Вы находитесь на странице: 1из 19

Materials Research Express

PAPER

Abrasive wear behavior of Nb-containing hypoeutectic Fe–Cr–C


hardfacing alloy under the dry-sand/rubber-wheel system
To cite this article: Runzhen Yu et al 2019 Mater. Res. Express 6 026535

View the article online for updates and enhancements.

This content was downloaded from IP address 128.143.7.175 on 15/11/2018 at 23:46


Mater. Res. Express 6 (2019) 026535 https://doi.org/10.1088/2053-1591/aaedf1

PAPER

Abrasive wear behavior of Nb-containing hypoeutectic Fe–Cr–C


RECEIVED
16 August 2018
hardfacing alloy under the dry-sand/rubber-wheel system
REVISED
22 October 2018
ACCEPTED FOR PUBLICATION
Runzhen Yu1, Yong Chen1, Shengxin Liu1,3 , Zhiquan Huang2,3 , Wei Yang2 and Wei Wei2
2 November 2018 1
School of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001 (People’s Republic of China)
PUBLISHED
2
Department of Special Welding Material, Zhengzhou Research Institute of Mechanical Engineering, Zhengzhou 450001 (People’s
14 November 2018 Republic of China)
3
Authors to whom any correspondence should be addressed.
E-mail: zdclyjs@163.com

Keywords: weld overlay, abrasive wear, three-body, rolling contact, Fe–Cr–C alloy, NbC

Abastract
Dry sand/rubber wheel test with different normal loads (70 N, 190 N) and rubber wheel rotational
speeds (100 rpm, 130 rpm, 160 rpm, 190 rpm, 220 rpm) was carried out on Nb-containing
hypoeutectic Fe–Cr–C hardfacing alloy. The change of alloy wear behavior with wear conditions was
investigated. The results showed that increasingly-severe wear condition deteriorated the uniformity
of macro wear scar. The wear behavior change of two regions on the wear scar affected the overall wear
loss variation. When abrasives driven by rubber wheel were gradually fed into the immediate-contact
position between rubber wheel and alloy surface, a fully-worn region was left on the alloy surface.
Afterwards, abrasives continued to be driven by rubber wheel away from the immediate-contact
position and tended to leave a weakly-worn region that can impede the increase in wear loss. The wear
mechanism of the fully-worn region was the dendritic austenitic micro-cutting and, the spalling of
NbC and eutectic-carbides. The weakly-worn region, however, mainly showed a mechanism of slight
cutting/ploughing along wear direction. Under the 70 N load, abrasives driven by increasing
rotational speed performed severer abrasion to make the wear scar mainly consist of the fully-worn
region, which resulted in a relatively uniform macro wear scar and the constant increase in wear loss.
Under the 190 N load, differently, the increasing rotational speed significantly intensified the work-
hardening of the fully-worn region and the fragmentation of abrasives, which was advantageous for
improving the wear-resistance of austenitic matrix and developing the area of weakly-worn region.
Therefore, as the rotational speed increased, the alloy wear loss showed a trend of first stepping up and
then decreasing.

1. Introduction

Fe–Cr–C hardfacing alloys are currently widely used for repairing and strengthening of various wear-resistance
workpieces in manufacturing and mining [1–4]. Hypoeutectic Fe–Cr–C alloy has the ductile austenite matrix
with cracks having a tendency to propagate along brittle eutectic-carbides in a small range. [5, 6]. And the strain-
induced martensite (SIM) is easily generated in austenite grain to harden the worn surface and improve the
wear-resistance to some extent [7–9]. Chen et al [5] reported that the dendritic γ-Fe with lower hardness than
abrasives was always susceptible to get cut/ploughed, moreover, the micro-cracks tended to indefinitely
propagate along brittle and reticular carbides in eutectic. Sabet et al [7] compared the difference in wear behavior
of Fe–Cr–C hardfacing alloy between hypoeutectic and hypereutectic through dry sand/rubber wheel test. The
resulted showed that a pronounced SIM layer generated in the worn subsurface of hypoeutectic alloy. And
profited from work-hardening, the hypoeutectic alloy exhibited less wear weightlessness than the hypereutectic
alloy which underwent significant brittle fracture.

© 2018 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 026535 R Yu et al

Table 1. Elemental composition of experimental hardfacing alloy. (wt. %).

C Cr Nb Mn Si Ni P, S Fe

1.0–2.5 „15.0 „15.0 „2.5 „2.0 0.3–0.5 „0.035 Bal.

Table 2. Welding conditions.

Parameters Value

Electric arc voltage (V) 30 to 32


Electric arc current (A) 280 to 320
Traveling velocity (mm/s) 2.48
Oscillating velocity (mm/s) 20
CO2 flowrate (L/min) 15 to 18
Stick-out (mm) 10 to 15
Pre-heating No
Inter-pass temperature (°C) Bottom layer:<370
Intermediate and top layer: 340
to 370
Width of bead (mm) 35 to 40
Bead step over (mm) 12
Cooling method Slow cooling (wrapped with asbes-
tos cloth)

Due to better bearing capacity of ductile austenitic matrix, alloying elements to form refractory carbides are
often added into the alloy to in situ synthesize carbides as hard phases during the metallurgical reaction [10–12].
The addition of Nb can refine the dendritic austenite due to the heterogeneous nucleation, and on the other
hand, Nb-carbides with high-hardness dispersed greatly in matrix can also effectively resist severe abrasive wear
[10, 13–15]. Filipovic et al [6] reported that due to their characteristic morphology, NbC showed a higher
dynamic fracture toughness than eutectic-carbides M7C3, moreover, a higher content of niobium addition
decreased the volume fraction of brittle M7C3, which improved the fracture toughness of alloy since the cracks
propagated easily through reticular eutectic-carbides.
Current researches about the wear-resistance of Fe–Cr–C hardfacing alloy focused mostly on the effect of
alloy composition adjustment and carbides morphology controlling, however, there are few investigations on
the wear behavior of a certain alloy under a certain wear condition [16]. In this study, dry sand/rubber wheel test
with different normal loads (70 N, 190 N) and rubber wheel rotational speeds (100 rpm, 130 rpm, 160 rpm,
190 rpm, 220 rpm) was performed on Nb-containing hypoeutectic Fe–Cr–C hardfacing alloy. Then abrasive
wear behavior of alloy with change of wear conditions was investigated through relevant characterizations and
tests for wear scars and their sub-surfaces.

