You are on page 1of 10

Wear 386–387 (2017) 129–138

Contents lists available at ScienceDirect

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

Enhancing the erosion-corrosion resistance of steel through friction stir MARK


processing

K. Selvama, A. Ayyagarib, H.S. Grewala, S. Mukherjeeb, H.S. Aroraa,
a
Department of Mechanical Engineering, School of Engineering, Shiv Nadar University, Uttar Pradesh 201314, India
b
Department of Materials Science and Engineering, University of North Texas, Denton, TX 76203, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Erosion-corrosion is a critical problem in marine components and has huge economic impact. In the current
Erosion-corrosion study, friction stir processing was utilized for enhancing the erosion-corrosion resistance of SS316L steel, most
Friction stir processing widely used material for marine applications. Compared to the unprocessed alloy, the friction-stir processed
Phase transformation specimen showed nearly 3.5 times and 5 times higher erosion and erosion-corrosion resistance respectively at
oblique impingement. In contrast, erosion and erosion-corrosion resistance for processed alloy was 2 times
higher at normal impingement. The unusual enhancement in erosion-corrosion resistance was attributed to
surface strengthening through grain-size refinement and martensite phase formation. In addition, higher cor-
rosion resistance due to stronger passivation also contributed towards superior performance of the processed
alloy.

1. Introduction material. Positive synergy indicates enhanced degradation due to


combination of erosion and corrosion, results in either an erosion
Material degradation in the form of corrosion, erosion and their dominant or a corrosion dominant process. In contrast, negative sy-
combination are critical problems in marine environment. nergy signifies that material degradation during erosion-corrosion is
Impingement of hard abrasive particles mixed in a corrosive medium is reduced compared to their standalone values.
the root cause for degradation of marine components exposed to hydro- Surface coatings are typically used to address erosion-corrosion
dynamic conditions. Typically, erosion involves repetitive impacts of problems in naval platforms [1–5]. Thermally sprayed nickel-alumi-
hard erodent particles which eventually leads to surface damage and nium bronze coating is the most commonly used alloy composition to
severe material loss. The presence of corrosive media further enhances address material degradation by erosion-corrosion. This composition
material deterioration by synergistic effects of erosion (mechanical shows high resistance to erosion-corrosion by formation of a stable and
process) and corrosion (electro-chemical process). Turbulent flow, high strongly adherent oxide film. However, inherent lamellar micro-
impact velocities and hard erosive particles can further exacerbate the structure of thermal sprayed coatings, presence of splat boundaries,
rate of material loss. pores and un-melted particles results in anisotropic behavior together
Austenitic stainless steels are widely used for various marine ap- with poor mechanical and tribological properties. This leads to pre-
plications. Amongst these, SS316L is the most commonly used material mature coating failure and limits its applicability. In contrast, tailoring
for propulsion, seawater handling systems and shipping industries. the surface properties of the parent material itself without changing the
Austenitic steels are known for their superior erosion and corrosion surface chemistry is an elegant route to overcome or avoid the afore-
resistance. However, the synergistic effect of erosion-corrosion leads to mentioned limitations. Severe plastic deformation techniques such as
significant mass loss in these materials. This is primarily due to re- friction stir processing (FSP) are widely used for tuning the surface as
peated breakdown of protective passive layer during the erosion pro- well as bulk properties through microstructural refinement in materials
cess. The total material loss during erosion-corrosion can be evaluated [6–13]. However, most of these studies are on light weight alloys and
using the expression: T = E + C + S , where T represents the total mass limited work has been done on high strength steels. Further, the studies
loss, C represents mass loss in pure corrosion, E represents mass loss in on high strength stainless steels are either on pure corrosion [14–18] or
pure erosion and S is the synergy component. Synergy plays a vital role pure erosion [19–25]. There are very limited studies on synergistic
in determining the overall erosion-corrosion characteristics of a erosion-corrosion behavior of stainless steel and the influence of severe


Corresponding author.
E-mail address: harpreet.arora@snu.edu.in (H.S. Arora).

