You are on page 1of 9

Wear 386-387 (2017) 230–238

Contents lists available at ScienceDirect

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

Slurry erosion behavior of high entropy alloys


R.B. Nair a, Karthikeyan Selvam a, H.S. Arora a, Sundeep Mukherjee b, H. Singh c, H.S. Grewal a,n
a
Department of Mechanical Engineering, Shiv Nadar University, Gautam Budh Nagar, India
b
Department of Materials Science and Engineering, University of North Texas, Denton, TX 76203,USA
c
Department of Mechanical Engineering, Indian Institute of Technology Ropar, Rupnagar, India

art ic l e i nf o a b s t r a c t

Article history: High entropy alloys (HEAs) represent a new paradigm of structural alloys comprising of multiple prin-
Received 8 September 2016 cipal elements in equimolar or near-equimolar concentration. The superior corrosion and oxidation re-
Received in revised form sistance of HEAs at high temperature make them attractive for several structural applications. In this
30 December 2016
context, erosion behavior of HEAs has been largely unexplored. In this study, the slurry erosion perfor-
Accepted 5 January 2017
mance of single phase Al0.1CoCrFeNi high entropy alloy was investigated. For comparison, mild steel and
stainless steel (SS316L) were also investigated under similar conditions. The slurry erosion was con-
Keywords: ducted at different impingement angles and at a constant velocity of 20 m/s. The microstructural and
High entropy alloy mechanical characterization were conducted using scanning electron microscope (SEM), energy dis-
Slurry erosion
persive spectroscopy (EDS), x-ray diffraction (XRD), nanoindentation and micro-hardness testing. Similar
Work hardening
to the mild steel and stainless steel, HEA also showed ductile mode of erosion. The erosion rate for HEA
Structural-property correlation
was found to be higher compared to stainless steel, however in spite of lower tensile strength and
hardness, HEA exhibited higher erosion resistance compared to mild steel. The high erosion resistance of
HEA compared to mild steel is explained on the basis of its work hardening behavior, low stacking fault
energy, and superior corrosion resistance. Erosion response of the investigated materials showed sig-
nificant correlation with ultimate strength and ultimate resilience. In depth analysis of the eroded HEA
samples showed ploughing as the prominent material removal mechanism at oblique angles compared
to micro-cutting for SS316L and mild steel. In contrast, highly deformed and work-hardened platelets
were observed at normal impingement angle for all materials.
& 2017 Elsevier B.V. All rights reserved.

1. Introduction entropy is a dominant factor which influences the mechanical and


tribological properties of HEAs [3]. The high mixing entropy in a
Materials used in fluid machineries such as pumps, turbines, multi-principal element system promotes the thermodynamically
propellers and valves typically show surface degradation in the stable solid solution, which otherwise results in a brittle inter-
form of erosion and result in significant economic impact [1]. metallic phase [4–6]. The high mixing entropy is also responsible
State-of-the-art materials used in these applications show limited for exceptional high temperature properties of HEAs [6]. The
resistance against wear, erosion, and corrosion. Therefore, ad- structure of HEAs is mainly composed of simple face centred cubic
vanced materials with superior surface properties are needed to (FCC), body centred cubic (BCC) and/or their combination, despite
counter these forms of degradation. Recently developed high en- being accompanied by elements with different structures. The
tropy alloys (HEAs) with exceptional properties have shown pro- properties of HEAs are influenced and regulated by interactions
mising results [2,3]. HEAs contain at least four to five principal among the constituent elements. Tung et al. [7] observed that
elements mixed in equi-molar or nearly equi-molar fraction. The increasing Al and Cr contents in AlCoCrCuFeNi, results in trans-
exceptional properties of HEAs are mainly attributed to four core formation from FCC to BCC phase, whereas, Co, Fe and Ni favours
effects: high configurational entropy, sluggish diffusion, lattice FCC. On the other hand, an increase in Cu content leads to the
distortion and cock tail effect [2–4]. HEAs may also contain minor formation of Cu rich inter-dendrite structure due to the low mix-
elements with atomic fraction less than 5 at%. The high mixing ing enthalpy with other constituent elements. Tong et al. [8] re-
ported increased hardness and elastic modulus of AlxCoCrCuFeNi
n HEA being related with the strong binding energy and large atomic
Corresponding author.
E-mail addresses: harpreetsg@iitrpr.ac.in, size of Al. Wang et al. [9] reported an increase in hardness of
harpreet.grewal@snu.edu.in (H.S. Grewal). AlxCrCoFeNi with the addition of Al due to transformation of

http://dx.doi.org/10.1016/j.wear.2017.01.020
0043-1648/& 2017 Elsevier B.V. All rights reserved.
R.B. Nair et al. / Wear 386-387 (2017) 230–238 231

