Materials and Design: Shaoning Geng, Junsheng Sun, Lingyu Guo

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Materials and Design 88 (2015) 17

Contents lists available at ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/jmad

Effect of sandblasting and subsequent acid pickling and passivation on


the microstructure and corrosion behavior of 316L stainless steel
Shaoning Geng, Junsheng Sun , Lingyu Guo
Key Laboratory for LiquidSolid Structural Evolution and Processing of Materials, Ministry of Education, Shandong University, Jinan 250061, P R China

a r t i c l e

i n f o

Article history:
Received 5 March 2015
Received in revised form 21 August 2015
Accepted 22 August 2015
Available online 30 August 2015
Keywords:
316L stainless steel
Microstructure
Corrosion behavior
Sandblasting
Acid pickling
Passivation

a b s t r a c t
The microstructure, surface morphologies and corrosion behavior of 316L stainless steel processed by
sandblasting (SB) and sandblasting + acid pickling + passivation (SBPP) were studied in this work. Microstructure of the surface layer was investigated by optical microscope, scanning electron microscope and X-ray diffraction etc. The corrosion behavior was studied by potentiodynamic polarization tests and the surface morphologies
were observed by scanning electron microscope. Results indicated that SB and SBPP samples had almost the same
microhardness and dislocation density which were higher than the as-received (AR) sample. Besides, straininduced -martensite phase was observed in SB and SBPP samples. After sandblasting, the corrosion resistance
decreased dramatically due to the surface morphologies, formation of -martensite phase and increase of
dislocation density, while it was improved to some extent by the subsequent acid pickling and passivation.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
316L austenitic stainless steel is widely used in exhaust manifolds,
oil and gas, nuclear industry, and paper and pulp industry due to its excellent corrosion and oxidation resistance, good formability, durability
and weldability [13]. However, on account of the chemical composition and austenite microstructure, austenitic stainless steel has low
mechanical strength and poor wear resistance, which have been the
main obstacles hindering its application [4]. In addition, oxide scale,
burr and oil contamination, which are formed during the machining
process, are also inconvenient for the subsequent operations such as
welding and assembly.
Sandblasting, as a surface treatment method, is commonly used for
surface modication [5], surface strengthening [6], surface clearing
and rust removal [7]. Compared with shot peening, it has smaller
abrasive particles and lower pressure of compressed air. Therefore, a
smoother surface can be obtained by sandblasting in general. During
the sandblasting process, the surface of specimens is blasted repeatedly
by sand particles or other hard particles with high speed, which leads to
the removal of oxide scale and generation of local plastic deformation in
the surface layer. Meanwhile, it often results in a compressive residual
stress in the subsurface region. Multigner et al. [8] reported that severe
plastic deformation and grain renement were generated after
sandblasting in the surface layer of Low Vacuum Melting 316 stainless
steel (316LVM). Besides, -martensite phase transformation was also
Corresponding author.
E-mail address: mstsun@sdu.edu.cn (J. Sun).

http://dx.doi.org/10.1016/j.matdes.2015.08.113
0264-1275/ 2015 Elsevier Ltd. All rights reserved.

discovered in the sandblasting affected zone. Wang et al. [9] investigated the mechanical and electrochemical behaviors of 304 stainless steel
after sandblasting. They found that nanocrystalline surface layer was
formed after sandblasting and the surface was signicantly enhanced.
However, the corrosion resistance decreased dramatically compared
with the as-received 304 stainless steel. As for 316L stainless steel, few
relevant investigations have been done in the past few years. According
to the previous study on 304 stainless steel, the sandblasting process
may also result in the decrease of corrosion resistance of 316L stainless
steel. Thus, how to improve the corrosion resistance is a critical issue for
the application of sandblasting in stainless steel.
It has been reported that a variety of factors, such as surface morphologies, strain-induced -martensite phase and dislocation density,
would inuence the corrosion resistance of stainless steel [1012].
Therefore, it is reasonable to assume that one method, which is able to
change the surface conditions, has the ability to change the corrosion resistance. Acid pickling and passivation are common surface treatment
methods applied in stainless steel. Acid pickling is used to remove the
surface impurities, such as rust, oxidation scale and stains [13,14]. And
passivation is the chemical treatment with a mild oxidant, such as a
nitric acid solution, for the purpose of the removal of free iron or other
foreign matters and enhancing the spontaneous formation of the
protective passive lm [13,14]. Both acid pickling and passivation
could inuence the surface conditions, hence, it is considered feasible
to change (improve) the corrosion resistance after sandblasting with
acid pickling and passivation. However, as far as the author knows, little
research has attempted to investigate this method to improve the
corrosion resistance of sandblasted 316L stainless steel by far.