http://dx.doi.org/10.1016/j.wear.2017.06.009
Received 31 March 2017; Received in revised form 5 June 2017; Accepted 21 June 2017
Available online 23 June 2017
0043-1648/ © 2017 Elsevier B.V. All rights reserved.
K. Selvam et al. Wear 386–387 (2017) 129–138

plastic deformation on the erosion-corrosion behavior has not been


evaluated.
In the current study, surface microstructure modification of SS316L
stainless steel was done using friction stir processing. FSP was done at
two different strain rates, one high and one low. Microstructure evo-
lution and phase transformation following FSP was studied using
electron-back-scatter diffraction (EBSD). Synergistic degradation by
erosion-corrosion as well as their independent contributions were in-
vestigated for the as-received and processed samples. Sample processed
at lower strain rate showed nearly five times higher resistance to ero-
sion-corrosion at oblique impingement angles. Remarkable improve-
ment in the erosion resistance was explained based on structural re-
juvenation resulting in surface strengthening and formation of stable
passive oxide layer on the processed samples.
Fig. 2. Schematic of slurry erosion test rig used in the current study. The test rig com-
2. Experimental prised of a diaphragm pump driven by compressed air. The premixed slurry in a container
is pumped using the diaphragm pump and made to impinge on a sample through a 2 mm
The material used in the current investigation was austenitic steel, diameter tungsten carbide nozzle.
SS316L. Samples with dimensions 70 mm × 50 mm × 3 mm were cut
from commercially available SS316L. FSP was performed on a universal
milling machine using a pin-less tool to tailor only the surface prop-
erties of the alloy. FSP was performed with rapid cooling, where the compressed air. The premixed slurry in a container was pumped using
sample was completely submerged in a liquid coolant (mixture of the diaphragm pump and impinged on the sample through a 2 mm
ethanol and distilled water) bath maintained at 0 °C. Two different diameter tungsten carbide nozzle. The velocity of slurry jet was con-
rotational speeds were used: 388 rpm (sample referred as 388C) and trolled by changing the pressure of the compressed air used for driving
1800 rpm (sample referred as 1800C) while keeping all other para- the pump. A magnetic stirrer prevented any sedimentation of the sand
meters constant. A special purpose FSP fixture was fabricated for through continuous stirring. A schematic representation of the test rig is
holding the sample while submerged in a pool of low temperature li- shown in Fig. 2. Prior to the testing, all the samples were polished down
quid. The FSP fixture was connected to the external chiller through inlet to 2000 grit emery paper. Erosion testing was done at a constant ve-
and outlet ports for the consistent flow of liquid. Fig. 1 shows a sche- locity of 20 m/s and 0.5 wt% mass flux rate of sand in distilled water.
matic representation of the FSP experimental set up. The processed Sand used for erosion tests was sieved to obtain particle size in range of
sample was sectioned along the cross-section using low speed diamond 75–150 µm. The total testing time was 2 h with weight loss measured
saw. The surface as well as cross-section samples were polished down to after each experiment. Each test was performed three times under si-
3000 grit followed by electro-polishing in 10% oxalic acid solution at milar conditions to ensure the repeatability. To prevent any occurrence
6 V for 2 min. Hardness of the processed sample were evaluated using of corrosion during pure erosion test, a cathodic protection of −1 V was
microhardness test along the cross-section, while elastic modulus was given to the sample surface. Dynamic corrosion tests were done in 3.5%
obtained from nano-indentation. The grain size, size distribution and NaCl solution without any abrasive particles while impinging the sur-
phase transformation for the as-received and processed samples were face at a constant impact velocity of 20 m/s. Erosion-corrosion tests
obtained using electron back scatter diffraction (EBSD). EBSD analysis were done in 3.5% NaCl solution containing abrasive particles without
was conducted using FEI Quanta 3D FEG using step size of 0.1 µm. any cathodic protection of the sample surface. Mass loss measurements
Erosion and erosion-corrosion tests were carried out in a custom- during erosion and erosion-corrosion tests were done using high pre-
built test rig per ASTM standard G134. It has built in flexibility for cision weighing balance with an accuracy of 0.01 mg. Static corrosion
controlling the impact velocity, stand-off distance between nozzle experiments were done in electrochemical cell using a standard three
outlet and the target surface, the impingement angle as well as the electrode cell configuration with high density graphite rod as counter
slurry medium. The test rig comprised of a diaphragm pump driven by electrode, saturated calomel electrode (SCE) as reference and sample as
working electrode. Cyclic polarization was done in the voltage range of
−0.4 V to 1.5 V vs EOCP with forward and reverse scan rate of
0.166 mV/s and peak limiting current density of 25 mA/cm2. The sur-
face of all eroded samples was studied using scanning electron micro-
scope (SEM) for understanding the erosion and erosion-corrosion me-
chanisms.