structure from FCC to BCC. The valence electron concentration Samples of HEA, MS and SS of size 10  10  5 mm3 were pre-
(VEC) has also been observed to influence transition in the lattice pared from ingots using conventional machining processes fol-
structure [6]. lowed by cutting on wire electric discharge machining (wire EDM).
Although limited, there are studies that have explored the wear All samples were polished and grounded using abrasive papers
and corrosion behavior of HEAs. Wu et al. [10] reported increase in down to 2000 grit size. Microstructural characterization of the
wear resistance of AlxCoCrCuFeNi HEA with the addition of Al samples was performed using electron backscatter diffraction
(x ¼3.0) mainly due to increased hardness. In addition, the wear (EBSD) and x-ray diffraction (XRD). Mechanical properties of the
mechanism changes from oxidation controlled to predominantly samples were measured using micro and nanoindentation. Elastic
delamination wear. Chuang et al. [11] observed that modulus of the test samples was calculated from the slope of
Co1.5CrFeNi1.5Ti and Al0.2Co1.5CrFeNi1.5Ti exhibited excellent wear unloading curve using Oliver-Pharr method. To avoid hardness and
resistance compared to conventional alloys due to higher re- elastic modulus measurement from the single grain (especially for
sistance to oxidation and lower high temperature softening. For HEA), micro and nano indentations were performed on two dif-
AlCoCrFexMo0.5Ni and AlCoCrCuFeNi, the wear resistance was re- ferent samples for each material, with at least 10 indentation
markably higher due to the formation of harder sigma (s) phase as performed on each sample with 1 mm spacing. Sand particles in
a result of replacing Cu with Mo [10,12]. With further increase in the size range of 75–150 mm were used for the slurry erosion
Fe (x 41.0), the hardness and wear resistance decreases mainly studies (Fig. 1).
due to decrease in s phase and higher oxidation rate [12]. Hsu
et al. [13] reported an increase in wear resistance and high tem- 2.2. Slurry erosion test rig
perature compressive strength of Al0.5CoCrCuFeNiBx with the ad-
dition of boron content mainly due to the formation of borides. Slurry erosion testing was performed using re-circulation type
HEAs also showed high corrosion and pitting resistance com- test rig shown in Fig. 2. The test rig comprised of a diaphragm
pared to 304 stainless steel [14,15]. Zhao et al. [16] reported im- pump driven by compressed air. The premixed slurry in a con-
proved tribo-corrosion behaviour for annealed AlCoCrCuFeNi at tainer is pumped using this diaphragm pump and made to im-
high temperature compared to 304L stainless steel. Low mixing pinge on a sample through 2 mm diameter tungsten carbide
enthalpy of Cu with constituent elements in AlxCoCrCuFeNi lead- nozzle. The velocity of slurry jet is controlled by changing the
ing to formation of Cu-rich interdendritic region resulted in in- pressure of the compressed air used for driving the pump. The test
creased wear rates and high corrosion rates due to galvanic action rig provides the flexibility to perform experiments at a range of
[8,10]. Replacing Cu with Mn in AlxCrFe1.5MnNi0.5, Lee et al. [17] different parameters such as impingement angle, particle size,
reported an increase in corrosion and pitting resistance. Further stand-off distance, working media and impact velocity. The sedi-
the corrosion and pitting resistance also increased with a decrease mentation of sand particles was prevented by continuous stirring.
in Al content. Lin et al. [18] showed that segregation of Al-Ni rich
phase leads to high hardness with nominal corrosion resistance of 2.3. Slurry erosion testing
as-cast and aged Al0.5CoCrFeNi HEA compared to conventional
steel. Ji et al. [19] showed Al3CoCrFeNi coating lowers the erosion Slurry erosion testing was conducted according to ASTM G-73
rates of 17-7 stainless steel. Although there are limited number of standard procedure. The test parameters used for slurry erosion
studies on wear and corrosion behaviour of HEAs, their erosion experiments are shown in Table 2. Slurry with concentration of
behavior has been largely unexplored. 5 kg/m3 (5000 ppm) was prepared using sand (75–150 mm), mixed
In the current work, slurry erosion behavior of Al0.1CrCoFeNi with tap water (pH 7.72). For each test sample, fresh slurry was
high entropy alloy was investigated. For comparison, two con- used. Each sample was tested for four hours with a cycle time of
ventionally used structural materials, mild steel and SS316L one hour. Gravimetric analysis was performed using high precision
stainless steel were tested under identical conditions. Despite weighing balance of 0.01 mg resolution. Prior to weight mea-
lower hardness, the HEA exhibited higher erosion resistance surement, samples were cleaned with acetone and dried using air
compared to mild steel, which is primarily attributed to higher stream. Slurry erosion experiments were performed at a constant
ductility and superior corrosion resistance. velocity of 20 m/s with samples positioned at different impinge-
ment angles i.e. 30°, 60° and 90°. The eroded surfaces were ana-
lyzed using scanning electron microscope (SEM) to investigate the
2. Experimental details erosion mechanism.