S. Geng et al. / Materials and Design 88 (2015) 17

Table 1
Chemical composition (wt.%) of the employed 316L stainless steel.
Component

Si

Mn

Cr

Ni

Mo

Fe

Base metal

0.023

0.54

1.65

0.032

0.01

19.00

13.60

2.48

Bal.

In the present work, the microstructure, surface morphologies


and corrosion behavior of different surface-treated 316L stainless steel
samples, i.e. as-received (AR) sample, sandblasted (SB) sample and
sandblasted + acid pickled + passivated (SBPP) sample, were investigated. Microstructure of the surface layer was studied by an optical
microscope, scanning electron microscope (SEM) and X-ray diffraction
(XRD). To evaluate the corrosion behavior, the potentiodynamic
polarization measurements of each sample were performed according
to ASTM standard G-61 [15]. Besides, surface morphologies before
and after the electrochemical measurements were studied by SEM.
This work provides a novel method to improve the corrosion resistance
of sandblasted 316L stainless steel with maintaining the benets
such as high microhardness brought by sandblasting. This method
(sandblasting + pickling + passivation) would be a promising
approach for surface strengthening in the future.
2. Materials and method
The base metal employed in this study was a commercial 316L stainless steel sheet of 4 mm thickness and the chemical composition is
shown in Table 1. The abrasive used for sandblasting was a mixture of
80% micro glass beads and 20% quartz sand within 80100 mesh. The
pressure of compressed air was about 320 psi. After sandblasting, the
workpiece was cut into two portions in equal-size by wire-cut Electrical
Discharge Machining (WEDM). Then, one part was pickled by a
commercial pickling solution (HNO3: 1525 vol%, HF: 510 vol%, H2O:
6580 vol%) within 1015 min and subsequently passivated by a

commercial passivation paste (HNO3: 30 mL, K2CrO7: 28 mg, bentonite: 100200 mesh) within 2030 min..
To observe the microstructure of the surface layer and cross-section,
each sample was grinded up to 1500 mesh abrasive paper and polished
by 1.5 m diamond paste. Then, all samples were electrochemically
etched by 10% oxalic acid electrolyte at an etching potential of 6 V(SCE)
for 60 s to reveal the grain boundaries. Before observation, all samples
were cleaned ultrasonically with ethanol, rinsed by distilled water and
dried in hot air. In addition, the microhardness on the surface of each
sample was tested by a Digital Microhardness Tester with a load of
200 g and a holding time of 10 s. Each sample was measured at least
ve times and the mean value was set as the nal result.
X-ray diffraction (XRD) line proles were measured on the
surface using an X-ray diffractometer with Cu K radiation ( =
0.154056 nm) and the scan rate was set as 6 /min with a scan step of
0.02.
For the electrochemical measurements, specimens with the dimensions of 10 10 mm were cut off from AR, SB and SBPP samples by
WEDM. Then each sample was embedded in polyester resin with an exposed area of 1 cm2 after establishing the electrical contact. A conventional three-electrode (reference, counter and working electrodes)
electrochemical cell was used in this work. The saturated calomel electrode (SCE) was used normally as the reference electrode and all potentials in this study referred to it. The working electrode was the specimen
and the platinum foil was used as the counter electrode. Potentiodynamic polarization measurements were performed in 3.5 wt.% NaCl
solution at 25 C. After immersion, the open circuit potential (OCP)
was obtained within 30 min and the sweep rate was set as 1 mV/s,
starting from a potential of 800 mV below OCP to the potential at
which the current density reached 10 mA/cm2. Each test was carried
out at least three times to ensure the reproducibility. Besides, surface
morphologies before and after the electrochemical measurements
were observed by SEM.

