Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Accepted Manuscript

Magnetoelectropolishing treatment for improving the oxidation resistance of 316L


stainless steel in pressurized water reactor primary water

Qian Xiao, Zhanpeng Lu, Junjie Chen, Jiarong Ma, Qi Xiong, Hongjuan Li, Jian Xu,
Tetsuo Shoji

PII: S0022-3115(18)30599-3
DOI: https://doi.org/10.1016/j.jnucmat.2019.03.028
Reference: NUMA 51520

To appear in: Journal of Nuclear Materials

Received Date: 26 April 2018


Revised Date: 17 February 2019
Accepted Date: 17 March 2019

Please cite this article as: Q. Xiao, Z. Lu, J. Chen, J. Ma, Q. Xiong, H. Li, J. Xu, T. Shoji,
Magnetoelectropolishing treatment for improving the oxidation resistance of 316L stainless steel
in pressurized water reactor primary water, Journal of Nuclear Materials (2019), doi: https://
doi.org/10.1016/j.jnucmat.2019.03.028.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Magnetoelectropolishing treatment for improving the oxidation resistance of


316L stainless steel in pressurized water reactor primary water

Qian Xiao 1,2, Zhanpeng Lu1,2,*, Junjie Chen1.2, Jiarong Ma1,2, Qi Xiong1,2, Hongjuan
Li1,2, Jian Xu3, Tetsuo Shoji3

PT
1. Institute of Materials, School of Materials Science and Engineering, Shanghai
University, 149 Yanchang Road, Shanghai 200072, China

RI
2. State Key Laboratory of Advanced Special Steels, Shanghai University, 149
Yanchang Road, Shanghai 200072, China

SC
3. New Industry Creation Hatchery Center, Tohoku University, Sendai 980-8579,

U
Japan AN
*Corresponding author: Zhanpeng Lu, zplu@t.shu.edu.cn
M
D
TE
C EP
AC

1
ACCEPTED MANUSCRIPT

ABSTRACT: The surfaces of 316L stainless steel with mechanical polishing (MP) or
electropolishing (EP) in the presence of a magnetic field (MEP) before and after
exposure to simulated pressurized water reactor (PWR) primary water were
characterized by various techniques such as SEM, Raman spectroscopy, XPS and
TEM measurements. The surface microhardness was lower and the surface roughness

PT
was higher for the as-treated MEP specimen than for the as-treated MP specimen.
The oxide films that formed on the as-treated MP and MEP surfaces were nanometer-

RI
thick and rich in chromium. The Cr/(Fe+Ni) ratio of the as-treated MEP surface film
was higher than that of the as-treated MP surface film. After exposure to simulated

SC
PWR primary water at 310 oC for 1050 h, a dark double oxide layer that was
approximately 662 nm thick formed on the MP surface, while the MEP surface

U
retained its original brightness as before exposure with a nanometer-thick oxide layer
AN
with a high Cr content. The MP and MEP specimens showed different oxidation
kinetics when exposed to high temperature water. The high Cr/(Fe+Ni) ratio of the
M

surface film on the MEP 316L SS contributed to its excellent oxidation resistance in
simulated PWR primary water.
D

KEYWORDS: Stainless steel, Pressurized water reactor, Primary water,


TE

Magnetoelectropolishing, Oxide film


C EP
AC

2
ACCEPTED MANUSCRIPT

1. Introduction
Many austenitic stainless steel (SS) components are used in contact with the
primary water in pressurized water reactor (PWR) plants. The stress corrosion
cracking (SCC) of stainless steel components in certain local areas in PWR has been
observed [1–7]. The nucleation and propagation of localized corrosion, particularly

PT
stress corrosion cracking, is generally recognized to be related to the properties of the
oxide film that formed on the metal surface [1,8]. Localized hardening and high

RI
residual stresses as a result of fabrication and the surface finish are thought to be
important factors affecting the SCC in these components. It is recognized that the

SC
surface properties, including the surface roughness, near-surface microstructure, and
residual surface stress, have important effects on the oxidation behavior and SCC