2.1. Materials
3. Results and discussions
The nominal composition of Al0.1CrCoFeNi high entropy alloy
used in present work is shown in Table 1. The composition of mild 3.1. Microstructural and mechanical characterization
steel (MS) and SS316L (SS) stainless steel used for comparison, is
also shown. HEA alloy was prepared using induction melting fol- The microstructure of the Al0.1CoCrFeNi HEA, SS316L stainless
lowed by hot isostatic pressing (HIP) for densification. HIP was steel and mild steel is shown in Fig. 3. The optical micrograph of
performed at 1473 K for 4 h at a pressure of 100 MPa, followed by mild steel (Fig. 3a) indicates average grain size of around 30 μm.
cooling. The electron back scatter diffraction (EBSD) map of the

Table 1
Nominal composition at wt% of Al0.1CoCrFeNi HEA, SS 316L stainless steel and mild steel used in experimentation.

Composition (wt%) Al Fe Cr C Mn Ni Mo Co S Si P Cu

Al0.1CrCoFeNi 1.2 24.5 22.8 – 25.7 – 25.8 – – – – –


SS 316L – bal. 18 0.035 2 14 3 – 0.03 1 0.045 –
Mild steel – bal. – 0.26 1.03 – – – 0.05 0.28 – 0.20
232 R.B. Nair et al. / Wear 386-387 (2017) 230–238

hand, microstructure of SS316L steel was comprised of large


number of twins. The large grain size for HEA is mainly due to hot
isostatic processing at high temperature causing a significant grain
growth. The abnormal grain growth is primarily responsible for
lower yield strength and hardness (Table 3). The x-ray diffraction
(XRD) results for all the samples is shown in Fig. 4. XRD results for
HEA indicate a single phase solid solution with face centre cubic
(FCC) structure. It is interesting to observe that despite of large
number of constituent elements, no intermetallic phases were
observed for HEA. As expected, mild steel and SS316L steels
showed BCC and FCC structures respectively. In addition to aus-
tenite FCC structure, microstructure of SS316L was also composed
of brittle martensitic phase along with Cr23C7 precipitates. How-
ever, microstructure of HEA showed no indication of such sec-
ondary phases, which is attributed to high configurational entropy,
Fig. 1. Scanning electron microscopy image of the sand used for slurry erosion
testing.
sluggish diffusion and cocktail effect [3]. The thermodynamic
stability of solid solution phase is ensured through use of equi-
molar concentration of constituent elements. The thermodynamic
stability of the system is measured using Gibbs free energy given
as
G = H − ∆TS

where G is Gibbs free energy, H is enthalpy, ΔT is change in


temperature and S is entropy [3–6]. Therefore, by increasing the
entropy, the free energy of the system can be lowered leading to a
stabilized system. The configurational entropy of the system can
be modulated by controlling the number of constituent elements
as
n
∆SConf = − R ∑ xi ln xi = R ln n
Fig. 2. Schematic illustrating the slurry erosion test rig used for experimentation.
I=1

Table 2 where n is the number of elements with xi mole fraction and R is


Value of test parameters used for slurry erosion testing. universal gas constant. For given number of elements, the con-
figurational entropy ( ∆SConf ) is highest when elements are mixed
Parameter Value
in equi-atomic proportion. However, the configurational entropy
Temperature (°C) 25 of the alloy system used in present study (1.47R) is lower than that
Impact velocity (m/s) 20 of five element system mixed in equi-molar fraction (1.61R). Even
Nozzle diameter (mm) 2.0 though the configurational entropy is lower, HEA forms a solid
Angle of impingement (deg.) 30, 60, 90 solution with FCC structure. This highlights the role of other core
Particle size (mm) 75–150
Stand-off distance (mm) 20
factors such as lattice distortion, sluggish diffusion and cock tail
effect.
The load-displacement plots obtained using nanoindentation
for the HEA and SS316L steel are shown in Fig. 5. The elastic
Al0.1CoCrFeNi HEA (Fig. 3b) shows significantly large sized grains modulus values are shown in Table 3. Hardness for all the samples
[20]. The average grain size calculated for HEA was around measured using microhardness testing are also shown in Table 3.
2000 mm. The average grain size for SS316L stainless steel was Amongst all, SS316L steel showed highest hardness and elastic
found to be around 22 μm (Fig. 3c). Compared to the micro- modulus, followed by mild steel and HEA. Accordingly, the yield
structure of SS316L, HEA showed no twinned regions. On the other and ultimate tensile strengths were also lowest for the

Fig. 3. (a) Optical micrograph showing microstructure of mild steel. Electron backscatter diffraction maps for (b) Al0.1CoCrFeNi high entropy alloy [20] (c) SS316L stainless
steel.
R.B. Nair et al. / Wear 386-387 (2017) 230–238 233

Table 3
Mechanical properties of mild steel (MS), SS316L stainless steel (SS) and Al0.1CoCrFeNi high entropy alloy.