3. Results and discussion

3.1. Microstructure and mechanical characterization

Fig. 3(a) is the EBSD map for as-received SS316L showing the grain
structure and grain size distribution. The microstructure shows twin
boundaries, attributed to low stacking fault energy for austenitic steels.
Fig. 1. Schematic of friction stir processing (FSP) set-up used in the current study. A Average grain size for the as-received alloy calculated from grain size
specially designed tool-fixture was used to clamp the workpiece of variable dimensions. distribution (Fig. 3(b)) is nearly 22 µm. Microstructural details for the
The fixture has a cooling tank located above the work-piece which is connected to the processed samples are shown in Fig. 3(c) to (f). Both the processed
external chiller unit through inlet and outlet ports. The coolant used in the chiller was
samples showed significant grain refinement compared to as-received
mixture of distilled water and ethanol.
sample, with an average grain size of 0.67 µm for 388C and 0.9 µm for
1800C. Dynamic recrystallization (DRX) comprising rearrangement of
dislocations into sub-grains and grain boundary migration are believed

130
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 3. EBSD map for (a) as-received austenitic steel SS316L, (c) friction stir processed SS316L at 388 rpm (388C) and (e) friction stir processed SS316L at 1800 rpm (1800C); Grain size
distribution for (b) as-received SS316L, (d) 388C and (f) 1800C.

to be the primary mechanisms for grain structure refinement during FSP expressed by Zener Holloman parameter, Z = ∈̇ exp (Q/ RT ) , where ∈̀ is
[26]. A pseudo heat index for the FSP process is given as: ω2/υ, where ω the strain rate, Q is the activation energy for deformation, T is the
is the tool rotational speed υ is the transverse speed. This implies that absolute temperature and R is the gas constant. Deformation at high
heat input and strain rate are both functions of tool rotational speed strain-rates leads to significant rise in temperature which may result in
[24,25]. Higher strain rate favors finer grain structure while increasing a coarse grain structure [7,18,27]. Hence it is critical to control tem-
temperature induces grain growth. Hence temperature and strain rate perature during high strain-rate deformation process to achieve ultra-
compete to control the final grain size during dynamic recrystallization fine grain structure [13,30,31]. The finer grain structure for 388C
[27–29]. The combined influence of strain rate and temperature is sample is attributed to lesser temperature rise during processing. EBSD

131
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 4. EBSD phase map showing distribution of austenite and martensite phases in (a) as-received austenitic steel, SS316L; sample friction stir processed at (b) 388 rpm (388 C); (c)
1800 rpm (1800C); (d) overall fraction of austenitic and martensitic phases in all the samples. In the phase map, green color indicate austenite phase while red color indicate martensite.

phase map showing austenite and martensite phase distribution in the both the processed samples. The as-received alloy shows a strong (111)
as-received and processed samples are given in Fig. 4(a) to (d). As-re- parallel ND texture while 388C and 1800C shows strong (100) parallel
ceived alloy had austenite as the primary phase (Fig. 4(a)). It is seen ND and (211) parallel ND texture respectively. Face-centred-cubic
that there is austenite to martensite phase transformation during pro- materials show higher elastic modulus along (111) plane compared to
cessing. Martensite phase is significantly higher for 388C compared to other crystallographic plane [33]. This is likely attributed to the closest
1800C. High strain rate is known to suppress austenite to martensite atomic packing on (111) plane resulting in strong inter-atomic bonding.
transformation due to more adiabatic heating at higher strain rate [32]. Therefore, higher elastic modulus for the as-received alloy is likely due
Therefore, lower strain rate deformation accompanied by rapid cooling to strong (111) texture which changes during processing resulting in
resulted in higher fraction of martensite phase for 388C. decrease in elastic modulus.
Fig. 5(a) shows the variation in microhardness for both the pro-
cessed samples as a function of depth from the top surface. Hardness 3.2. Erosion-corrosion behavior
was maximum at the surface and decreased along the depth. 388C
showed maximum hardness of nearly 1.8 times compared to the as- The erosion and erosion-corrosion depth for the processed samples
received alloy, while it was nearly 1.5 times for 1800C sample. The was measured to be nearly 20–25 µm and 60–70 µm respectively. Since
significant rise in hardness after processing was primarily attributed to the processed depth was nearly 150 µm, the material removal was
grain boundary strengthening and austenite to martensite transforma- completely from the processed region only. Fig. 6(a) to (c) show results
tion as shown by the EBSD results. The hardness of both the processed for pure erosion test. It is seen that erosion rate was higher at 30° im-
samples decreased to the value for as-received alloy at about 150 µm, pingement compared to normal impingement for both the as-received
corresponds to the depth of the processed zone. The elastic modulus for as well as processed samples. This implies similarity of erosion beha-
all the samples was obtained using nano-indentation. The load-dis- vior/mechanism for the unprocessed and processed samples. Lower
placement curves for the as-received alloy and both the processed erosion rate at normal impingement and higher at oblique angles im-
samples are shown in Fig. 5(b). The modulus values are shown along- plies ductile mode of material removal for austenitic steel, remains the
side the load-displacement curves. Hardness increased after processing, same for processed samples as well. Further, mass loss is significantly
while elastic modulus decreased. The decrease in elastic modulus values lower for both the processed samples. At 30° impingement angle, the
after processing is likely due to texture changes during friction stir erosion resistance for 388C sample was nearly 3.5 times higher com-
processing. Fig. 5(c) shows inverse pole figures (IPF) for as-received and pared to the unprocessed alloy, while it was nearly 1.6 times higher for