U
behavior of stainless steels in high temperature water environments [5,9,10]. Many
AN
studies have been conducted on the effects of a variety of surface treatments that were
performed on stainless steels and alloys [8–14]. Electropolishing (EP) has been used
M

to smooth the surface and to perform surface passivation [11–14]. The passivity can
be enhanced by modifying the thickness, morphology, and chemical composition of
D

the surface oxide layer via different treatments [9,14–20]. It has been reported that
TE

electropolishing can remove surface microstrains, thereby benefitting the


homogeneous nucleation of oxides in high temperature water [10,12,15]. Hryniewicz
EP

et al. [16,21] described that EP was usually selected as an extensive finishing


treatment because it leads to the formation of a particular surface film due to the
selective dissolution of specific elements. Selective dissolution enriches the passive
C

layer in the element in which oxide is the most corrosion resistant. EP has been used
AC

to improve the oxidation resistance of stainless steel in a PWR environment


[10,11,15]. The previous experimental results [10] and reported data [12,15,20] show
that EP surfaces were more resistant to oxidation than ground or mechanically
polished (MP) specimens during their exposure to simulated PWR primary water
environments.
Magnetic fields have been introduced in order to solve multiple issues in
electrochemistry [22–26]. Electropolishing in the presence of a magnetic field, which
3
ACCEPTED MANUSCRIPT

is called magnetoelectropolishing (MEP), has been used to further improve the


surface properties of alloys. Hryniewicz et al. [17,21,27–32] investigated the
corrosion behaviors of medical austenitic stainless steel after MEP and other surface
treatments. They showed that MEP increased the Cr content of the surface and
improved the corrosion resistance. It has been reported that the oxide film that

PT
formed on 316L SS in simulated PWR primary water has a double layer structure
with iron-rich spinel as the outer layer and Cr-rich spinel oxide as the inner layer

RI
[1,10,12,19,20,33,34]. Increasing the chromium content produced fine oxide particles
with a higher chromium concentration in the inner layer and decreased the oxidation

SC
rate constant [35]. One way to improve the oxidation resistance of stainless steel in
simulated PWR primary water is to modify the surface chemical compositions, such

U
as increasing the Cr content. Higher Cr content in the surface layer can be obtained
AN
using MEP surface treatment. MEP is used in this work with the goal of achieving
better oxidation resistance of stainless steel in simulated PWR primary water over
M

long-term exposure. The work aims at establishing a new method for mitigating SCC
that is determined by the crack tip oxidation rate kinetics and the crack tip
D

mechanical fields [33,36]. This research investigated the surface morphologies and
TE

the chemical compositions of the oxide films that formed on MP, EP and MEP 316L
SS specimens that were exposed to simulated PWR primary water for 1050 h. The
EP

surfaces of the exposed specimens were characterized using scanning electron


microscopy (SEM), Raman spectroscopy, X-ray photoelectron spectroscopy (XPS)
and transmission electron microscopy (TEM). The as-treated MP and MEP surfaces
C

were characterized in order to be compared to the surfaces that were exposed to


AC

simulated PWR primary water.

2. Experimental
2.1. Materials and specimen preparation
The chemical composition of the 316L SS that is used in this work is given in
Table 1. The samples were cut from a 22-mm-thick plate. The plate was solution-
annealed (SA) at 1100 °C for 2 h, which was followed by water quenching. The size
4
ACCEPTED MANUSCRIPT

of the specimens for exposure was 10 mm × 5 mm × 3 mm. The MP specimens were


prepared by grinding using SiC papers that were up to 3000 grit and were
mechanically polished thereafter using 1-µm diamond paste. Hryniewicz et al. [21,28]
found that the specimen surfaces after magnetoelectropolishing in an electrolyte that
was a mixture of sulfuric acid and orthophosphoric acid under 0.25, 0.3 and 0.35 T

PT
exhibited a higher Cr content than the specimen surfaces after electropolishing. In
this work, the electropolishing and magnetoelectropolishing surface treatments were

RI
performed in a 20% HClO4 + 80% CH3COOH (vol.%) solution. It is expected that
the 0.3 T magnetic field would have a significant effect on the surface properties of

SC
MEP specimens, given the results of Hryniewicz et al. [21,28]. MEP was performed
in a rectangular cell with an external circulation cooling jacket and an applied

U
magnetic field. A magnetic field (B) with a flux density of 0.3 T was generated using
AN
an electromagnet with a DC galvanostatic power source. The magnetic field was
introduced only to the sample side of the rectangular electrochemical cell and was
M

directed parallel to the electropolishing workpiece surface, as shown in Fig. 1. During


the EP and MEP processes, the solution was not stirred, and the solution temperature
D

was kept at 10-20 °C. The electropolishing voltage was kept at 40 V for 60 s during
TE

the EP and MEP, which was according to the electropolishing electrolyte and
parameters in previous works [10,19]. The surface of the 316L SS sample was spot
EP

welded using a polytetrafluoroethylene-(PTFE) insulated 316L SS lead wire for


hanging in the autoclave. Prior to the measurements, all of the specimens were rinsed
in acetone. After the surface treatments without any exposure tests in water, the
C

specimens are named the as-treated specimens.