Fig. 1. Microstructure of the surface layer in different samples: (a) AR, (b) SB, (c) SBPP, and (d) SEM image of SB sample.

S. Geng et al. / Materials and Design 88 (2015) 17

3. Results and discussion


3.1. Microstructure characterization
Fig. 1(a, b and c) show the microscopic images of AR, SB and SBPP
samples taken by optical microscopy. As shown in these gures, all samples had approximately the same grain sizes, while SB and SBPP samples
contained a large number of slip bands in the grains. It should be noted
that there are a few slip bands in the AS sample and this may be attributed to the mechanical polishing during the preparation of metallographic sample. Compared with the surface of 316L stainless steel
resulted from the shot peening process in the previous work [16], the
slip bands were much fewer because of a lower impact force in the
sandblasting process. Furthermore, it was observed that the amount of
slip bands in the grains in SBPP sample is almost as same as those in
SB sample, indicating the subsequent acid pickling and passivation
would not lead to the disappearance of slip bands. Besides, no obvious
strain-induced -martensite was observed in all samples and this
would be discussed in detail in the following XRD analysis. The SEM
image of slip bands after sandblasting was shown in Fig. 1d. It was
found that the number of activated slip systems was distinguished in
different areas. As Fig. 1d shows, only one slip system was activated in
area 1, while three slip systems were activated in area 2, which primarily owed to the uneven distributions of impact force and local plastic
deformation [16].
To evaluate the thickness of deformed region in different samples,
the SEM images were taken at the cross-section. As shown in Fig. 2a,
no obvious deformation was observed in AR sample, whereas severely
deformed regions of about 55 m and 45 m thickness were generated
in SB (Fig. 2b) and SBPP (Fig. 2c) samples respectively. This difference in
thickness (about 10 m) between SB and SBPP samples was obviously
caused by the subsequent acid pickling and passivation. During the
acid pickling process, the outer layer of the deformed region was eroded
and consequently led to the decrease in thickness. It should be
noted that the concentration of acid solution, pickling time and temperature would affect the thickness of corroded layer. However, it would
not discuss in detail in this work and further study would be done in
the future.
3.2. Microhardness
The microhardness in the surface layer of different samples is shown
in Fig. 3. As can be seen, SB and SBPP samples had approximately the
equal microhardness which was signicantly higher than AR sample.
Microhardness values are very sensitive to dislocation density which
increases after plastic deformation [17]. In the previous work, the
microhardness can be expressed in terms of dislocation density and
the microhardness value H can be calculated by the following equation
[18,19]:

p
H H0 aGb

where is the average dislocation density, H0 is the hardness of the corresponding ideal material without any defects, G is the shear modulus,
and b is the scalar value of Burgers vector. It has to be noted that H0, a,
G and b are the material's constants. According to Eq. (1), the microhardness and dislocation density have positive correlation. Thus, the
higher microhardness of SB and SBPP samples signied the higher dislocation density in the surface layer compared with AS sample, which was
in good accordance with the microstructure discussed above. Moreover,
the microhardness of SBPP sample was a little lower than that of SB
sample and this was mainly attributed to a lower dislocation density
in SBPP sample.

Fig. 2. SEM images of the cross-section in different samples: (a) AR, (b) SB, (c) SBPP.