132
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 5. (a) Hardness as function of depth for the as-received austenitic stainless steel, SS316L, friction stir processed SS316L at 388 rpm (388C) and friction stir processed SS316L at
1800 rpm (1800C); (b) load-displacement curves for the as-received austenitic stainless steel, SS316L, 388C and 1800C specimens obtained using nano-indentation; (c) Normal direction
(ND) inverse pole figure (IPF) for as-received, 388C and 1800C specimens. The color bar shows the texture intensity in each case.

1800C sample. At 90° impingement, both the processed samples corrosion and synergy are summarised in Table 1. Higher synergy for
showed nearly similar behavior with nearly two times higher erosion the as-received alloy compared to processed samples indicate its lower
resistance compared to the as-received alloy. Phenomenal increase in resistance to material loss under simultaneous influence of erosion and
erosion resistance for 388C is attributed primarily to its higher hard- corrosion. Interestingly, synergy is higher for 388C compared to 1800C.
ness, as a consequence of grain size refinement and higher fraction of This is likely due to higher fraction of martensite in 388C which has
martensite phase formed during processing (Fig. 4(b)). inferior corrosion properties compared to austenite [32,34].
Results for erosion-corrosion are also shown in Fig. 6(a) to (c). For To understand the superior erosion-corrosion resistance of pro-
all samples, the material loss was higher in erosion-corrosion compared cessed samples, standalone electrochemical corrosion tests were done.
to pure erosion. The increase in material loss during erosion-corrosion These include Tafel and cyclic polarization in 3.5% NaCl solution. The
may be attributed to the following factors: (i) removal of hardened cathodic and anodic polarization curves obtained using cyclic polar-
surface by corrosion, exposing relatively soft surface to further erosion; ization for all the specimens are shown in Fig. 7(a) to (d). The corrosion
(ii) corrosion attack at grain boundaries resulting in enhanced material current (Icorr) was evaluated using Tafel exploration from the polar-
removal; (iii) detachment of raised lips due to corrosion; (iv) micro- ization curves and values are shown in Table 2. Icorr is the highest for as-
pitting resulting in increase in number of stress concentration defects. received alloy followed by 1800C and lowest for 388C indicating higher
In contrast to the as-received alloy, the mass loss during erosion-cor- corrosion resistance of processed specimens. Interestingly, the grain
rosion is much smaller for the processed samples. For 30° impingement, size also follows similar order- largest for as-received, followed by
the mass loss for 1800C was nearly 3.5 times lower, while for 388C it 1800C and 388C. Higher corrosion resistance for the processed samples
was lower by nearly 5 times. At normal impingement, the performance can be attributed to their finer grain structure. Typically, the oxide
of both the processed specimens was nearly two times better compared layer formed on stainless steel has a duplex structure with Cr2O3
to the as-received alloy. The 388C sample showed unusually high ero- (having p-type semiconductor characteristics) as an inner layer and
sion-corrosion resistance, particularly at oblique angle, the first report Fe2O3 (having n-type semiconductor characteristics) as an outer layer
of its kind. The mass loss during dynamic corrosion was insignificant for [35]. The donor to acceptor ratio in the passive layer has significant
all the samples (0.25 mg/h, 0.155 mg/h and 0.215 mg/h for the as-re- influence on the oxide layer characteristics. In general, the donor to
ceived, 388C and 1800C specimens respectively). As mentioned earlier, acceptor density in passive layer decreases with the grain refinement
erosion-corrosion involves the synergy effect as expressed by the rela- [35], causing reduction in the available charge carriers and improving
tion: T = E + C + S where synergy (S) could be either positive (in- passivity. The donor to acceptor density further reduces in the chloride
crease in degradation) or negative (decrease in degradation). Fig. 6(d) environment where chloride ion tends to occupy the oxygen vacancies
shows that the synergy obtained in the current study is positive. For the in the passive film [35]. Thus, the finer grain structure of processed
as-received alloy, synergy component is the highest (~52%) followed samples contributed towards faster passivation kinetics, better passive
by 388C (~47%) and 1800C (~27%). The results for erosion, erosion- layer stability, and superior corrosion resistance.