AC

2.2. Exposure tests in simulated PWR primary water


Both MP and MEP specimens were used for the exposure tests in simulated
PWR primary water. All tests were conducted in a 316 SS autoclave that was
equipped with a water recirculating loop system that has been described in a previous
work [33]. The test solution was simulated PWR primary water containing 1200 ppm
B in H3BO3 and 2 ppm Li in LiOH (ppm refers to the weight percentage, wt.%) at
310 °C and a pressure of ~ 12.2 MPa. The flow rate was ~5 L/h. The simulated PWR
5
ACCEPTED MANUSCRIPT

primary water with DO < 5 ppb and DH ~ 2.65 ppm (30 ml/kg water STP) was
achieved using H2 purging in the make-up water tank and maintaining an H2
overpressure of approximately 0.065 MPa. The specimens after the oxidation tests in
simulated PWR primary water at 310 oC for 1050 h were referred to as the exposed
specimens.

PT
2.3. Surface characterization of as-treated and exposed specimens
The surface roughness of the specimens before exposure was measured by

RI
profilometry using a Surfcorder SE300 surface roughness measuring instrument from
Kosaka Laboratory Ltd. Two measuring directions were examined: the length of the

SC
specimen surface (L-direction) and the width of the specimen surface (W-direction).
More than three lines with lengths of 3.2 mm were measured from each direction in

U
order to obtain the average roughness Ra. The Ra values were determined according
AN
to Japanese Standards Association JIS B0601-1982 [37]. The Vickers hardness (HV)
of the as-treated specimens was measured using a FISCHERSCOPE HM2000
M

microhardness measurement system. To minimize the indentation depth, a load of 5


mN was applied for 20 s. The Vickers hardness of every specimen was measured five
D

times in different regions. The average data were calculated by removing the
TE

maximum and minimum values. The indentation depths of the specimens were 0.16-
0.26 µm. The surface morphologies of the as-treated and exposed specimens were
EP

observed using a Hitachi SU-70 thermal field emission SEM and a JSM-7500F cold
field emission SEM, respectively. The surfaces of the exposed specimens were
analyzed using a LabRam HR800 Raman spectrometer with a laser wavelength of
C

532 nm.
AC

The XPS measurements of the elemental depth profiles of the oxide films were
performed using a PHI Quantum 2000 X-ray photoelectron spectrometer with a
monochromatic Al Kα source. The C1s peak from the adventitious carbon at 285 eV
was used as a reference to correct the charging shifts. Depth profiling was performed
over an area of 2×2 mm under 2 KeV Ar-ion sputtering, and the spectra were
obtained over a 100-µm spot using a focusing X-ray monochromator. The sputtering
rate for the as-treated specimen surfaces and the exposed MEP specimen surface were
6
ACCEPTED MANUSCRIPT

set at 1 nm/min because of the thin surface film, which was 3 nm/min (vs. SiO2) for
the exposed MP surface. The XPS data were calculated using the Multipak Software
from ULVAC-PHI Co. Ltd. The high-resolution XPS spectra were processed using
the xpspeak 4.1 software package.
The cross-section of the oxide film on the exposed MP 316L SS was examined

PT
using TEM. The TEM specimen was prepared using an FEI Helios 600i dual-beam
focused beam (DB-FIB) with Ga ion sputtering after a protective Pt strap was

RI
deposited on the selected oxide films. The TEM and selected area electron diffraction
analyses (SAED) were conducted using a JEOL JEM-2010F TEM equipped with an

SC
energy dispersive spectrometer (EDS) system operating at 200 kV. The TEM has a
high angle annular dark-field (HAADF) detector.