3.3. Line prole analyses


The XRD patterns of AR, SB and SBPP samples are illustrated in Fig. 4,
and the peaks of all patterns have been indexed. For all diffraction proles, the backgrounds were corrected and K2 proles were subtracted.
It can be seen that beside the diffraction peaks of austenite (111) crystal
plane, the peaks of -martensite phase existed in SB and SBPP samples,

S. Geng et al. / Materials and Design 88 (2015) 17

Fig. 3. Microhardness on the surface in different samples: (1) AR, (2) SB, (3) SBPP.

which indicated that the strain-induced -martensite was formed in


the surface layer of SB and SBPP samples.
The full-width at half-maximum (FWHM) in line proles was used
to investigate the dislocation density in many previous works [2022].
In this study, the strongest XRD peak with (111) crystal planes were
chosen for detailed analysis. As shown in Fig. 5, the FWHM of the proles of SB (0.513) and SBPP (0.505) samples were considerably
broader than that of AR (0.421) sample, which indicated that severe
plastic deformations were induced in SB and SBPP samples. Moreover,
the FWHM of SBPP sample was almost the same as SB sample, implying
that they had approximately equivalent dislocation density. However,
the slight difference in FWHM of SB and SBPP samples (about 0.08)
showed that the dislocation density of SB sample was slightly higher
than that of SBPP sample, which would lead to the different microhardness for different samples. This result highly agreed with the above
microstructure and microhardness discussion. In addition, it was
found that the peak position of SB and SBPP samples obviously shifted
to the higher 2 angle compared with AR sample due to the residual
compressive stress which arose from sandblasting. Similar results
have been found in a low-alloy steel after shot peening in Trko's
study [17]. In particular, the shift of SBPP sample was less than SB

Fig. 4. XRD patterns of AR, SB and SBPP samples.

Fig. 5. Normalized k1 diffraction prole of (111) peak of all samples.

sample, indicating a lower residual compressive stress after acid


pickling and passivation. To ensure that the peak shift was not caused
by the measurement error, each sample was measured at least three
times.
3.4. Pitting corrosion resistance
To investigate the corrosion behavior of different samples, the
potentiodynamic polarization measurements were carried out in
3.5 wt.% NaCl solution at 25 C. The potentiodynamic polarization curves
are shown in Fig. 6 and the corrosion parameters such as corrosion
potential (Ecorr), passivation range, pitting potential (Ep) and average
passive current density (ipass), are listed in Table 2.
Ecorr was used to reect the corrosion tendency of different samples
and generally, a higher corrosion potential implied a lower corrosion
tendency. As shown in Table 2, the Ecorr of AR, SB and SBPP samples
was 432, 556 and 530 mV(SCE) respectively. Compared to AR
sample, the Ecorr of SB and SBPP samples exhibited a cathodic shift,
showing a deleterious inuence of sandblasting on the corrosion resistance of 316L stainless steel. Parallel phenomena have been reported
in various materials after shot peening [2325]. However, there was
an anodic shift in Ecorr of SBPP sample with respect to SB sample. From
the perspective of passivation range and pitting potential, the corrosion
resistance of SB sample was considerably poorer than AR sample, while
as for SBPP sample, the performance was superior to AR sample unexpectedly. The ipass value is the average current density in the passivation
region. In particular, it was calculated by the passive current whose potential ranged from 0.2 V(SCE) to 0.2 V(SCE) in this study (shown in
Fig. 6). The ipass results suggested that AR sample had the best performance, followed by SBPP sample, and SB sample was still the worst. In
conclusion, SB sample had the poorest corrosion resistance denitely,
while the comparison of AR and SBPP samples was controversial due
to the diverse performances from different perspectives.
It is well-known that surface morphologies have a great inuence on
corrosion resistance and especially, a smoother surface is considered to
offer a higher corrosion resistance [2628]. The surface morphologies of
different samples before and after the potentiodynamic polarization
measurements are shown in Fig. 7. As can be seen, the surface of AR
sample was consisted of approximately the equal area of concave and
convex (Fig. 7a). Nevertheless, these concave and convex disappeared
after sandblasting because of the severe plastic deformation during
the sandblasting process. Besides, a large amount of foreign matters,
which was embedded in the 316L stainless steel matrix, was observed
as shown in Fig. 7c. Energy dispersive spectroscopy (EDS) linked to a
SEM system showed that they had the same chemical composition as
the employed glass beads. Therefore, we believed that the foreign
matters, embedded in the 316L stainless steel matrix, were glass
chippings which were generated at the moment the high-speed glass
beads impacted on the matrix. Fig. 7e shows the surface morphologies