133
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 6. Material removal rate (mg/h) as a function of impingement angle during pure erosion and erosion-corrosion test for (a) as-received austenitic stainless steel, SS316L, (b) friction
stir processed SS316L at 388 rpm (388C), (c) friction stir processed SS316L at 1800 rpm (1800C), (d) pie-chart showing erosion, corrosion and synergy components for as-received, 388C
and 1800C specimens.

Table 1 with the austenite phase promoting localized pitting.


Mass loss under pure erosion, pure corrosion and erosion-corrosion for all the samples. Fig. 8 shows SEM images of processed and as-received alloy tested
Synergy between erosion-corrosion and mass loss due to each component is also given.
under pure erosion. As seen from these images, the primary erosion
Sample Total Erosion (E) Corrosion (C) Synergism % of mechanism at 30° impact angle was combination of micro-cutting,
(SS316L) mass loss [mg/h] [mg/h] total degradation ploughing and platelets. The sharp abrasive particles remove material
(T) [mg/ S by cutting action as it impinges at oblique angles resulting in significant
h] micro-cutting features on the eroded surface. Some raised-up lips,
As-received 8.915 4.04 0.255 51.82
produced by material ploughing can also be seen in Fig. 8(a) through
388C 1.9 0.84 0.155 47.63 (c). Platelets are formed due to continuous extrusion-forging on the
1800C 2.7 1.655 0.215 27.04 ductile sample surface. Initially, the surface is plastically deformed
through extrusion process and subsequent impact of erodent particles
leads to forging of extruded metal forming pile up of platelets. Fig. 8(d)
In addition to general corrosion, localized pitting behavior of all the through (f) show eroded surfaces for the as-received, 388C and 1800C
samples was obtained by evaluating the pitting potential (Epit) from at 90° impact angle respectively. Severe plastic deformation is the pri-
cyclic polarization test (Fig. 7(a) to (c) and Table 2). The resistance to mary erosion mechanisms at normal impingement. In addition, all
pit initiation is given by Epit-Ecorr and is shown in Fig. 7(d) for all the samples show features of material removal through indentation by
specimens. 1800C showed higher resistance to pit initiation compared erosive particles, where accelerated erodent particles impacts the spe-
to 388C followed by the as-received alloy. Further, the severity of pit- cimen normally, leading to the material removal in the form of debris of
ting corrosion relates to the loop area above Epp which is largest for the plastically deformed material. SEM images of samples subjected to
as-received alloy and least for 1800C specimen (Table 2). Thus, both erosion-corrosion are shown in Fig. 9. It is seen that the material re-
the processed specimens showed stronger passivation compared to the moval mechanism for erosion-corrosion are similar to that of pure
as-received alloy. Enhancement in pitting potential for processed sam- erosion with micro-cutting, ploughing and platelets at oblique angles
ples may be attributed to increase in Cr diffusion at grain boundaries and micro indent at normal impingement.
which promotes quicker passivation [15,16,18]. In addition, both the As discussed, the material removal during pure erosion at oblique
processed samples have more refined grain structure which is likely to impingement is primarily by the cutting phenomenon. Material's
enhance the adherence of passive oxide layer to the surface by oxide- hardness play a more significant role in controlling material removal by
pegging effect [36]. Lower pitting resistance for 388C compared to cutting and hence is the controlling factor at oblique impingement.
1800C is attributed to presence of martensite phase in the austenite Therefore, the erosion resistance followed the same trend as hardness
matrix (Fig. 4(b)). Martensite is known to have more negative galvanic i.e. 388C > 1800C > As-received. During erosion-corrosion at oblique
potential compared to austenite [37]. It forms a micro-galvanic couple