2. Experimental

2.1. Welding process and specimen preparation


The base metal used for surfacing was the 220 mm×160 mm×25 mm plate prepared from low carbon steel of
Q235. And the self-made flux-cored wires of Nb-containing hypoeutectic Fe–Cr–C alloy were utilized to deposit
on base metal through GMAW (CO2 gas protection). The elemental composition of hardfacing alloy obtained
finally is given in table 1. And the welding conditions are listed in table 2. The welding deposition (height of
8∼12 mm) consisted of three beads and three layers overlapped. Continuous multi-beam multi-layer
automatic welding method with high inter-pass temperature was performed. And each bead was surfacing
through oscillating traverse method. Welding directions of two adjacent beads were completely opposite to
maintain a uniform distribution of internal stress. The test plate completed by surfacing welding is shown in
figure 1(a).
Sampling using mechanical methods. The upper surface of deposition metal was first finely machined with
a grinder machine to obtain the pre-wear surface (Ra<0.75 μm) as shown in figure 1(c). Then the 71 mm×
25 mm×13 mm wear specimen was fabricated through wire electrical-discharge machining at the position
displayed in figure 1(b). Figure 1(c) also reveals the macro morphology of 76 mm×13 mm sectional surface of
wear specimen etched by 4% nitric acid alcohol. It can be seen that the hardfacing alloy with different distance to
the fusion line was subjected to different welding thermal cycles and resulted in a typical delamination. Three
layers from bottom to top respectively are: base metal, hardfacing alloy composed of coarse equiaxed austenitic

2
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 1. (a) Test plate after welding surfacing. (b) Obtaining position of wear specimen. (b) Wear specimen.

Figure 2. (a) Schematic of dry sand/rubber wheel testing machine according to ASTM 65 [7, 19]. (b) Foundry quartz sand abrasives.
(c) Size distribution of fresh abrasives. (d) EDS analysis result of quartz abrasive medias.

grains, and, the hardfacing alloy (investigated portion in this paper) with typical primary austenitic dendrites
[17, 18]. The thickness of three layers was controlled to be (from the bottom to the top) 2.8–3.2 mm,
3.5–4.1 mm, 5.7∼6.7 mm respectively in order to keep each specimen at the same level.

2.2. Dry sand/rubber wheel test


The three-body abrasive wear test was performed using a dry sand/rubber wheel apparatus (see figure 2(a))
according to the procedure A of ASTM-G65. In recent researches and industrial applications, though 130 N is
recommended as a relatively stable load parameter in dry sand/rubber wheel test, different normal loads of
20–250 N are also used widely [7, 19–21]. In this study, in order to investigate the wear behavior under a low
normal load and a high normal load more targetedly and properly, 70 N and 190 N were selected. And the
rubber wheel rotational speed series of 100 rpm, 130 rpm, 160 rpm, 190 rpm, 220 rpm was used to
systematically estimate the effects of different abrasives cutting-forces [20, 22]. In addition, since the
experimental hardfacing alloy is often applied to the surface repairing and strengthening of the massive-
materials grinding workpieces, in order to better simulate the harsh wear conditions with large-sized abrasive
medias in practical applications, the 460–540 μm foundry quartz sands (see figures 2(b), (c)) was used as the
experimental abrasives at an average flowing rate of 276–288 g min−1. The calculated wear distance was set to
4314 m. The rubber surface of rubber wheel was made of chlorabutyl with the hardness of A-60. The dimension

3
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 3. XRD diffraction patterns of experimental hardfacing alloy.

of wear specimen is shown in figures 1(b), (c). Weight loss (accurate to 0.0001 g) was used to evaluate the wear
degree. One normal load was matched with one rubber wheel rotational speed as one wear condition (e.g. 70 N/
100 rpm). And five specimens were tested under each wear condition for statistical calculation.

2.3. Characterization
FEI QUANTA 200 scanning electron microscope (SEM) and EDAX Genesis 2000 xms energy dispersive
spectrometer (EDS) were used to characterize the microstructure and wear scar characteristics of hardfacing
alloy. MAXima XRD-7000 x-ray diffractometer (XRD) was utilized to analyze the phase composition (with Cu-
Kα radiation and scanning speed of 5°/min). The granulometry of the abrasives before and after the tests were
evaluated using a laser granulometer Sympatec (model Helos/BF) to evaluate the breakage tendency.
The macro-morphology of wear scar were photographed using ZEISS SteREO Discovery.V8
stereomicroscope with the MosaiX stitching mode. Axio Vision SE64 software was used to measure the area
value of the macro wear scar. The laser scanning confocal microscope (LSCM, equipment: OLYMPUS LEXT
OLS4100) was used to obtain the 3D-morphology and longitudinal-sectional profile of the wear scar. And using
the matching LEXT software to measure the maximum depth of wear scar (μm). For one specimen, the depth
value of ten points at the deepest part of the wear scar was measured for statistical calculation. The macro-
hardness of the wear scar was characterized using a Wolpert 401 MVD digital Vickers-hardness tester (load of
300 gf, dwell time of 15 s). The testing region of each specimen was carried out hardness tests in twenty-five-
times for statistical calculations. In addition, for the experimental results of one specific wear condition, such as
the area value of the wear scar, the maximum depth value of the wear scar, the hardness of the wear scar, etc, the
results of five specimens under this wear condition were performed statistical calculation.

3. Results and discussion

3.1. Microstructure and hardness of experimental hardfacing alloy


Figure 3 reveals the XRD pattern of experimental hardfacing alloy. Peaks of γ-Fe, α-Fe, NbC and M7C3 can be
detected. Figure 4 shows the SEM images of alloy and x-ray mapping of Fe, Cr, C, Nb elements. The white
irregular-shaped NbC are distributed on the dendritic austenite matrix with distinct profile. And the eutectic
containing reticulated M7C3 carbides tended to finally solidify through N-type and be distributed around the
dendrites [23, 24]. In addition, since the solid-state phase transformation occurred in hardfacing alloy during
the slow cooling, acicular ferrite was precipitated in the dendritic austenite grains (see figure 4(b)). And the
corresponding α-Fe with bcc structure was also detected through XRD (see figure 3) [25, 26].
As shown in figure 4(c), Fe atoms concentrated in primary austenitic dendrites. Cr content in eutectic is
higher than that in primary dendrites, which can be ascribed to the low solubility of Cr in γ-Fe [6, 27, 28]. Nb
mainly involved in the formation of NbC shows more distribution in white phases. In addition, extremely low
content of Fe in NbC can be observed while a small amount of Nb is in a position to dissolve in solid solutions
and eutectic carbides [29, 30]. Figure 5 reveals the macro-hardness comparison between the experimental alloy
and the quartz abrasives, and the hardness ratio (hardfacing alloy to abrasives) is 0.816 calculated. Due to the
presence of a large number of primary austenite dendrites with low hardness and relatively good ductility,
hardfacing alloy shows a lower macro-hardness level. However, as the main wear-resistance phases,
eutectic-M7C3 and NbC with high-hardness can effectively resist the local abrasion of quartz sand with
low-stress [31].