S. Geng et al. / Materials and Design 88 (2015) 17

Fig. 6. Potentiodynamic polarization curves of different samples in 3.5 wt.% NaCl solution
at 25 C.

of SBPP sample and it was found that there were few glass chippings remaining on the surface after acid pickling and passivation. The surface
morphologies of AR, SB and SBPP samples after electrochemical tests
are shown in Fig. 7(b, d and f) respectively. As for AR sample, the pitting
was located at the boundaries of concave and convex (Fig. 7b). However, the glass chippings were considered as the main factor leading to
pitting in SB sample (Fig. 7d). To verify it, the chemical composition of
the unknown substance located at the center of the pit was tested by
EDS and results revealed that it had the similar chemical composition
as the glass beads used in this work. For SBPP sample, the glass chippings lost their leading role on the corrosion resistance because few of
them remained after the subsequent acid pickling and passivation.
According to previous study, the presence of certain anions such as
chloride incorporated in the passive lm or adsorbed on the surface
layer could have negative effects on the lm stability and results in pit
initiation [2931]. Several theories have been proposed to account for
the breakdown of the passive lm, such as complex ion formation theory, ion penetration theory and chemico-mechanical theory [32]. Take
the complex ion formation theory for example, it was assumed that a
small number of Cl ions jointly adsorb around a cation in the lm surface and formed a high energy complex which would readily dissolve
into the solution. The lm was thus made thinner locally. To better understand the effects of surface morphologies on the corrosion behavior,
the schematic illustration of the process triggering the pitting corrosion
of different samples was shown in Fig. 8. As discussed above, the surface
of AR sample was composed of concave and convex (Fig. 8a), leading to
the boundaries between them relatively protruding sites which were
supposed to easily result in the concentration of Cl. This may bring
about the breakdown of passive lm. As to SB sample, the interface
between the glass chippings and 316L stainless steel matrix was easier
to attack pitting due to the concentration of Cl. Unlike the boundary
between concave and convex in AR sample, the gaps between the
glass chippings and matrix could be regarded as earlier defects which
had broken the continuity and integrity of the passive lm before
pitting. The existence of these defects may lead to a high ipass and a
low Ep during the anodic polarization. In SBPP sample, there were few
so-called boundaries or gaps on the surface. Therefore, the passive
lm was expected to have a better stability which would result in a
low ipass and a high Ep during the anodic polarization theoretically. But
in fact, the ipass in SBPP sample was higher than that of AR sample,