134
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 7. Cyclic polarization curves for (a) as-received austenitic stainless steel, SS316L, (b) friction stir processed SS316L at 388 rpm (388 C), (c) friction stir processed SS316L at 1800 rpm
(1800C), (d) pit initiation resistance (Epit - Ecorr) for as-received, 388 C and 1800C specimens.

sample again showed two times higher resistance, similar to pure ero-
Table 2
sion, which can be attributed to brittle fracture of the passive oxide
Results of electrochemical corrosion test for all the samples. layer during normal impingement. Thus, the passive oxide layer could
not contribute to further enhance the erosion-corrosion resistance of
Sample Icorr (µA) Ecorr (mV) Epit (mV) Epp (mV) Epit Area processed specimens. Further, comparison of Figs. 8 and 9 indicates
–Ecorr under
that platelet formation for pure erosion is more severe compared to
(mV) the loop
(mV.A/ erosion-corrosion. Slurry impingement results in material extrusion
cm2) which is weakened by the corrosive media present in the slurry. This
leads to rapid loss of material as debris, instead of pile up as in the case
As-received 19.3 −319 168.4 −18.66 487.4 7.035
of pure erosion and likely explains increased material loss in erosion-
388-C 5.23 −368 303.4 −41.06 671.4 6.35
1800-C 9.49 −326 387.1 −26.57 713.1 4.245 corrosion.

4. Conclusions

impingement, material hardness as well as passive layer stability are the Surface properties of SS316L stainless steel were tailored using
controlling factors. 388C specimen has the highest hardness as well as submerged friction stir processing and their erosion and erosion-cor-
better passive layer stability than the as-received alloy which enhanced rosion behavior was evaluated. Major findings of the current study are:
its erosion-corrosion resistance by five times. In contrast, the material
removal at normal impingement is dominated by plastic deformation 1. Submerged friction stir processing refined the grain structure from
and indentation. Therefore, material plasticity plays a more dominant 22 µm for the as-received alloy to as low as 0.67 −0.89 µm for the
role at normal impingement. Greenwood and Williamson [38] proposed processed specimen.
that plasticity index for the material is inversely proportional to hard- 2. The processed samples showed significant increase in hardness.
ness to modulus (H/E) ratio. H/E ratio for the processed materials was Compared to as-received alloy (220 HV), the hardness of the pro-
calculated to be 0.022 and 0.02 for 388C and 1800C specimens re- cessed samples was around 420 HV for 388C processing condition
spectively. These values are close enough which explains similar ero- and 350 HV for 1800C processing condition. Increase in hardness is
sion behavior for both the processed samples at normal impingement. attributed to grain refinement and austenite to martensite trans-
During erosion-corrosion at normal impingement, both processed formation.

135
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 8. Scanning electron images for as-received SS316L subjected to pure erosion at (a) 30° impingement and (b) at 90° impingement; SS316L friction stir processed at 388 rpm (388C)
and subjected to pure erosion at (c) 30° impingement and (d) at 90° impingement; SS316L friction stir processed at 1800 rpm (1800C) and subjected to pure erosion at (e) 30°
impingement and (f) at 90° impingement.

3. The processed samples demonstrated significantly higher erosion 4. The processed samples also showed superior corrosion properties
and erosion-corrosion resistance. At oblique impingement, 388C which is due to higher stability of the passive layer as a result of
shows around 3.5 times higher erosion resistance and about 5 times pegging of the oxide layer and faster passivation kinetics.
higher erosion-corrosion resistance. At normal impingement, pro- 5. Erosion-corrosion process was found to be erosion dominant, how-
cessed samples showed two times higher erosion as well as erosion- ever, synergy with corrosion was also observed to be contributing
corrosion resistance. significantly.