3. Results
U
AN
3.1. As-treated specimen surfaces
The morphologies of the 316L SS surface with the MP, EP and MEP surface
M

treatments before exposure are shown in Fig. 2. The MP specimen surface was
D

smooth with only slight polishing scratches. The grains and grain boundaries were
clearly observed on the EP and MEP specimen surfaces. Figure 3 shows that the
TE

surface roughness Ra increased in the order of MP, EP, and MEP, and the
microhardness decreased in the order of MP, EP, and MEP.
EP

The XPS depth profiles for the as-treated MP, EP and MEP 316L SS surfaces are
shown in Fig. 4. The plots for the depth of 0 were not included in the calculation of
C

the oxide film thickness in order to avoid the effects of air on the oxide surface. The
AC

region where the oxygen level decreased to 50% of its initial value (starting from the
second point) was treated as the interface between the surface film and metal matrix
[10,38]. From this method, thin oxide films were formed on the as-treated MP, EP
and MEP specimens with thicknesses of 2.2 ± 0.2 nm. The surface films on the as-
treated EP and MEP specimens had a higher Cr content and a lower Fe content than
those for the as-treated MP specimen. The elemental relationship was Fe%>Cr%>>Ni%
(at.) in the MP surface oxide. There were Fe%≈Cr%>>Ni% (at.) in the EP and MEP
7
ACCEPTED MANUSCRIPT

surface oxides. The Cr content in the oxide film of the MEP specimen was slightly
higher than that of the EP specimen. There was a Ni-enriched area in the matrix near
the oxide film for each specimen.
Figure 5 shows the Cr/Fe and Cr/(Fe+Ni) atomic percent ratios as a function of
the depth in the surface films on the as-treated specimens. The outermost points were

PT
not included in order to avoid the effects of air on the oxide surface. The Cr/Fe
atomic ratio in the oxide film of the as-treated MEP specimen could exceed 1.0, and

RI
that of the Cr/(Fe+Ni) could exceed 0.8; these ratios were slightly higher than that in
the as-treated EP specimen in the oxide depth range of 1-3.5 nm and considerably

SC
higher than those in the alloy matrix and in the oxide film of the as-treated MP
specimen.

U
Figure 6 shows the high-resolution XPS spectra for the as-treated MP and MEP
AN
specimens at a sputter depth of 1 nm, respectively. The sputtering depth of 1 nm,
which is near the middle of the surface film, was considered to be more
M

representative of the characteristics of the thin surface film that formed on the as-
treated specimens. The peak decompositions of the O1s, Cr2p3/2, Fe2p3/2, Ni2p3/2
D

core level spectra are in agreement with the X-ray photoelectron spectroscopy
TE

handbook [39] and other investigations [20,38–41]. For the O1s peak decomposition,
the binding energies (BEs) at 531.6 eV and 530.4±0.2 eV are assigned to OH- and O2-,
EP

respectively [41]. The O1s core spectra for the as-treated MP and MEP can be
systematically decomposed into OH- and O2- components at a sputter depth of 1 nm.
The relative intensity of the OH- peak in the as-treated MEP surface film decreases
C

compared to the as-treated MP surface film. For the Cr2p3/2 peak decomposition, the
AC

BEs at 576.7±0.1 eV and 574.0 eV are considered to be Cr3+ and Cr0, respectively
[38–40]. The Cr3+ with a BE of 576.7±0.1 eV may be chromia (Cr2O3) [39,42]. For
the Fe2p3/2 peak decomposition, the BE at 709.6±0.4 eV is considered to be Fe2+/
Fe3+, which may be related to a mixed oxide of FeO and Fe2O3 (or Fe3O4), as well as
FeOOH [42]. The BE at 706.9±0.1 eV is considered to be Fe0 [20]. For the Ni2p3/2
peak decomposition, the BEs at 855.3±0.2 eV and 852.9±0.1 eV are considered to be
Ni2+ and Ni0, respectively [40,41]. The weak Ni2+ peak was likely a result of the
8
ACCEPTED MANUSCRIPT

Ni(OH)2 and NiO [39,41]. A strong metallic Ni0 peak was detected on both as-treated
MP and MEP specimens. In conclusion, the surface films of the as-treated MP and
MEP 316L SS are primarily composed of Cr2O3, Fe3O4, and several hydroxides.
3.2. Surfaces of the specimens exposed to PWR primary water
Figures 7 and 8 show the optical and SEM morphologies of the surfaces of the

PT
MP, EP and MEP 316L SS specimens after exposure to simulated PWR primary
water at 310 oC for 1050 h, respectively. The entire surfaces of the exposed MP

