heat treated 27% Cr eutectic (600þHBW) and erosion, the laboratory wear test results compared were a
white iron to ISO21988/JN/HBW555XCr27 [4] and a nominal 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]
120
for similar materials were also compared.
100
3.1. Jet tester (eductor)
Cum. % Passing

80
The jet test (eductor) is described by Walker and Bodkin [6]
60
and consists of municipal (tap) water with solid particles
entrained being blasted onto the surface of the sample. The
40 sample holder can rotate to give a variety of impingement angles
between 151 and 901. The water flow rate can be metered and the
20 resultant slurry jet velocity can be varied between 10 and 20 m/s.
Both the solid particle sizing and the liquid velocity determine the
0 concentration of the slurry in the nozzle. Particles are passed
1 10 100 1000 10000 100000 through the rig only once to ensure no degradation of shape. The
Size (microns) test conditions used which most closely represent the slurry
conditions of the field trial are shown in Table 2.
Fig. 5. Slurry particle size distribution from mill circuit pump.
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 metal in all cases. The wear rate of the 35% Cr iron is greater than
template that of the 27% Cr for the sharp particles, but less for the rounded
particles.
worn
material

rubber Table 2
Jet test (eductor) test conditions.

Impingement angle (deg.) 30


Jet velocity (m/s) 20
Duration (min) 10
Sand flow (g/min) 1400
Nozzle dia. (mm) 4.5
Stand-off distance (mm) 20
Sharp particle (mm) Alumina 300–600
Fig. 6. Throatbush showing typical wear pattern with 40 mm of material Round particle (mm) Silica sand 250–500
removed.

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

Material description Mean gouging wear rate Coefficient of No. of data points Wear rate relative
(mm/day) variance to 27% Cr

50 Duro natural rubber 2.26 0.660 25 1.45


Composite mix of natural rubber and 27% Cr white iron insert 1.56 0.306 9 1.00
Composite mix of natural rubber and 35% Cr white iron insert 0.888 0.347 10 0.57
C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034 1029

Table 3 steel) reference. Additional variants of each composition were


Comparative wear rates of different materials [6]. tested–with and without heat treatment and with microstructural
refinement for the 35% Cr material. Results are shown in Table 5.
Material Sharp particle Round particle Wear rate Wear rate
description wear rate wear rate relative to 27% relative to As can be seen the 35% Cr material that performed so well in
(mm3/min) (mm3/min) Cr (sharp) 27% Cr the field was the same as or worse in this test than the 27% Cr
(round) material. Heat treatment (and higher hardness) seemed to
improve wear resistance, while the finer carbide 35% Cr iron
40 Duro 3.53 0.605 0.58 0.31
natural
performed better than the coarser carbide material (when both
rubber were not heat treated).
27% Cr 6.05 1.95 1.0 1.0
white
3.3. Slurry jet erosion (SJE)
iron
35% Cr 7.44 1.49 1.2 0.76
white The SJE is a recirculating slurry loop with an air operated
iron, heat diaphragm pump that is controlled to ensure constant velocity of
treated flow 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
Table 4 profilometer. A photo of the rig and typical sample profile are
DSRW test conditions.
shown in Figs. 8 and 9, respectively.
Test load (N) 130 Results for the SJE test are shown in Table 7. Although very
Wheel speed (r/min) 200 similar test rigs, the wear rates for the SJE are much lower than
Surface speed (m/s) 2.3 those for the jet tester. Overall, the 35% Cr materials did not wear
Duration (min) 30 as well as the 27% Cr material. Wear rates for the rubber were
Sand flow (g/min) 465
Particle (mm) Silica sand 150–300
very low compared to the white irons.
Rubber wheel 223 mm, 60 Shore A
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 pre-
weighed and dispensed continuously into a plastic funnel where
it is mixed with metered municipal water to produce a 10 wt%
slurry that flows 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 cen-
trifugal 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 212–300 mm (similar to the SJE test) and a SiC

Table 5
DSRW laboratory wear test results on white irons.