4
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 4. (a) SEM images of experimental hardfacing alloy. (b) Enlarged view of area A. (c) x-ray mapping for Fe, Cr, Nb of figure 4(b).

Figure 5. Hardness comparison between hardfacing alloy and quartz abrasives. The hardness ratio (hardfacing alloy to abrasives) is
0.816 calculated.

3.2. Wear loss


Figure 6 reveals the results of abrasive wear test in form of wear loss versus rotational speed. The change trend of
wear loss with the rotational speed is significantly different between two normal loads. And wear loss at 190 N
shows a much higher level than 70 N. The increasingly severe wear condition (rotational speed, normal load) did
not always lead to a growth in the wear loss. As shown in figure 6, at a low normal load of 70 N, wear loss
increases continuously with the rotational speed, while at 190 N the change slope of wear loss gradually decreases
to less than zero indicating a trend of increasing first and then decreasing.

3.3. Macro-characteristics of wear scar


3.3.1. Macro-morphology
Figure 7 reveals the area value change of macro wear scar, and figure 8 shows the macro-morphologies of wear
scars obtained at six typical wear conditions (70 N/100 rpm, 70 N/160 rpm, 70 N/220 rpm, 190 N/100 rpm,
190 N/160 rpn, 190 N/220 rpm). Typical rectangular wear scar was generated through dry sand/rubber wheel

5
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 6. Wear loss change with various rotational speeds.

Figure 7. Area value change of macroscopic wear scar with various rotational speeds.

Figure 8. Macro-morphologies of wear scars obtained at six typical wear conditions: (a) 70 N/100 rpm, (b) 70 N/160 rpm,
(c) 70 N/220 rpm, (d) 190 N/100 rpm, (e) 190 N/160 rpn, (f) 190 N/220 rpm.

test [19, 32]. And increasingly severe wear condition inevitably grows the area value and deteriorated the
uniformity degree of wear scar. Taking the wear scar obtained at 190 N/160 rpm for an example (see figure 8(e)),
there are up to four distinct regions generated on wear scar after the test: brighter impact-erosion region, Region
A with a lower brightness, dark Region B, and, Region C produced exactly after the generation of Region B.
Abrasive particles tended to rebound and/or collide between the rubber wheel and the specimen after
dropping down from sand nozzle and firstly produced an impact-erosion region on the pre-wear surface of
hardfacing alloy [32]. It can be observed from figure 8 that the increasing rotational speed contributed to the area
growth of impact-erosion region, which can be attributed that the abrasives hitting the high-speed rotating
rubber wheel were given a higher deflective momentum to perform more significant impact-erosion action. In
addition, it can be noted that approximately symmetrical Region B tended to connect from two long sides to the

6
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 9. (a) Position of X′Y′Z′ axis relative to both wear specimen and wear direction. Longitudinal-sectional profile (Y′ axis as
abscissa) of macro wear scar with various rotational speeds under the normal loads of (b) 70 N and (c) 190 N.

middle position of wear scar to get significant area as the rotational speed increased. And the increase in
rotational speed at the high normal load (190 N) could more easily facilitate the generation of Region B. As
shown in figures 8(c), (e), (f), the formation of Region B shows relatively slight at 70 N/220 rpm but extremely
significant at 190 N/160 rpm and 190 N/220 rpm.

3.3.2. Longitudinal-sectional profile


The longitudinal-sectional profiles of wear scars obtained at six typical wear conditions are shown in figure 9. It
can be respected that the rolling-contact caused the profile of the wear scar to exhibit a significant concave arc
[33, 34]. Moreover, the arc bottoms (namely, the deepest parts of the wear scar, calling DW) at six wear
conditions are all relatively closer to the left side, which implies the abrasion that alloy underwent did not
proceed completely along the trajectory of the rubber wheel rolling, causing the wear scar to exhibit varying
degrees of non-uniformity.
As displayed in both figures 8 and 9(b), (c), the impact-erosion region and the Region C are respectively
located at two ends of the wear scars, and their profiles are substantially at the same height as the horizontal pre-
wear surface. So it can be considered that the wear loss these two regions brought about was extremely small. In
contrast, since Region A and Region B are within the significant concave arc, their contribution to the wear loss
cannot be ignored. Due to the immediate-contact between the rubber wheel and the alloy surface, severe
abrasive wear the DW underwent resulted in a considerable materials loss. However, as can be seen from the
circularly framed profile parts (see figures 9(b), (c)), which are correspond to the dark Region B shown in
figures 8(c), (e), (f), the Region B with shallower depth than the DW implies that the abrasive wear occurred here
was relatively weak to cause a low degree of materials removal. So it can be reasonably conjectured that the large
area growth of weakly-worn Region B was beneficial for reducing the wear loss, which was consistent with the
fact that under the 190 N load, as the rotational speed increases, the area of Region B increases significantly and
the growth of the wear loss is severely slowed down (see figures 6 and 8(e), (f)). Under the 70 N load, nevertheless,
the weakly-worn Region B shows slight even at 220 rpm (see figure 8(c)) and the longitudinal-sectional profiles
of wear scars are all relatively uniform (see figure 9(b)), which means that the abrasive wear proceeded relatively
evenly and thoroughly and tended to bring about more and more wear loss as the rotational speed increased.
Change of maximum wear scar depth is displayed in figure 10. At 190 N/220 rpm, wear scar reaches a depth
within 800 μm suggesting that the abrasive wear only took place in dendritic austenite layer of alloy far away
from the fusion line (see figure 1). As can be seen from figure 10, under the 70 N normal load, the maximum
wear scar depth increases with the rotational speed similarly to the change trend of wear loss (see figure 6).
However, under the 190 N load, although the increasing rotational speed always grows the maximum depth of
the wear scar, it slows down the growth, which means the abrasive wear that the DW underwent was gradually
weakened as the rotational speed increased.
Therefore, as analyzed above, it can be preliminarily considered that under the 190 N normal load, as
rotational speed increased, the generation of weakly-worn Region B and the attenuation of abrasion degree at
DW could be the main factors impeding the wear loss growth (see figure 6).

3.3.3. 3D-morphology
For the DW and the Region B, which affected the overall wear behavior of the alloy, their 3D-morphologies were
characterized through LSCM furtherly.
Figure 11 shows the 3D-morphologies of the DW obtained at six typical wear conditions. It can be observed
that the generation of wear scar was accompanied by lots of notable pits connecting in wear direction, which
resulted from the comprehensive action of normal pressing and tangential cutting/ploughing of the third-body

7
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 10. Maximum wear scar depth with various rotational speeds.