which may arise from the formation of strain-induced -martensite


phase and the increase of dislocation density.
It is clear that the corrosion resistance of different samples cannot be
explained only in terms of surface morphologies and other factors such
as strain-induced -martensite phase and dislocation density induced
during the sandblasting should also be considered [23,33]. The inuence
of martensite phase on the corrosion resistance of stainless steel has
been investigated by many researchers [11,34,35]. Balusamy et al. [24]
pointed out that the formation of martensite was likely to cause a
cathodic shift in Ecorr and this trend was in good accordance with our
observations. Alvarez et al. [36] have studied the effect of straininduced martensite on the anodic dissolution of austenite stainless
steels (304 and 316 stainless steel) in 0.5 M HCl + 2 M H2SO4. They
observed that selective anodic dissolution of the martensite phase happened at potentials closer to Ecorr, but it could no longer be observed at
higher anodic potentials. Thus, it was believed that martensite phase
seemed to have a lesser effect on corrosion resistance at potentials
higher than Ecorr. In other words, the martensite phase had no signicant inuence on the increase of ipass in SB and SBPP samples. Dislocations generated during deformation were also a crucial factor which
inuenced the corrosion resistance in stainless steel. It had been
demonstrated that dislocations decreased the electron work function
and reduced the energy barrier for electrochemical reactions. Furthermore, the increase of dislocation density would lead to a large number
of active sites and promote the corrosion rate [37,38]. It should be
noted that the corrosion rate was reected by ipass in this work.
Therefore, the ipass of SB and SBPP samples was higher than AR sample
primarily due to the increase of dislocation density after sandblasting.
4. Conclusions
In the present work, the microstructure, surface morphologies and
corrosion behavior of 316L stainless steel processed by SB and SBPP
were investigated. The conclusions can be given as follows:
The grain size in different samples was almost the same, while
substantial slip bands were generated in SB and SBPP samples due to
the severe plastic deformation in sandblasting process. Strain-induced
-martensite phase was also observed in SB and SBPP sample. Besides,
SB and SBPP samples had approximately the same microhardness
which was signicantly higher than the AR sample, which was mainly
attributed to the increase in dislocation density after sandblasting. As
for SBPP samples, the subsequent acid pickling and passivation after
sandblasting had little inuence on the microhardness and dislocation
density of the surface layer.
The corrosion resistance of different samples was greatly inuenced
by surface morphologies, formation of strain-induced -martensite
phase and dislocation density. The process triggering the pitting corrosion in different samples had a signicant difference. In AR sample,
boundaries between concave and convex were the weaker sites which
were easy to suffer pitting attack. As for SB sample, the corrosion resistance was severely impaired by the glass chippings embedded in the
316L stainless steel due to their deleterious effect on the integrality of
passive lm. However, the glass chippings lost their leading role on
corrosion resistance because few of them remained after acid pickling
and passivation. In addition, the change in corrosion potential and
average passive current density primarily owed to the formation
of strain-induced -martensite phase and the dislocation density
respectively according to this work and a previous work.

Table 2
Corrosion parameters of different samples evaluated by potentiodynamic polarization tests in 3.5 wt.% NaCl solution at 25 C.
Samples

Corrosion potential Ecorr (mV(SCE))

Passivation range (mV(SCE))

Pitting potential Ep (mV(SCE))

Average passive current density ipass (A/cm2)

AR
SB
SBPP

432 7
556 10
530 9

0.295 to +322 (617 mV)


0.286 to +291 (577 mV)
291 to +351 (642 mV)

322 9
291 13
351 10

11.57
12.84
12.18

S. Geng et al. / Materials and Design 88 (2015) 17

Fig. 7. Surface morphologies of different samples: (a) AR before electrochemical test, (b) AR after electrochemical test, (c) SB before electrochemical test, (d) SB after electrochemical test,
(e) SBPP before electrochemical test, (f) SBPP after electrochemical test.

Fig. 8. Schematic illustration of the process triggering pitting corrosion in different samples: (a) AR, (b) SB.