Test material Mass loss relative Wear rate relative


to K110 steel to 27% Cr

27% Cr (heat treated) 0.94 1.00


35% Cr (heat treated) 0.89 0.95
35% Cr coarse carbide (no H/T) 1.61 1.71
35% Cr fine carbide (no H/T) 0.91 0.97

Fig. 7. DSRW tester. Table 6


SJE test conditions.

3.2. Dry sand rubber wheel (DSRW)


Impingement angle (deg.) 20
Jet velocity (m/s) 16
The two metal samples were also tested using an ASTM G65 Duration (min) 120
DSRW test under the conditions outlined in Table 4. A photo of Sand concentration (w/w %) 10
Particle detail (mm) Silica sand 212–300
the rig is shown in Fig. 7.
Nozzle dia. (mm) 5
Each sample was tested twice, and the 35% Cr samples 4 times. Stand-off distance (mm) 100
Sample mass loss was measured relative to a K110 (heat treated
1030 C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034

angular particle with a size range of 205–365 mm. Test conditions resolution is 26 mm for all acquired maps. A countor representa-
are summarised in Table 8 and a schematic of the test rig is tion of a wear scar with X and Y line profiles is illustrated in
shown in Fig. 10. Fig. 11.
Volume loss determinations for the wear scars were carried The wear test results are shown in Table 9. In all cases the
out with a non-contacting optical 3D surface imaging system. wear rates of the Hyperchromes materials were less than that of
The vertical resolution of this instrument is 3 nm, and the spatial the 27% Cr iron, wear rate decreasing with increasing carbide
volume. The impact of carbide size was very apparent, with very
fine 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.

Test material Wear rate Wear rate relative


(mm3/min) to 27% Cr

40 Duro natural rubber 0.0017 0.03


50 Duro natural rubber 0.00275 0.05
27% Cr (heat treated) 0.0561 1.00
35% Cr coarse carbide 0.0813 1.86
35% Cr fine carbide 0.105 1.45
Fig. 8. Slurry jet erosion tester.

Fig. 9. Typical SJE white iron worn sample.


C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034 1031

4. Discussion of results the 35% Cr iron with a higher wear rate than the 27% Cr iron
which is the opposite of what was seen in the field test. The SJE,
The relative wear rates for all the materials on different testers jet eductor and Coriolis test have the rubber performing better
are shown in Table 10. In general the DSRW and the SJE test have than the 27% Cr iron which again is the opposite of what was seen
in the field test. The Coriolis test and the jet test have the 35% Cr
Table 8 iron with a similar lower wear rate compared to the field test;
Coriolis test conditions [7,8]. however predictive accuracy was relatively poor.

Impingement angle (deg.) 0–8


Slurry velocityapprox. (m/s) 14–24 4.1. Comparison of test conditions
Duration (min) 6
Solidsconcentration (w/w %) 10
Round particle detail (mm) Silica sand 212–300
Mean particle sizes are similar for all laboratory tests
Sharp particle detail (mm) Silicon carbide 205–365 (150–500 mm) and also similar to the mean particle size of the field
Wear scar width (mm) 1 test (300 m). The major difference is with particle size distribution,
Stand-off distance (mm) 40 where the field 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 field tests. In the field
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 field test. The quantity of larger particles seen in the field 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 field test and the sand laboratory tests, but the sharp SiC

Table 9
Coriolis laboratory wear test results on white irons.

Test material Wear rate relative Wear rate relative


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

40 Shore A natural rubber 0.22 –


60 Shore A natural rubber 0.86 –
27% Cr (heat treated) 1.00 1.00
27% Cr—very fine carbide (H/T) – 0.3
35% Cr coarse carbide (heat treated) 0.54 0.64
35% Cr coarse carbide (no H/T) – 0.43
35% Cr—very fine carbide (no H/T) – 0.11
Fig. 10. Coriolis wear tester arrangement [7].

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


1032 C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034

Table 10
Relative wear rates of different materials–compared to 27% Cr white iron for both silica sand (round) and SiC (sharp) 150–500 mm particles.