Figure 11. 3D-morphologies of DW obtained at six typical wear conditions: (a) 70 N/100 rpm, (b) 70 N/160 rpm, (c) 70 N/220 rpm,
(d) 190 N/100 rpm, (e) 190 N/160 rpn, (f) 190 N/220 rpm. DW: the deepest part of wear scar.

quartz abrasives. As can be seen from figures 11(d)–(f), the DW morphologies of 190 N show more dark
concaves compared with the relatively flat morphologies of 70 N, which implies the uneven abrasion was more
likely to occur at DW under the 190 N normal load. In addition, the 3D-morphologies of 190 N/160 rpm and
190 N/220 rpm (see figures 11(e), (f)) have more pronounced concaves in larger number than 190 N/100 rpm.
And there are not any observable differences in 3D-morphologies between 190 N/160 rpm and 190 N/220 rpm.
This may mean that under the 190 N normal load, as rotational speed increased, the uneven abrasion at DW was
exacerbated to a certain degree and then remained.
Figure 12 reveals the 3D-morphologies of pronounced Region B on wear scars obtained at three wear
conditions (70 N/220 rpm, 190 N/160 rpm, 190 N/220 rpm, see figures 8(c), (e), (f)). Almost no significant
differences can be noted between three morphologies. However, compared with figure 11, it can be seen that
Region B has the features obviously different from the DW. As shown in figure 12, the pits of Region B are too
small to be clearly observed, and grooves along wear direction formed by the connected small pits are extremely
straight and significant. Furthermore, there are visible burrs at the edge of deep grooves, which implies the alloy
underwent a lesser degree of plastic deformation at the location where the abrasives performed uniform
ploughing action [35, 36]. Therefore, it can be inferred that Region B actually was a worn surface dominated by
cutting/ploughing mechanism. Namely, abrasives at Region B tended to produce notable tangential cutting/
ploughing and extremely-weak normal pressing.

3.4. Micro-characterization of worn surface


Both the worn morphology and the cross-sectional sub-surface of DW and Region B were characterized through
SEM/EDS.

8
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 12. 3D-morphologies of Region B obtained at three typical wear conditions: (a) 70 N/220 rpm, (b) 190 N/160 rpm,
(c) 190 N/220 rpm.

Figure 13. SEM images (at high-power field) and x-ray mapping of DW obtained at (a) 70 N/160 rpm, (b) 190 N/160 rpm.

3.4.1. Microscopic worn morphology


Figure 13 shows the SEM images at high-power field of DW obtained at 70 N/160 rpm, 190 N/160 rpm, and the
x-ray mapping of Fe, Cr, Nb, Si, O, Al element. The large difference in thermal expansion coefficient and plastic-
strain rate between γ-Fe, eutectic-carbides and NbC resulted in more stress concentration at the phase interface,
which could induce the fracture and spalling of NbC and eutectic-carbides during the abrasive wear [7, 37, 38].
NbC with extremely high hardness played a main wear-resistant role and, became exposed and unstable when
the austenitic matrix was washed away [6, 10]. On the other hand, since the microcracks tended to propagate
along the brittle reticular carbides, when abraded by quartz the eutectic with higher content of Cr was prone to
localized flaky fracture and left craters. For the dendritic austenite matrix with low Cr content and relatively
good ductility, the hard and edgy abrasives were more inclined to leave visible grooves through the severe cutting
action. In addition, it can also be seen from figure 13 that there is a high content of O, Si, and Al detected in

9
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 14. SEM images (at relatively-low-power field) of DW obtained at six typical wear conditions: (a) 70 N/100 rpm,
(b) 70 N/160 rpm, (c) 70 N/220 rpm, (d) 190 N/100 rpm, (e) 190 N/160 rpn, (f) 190 N/220 rpm.

craters, which means small-sized quartz abrasives containing Al impurities were easily adhered here during the
abrasion [39, 40].
Figure 14 reveals the SEM images at relatively-low-power field of DW obtained at six typical wear conditions.
The random cutting/ploughing and pressing caused by loose third-body abrasives left lots of black craters and
non-directional grooves at DW.
Under the normal load of 70 N, it can be seen that at the worn morphology of 70 N/100 rpm (see
figure 14(a)), NbC are relatively well bonded to the matrix, and the isolated craters are distributed mainly on gray
austenite matrix and/or eutectic. As the rotational speed increased, the abrasives driven by high-speed rotating
rubber wheel got stronger tangential stress and produced more and deeper long grooves on worn morphologies
(see figures 14(b), (c)). Furthermore, the NbC spalling degree was also exacerbated with increasing rotational
speed. More and larger black spalling pits around the NbC communities can be seen at the worn morphologies
of 70 N/160 rpm and 70 N/220 rpm, that is, the preferentially spalling NbC due to the poor interface stability.
The spalled NbC could become new third-body abrasives to furtherly intensify the abrasion [41]. Therefore, it
can be inferred that the increase in rotational speed under the 70 N normal load improved the wear degree at
DW, which is manifested by the enhanced abrasives cutting action and the exacerbated NbC spalling tendency.
The worn morphologies of DW under the 190 N normal load show a significant difference compared with
70 N. The tendency of long grooves to become notable cannot be observed with the increase of rotational speed,
though some pronounced long grooves have already been produced at 190 N/100 rpm (see figure 14(d)). This
implies that under the normal load 190 N, the austenitic matrix gradually exhibited a better wear-resistance to
quartz abrasives as rotational speed increased. Moreover, a large number of small-sized craters that are not
concentrated around the NbC communities can be seen from the morphologies of 190 N/160 rpm and 190 N/
220 rpm (see figures 14(e), (f)). And the NbC communities visible on morphologies of 190 N/160 rpm and
190 N/220 rpm were relatively sparse compared with 70 N/160 rpm and 70 N/220 rpm. This indicates an
extremely severe spalling tendency of both NbC and eutectic-carbides. In addition, consistent with the 3D-
morphologies shown in figures 11(e), (f), there is also no significant difference in the SEM worn morphologies
between 190 N/160 rpm and 190 N/220 rpm (see figures 14(e), (f)). Therefore, it can be deduced that at the DW
of 190 N normal load, although both NbC and eutectic-carbides have a strong spalling tendency, the increase in
rotational speed did not always improve the degree of abrasion and materials removal. And as the rotational
speed increased, the wear-resistance of austenitic matrix appeared to be improved, whereas the spalling
tendency of materials was increased to a certain level and then substantially remained.
Figure 15 displays the SEM images of pronounced Region B on wear scars obtained at three typical wear
conditions. Totally different characteristics can be seen compared with the DW. A relatively large area of NbC
communities were exposed and well bonded to the matrix indicating that the NbC did not fully exert its superior

10
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 15. SEM images of Region B obtained at three typical wear conditions: (a) 70 N/220 rpm, (b) 190 N/160 rpm,
(c) 190 N/220 rpm.