RI
specimens were covered with dark oxide films. The exposed EP specimens exhibited
uneven surfaces that were partly bright and partly covered with dark oxides. The

SC
exposed MEP surface retained its original color, showing only localized dark regions
on the edges and side areas. The dark oxide films were observed in the regions that

U
were affected by the spot welding with the lead wire and the clamping of the tweezers
AN
during magnetoelectropolishing. It is supposed that the dark oxides could only
nucleate from these defective regions on the MEP specimens. No dark oxide films
M

could be observed in the other regions on the exposed MEP samples. It is necessary
to control the MEP process in order to reduce the nucleation sites for the dark oxide
D

film. These results showed that the oxidation resistance of MEP specimens in
TE

simulated PWR primary water was higher than that of MP specimens and EP
specimens, thereby showing the advantage of the MEP surface treatment. Grain
EP

boundaries could be observed on the exposed MEP surface and the bright area on the
exposed EP surface, as shown in Figs. 8e and g. The change in contrast over different
the grain orientations after exposure mainly originated from the different grain
C

orientations of the as-treated EP and MEP specimens. The exposed MP specimens


AC

and the dark area on the exposed EP specimens were covered with randomly
distributed large particles and small oxide particles in the outer layer (Fig. 8b and d).
Figure 9 shows the Raman spectrum of the oxide films that formed on 316L SS
with various surface treatments after exposure to hydrogenated PWR primary water
at 310 oC for 1050 h. The Raman spectrum peaks at 328, 478, 570 and 691 cm-1 are
identified to be the characteristic of iron-nickel spinel [33,43]. An analysis of the
bright areas on the exposed EP and MEP specimens did not generate any
9
ACCEPTED MANUSCRIPT

characteristic peak signal in the Raman spectrum.


3.3. Depth profiles and cross-section of the surface films on the exposed
specimens
The XPS depth profiles of the surface films of the MP, EP and MEP specimens
after exposure to PWR primary water at 310 oC for 1050 h are shown in Fig. 10. An

PT
oxide film with a thickness of 662 nm was observed on the exposed MP 316L SS,
which exhibited a higher Fe content and a lower Cr content in the outer oxide film

RI
and a lower Fe content and a higher Cr content in the inner oxide film. The oxide film
on the exposed MEP specimen was approximately 4.0 nm thick, thus permitting

SC
enrichment of the Cr, depletion of the Ni and Fe. An oxide film with a thickness of
618 nm was observed on the dark area of the exposed EP 316L SS while a thin oxide

U
film with a thickness of 5.4 nm was observed on the light area.
AN
Figure 11 shows the TEM cross-sectional morphologies, EDS maps and EDS
line profiles of the oxide film that formed on the MP 316L SS after exposure to
M

simulated PWR primary water at 310 oC for 1050 h. The oxide film has double layers,
as shown in Fig. 11b. The oxide particles in the outer layer had enriched Fe and Ni
D

and depleted Cr. The inner layer had enriched Cr, and there were two sublayers, the
TE

sublayer near the matrix (known as sublayer M) and the sublayer near the outer layer
(known as sublayer O), as shown in Fig. 11c. Sublayer M had a lower Cr content and
EP

a higher Fe content than sublayer O. There was a Ni-depleted region at the interface
of the matrix/oxide. Figure 12 shows the TEM images and the corresponding SAED
patterns of the film on the MP 316L SS after exposure to simulated PWR primary
C

water at 310 oC for 1050 h. The outer oxide particles were spinel, and the inner layer
AC

was nanosized spinel oxides, as shown in Fig. 12a. Oxides with a twin structure were
observed in sublayer M, as shown in pattern III.
Figure 13 shows the high-resolution XPS spectra for the MEP specimens after
exposure to simulated PWR primary water at 310 oC for 1050 h at a sputtering depth
of 2 nm. The sputtering depth of 2 nm was near the middle position of the oxide film
that formed on the exposed MEP specimen. The peak decompositions of the O1s,
Cr2p3/2, Fe2p3/2, and Ni2p3/2 core level spectra were performed as described above.
10
ACCEPTED MANUSCRIPT

The main peak in O1s can be decomposed into two possible peaks, OH- for hydroxide
and O2- for oxide. The Cr2p3/2 spectrum exhibited the main peak related to Cr3+. This
suggests that the oxide is chromia (Cr2O3). The main peak in Fe2p3/2 can be
decomposed into two possible peaks: a strong Fe2+/Fe3+ peak related to Fe3O4 and
FeOOH, and a weak metallic Ni0 peak. The main peak in Ni2p3/2 can be decomposed