S. Geng et al. / Materials and Design 88 (2015) 17

Acknowledgments
The authors gratefully acknowledge the support provided by the
Natural Science Foundation of Shandong Province (ZR2010EM067)
and Innovation Fund for Technology Based Firms (11C26213702324).
References
[1] D. Kianersi, A. Mostafaei, A.A. Amadeh, Resistance spot welding joints of AISI 316L
austenitic stainless steel sheets: phase transformations, mechanical properties and
microstructure characterizations, Mater. Des. 61 (2014) 251263.
[2] D. Katherasan, P. Sathiya, A. Raja, Shielding gas effects on ux cored arc welding of
AISI 316L (N) austenitic stainless steel joints, Mater. Des. 45 (2013) 4351.
[3] M. Eskandari, A. Zarei-Hanzaki, H.R. Abedi, An investigation into the room
temperature mechanical properties of nanocrystalline austenitic stainless steels,
Mater. Des. 45 (2013) 674681.
[4] X.H. Chen, J. Lu, L. Lu, K. Lu, Tensile properties of a nanocrystalline 316L austenitic
stainless steel, Scr. Mater. 52 (2005) 10391044.
[5] R.K. Chintapalli, F.G. Marro, E. Jimenez-Pique, M. Anglada, Phase transformation and
subsurface damage in 3Y-TZP after sandblasting, Dent. Mater. 29 (2013) 566572.
[6] R.K. Chintapalli, A. Mestra Rodriguez, F. Garcia Marro, M. Anglada, Effect of
sandblasting and residual stress on strength of zirconia for restorative dentistry
applications, J. Mech. Behav. Biomed. 29 (2014) 126137.
[7] A. Raykowski, M. Hader, B. Maragno, J.K. Spelt, Blast cleaning of gas turbine
components: deposit removal and substrate deformation, Wear 249 (2001)
126131.
[8] M. Multigner, E. Frutos, J.L. Gonzlez-Carrasco, J.A. Jimnez, P. Marn, E.J. Ib,
Inuence of the sandblasting on the subsurface microstructure of 316LVM stainless
steel: implications on the magnetic and mechanical properties, Mater. Sci. Eng. C 29
(2009) 13571360.
[9] X.Y. Wang, D.Y. Li, Mechanical and electrochemical behavior of nanocrystalline
surface of 304 stainless steel, Electrochim. Acta 47 (2002) 39393947.
[10] M. Laleh, F. Kargar, Suppression of chromium depletion and sensitization in
austenitic stainless steel by surface mechanical attrition treatment, Mater. Lett. 65
(2011) 19351937.
[11] L. Peguet, B. Malki, B. Baroux, Effect of austenite stability on the pitting corrosion
resistance of cold worked stainless steels, Corros. Sci. 51 (2009) 493498.
[12] T. Bai, K. Guan, Evaluation of stress corrosion cracking susceptibility of
nanocrystallized stainless steel 304L welded joint by small punch test, Mater. Des.
52 (2013) 849860.
[13] A A. Standard Specication for Chemical Passivation Treatments for Stainless Steel
Parts, 2005.
[14] A A. Standard Practice for Cleaning, Descaling, and Passivation of Stainless Steel
Parts, Equipment, and Systems, 2005.
[15] G A. Standard Test Method for Conducting Cyclic Potentiodynamic Polarization
Measurements for Localized Corrosion Susceptibility of Iron-, Nickel-, or
Cobalt-Based Alloys, 2009.
[16] P. Peyre, X. Scherpereel, L. Berthe, C. Carboni, R. Fabbro, G. Branger, et al., Surface
modications induced in 316L steel by laser peening and shot-peening. Inuence
on pitting corrosion resistance, Mater. Sci. Eng. A 280 (2000) 294302.
[17] L. Trko, O. Bokvka, F. Nov, M. Guagliano, Effect of severe shot peening on
ultra-high-cycle fatigue of a low-alloy steel, Mater. Des. 57 (2014) 103113.