Figure 16. (a) Vickers-hardness test results of both DW and Region B obtained at typical wear conditions, (b) corresponding calculated
hardness ratio (testing area to abrasive medias).

wear-resistance due to the weak abrasion occurred here. Moreover, visible grooves are almost parallel to the
wear direction, and fine/shallow abrasion pits generated from the weak pressing action of abrasives can be seen.
This is in agreement with the results and analysis of figure 12. In other words, at Region B, abrasives mainly
produced a strong and stable tangential cutting/ploughing action parallel to the wear direction. In addition,
wide grooves can be seen at the morphologies of 70 N/100 rpm (see figure 14(a)), while the grooves in
figures 14(b), (c) are relatively narrow and small in number, which means that abrasives tended to act a severer
cutting/ploughing at 70 N/100 rpm.

3.4.2. Hardness test


The Vickers-hardness test was carried out on both the DW and the Region B obtained at the typical wear
conditions, and the results are shown in figure 16(a). It can be respected that the weakly-worn Region B is at the
same level as the pre-wear surface. And in contrast, the severe strain resulted from fully abrasion caused the DW
to undergo a significant work-hardening and exhibit a higher hardness level than pre-ground surface [7, 42]. The
increase in normal load/rotational speed was helpful to improve the DW hardness. However, it can also be

11
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 17. SEM images of cross-sectional sub-surface of DW obtained at six typical wear conditions: (a) 70 N/100 rpm,
(b) 70 N/160 rpm, (c) 70 N/220 rpm, (d) 190 N/100 rpm, (e) 190 N/160 rpn, (f) 190 N/220 rpm.

observed from figure 16(a) that under the 190 N normal load, the growth of DW hardness is slowed down with
the increase in rotational speed.
Figure 16(b) reveals the calculated hardness ratio of the testing wear scar area to the abrasives, of which the
change trend is in line with the hardness result. During the abrasive wear, the higher normal load and rotational
speed tended to increase the hardness ratio of DW to abrasive medias. At 190 N/160 rpm, the hardness ratio is
greater than 1, which implies a difficulty for quartz abrasives to effectively abrade at DW [43]. Therefore,
although the rotational speed continued to increase from 160 rpm, the degree of both abrasion and materials
removal was hard to get improved remarkably. This could be responsible for the phenomenon that the
maximum wear scar depth develops slowly with the rotation speed from 190 N/160 rpm to 190 N/220 rpm
(see figure 10), and the fact that there are no observable differences in 3D and SEM morphologies between
190 N/160 rpm and 190 N/220 rpm (see figures 11(e), (f) and 14(e), (f)).

3.4.3. Microscopic morphologies of cross-sectional sub-surface


The cross-sectional sub-surface of DW obtained at six typical wear conditions are illustrated in figure 17. The
acicular SIM in dendritic austenite near the worn surface proves the fact that work-hardening took place at DW
during the abrasive wear, which is similar to the results reported in [7]. For the hypoeutectic Fe–Cr–C alloy, the
high content of Cr and the addition of Mn reduced the stacking fault energy of γ-Fe leading to a difficulty of
dislocation slip [44, 45], which is advantageous for the twins generation in austenite grains. During the abrasion
process, the alternating stress of abrasives and the deformation of austenite matrix led to the formation of shear
bands composed of stacking faults and twins in austenite grains at sub-surface. This furtherly caused the strain-
induced α-martensite to nucleate and grow at the intersection of the shear bands [46, 47].
As shown in figures 17(a)–(c), although the hardening-layer thickness of 70 N is relatively small, increasing
rotational speed constantly facilitates the SIM generation in austenite grains. Differently, the hardening-layer of
190 N normal load has a greater thickness and a denser SIM distribution than that of 70 N. And as can be seen
from figures 17(e), (f), the sub-surfaces at 190 N/160 rpm and 190 N/220 rpm have almost the same work-
hardening degree, which also demonstrates that under the 190 N normal load, the increase of rotational speed
from 160 rpm to 220 rpm failed to furtherly exacerbate abrasion and work-hardening at DW. In addition, SIM
formation may deteriorate the heterogeneity of austenite matrix to some extent. Due to the difference in thermal
expansion coefficient and plastic-strain rate between martensite and austenite [7, 37, 38, 42], stress
concentration tended to take place at the martensite/austenite interface to induce materials spalling in a
small scale randomly. Therefore, for the DW with severe work-hardening obtained at 190 N/160 rpm and
190 N/220 rpm, as presented in figures 14(e), (f), a pronounced tendency of materials spalling can be observed
on their SEM morphologies.

12
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 18. SEM images of cross-sectional sub-surface of Region B obtained at three typical wear conditions: (a) 70 N/220 rpm,
(b) 190 N/160 rpm, (c) 190 N/220 rpm.

Figure 19. Size distribution of worn abrasives: (a) 70 N/100 rpm, (b) 70 N/160 rpm, (c) 70 N/220 rpm, (d) 190 N/100 rpm,
(e) 190 N/160 rpn, (f) 190 N/220 rpm.

Figure 18 shows the cross-sectional sub-surface of Region B obtained at three typical wear conditions. The
notable work-hardening phenomenon is even difficult to be observed. Under the cutting/ploughing-dominated
wear mechanism, the degree of deformation and the normal stress of the abrasives were small, which caused the
sub-surface of Region B to be subjected to weak strain and difficult to undergo a remarkable work-hardening.