Material description Field test DSRW test SJE test Coriolis test (round) Coriolis test (sharp) Jet test (round) Jet test (sharp)

40 Duro natural rubber 0.03 0.22 0.31 0.58


50 Duro natural rubber 1.45 0.05
60 Duro natural rubber 0.86
35% Cr coarse carbide (heat treated) 0.95 0.64 0.54 0.76 1.2
35% Cr coarse carbide (no H/T) 1.71 1.86 0.43
35% Cr fine carbide (no H/T) 0.57 0.97 1.45 0.11

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 (1400–1900 HV)
relatively more than that of the silica sand (1000 HV). Fig. 13. Photo of worn surface of 35% Cr iron following field test.
Relative slurry velocities in the laboratory erosion tests (10–20
m/s) were similar to those of the field test (16 m/s estimated) while be seen in Fig. 14. In both laboratory test cases the carbides and
the DSRW test rubbing speed was considerably lower at 2.3 m/s. matrix of the material seem to be worn at the same rate,
While the slurry concentrations of the laboratory tests (10 wt%) indicating greater particle impact forces (i.e. higher hardness of
were lower than the field condition (55 wt%) it is thought not to be a the carbides has not helped). A further indicator of the high
significant influence on the relative performance of various materials. stresses seen in the laboratory test wear is that the carbide
trailing edges are cracked at right angles to the wear direction.
4.2. Effect of microstructure on white iron results Wear direction is that of the rubber wheel rotation in the DSRW
and the particle direction in the jet eductor test.
It is well known that the finer the microstructure, the smaller In the jet eductor test (Fig. 14b) significant ploughing is evident
the inter-carbide distance, and the more effective the carbide along with carbide spalling and breakout from the matrix surface.
volume can be in preventing wear [7]. In all of the tests under- Fig. 15 shows the worn 35% Cr white iron sample from the
taken and reviewed here (Coriolis, jet, SJE and DSRW) the finer Coriolis test work of Jones [8]. As can be seen in Fig. 15(a) for the
carbide material performed better. rounded particle wear, the bulk of the material removal occurs in
As mentioned previously, Llewellyn et al. [7] found that the the matrix due to micro-ploughing while the carbides have edge
chill cast surface wear in the Coriolis test had mean wear rate rounding and stand out from the matrix. In Fig. 15(b) with the
values some 3–4  greater than those for coarser structures of the hard and sharp SiC particles, the wear is considerably different,
same casting composition (see Fig. 12). with ploughing and micro-machining occurring uniformly across
the surface, regardless of the carbide structure.
4.3. Wear mechanisms There is no carbide fracturing evident with the Coriolis test
like that seen from the field sample (viz. Fig. 13) and this may be
The worn 35% Cr iron throatbush surface from the field trial is due to the much smaller top size of the particles in the laboratory
shown magnified in Fig. 13. The wear mechanisms seem to be a test (300 mm) compared to the field test (10,000 mm).
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 4.4. Predictive ability of different tests
mostly in the same direction and could potentially be a result of
the 3 body abrasion (particles jammed between rotating impeller Llewellyn et al. [7] believe that the Coriolis test provides
and throatbush) or directional erosion due to some of the larger discriminative power that is almost two orders of magnitude
particles. The matrix has been worn at a large rate, leaving the greater than those of traditional slurry jet tests on white iron
carbides standing ‘‘proud’’ of the surface. materials. In jet tests, the damage is mainly caused by high
The worn surface morphology of the DSRW and jet eductor normal impacts; thus, the wear rate mainly reflects the material’s
test samples are quite different from the field test surface as can resistance to plastic deformation and carbide cracking. In Coriolis
C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034 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 finer microstructure than would be seen with large thickness
material properties also play major roles in the extent of damage. cast parts.
The challenge in using the jet test and the Coriolis test is to  One of the major logistics problems with a laboratory wear
ensure that the particle size, shape and velocity are representa- test is trying to get a result in a reasonable period of time,
tive, so that ensuring the wear mechanisms with both white iron particularly with elastomeric materials. Inevitably speeding up
and rubber is similar to what is happening in the field. This seems the test (by increasing velocity and using sharper, larger
to be particularly the case with rubber, where exceeding a particles) changes the wear mechanisms, invalidating the
threshold particle shape and size can radically alter the wear rate result.
as contact conditions change from purely elastic collisions to  The Coriolis test looks most promising in simulating the wear
cutting or tearing. conditions but suffers because of the limit on maximum
The DSRW abrasion test does not appear to be that helpful in particle size ( o1 mm). A modified tester design that could
predicting predominantly erosion wear in a throatbush. handle a 3–6 mm particle top size may give more accurate
The SJE recirculating slurry jet test is problematic due to the wear results for mill circuit slurries.
attrition of the particles over the test period. This results in much
lower wear rates than would be seen in a field situation,
particularly with rubber materials that have extremely low wear 5. Conclusions
with rounded particles. The larger stand-off distance of the
sample (100 mm) in the SJE test also allows more diffusion of 1 The field test of a slurry pump throatbush found that the wear
the slurry stream than in the jet eductor tester (20 mm stand-off). rate on natural rubber was 45% greater and the wear rate of a
This will further reduce the wear rate in the SJE test relative to the hypereutectic white iron was 43% less than that for a standard
jet test. high chrome eutectic white iron (ISO21988/JN/HBW555XCr);
A couple of final comments: 2 coriolis and jet eductor lab wear tests gave same order of
magnitude relative results as did the field test for the hyper-
 Cast white iron microstructure is critical in determining wear eutectic white irons; however accuracy was poor;
rate and any laboratory sample must be the same as the field 3 the DSRW and SJE lab test found that 35% Cr hypereutectic
part to ensure reasonable accuracy of the laboratory tests white iron wear rate was higher than that of the standard 27%
(reducing carbide size and increasing carbide volume in Cr eutectic white iron (i.e. opposite of the field trial);
hypereutectic white iron can reduce wear rate by up to 10  4 the SJE, Coriolis and jet eductor lab test found that the natural
over that of standard eutectic white iron). Producing labora- rubber wear rates were much lower than that of the standard
tory (small) scale samples of white iron automatically leads to 27% Cr eutectic white iron (i.e. opposite of the field trial);
1034 C.I. Walker, P. Robbie / Wear 302 (2013) 1026–1034