136
K. Selvam et al. Wear 386–387 (2017) 129–138

Fig. 9. Scanning electron images for as-received SS316L subjected to erosion-corrosion at (a) 30° impingement and (b) at 90° impingement; SS316L friction stir processed at 388 rpm
(388C) and subjected to erosion-corrosion at (c) 30° impingement and (d) at 90° impingement; SS316L friction stir processed at 1800 rpm (1800C) and subjected to erosion-corrosion at
(e) 30° impingement and (f) at 90° impingement.

References [5] H.J.C. Voorwald, L.F.S. Vieira, M.O.H. Cioffi, Evaluation of WC-10Ni thermal
spraying coating by HVOF on the fatigue and corrosion AISI 4340 steel, Procedia
Eng. 2 (2010) 331–340.
[1] C. Liu, A. Leyland, Q. Bi, A. Matthews, Corrosion resistance of multi-layered [6] R.Z. Valiev, I.V. Alexandrov, Nanostructured materials from severe plastic de-
plasma-assisted physical vapour deposition tin and CrN coatings, Surf. Coat. formation, Nanostruct. Mater. 12 (1999) 35–40.
Technol. 141 (2001) 164–173. [7] I. Charit, R.S. Mishra, High strain rate superplasticity in a commercial 2024 Al alloy
[2] D. Toma, W. Brandl, G. Marginean, Wear and corrosion behaviour of thermally via friction stir processing, Mater. Sci. Eng.: A 359 (2003) 290–296.
sprayed cermet coatings, Surf. Coat. Technol. 138 (2001) 149–158. [8] J.-Q. Su, T.W. Nelson, C.J. Sterling, Friction stir processing of large-area bulk UFG
[3] K.L. Choy, Chemical vapour deposition of coatings, Prog. Mater. Sci. 48 (2003) aluminum alloys, Scr. Mater. 52 (2005) 135–140.
57–170. [9] C.J. Hsu, C.Y. Chang, P.W. Kao, N.J. Ho, C.P. Chang, Al–Al3Ti nanocomposites
[4] V.A.D. Souza, A. Neville, Aspects of microstructure on the synergy and overall produced in situ by friction stir processing, Acta Mater. 54 (2006) 5241–5249.
material loss of thermal spray coatings in erosion–corrosion environments, Wear [10] M.A. García-Bernal, R.S. Mishra, R. Verma, D. Hernández-Silva, High strain rate
263 (2007) 339–346. superplasticity in continuous cast Al–Mg alloys prepared via friction stir processing,