3.5. Size distribution of worn abrasives


The abrasive medias were also damaged in turn when they abraded the alloy [48, 49]. Especially at DW, due to
the direct action of the normal load, abrasive particles were more susceptible to damage into fragmentation.
Figure 19 reveals the size distribution of the worn abrasives at six typical wear conditions. It can be known from
figures 2(c) and 19(a)–(c) that the size distribution of abrasives after worn at 70 N load is essentially the same as
the fresh abrasives. Namely, the frequency of abrasive particles with a size of 490–510 μm is the highest.
However, as displayed in figures 19(d)–(f), after worn under 190 N normal load, the highest frequency of
abrasives size is below 490–510 μm, which signifies the abrasive medias underwent a nonnegligible
fragmentation. Moreover, it can also be seen that the size of worn abrasives under the 190 N normal load shows
an obvious decreasing tendency with the increase of rotational speed.
The austenite matrix with significant work-hardening under the 190 N normal load shows better wear-
resistance to quartz abrasive medias, which caused the abrasives to be more vulnerable to crushing into small-
sized at DW where the normal load acted directly. So it can be respected from figures 14(d)–(f) that due to the
weak cutting action produced by small-sized abrasives, increasing rotational speed did not result in more
pronounced long grooves left at DW. In addition, although a severe tendency of materials spalling can be seen at
the SEM images of DW obtained at 190 N/160 rpm and 190 N/220 rpm, the energy needed to cause wear loss
went more into damaging the abrasive medias than into abrading the alloy surface, which means the relatively

13
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 20. Schematic diagram: (a) Dry sand/rubber wheel abrasive wear system. (b) Enlarged view of the frame in figure 19(a). (c) Top
view of figure 19(a).

uneven abrasion inevitably occurred to the alloy surface [49, 50]. And therefore as displayed in the 3D-
morphlogies of DW obtained at 190 N normal load (see figures 11(d)–(f)), more dark concaves were produced
resulted from the uneven abrasion. This also explains from another angle that the extremely high work-
hardening degree of DW at 190 N/160 rpm and 190 N/220 rpm caused the abrasives to insufficiently abrade the
DW in the later stages of the abrasion process, so the maximum wear scar depth of 190 N develops slowly with
the increase of rotational speed from 160 rpm to 220 rpm (see figure 10).

3.6. Analysis on wear mechanism and formative mechanism of non-uniform macro wear scar
Although the wear characteristics of DW at different wear conditions show diversity, under the dry sand/rubber
wheel system, Region B with relatively weak abrasion was also involved in the overall wear behavior and affected
the wear loss. Based on the above test results and analysis, the schematic diagram of figure 19 can be drawn to
furtherly conjecture the formative mechanism of Region B and non-uniform macro wear scar.
After falling into the position between the rubber wheel surface and the alloy pre-wear surface from sand
nozzle, abrasives rebounded and/or collided to erode the pre-wear surface firstly and produce the impact-
erosion region (see figure 8). Then the abrasives were driven by the rubber wheel to abrade the pre-wear surface
and produce Region A including DW. As shown in figure 20(b), after producing sufficient abrasion to leave DW,
some abrasives driven by rubber wheel gained a momentum along the direction I and tended to rush to the
rubber wheel surface or run away. Meanwhile, due to the deviation from the position where the normal load
directly acted (i.e. the immediate-contact position between rubber wheel and alloy surface), when moving in
direction I the abrasives performed a gradually-weakened normal stress and the dominant tangential stress on
the alloy surface to generate Region B. After hitting the rubber wheel surface, a small portion of the abrasives
rebounded in direction II to run away or rush to the specimen surface to generate a second impact-erosion
region, i.e. the Region C. And finally abrasives rebounded again in direction III and ran away.
Figure 21 reveals the 3D-morphologies and SEM images of the impact-erosion region and the Region C obtained
at 190 N/160 rpm. No difference can be observed between two 3D-morphologies. Slightly eroded by abrasives, the
impact-erosion region and Region C all show a smooth and flatter morphology with lots of extremely shallow pits,
which was totally different from the DW and the Region B. However, as shown in the SEM morphologies of
figure 21, it can be seen that the gouging action caused by the impact of abrasives left a large number of black craters

14
Mater. Res. Express 6 (2019) 026535 R Yu et al

Figure 21. 3D-morphology and SEM images of impact-erosion region and Region C of wear scar obtained at 190 N/160 rpm:
(a) impact-erosion region, (b) Region C.

[51, 52]. In SEM morphology of the impact-erosion region (see figure 21(a)), a large amount of white NbC well
bonded to the matrix can be seen, and the black pits have no obvious directivity. This is because the abrasives were
extremely loose when they just fell between rubber wheel/alloy specimen and were easy to lose at the both sides of the
rubber wheel, which led to a relatively insufficient impact-erosion. Differently, the directional black pits parallel to
wear direction and the severe spalling of NbC communities can be seen from the SEM morphology of Region C (see
figure 21(b)), which can be attributed to the relatively concentrated and sufficient impact-erosion action occurred at
Region C. Abrasion degree at the middle and the marginal position of rubber surface is inconsistent. As shown in
figure 20(c), abrasives can be better constrained at the middle position to produce relatively fully abrasion with more
normal stress. However, near the edge of the rubber wheel, loose abrasive grits were less constrained and easily lost
performing unidirectional-tangential-stress dominated abrasion on the alloy pre-wear surface. Therefore, Region B
preferentially formed at the both long sides of macro wear scar and tended to connect to the middle position of the
wear scar, exhibiting a symmetrical shape (see figure 8). And for Region C, well constrained abrasives bundle also
tended to perform a relatively concentrated and sufficient impact-erosion action on the middle position of Region C
when moving in direction I and rebounding in direction II (see figure 20(b)).
Under the 70 N normal load, the large-sized abrasives not significantly crushed had always shown strong
destructive potential to alloy and could uniformly abrade the alloy surface along the rubber wheel trajectory
leaving a uniform wear scar macroscopically (see figures 8(a)–(c) and 9(b)). Only when abrasives got a
sufficiently large velocity moving in direction I and produced a sufficiently small normal stress, i.e. when
abrasives were about to leave from the wear scar concave in direction I at 220 rpm condition, could the abrasives
performed a cutting/ploughing-dominated action on alloy surface in a small scale. Therefore, for the 70 N
normal load, a slight formation of Region B at 220 rpm can be seen macroscopically on the wear scar (see
figure 8(c) and the profile circularly framed in figure 9(b)). And the low formative tendency of weakly-worn
Region B indirectly caused the overall wear loss and the maximum wear scar depth to constantly and
substantially increase with the rotational speed (see figures 6 and 10).
Under the 190 N normal load, abrasives subjected to high normal load tended to perform deep cutting and
strong pressing on DW at the early stages of abrasion process, resulting in severe deformation and materials
removal. So the wear loss, the maximum wear scar depth and the work-hardening degree of 190 N were all
much greater than that of 70 N. The increasing rotational speed under the 190 N load also intensified the

15
Mater. Res. Express 6 (2019) 026535 R Yu et al

work-hardening degree at DW and the fragmentation of abrasive medias, which made the maximum wear depth
show a slow growth trend as the rotational speed increased. On the other hand, although higher rotational speed
was beneficial to enhance the tangential cutting/ploughing action of the abrasives as they moved in direction I,
NbC and eutectic-carbides showed well stability and wear-resistance to small-sized crushed abrasives. It can be
seen from figure 15 that with respect to 70 N/220 rpm, at 190 N/1 60 rpm and 190 N/220 rpm the cutting/
ploughing action performed by small-sized abrasive grits was relatively weak to leave fewer and slighter grooves,
and NbC was better combined with the matrix showing less spalling tendency. Abrasives crushed significantly at
DW could produce cutting/ploughing-dominated abrasive wear earlier to leave weakly-worn Region B on the
wear scar. And as the rotational speed increased, smaller-sized abrasives tended to leave a larger area of Region B
(see figures 8(d)–(f)) and effectively impede the further growth in wear loss. Therefore, under the normal load of
190 N, with the increase in rotational speed, the relatively weak abrasion at DW and the area increase of weakly-
worn Region B together slowed the growth of the overall wear loss severely (see figures 6 and 10).