5 the differences in the laboratory wear test are largely attrib- [2] C.I. Walker, Slurry pump wear life uncertainty analysis, in: Proceedings of
uted to the wear mechanisms not being representative of Hydrotransport 14, BHR Fluid Engineering, Maastricht, Holland, 1999,
pp. 663–679.
those seen in the field situation (due to particle impact energy [3] C.I. Walker, Aspect of wear in mill circuit pumps, in: Proceedings of SAG 2011,
and/or impingement angle and particle shape and/or size). Vancouver, Canada, 2012.
[4] International Standard ISO21988: 2006 (E), Abrasion resistant cast irons–
classification.
Acknowledgements [5] K.F. Dolman, C.I. Walker, C.P. Harris, A.W. Thomson, Microstructurally refined
multi-phase castings, US Patent 5,803,152, 1998.
[6] C.I. Walker, G.C. Bodkin, Erosive wear characteristics of various materials, in:
The permission of The Weir Group PLC to publish this paper Proceeding of Hydrotransport 12, BHR Fluid Engineering, Brugge, Belgium,
is gratefully acknowledged. Warmans and Hyperchromes are 1993, pp. 191–210.
registered trademarks of Weir Minerals Australia Ltd. [7] R.J. Llewellyn, S.K. Yick, K.F. Dolman, Scouring erosion resistance of metallic
materials used in slurry pump service, Wear 256 (2004) 592–599.
[8] L.C. Jones, Low angle scouring erosion behaviour of elastomeric materials,
References Wear 271 (2011) 1411–1417.

[1] C.I. Walker, Slurry pump side-liner wear: comparison of some laboratory and
field results, Wear 250 (2001) 81–87.

You might also like