Material Density, ρ (kg/m3) Hardness H (Hv) Elastic modulus E (GPa) Yield strength ry (MPa) Ultimate Strength ru (MPa) True failure strain εf

HEA 7450 150 186 160 389 0.58


SS316L 8000 226 210 290 580 0.35
Mild steel 7800 205 200 250 410 0.25

Fig. 4. X ray diffraction (XRD) plots of Al0.1CrCoFeNi high entropy alloy, mild steel
and SS316L stainless steel. Fig. 6. The cumulative volume loss/hour as the function of impact angle.

erosion mode [22,23]. It is interesting to observe that at oblique


angles, despite of lowest hardness amongst all the tested materi-
als, the erosion rate for Al0.1CoCrFeNi HEA were similar to mild
steel. At normal impingement, the HEA performed better than
mild steel, with 20% lower erosion rate. The lowest erosion rates
were observed for SS316L stainless steel which is attributed to its
high hardness and tensile strength. For oblique impacts, hardness
is a dominant mechanical property that modulates the erosion
rates for a given class of materials [24]. Therefore, lowest erosion
rates for SS316L steel is highly likely. However, the similar argu-
ment is implausible in case of HEA compared with mild steel. Even
though, the HEA had around two-third the hardness of mild steel,
the erosion rates were remarkably similar at oblique impingement
angle.
For normal impingement, HEA showed significantly lower
Fig. 5. Load displacement plots for Al0.1CoCrFeNi high entropy alloy and SS316L erosion rates compared to mild steel sample. The reduced erosion
stainless steel. rates for HEA can be attributed to its high ductility and work
hardening capability compared with mild steel. Contribution of
Al0.1CrCoFeNi HEA which can be attributed to its large grain size ductility towards increased erosion resistance has been reported
[20,21]. In contrast, failure strain required for fracture is highest previously [25,26]. High ductility increases the critical strain for
for HEA. The ratio of hardness to elastic modulus, related to failure, εf as a result of which, number of impacts required for
plasticity index is lowest for HEA. This indicates early onset of removal of the material would also increases, contributing towards
plastic deformation for HEA compared to steels. However, con- an increase in erosion resistance [27,28]. Further, at normal im-
siderable increase in hardness following plastic deformation is pingement, plastic deformation is primarily the governing phe-
expected considering exceptionally high work hardening rate of nomenon operational for ductile materials [29]. The work hard-
Al0.1CoCrFeNi HEA [20]. ening is related with the stacking fault energy (SFE) of the material
[30]. A decrease in SFE results in increase in work hardening
3.2. Slurry erosion studies capability of the material, thereby lowering the material removal
rate. Zhang and Fang [31] reported a direct correlation between
The results of slurry erosion tests are shown in Fig. 6 as a cavitation erosion resistance and SFE. Using DFT calculation and
function of impingement angle. The measured mass loss rates XRD analysis, Zaddach et al. [32] reported SFE for CoCrFeNi alloy to
were converted to corresponding volume loss (erosion rate) using be as low as 17 mJ/m2. Whereas, for mild steel and SS316L stainless
the density values given in Table 3. The reported values are aver- steel, SFE is of the order of 70–80 mJ/m2 [33]. Further, discussion
age steady state erosion rates. All the materials showed higher on the correlation with different parameters is given in the sub-
erosion rates at oblique angle. The erosion rates decreased with an sequent section (Section 3.4).
increase in impingement angle which indicates ductile mode of
erosion for Al0.1CoCrFeNi HEA, as well as for the other two con- 3.3. Erosion mechanism
ventional materials. Compared to ductile erosion mode of single
phase HEA, single phase amorphous metallic glass showed brittle The scanning electron microscopy (SEM) images illustrating
234 R.B. Nair et al. / Wear 386-387 (2017) 230–238

Fig. 7. The low magnification scanning electron microscopy images showing surface morphology of (a), (c), and (e) stainless steel SS316L and (b), (d), and (f) Al0.1CoCrFeNi
high entropy alloy after slurry erosion testing at (a) and (b) 30°, (c) and (d) 60°, and (e) and (f) 90° impingement angles. For convenience impingement angles are marked on
the images.