137
K. Selvam et al. Wear 386–387 (2017) 129–138

Scr. Mater. 60 (2009) 850–853. stainless steel by friction stir processing, Appl. Surf. Sci. 308 (2014) 184–192.
[11] L. Fratini, G. Buffa, R. Shivpuri, Mechanical and metallurgical effects of in process [25] M. Hajian, A. Abdollah-Zadeh, S. Rezaei-Nejad, H. Assadi, S. Hadavi, K. Chung,
cooling during friction stir welding of AA7075-T6 butt joints, Acta Mater. 58 (2010) M. Shokouhimehr, Microstructure and mechanical properties of friction stir pro-
2056–2067. cessed AISI 316L stainless steel, Mater. Des. 67 (2015) 82–94.
[12] S. Ramesh Babu, V.S.S. Kumar, G.M. Reddy, L. Karunamoorthy, Microstructural [26] A. Rollett, F. Humphreys, G.S. Rohrer, M. Hatherly, Recrystallization and related
changes and mechanical properties of friction stir processed extruded AZ31B alloy, annealing phenomena, Elsevier, 2004.
Procedia Eng. 38 (2012) 2956–2966. [27] C. Chang, C. Lee, J. Huang, Relationship between grain size and Zener–Holloman
[13] C. Sharma, D. Dwivedi, P. Kumar, Influence of in-process cooling on tensile beha- parameter during friction stir processing in AZ31 Mg alloys, Scr. Mater. 51 (2004)
viour of friction stir welded joints of AA7039, Mater. Sci. Eng.: A 556 (2012) 509–514.
479–487. [28] R.S. Mishra, Z.Y. Ma, Friction stir welding and processing, Mater. Sci. Eng.: R: Rep.
[14] M. Hasegawa, M. Osawa, Corrosion behavior of ultrafine grained austenitic stain- 50 (2005) 1–78.
less steel, Corrosion 40 (1984) 371–374. [29] T. McNelley, S. Swaminathan, J. Su, Recrystallization mechanisms during friction
[15] M. Pisarek, P. Keedzierzawski, M. Janik-Czachor, K. Kurzydlowski, Effect of hy- stir welding/processing of aluminum alloys, Scr. Mater. 58 (2008) 349–354.
drostatic extrusion on the corrosion resistance of type 316 stainless steel, Corrosion [30] D.C. Hofmann, K.S. Vecchio, Submerged friction stir processing (SFSP): an im-
64 (2008) 131–137. proved method for creating ultra-fine-grained bulk materials, Mater. Sci. Eng.: A
[16] Z.J. Zheng, Y. Gao, Y. Gui, M. Zhu, Corrosion behaviour of nanocrystalline 304 402 (2005) 234–241.
stainless steel prepared by equal channel angular pressing, Corros. Sci. 54 (2012) [31] C. Lee, J. Huang, P. Hsieh, Mg based nano-composites fabricated by friction stir
60–67. processing, Scr. Mater. 54 (2006) 1415–1420.
[17] A.T. Krawczynska, M. Gloc, K. Lublinska, Intergranular corrosion resistance of na- [32] H. Nakagawa, T. Miyazaki, Effect of retained austenite on the microstructure and
nostructured austenitic stainless steel, J. Mater. Sci. 48 (2013) 4517–4523. mechanical properties of martensitic precipitation hardening stainless steel, J.
[18] M. Rifai, H. Miyamoto, H. Fujiwara, Effects of strain energy and grain size on Mater. Sci. 34 (1999) 3901–3908.
corrosion resistance of ultrafine grained Fe-20% Cr steels with extremely low C and [33] S. Pugh, XCII. Relations between the elastic moduli and the plastic properties of
N fabricated by ECAP, Int. J. Corros. 2015 (2015). polycrystalline pure metals, Lond. Edinb. Dublin Philos. Mag. J. Sci. 45 (1954)
[19] C. Heathcock, B. Protheroe, A. Ball, Cavitation erosion of stainless steels, Wear 81 823–843.
(1982) 311–327. [34] M. Isakov, S. Hiermaier, V.-T. Kuokkala, Effect of strain rate on the martensitic
[20] G. Stachowiak, Particle angularity and its relationship to abrasive and erosive wear, transformation during plastic deformation of an Austenitic stainless steel, Metall.
Wear 241 (2000) 214–219. Mater. Trans. A 46 (2015) 2352–2355.
[21] D. Lopez, J. Congote, J. Cano, A. Toro, A. Tschiptschin, Effect of particle velocity [35] L. Jinlong, L. Hongyun, Effect of temperature and chloride ion concentration on
and impact angle on the corrosion–erosion of AISI 304 and AISI 420 stainless steels, corrosion of passive films on nano/ultrafine grained stainless steels, J. Mater. Eng.
Wear 259 (2005) 118–124. Perform. 23 (2014) 4223–4229.
[22] J. Escobar, E. Velásquez, T. Santos, A. Ramirez, D. López, Improvement of cavita- [36] J.K. Tien, F.S. Pettit, Mechanism of oxide adherence on Fe-25Cr-4Al (Y or Sc) alloys,
tion erosion resistance of a duplex stainless steel through friction stir processing Metall. Trans. 3 (1972) 1587–1599.
(FSP), Wear 297 (2013) 998–1005. [37] A.S. Hamada, L.P. Karjalainen, M.C. Somani, Electrochemical corrosion behaviour
[23] H. Grewal, H. Arora, H. Singh, A. Agrawal, Surface modification of hydroturbine of a novel submicron-grained austenitic stainless steel in an acidic NaCl solution,
steel using friction stir processing, Appl. Surf. Sci. 268 (2013) 547–555. Mater. Sci. Eng.: A 431 (2006) 211–217.
[24] M. Hajian, A. Abdollah-Zadeh, S. Rezaei-Nejad, H. Assadi, S. Hadavi, K. Chung, [38] J.A. Greenwood, J.B.P. Williamson, Contact of nominally flat surfaces, Proc. R. Soc.
M. Shokouhimehr, Improvement in cavitation erosion resistance of AISI 316L Lond. Ser. A. Math. Phys. Sci. 295 (1966) 300–319.

138

You might also like