4. Conclusion

• The increase in normal load and rubber wheel rotational speed deteriorated the uniformity of macro wear
scar. The change trend of wear loss was mainly dominated by the wear behavior of two regions on the wear
scar, i.e. a fully-worn region and a weakly-worn region.
• When the abrasive medias driven by rubber wheel were gradually fed into the immediate-contact position
between rubber wheel and alloy surface, the deep cutting and strong pressing produced by abrasives caused
the fully-worn region to be left on the alloy surface.
• Afterwards, abrasives continued to be driven by rubber wheel away from the immediate-contact position.
And when abrasives were significantly crushed or had a high tangential moving velocity, they tended to leave
the cutting/ploughing-dominated weakly-worn region on alloy surface.
• The wear mechanism of the fully-worn region was mainly the austenitic matrix cutting and, the spalling of
NbC and eutectic-carbides. Strong normal stress of abrasives and the severe alloy deformation resulted in the
work-hardening of fully-worn region.
• Due to the weak normal stress action and pronounced tangential stress action produced by the abrasives, the
wear mechanism of the weakly-worn region was mainly the straight cutting/ploughing. No work-hardening
occurred at weakly-worn region. The formation of weakly-worn region impeded the growth of overall wear loss.
• Under the 70 N normal load, the abrasive grits had always shown strong destructive potential, which resulted
in a relatively uniform macro wear scar and a less tendency to form weakly-worn region. The increasing
rotational speed intensified the abrasives cutting action and the NbC spalling tendency to constantly and
significantly grow the overall wear loss.
• Under the 190 N normal load, the work-hardening degree of fully-worn region was extremely significant,
which greatly improved the wear-resistance of austenitic matrix and caused the severe fragmentation of
abrasives. As the rotational speed increased, the gradual weak abrasion of fully-worn region and the large area
increase of weakly-worn region together led to a wear loss trend of increasing first and then decreasing.

Acknowledgments

The authors are grateful to the China National Training Program of Innovation and Entrepreneurship for
Undergraduates (No.2017cxcy213) and Major Scientific and Technological Innovation Project of Zhengzhou
(No.188PCXZX784) for providing financial support.

ORCID iDs

Shengxin Liu https://orcid.org/0000-0002-4071-3174


Zhiquan Huang https://orcid.org/0000-0002-7426-9921

References
[1] Chang C M, Lin C M, Hsieh C C, Chen J H and Wu W 2009 Micro-structural characteristics of Fe-40 wt%Cr-xC hardfacing alloys with
[1.0-4.0 wt%] carbon content J Alloy Compd 487 83–9