surface morphology of the eroded samples is shown in Fig. 7. irrespective of the impingement angle [34]. The high ductility
Detailed examination of these images indicates that micro-cutting assists the stressed platelets to undergo larger degree of elonga-
and ploughing are the two primary erosion mechanisms re- tion, compared to material with low ductility. As a result large
sponsible for the loss of the materials at oblique angle. Severe number of stressed and elongated platelets can be observed cov-
micro-cutting and ploughing marks were observed for the SS316L ering the surface of HEA eroded at 30° (Fig. 7b). Contrarily, the
and HEA samples, respectively. Micro-cutting was the dominant eroded surface of SS316L steel was primarily obscured with micro-
mechanism contributing to the removal of the material for the cutting marks.
SS316L steel (Fig. 7a). However, for HEA, ploughing was observed With the further increase in impingement angle to 60°, both
to be the prevalent material removal mechanism (Fig. 7b). Nu- micro-cutting and ploughing mechanisms were observed to be
merous raised lips and distressed platelets were observed for HEA operational for SS316L steel (Fig. 7c). However, for HEA, the for-
compared to SS316L steel. Few micro-cutting and micro indenta- mation and removal of platelets was the most preferential mode of
tion marks were also observed for HEA. The difference in primary material removal (Fig. 7d). Some micro-cutting marks were also
mechanism at 30° for HEA and SS may be related to their differ- observed however, ductility controlled platelet mode of material
ence in ductility. Although, the hardness of HEA (150 HV) was removal was still observed to be the primary material removal
moderately lower that of SS316L (226 HV), the ductility of the mechanism for HEA. Material removed through micro-cutting can
former ( εf = 0.58) was more than thrice to that of the latter be observed to attach with the surface in form of a platelet
( εf = 0.16) (Table 3). For highly ductile materials, the primary (Fig. 8a). Further, straining of this deformed platelet can also be
mode of material disintegration is platelet mechanism, observed. The formation of platelet was mainly through ploughing
R.B. Nair et al. / Wear 386-387 (2017) 230–238 235

Fig. 8. The high magnification scanning electron microscopy image showing surface morphology of Al0.1CoCrFeNi high entropy alloy after slurry erosion testing at (a) and
(b) 90°, (c) 60° impingement angle (d) work hardening rate for different materials.

and indentation mechanism. The extruded lips formed around the mechanical properties was investigated. This will help in determin-
ploughing craters and indents would undergo further straining ing critical properties controlling the erosion behavior of all, parti-
until they are finally fractured. cularly high entropy alloys. Fig. 9(a)–(e) illustrate erosion rate as a
For samples tested at normal impingement angle the erosion function of different mechanical properties. The correlation para-
mechanism was almost identical both for HEA and SS316L steel meter, adj. R2 shown in Fig. 9 are for the least square fitted lines
(Fig. 7e and f). Due to normal impact of sand particles the material relating the erosion rates at 30°. For 90° impingement angle, neither
extruded from the craters accumulated around the periphery in of the explored parameter showed good correlation with erosion
form of platelets. With subsequent impacts of the sand particles, rate. In contradiction to the literature [35,36], the erosion rates for
the accumulated materials was further extruded and strained. oblique angle (30°) showed no correlation with hardness. In general,
Magnified view in Fig. 8b and c shows highly deformed layer for a same class of materials, erosion rates is expected to show direct
formed with extruded and superimposed platelets. Highly de- correlation with ratio of particulate to target hardness (hp/ht) [36].
formed surface of HEA and SS316L steel at 90° impingement angle However, it is to be noted that all three investigated materials re-
indicates significant work hardening. Due to lower SFE, HEA has presents different class of materials with unique characteristics.
comparatively higher work-hardenability than SS316L steel. Stainless steel SS316L is a high alloy steel with high corrosion re-
However, the low ultimate strength of the HEA (389 MPa) com- sistance, and significantly different microstructure compared with
pared with SS316L steel (580 MPa) limits the erosion resistance of mild steel. Further, high entropy alloys also possesses unique prop-
the former. It is to be noted that HEA has significantly large grain erties such as high temperature strength, and high corrosion re-
size of around 2000 μm compared with 22 μm for SS316L steel. sistance with constituent elements in solid solution state. Hence, a
The large grain size of HEA was responsible for its low strength. direct correlation of erosion rate with hardness was likely
Besides ploughing and platelet mechanisms playing a domi- unexpected.
nant role in erosion of HEA, micro indentation was also observed For erosion at oblique angle (30°), three parameters, ultimate
to contribute significantly at all impingement angles. Compared to strength, ultimate resilience, UR and modified ultimate resilience,
SS316L, micro-indentation was highly profound for HEA which MUR showed good correlation with erosion rate (adj. R2 490%).
may be due to its lower hardness and yield strength. Fig. 8c shows Ultimate resilience showed highest correlation followed by ulti-
magnified view of one such micro indent with highly distressed mate strength and MUR. It is to be noted that both UR and MUR,
and raised platelets around its periphery. These platelets would are basically a function of ultimate strength. The ultimate resi-
essentially be disintegrated from the surface with subsequent lience, given as
impacts of the abrasive particles. 2
σut
UR =
2E
3.4. Correlation parameters
where σut is ultimate strength of the material and E is elastic
To understand the structure-property correlation of Al0.1CrCoFeNi modulus [37]. During discontinuous plastic deformation accom-
HEA, relationship between erosion resistance with different panied by considerable work-hardening, the maximum yield
236 R.B. Nair et al. / Wear 386-387 (2017) 230–238