[18] E. Ganin, Y. Komem, A. Rosen, Shock induced hardness in -iron, Mater. Sci. Eng. 33
(1978) 14.
[19] Chu JP, Rigsbee JM, Bana G, Elsayed-Ali HE. Laser-shock processing effects on
surface microstructure and mechanical properties of low carbon steel. Mater. Sci.
Eng. A 1999;260:2608.
[20] T. Ungr, A. Borbly, The effect of dislocation contrast on x-ray line broadening: a
new approach to line prole analysis, Appl. Phys. Lett. 69 (1996) 31733175.
[21] T. Ungr, G. Tichy, the effect of dislocation contrast on X-ray line proles in
untextured polycrystals, Phys. Status Solidi 171 (1999) 425434.
[22] M. Kumagai, K. Akita, Y. Itano, M. Imafuku, S. Ohya, X-ray diffraction study on
microstructures of shot/laser-peened AISI316 stainless steel, J. Nucl. Mater. 443
(2013) 107111.
[23] T. Balusamy, S. Kumar, T.S.N. Sankara Narayanan, Effect of surface nanocrystallization
on the corrosion behaviour of AISI 409 stainless steel, Corros. Sci. 52 (2010) 38263834.
[24] T. Balusamy, T. Sankara Narayanan, K. Ravichandran, I.S. Park, M.H. Lee, Inuence of
surface mechanical attrition treatment (SMAT) on the corrosion behaviour of AISI
304 stainless steel, Corros. Sci. 74 (2013) 332344.
[25] J.D.S. Peltz, L.V.R. Beltrami, S.R. Kunst, C. Brandolt, C.D.F. Malfatti, Effect of the shot
peening process on the corrosion and oxidation resistance of AISI430 stainless
steel, Mater. Res. 18 (2015) 538545.
[26] V. Barranco, E. Onofre, M.L. Escudero, M.C. Garca-Alonso, Characterization of
roughness and pitting corrosion of surfaces modied by blasting and thermal
oxidation, Surf. Coat. Technol. 204 (2010) 37833793.
[27] V. Azar, B. Hashemi, Y.M. Rezaee, The effect of shot peening on fatigue and corrosion
behavior of 316L stainless steel in ringer's solution, Surf. Coat. Technol. 204 (2010)
35463551.
[28] P. Chui, K. Sun, C. Sun, X. Yang, T. Shan, Effect of surface nanocrystallization induced
by fast multiple rotation rolling on hardness and corrosion behavior of 316L
stainless steel, Appl. Surf. Sci. 257 (2011) 67876791.
[29] C. Olsson, D. Landolt, Passive lms on stainless steelschemistry, structure and
growth, Electrochim. Acta 48 (2003) 10931104.
[30] I. Olefjord, L. Wegrelius, Passivation of metals and semiconductors surface analysis
of passive state, Corros. Sci. 31 (1990) 8998.
[31] V. Mitrovic-Scepanovic, B. MacDougall, M.J. Graham, Nature of passive lms on
Fe26Cr alloy, Corros. Sci. 24 (1984) 479490.
[32] L.F. Lin, C.Y. Chao, D.D. Macdonald, A point defect model for anodic passive lms II.
Chemical breakdown and pit initiation, J. Electrochem. Soc. 128 (1981) 11941198.
[33] B. Hashemi, M. Rezaee Yazdi, V. Azar, The wear and corrosion resistance of shot
peenednitrided 316L austenitic stainless steel, Mater. Des. 32 (2011) 32873292.
[34] S. Ningshen, U.K. Mudali, Pitting and intergranular corrosion resistance of AISI type
301LN stainless steels, J. Mater. Eng. Perform. 19 (2010) 274281.
[35] A. Barbucci, M. Delucchi, M. Panizza, M. Sacco, G. Cerisola, Electrochemical and
corrosion behaviour of cold rolled AISI 301 in 1 M H2SO4, J. Alloys Compd.
317318 (2001) 607611.
[36] S.M. Alvarez, A. Bautista, F. Velasco, Inuence of strain-induced martensite in the
anodic dissolution of austenitic stainless steels in acid medium, Corros. Sci. 69
(2013) 130138.
[37] S. Yin, D.Y. Li, R. Bouchard, Effects of strain rate of prior deformation on corrosion
and corrosive wear of AISI 1045 steel in a 3.5 Pct NaCl solution, Metall. Mater.
Trans. A 38 (2007) 10321040.
[38] W. Li, D.Y. Li, Variations of work function and corrosion behaviors of deformed
copper surfaces, Appl. Surf. Sci. 240 (2005) 388395.

You might also like