16
Mater. Res. Express 6 (2019) 026535 R Yu et al

[2] Liu D S, Wei Y H and Liu R P 2015 Microstructure and wear mechanism change by Nb added in slag free self-shielded flux cored wire
Sci Technol Weld Joi 20 693–701
[3] Filippov M A, Shumyakov V I, Balin S A, Zhilin A S, Lehchilo V V and Rimer G A 2015 Structure and wear resistance of deposited alloys
based on metastable chromium-carbon austenite Weld. Int. 29 819–22
[4] Li D R, Liu Z J and Su Y H 2018 Effect of TiN on microstructure and wear resistance of Fe–Cr–C hardfacing alloy: experimental
research and first-principles calculation Mater. Res. Express 5 P116513
[5] Chen J H et al 2013 Microstructure and abrasive wear properties of Fe–Cr–C Hardfacing alloy cladding manufactured by gas tungsten
arc welding (GTAW) Met Mater Int 19 93–8
[6] Filipovic M, Kamberovic Z, Korac M and Gavrilovski M 2013 Microstructure and mechanical properties of Fe–Cr–C–Nb white cast
irons Mater Design 47 41–8
[7] Sabet H et al 2011 The microstructure and abrasive wear resistance of Fe–Cr–C hardfacing alloys with the composition of hypoeutectic,
eutectic, and hypereutectic at Cr/C=6 Tribol. Lett. 44 237–45
[8] Xu L X, Yu Z W, Ma Y Q, Wang X and Shi Y Q 2002 Martensitic transformation and work-hardening of metastable austenite induced
by abrasion in austenitic Fe–C–Cr–Mn–B Alloy-a TEM study Mater Sci Tech-Lond 18 1561–4
[9] Zhao Z C, Su Y H and Huang H J 2015 Research on effect of chromium-carbon compounds on abrasion resistance of Fe–Cr–C
surfacing layer Hot Working Technology 44 60–2
[10] Yang J, Huang J, Fan D and Chen S 2015 Microstructure and wear properties of Fe-6 wt.%Cr-0.55 wt.%C-Xwt.%Nb laser cladding
coating and the mechanism analysis Mater Design 88 1031–41
[11] Bedolla-Jacuinde A 2001 Microstructure of vanadium-, niobium- and titaniumalloyed high-chromium white cast iron Cast Metal 13
343–61
[12] Lv Y Z, Sun Y F, Zhao J Y, Yu G W, Shen J J and Hu S M 2012 Effect of tungsten on microstructure and properties of high chromium cast
iron Mater Design 39 303–8
[13] Correa E O, Alcantara N G, Valeriano L C, Barbedo N D and Chaves R R 2015 The effect of microstructure on abrasive wear of a
Fe–Cr–C–Nb hardfacing alloy deposited by the open arc welding process Surf Coat Tech 276 479–84
[14] Zhang L M, Sun D B and Yu H Y 2008 Effect of niobium on the microstructure and wear resistance of iron-based alloy coating
produced by plasma cladding Mat Sci Eng A-Struct 490 57–61
[15] Chung R J, Tang X, Li D Y, Hinckley B and Dolman K 2013 Microstructure refinement of hypereutectic high Cr cast irons using hard
carbide-forming elements for improved wear resistance Wear 301 695–706
[16] Stevenson A N J and Hutchings I M 1996 Development of the dry sand/rubber wheel abrasion test Wear 195 232–40
[17] Yang Q 2015 Study on Fatigue Resistance of Hardfacing Layer Metals of Roller Press Jilin University
[18] Zuo Y T 2013 Study on the Bonding Strength between Hardfacing Alloys Used for Repairing Extrusion Rollers and 45 Steel Parent Metal Jilin
University
[19] ASTM. G65-04 Standard test method for measuring abrasion resistance using the dry sand/rubber wheel apparaturs Book of Standards
(Ohio: ASTM International) (https://doi.org/10.1520/G0065-04)
[20] Ma X, Liu R and Li D Y 2000 Abrasive wear behavior of D2 tool steel with respect to load and sliding speed under dry sandrrubber wheel
abrasion condition Wear 241 79–85
[21] Vostrak M, Houdkova S and Bystriansky M 2018 The influence of process parameters on structure and abrasive wear resistance of laser
clad WC-NiCrBSi coatings Mater. Res. Express 5 P096522
[22] Petrica M, Katsich C, Badisch E and Kremsner F 2013 Study of abrasive wear phenomena in dry and slurry 3-body conditions Tribol.
Int. 64 196–203
[23] Kirchgaßner M and Badisch E 2008 Behaviour of iron-based hardfacing alloys under abrasion and impact Wear 265 772–9
[24] Berns H and Fischer A 1987 Microstructure of Fe-Cr-C hardfacing alloys with additions of Nb, Ti and, B Metallography 40 401–29
[25] Cui Z Q and Liu B X 2007 Metallurgy and Heat Treatment Theory 3rd edn (Harbin: Harbin Institute of Technology Press)
[26] Zhang W Y 1999 Welding Metallurgy 1st edn (Beijing: Mechanical Industry Press)
[27] Cai Q Z and Wu S S 2010 Casting and Melting of Alloys 1st edn (Beijing: Chemical Industry Press)
[28] Doğan Ö N 1997 Solidification structure and abrasion resistance of high chromium white irons Metall Mater Trans A 28 1315–28
[29] Kesri R and Durand-Charre M 1987 Phase equilibria, solidification and solid-state transformations of white cast irons containing
niobium J. Mater. Sci. 22 2959–64
[30] Muthaiah V M S, Koch C C and Mula S 2018 Effect of Nb addition on Fe–7Cr–Nb and Fe–15Cr–Nb metastable alloy formation and
their thermal stability Mater. Res. Express 5 P056534
[31] Tabrett C P, Sare I R and Ghomashchi M R 1996 Microstructure-property relationships in high chromium white iron alloys Int. Mater.
Rev. 41 59–82
[32] Gahr K H Z and Eldis G T 1980 Abrasive wear of white cast irons Wear 64 175–94
[33] KenchiReddy K M and Jayadeva C T 2004 The effect of microstructure on 3 body abrasive wear behavior of hardfacing alloy Bonfring
International Journal of Industrial Engineering and Management Science 4 14–23
[34] KenchiReddy K M and Jayadeva C T 2016 The effects of welding processes and microstructure on 3 body abrasive wear resistances for
hardfacing deposit Transactions on Engineering Technologies 155–67
[35] Lim C Y H, Leo D K, Ang J J S and Gupta M 2005 Wear of magnesium composites reinforced with nano-sized alumina particulates
Wear 259 620–5
[36] Hassan S, Al-Qutub A, Zabiullah S, Tun K and Gupta M 2016 Effect of increasingly metallized hybrid reinforcement on the wear
mechanisms of magnesium nanocomposite B Mater Sci 39 1101–7
[37] Choteborsky R et al 2008 Abrasive wear of high chromium Fe–Cr–C hardfacing alloys Res. Agr. Eng 54 192–8
[38] Buchanan V E, Mccartney D G and Shipway P H 2008 A comparison of the abrasive wear behavior of iron-chromium based hardfaced
coatings deposited by SMAW and electric arc spraying Wear 264 542–9
[39] Ma X R 2009 Determination of Fe, Al, Ca, Ti, B, P in quartz sand by ICP-AES Chem. Eng. 1 19–20
[40] Zhang S F, Qin M and Qv L Q 2002 The simultaneous determination of iron (III) and aluminium(III) in quartz sand by first derivative
spectrophotometry Spectroscopy and Spectral Analysis 22 113–5
[41] Zhao H, Barber G C and Liu J 2001 Friction and wear in high speed sliding with and without electrical current Wear 249 409–14
[42] Correa E O, Alcantara N G, Tecco D G and Kumar R V 2007 The relationship between the microstructure and abrasive resistance of a
hardfacing alloy in the Fe–Cr–C-Nb-V syst2em Metall Mater Trans A 38 1671–80
[43] Wang Z T and Meng S J 2013 Friction & Wear & Wear-resistant Materials 1st edn (Harbin: Harbin Institute of Technology Press)
[44] Park M C et al 2014 Damage mechanism of cavitation erosion in austenite-martensite phase transformable Fe–Cr–C-Mn/Ni alloys
Wear 310 27–32

17
Mater. Res. Express 6 (2019) 026535 R Yu et al

[45] Wu Y J 2011 Metal Material Science. 1st edn (Beijing: Peking University Press)
[46] Murr L E, Staudhammer K P and Hecker S S 1982 Effects of strain state and strain rate on deformation-induced transformation in 304
stainless steel: part II, micro-struct study Metal Trans A 13 627–35
[47] Park M et al 2013 Effects of Ni and Mn on the cavitation erosion resistance of Fe–Cr–C–Ni/Mn austenitic alloys Tribol. Lett. 52 477–84
[48] Ramalho A and Miranda J C 2006 The relationship between wear and dissipated energy in sliding systems Wear 260 361–7
[49] Petrica M, Katsich C and Badisch E 2013 Study of abrasive wear phenomena in dry and slurry 3-body conditions Tribol. Int. 64 196–203
[50] Katsich C and Badisch E 2011 Effect of carbide degradation in a Ni-based hardfacing under abrasive and combined impact/abrasive
conditions Surf Coat Tech 206 1062–8
[51] Surzhenkov A et al 2017 Wear resistance and mechanisms of composite hardfacings at abrasive impact erosion wear Journal of Physics
843 P012060
[52] Islam M A and Farhat Z N 2014 Effect of impact angle and velocity on erosion of API X42 pipeline steel under high abrasive feed rate
Wear 311 180–90

18

Вам также может понравиться