Fig. 9. The correlation of erosion rate of Al0.1CoCrFeNi high entropy alloy (HEA), mild steel (MS), and SS316L stainless steel (SS) representing different class of materials with
(a) hardness, (b) ratio of hardness to elastic modulus (H/E) (c) ultimate resilience (UR) (d) modified ultimate resilience (e) ultimate strength.

strength expected for the deformed material would be equivalent σut H


MUR =
to ultimate strength [37]. Therefore, ultimate resilience (maximum 2E
impact energy) can predict the energy required to detach the de- Significant correlation of erosion rate with ultimate strength or
formed material from surface. Previous studies [38,39] showed related parameters for different class of materials at oblique angle,
strong correlation of ultimate resilience with erosion resistance in is related with prevalent erosion mechanisms. For both micro-
case of cavitation and liquid impingement erosion studies. Further, cutting and ploughing phenomena, removal of material takes
Rao et al. [40] proposed a modified form of ultimate resilience place through shearing process. Thus, ultimate shear strength
given as which essentially is related to ultimate tensile strength is likely to
R.B. Nair et al. / Wear 386-387 (2017) 230–238 237

4. Conclusion

In the present work, the slurry erosion behaviour of


Al0.1CoCrFeNi HEA was investigated and compared with conven-
tional structural materials, mild steel and SS316L stainless steel.
The results showed that despite of lowest hardness (two-third of
that of mild steel) due to macroscale grain size (  2 mm), HEA
showed similar erosion rate at oblique angles. At normal im-
pingement angle, HEA showed higher erosion resistance compared
to mild steel. Although SS316L stainless steel showed remarkably
lower erosion rates, however, with refined microstructure, erosion
resistance of Al0.1CoCrFeNi HEA is expected to improve. At oblique
impingement angles, ploughing was observed to be the dominant
erosion mechanism for HEA compared to micro-cutting being
primary material removal phenomenon for mild steel and SS316L
stainless steel. At normal impingement angle, the dominant ero-
sion mechanism for all materials was formation and removal of
Fig. 10. The electrochemical potential test on Al0.1CoCrFeNi high entropy alloy
(HEA), mild steel (MS), and stainless steel SS316L (SS).
platelets followed by severe plastic deformation. Ultimate resi-
lience showed significantly high correlation with the erosion re-
sistance of the investigated materials. Correlation with different
play a dominant role, especially, when comparing different class of parameter was satisfactorily high at oblique angle with less than
materials. Further, a good correlation with UR and MUR indicate 5% tolerance range, whereas, tolerance range increased to 20% at
erosion behavior of a material related with the maximum energy a normal impingement angle. The deviation from expected trend
material can absorb before fracture and/or work-hardening rate. At was attributed to high work hardenability of Al0.1CoCrFeNi HEA
normal impingement angle, the erosion behavior showed lack of and low corrosion resistance of mild steel.
correlation with any of the explored parameter. Although, UR and
MUR showed nominal correlation with erosion rate, the tolerance
range for both parameters is considerably higher. The lower ero- References
sion rate for HEA compared to mild steel at normal impingement
angle is related with the high work-hardening rate of the former [1] S. Lathabai, D.C. Pender, Microstructural influence in slurry erosion of cera-
(Fig. 8d). The dominant material removal mechanism at normal mics, Wear 189 (1995) 122–135.
[2] J.-W. Yeh, Recent progress in high-entropy alloys, Ann. Chim. Sci. Matér. 31
impingement angle was observed to be platelet phenomenon,
(2006) 633–648.
wherein material is deformed plastically until fractured. The high [3] M.-H. Tsai, J.-W. Yeh, High-entropy alloys: a critical review, Mater. Res. Lett. 2
work hardening rate enhances its energy absorption capacity (2014) 107–123.
(toughness/resilience), and hence provides better resistance to [4] F. Otto, Y. Yang, H. Bei, E.P. George, Relative effects of enthalpy and entropy on
the phase stability of equiatomic high-entropy alloys, Acta Mater. 61 (2013)
degradation through erosion. However, it is interesting to observe 2628–2638.
that even though, both UR and MUR, showed reasonable correla- [5] J.-W. Yeh, Alloy design strategies and future trends in high-entropy alloys, JOM
tion with erosion rates, the tolerance range were unexpectedly 65 (2013) 1759–1771.
[6] M.-H. Tsai, Physical properties of high entropy alloys, Entropy 15 (2013)
higher. The high material removal rates for mild steel, might be 5338–5345.
related with its corrosion resistance. Poor corrosion resistance for [7] C.-C. Tung, J.-W. Yeh, T.-t Shun, S.-K. Chen, Y.-S. Huang, H.-C. Chen, On the
mild steel in tap water has been reported in literature previously elemental effect of AlCoCrCuFeNi high-entropy alloy system, Mater. Lett. 61
(2007) 1–5.
[41–43]. Flow corrosion (tap water without sand particles) and [8] C.-J. Tong, Y.-L. Chen, J.-W. Yeh, S.-J. Lin, S.-K. Chen, T.-T. Shun, C.-H. Tsau, S.-
electrochemical experiments were performed to investigate the Y. Chang, Microstructure characterization of AlxCoCrCuFeNi high-entropy alloy
contribution to the degradation due to corrosion. In flow corrosion system with multiprincipal elements, Metall. Mater. Trans. A 36 (2005) 881–893.
[9] W.-R. Wang, W.-L. Wang, S.-C. Wang, Y.-C. Tsai, C.-H. Lai, J.-W. Yeh, Effects of Al
experiments, mild steel (1.4 mg) showed significant mass loss
addition on the microstructure and mechanical property of AlxCoCrFeNi high-
compared to unappreciable mass loss of stainless steel and HEA. entropy alloys, Intermetallics 26 (2012) 44–51.
Further, the electrochemical experiments with tap water also [10] J.-M. Wu, S.-J. Lin, J.-W. Yeh, S.-K. Chen, Y.-S. Huang, H.-C. Chen, Adhesive wear
behavior of AlxCoCrCuFeNi high-entropy alloys as a function of aluminum
showed poor corrosion resistance of mild steel compared with
content, Wear 261 (2006) 513–519.
HEA and SS. HEA showed highest resistance against corrosion with [11] M.-H. Chuang, M.-H. Tsai, W.-R. Wang, S.-J. Lin, J.-W. Yeh, Microstructure and
Icorr and Ecorr values considerably lower than mild steel (Fig. 10). wear behavior of AlxCo1.5CrFeNi1.5Tiy high-entropy alloys, Acta Mater. 59
Therefore, the lack of correlation of erosion rate with UR and MUR (2011) 6308–6317.
[12] C.-Y. Hsu, T.-S. Sheu, J.-W. Yeh, S.-K. Chen, Effect of iron content on wear be-
at normal impingement angle can be attributed to the difference havior of AlCoCrFexMo0.5Ni high-entropy alloys, Wear 268 (2010) 653–659.
in corrosion resistance of tested materials. However, the corrosion [13] C.-Y. Hsu, J.-W. Yeh, S.-K. Chen, T.-T. Shun, Wear resistance and high-tem-
showed negligible influence at oblique angle primarily due to high perature compression strength of Fcc CuCoNiCrAl0.5Fe alloy with boron ad-
dition, Metall. Mater. Trans. A 35 (2004) 1465–1469.
dominance of erosion [41]. However, for mild steel tested at nor- [14] Y.Y. Chen, T. Duval, U.D. Hung, J.W. Yeh, H.C. Shih, Microstructure and elec-
mal impingement angle, the contribution from corrosion en- trochemical properties of high entropy alloys—a comparison with type-304
hanced because of presence of platelet mechanism as the material stainless steel, Corros. Sci. 47 (2005) 2257–2279.
[15] Y.L. Chou, Y.C. Wang, J.W. Yeh, H.C. Shih, Pitting corrosion of the high-entropy
removal mode, which relatively is a slower process. Jana and Stack alloy Co1.5CrFeNi1.5Ti0.5Mo0.1 in chloride-containing sulphate solutions, Corros.
[41] also showed that for iron at a given impact velocity, con- Sci. 52 (2010) 3481–3491.
tribution of corrosion is lower at oblique angles compared to [16] J.H. Zhao, X.L. Ji, Y.P. Shan, Y. Fu, Z. Yao, On the microstructure and erosion–
corrosion resistance of AlCrFeCoNiCu high-entropy alloy via annealing treat-
normal impingement angle. Compared to erosion dominated re-
ment, Mater. Sci. Technol. (2016) 1–5.
gime prevalent at oblique angle, the material removal at normal [17] C. Lee, C. Chang, Y. Chen, J. Yeh, H. Shih, Effect of the aluminium content of
impingement angle for mild steel is comparatively, erosion-dis- AlxCrFe1.5MnNi0.5 high-entropy alloys on the corrosion behaviour in aqueous
solution dominated regime. This explains the higher material re- environments, Corros. Sci. 50 (2008) 2053–2060.
[18] C.-M. Lin, H.-L. Tsai, Evolution of microstructure, hardness, and corrosion
moval rates for mild steel at normal impingement angle which properties of high-entropy Al0.5CoCrFeNi alloy, Intermetallics 19 (2011)
further influenced the correlation with UR and MUR. 288–294.
238 R.B. Nair et al. / Wear 386-387 (2017) 230–238

[19] X. Ji, H. Duan, H. Zhang, J. Ma, Slurry erosion resistance of laser clad fault energies of NiFeCrCoMn high-entropy alloy, JOM 65 (2013) 1780–1789.
NiCoCrFeAl3 high-entropy alloy coatings, Tribol. Trans. 58 (2015) 1119–1123. [33] A. Das, Revisiting stacking fault energy of steels, Metall. Mater. Trans. A 47
[20] N. Kumar, M. Komarasamy, P. Nelaturu, Z. Tang, P.K. Liaw, R.S. Mishra, Friction (2016) 748–768.
stir processing of a high entropy alloy Al0.1CoCrFeNi, JOM 67 (2015) [34] R. Bellman, A. Levy, Erosion mechanism in ductile metals, Wear 70 (1981)
1007–1013. 1–27.
[21] ASM International, Atlas of Stress-Strain Curves, ASM International, 2002. [35] I. Finnie, J. Wolak, Y. Kabil, Erosion of metals by solid particles, J. Mater. 2
[22] H.S. Arora, H.S. Grewal, H. Singh, S. Mukherjee, Zirconium based bulk metallic (1967) 682.
glass—better resistance to slurry erosion compared to hydroturbine steel, [36] P.H. Shipway, I.M. Hutchings, The rôle of particle properties in the erosion of
Wear 307 (2013) 28–34. brittle materials, Wear 193 (1996) 105–113.
[23] H.S. Arora, H.S. Grewal, H. Singh, B.K. Dhindaw, S. Mukherjee, Unusually high [37] J. Hobbs, Experience with a 20-kc cavitation erosion test, in: Erosion by Ca-
erosion resistance of zirconium-based bulk metallic glass, J. Mater. Res. 28 vitation or Impingement, ASTM International, 1967.
(2013) 3185–3189. [38] F.G.H.R. Garcia, Cavitation damage and correlation with material and fluid
[24] A.V. Levy, G. Hickey, Liquid-solid particle slurry erosion of steels, Wear 117 properties, 5031-II-I, 1966.
(1987) 129–146. [39] B.S. Mann, V. Arya, An experimental study to corelate water jet impingement
[25] I.M. Hutchings, R.E. Winter, Particle erosion of ductile metals: a mechanism of erosion resistance and properties of metallic materials and coatings, Wear 253
material removal, Wear 27 (1974) 121–128. (2002) 650–661.
[26] T. Foley, A. Levy, The erosion of heat-treated steels, Wear 91 (1983) 45–64. [40] B.C. Syamala Rao, N.S. Lakshmana Rao, K. Seetharamiah, Cavitation erosion
[27] I.M. Hutchings, A model for the erosion of metals by spherical particles at studies with venturi and rotating disk in water, J. Basic Eng. 92 (1970)
normal incidence, Wear 70 (1981) 269–281. 563–573.
[28] G. Sundararajan, P.G. Shewmon, A new model for the erosion of metals at [41] B. Jana, M. Stack, Modelling impact angle effects on erosion–corrosion of pure
normal incidence, Wear 84 (1983) 237–258. metals: construction of materials performance maps, Wear (2005) 243–255.
[29] I.Finnie, Erosion of surface by solid particles, Wear. [42] Y.-S. Choi, J.-J. Shim, J.-G. Kim, Effects of Cr, Cu, Ni and Ca on the corrosion
[30] G.E. Dieter, D.J. Bacon, Mechanical Metallurgy, McGraw-Hill, London; New behavior of low carbon steel in synthetic tap water, J. Alloy. Compd. 391
York, 1988. (2005) 162–169.
[31] X.-F. Zhang, L. Fang, The effect of stacking fault energy on the cavitation ero- [43] M.A. Moore, Laboratory simulation testing for service abrasive wear en-
sion resistance of α-phase aluminum bronzes, Wear 253 (2002) 1105–1110. vironments, in: K.C. Ludema (Ed.), Wear of Materials, ASME, New York, 1987,
[32] A.J. Zaddach, C. Niu, C.C. Koch, D.L. Irving, Mechanical properties and stacking pp. 673–687.

You might also like