PT
into two possible peaks: a weak Ni2+ peak, likely Ni(OH)2 and NiO; and a strong
metallic Ni0 peak. The composition of the oxide film that formed on the MEP 316L

RI
SS after 1050 h exposure to simulated PWR primary water was similar to the surface
film of the as-treated MEP 316L SS and mainly contained Cr2O3, Fe3O4, and some

SC
hydroxides. A weak metallic Fe0 peak was detected on the surface of the as-treated
MEP 316L SS but not in the oxide film of the exposed MEP 316L SS.

U
AN
4. Discussion
The formation mechanisms of a double layer oxide films on stainless steels in
M

high temperature water have been reported [20,44–46]. The schematic diagrams of
the double layer oxide film that was grown on the MP 316L SS and the single layer
D

oxide film that was grown on the MEP 316L SS in simulated PWR primary water are
TE

shown in Fig. 14 with the reported parabolic rate law and the logarithmic rate law for
oxide film growth [35,47]. The outer layer in the double layer oxide is composed of
EP

large spinel oxide particles that are rich in Fe, likely in the form of (Ni, Fe)Fe2O4. The
inner layer with two sublayers (sublayer M and sublayer O) is compact and composed
of nanosized spinel oxides that are rich in Cr, likely in the form of FexCryNizO4. The
C

high Cr content in the sublayer M is due to the higher oxygen affinity of Cr [10,48]
AC

and the preferential outward diffusion of Cr from the matrix to the metal/oxide
interface [49]. The high Fe and low Cr contents in the sublayer O are thought to result
from the preferential diffusion of Fe cations in oxides. The diffusion rate of the
metallic cations in the oxides decreases from Fe2+>Ni2+>>Cr3+ [44]. The Fe and Ni
are preferentially diffused through the inner oxide films to form the outer oxide
particles. As shown in Fig. 11, the sublayer O with a higher Cr content exhibited as a
relatively dark region compared to sublayer M with a lower Cr content, which is
11
ACCEPTED MANUSCRIPT

similar to the dark region with a higher Cr content in the inner oxide film on the
ground and MP 316L SS in simulated PWR primary water at 310 oC that was
reported by Han et al. [10]. The Ni-enrichment at the oxide/matrix interface of the
exposed MP 316L SS is consistent with the reported results [10,19,20]. Ni has the
lowest oxygen affinity and the lowest diffusion rate in the matrix, meaning that the Ni

PT
tends to remain at the oxide/matrix interface [19,48].
The MEP specimen retained its original brightness after 1050 h of exposure to

RI
simulated PWR primary water. The oxide film is thin and estimated as a single layer
film. The high-resolution XPS spectra for the MEP specimens after exposure to

SC
simulated PWR primary water at 310 oC for 1050 h show that the oxide film was a
Cr-rich oxide.

U
The surface film on the as-treated MP 316LSS was a single layer oxide film with
AN
a thickness of 2.1 nm, which changed to a double layer oxide film that was 662 nm
thick after exposure to simulated PWR primary water at 310 oC for 1050 h, as shown
M

in Figs. 4a, 10a, 11 and 12. The surface film on the as-treated MEP specimen was a
single layer oxide film, which grew to 4.0 nm, but retained its single layer
D

characteristics after exposure to simulated PWR primary water at 310 oC for 1050 h,
TE

as shown in Figs. 4b and 10b. The surface roughness can affect the oxidation
behavior but does not significantly contribute to the high oxidation resistance of the
EP

MEP surface due to the higher surface roughness compared to the MP surface with a
low oxidation resistance. There is a relatively low surface cold work for the MEP
surface compared to the MP surface, which also improves the oxidation resistance.
C

However, this is thought to be less significant than that of the surface chemical
AC

composition. EP surface treatment can reduce the surface cold work, but EP
specimens do not have the excellent corrosion resistance of the MEP specimens. The
unique characteristics of the chemical composition of the MEP surface are thought to
be crucial to the high oxidation resistance in simulated PWR primary water.
The Fe content is lower and the Cr content is higher in the surface film on the
as-treated EP 316L SS specimen than that on the as-treated MP 316L SS specimens,
which contributed to the elevated oxidation resistance in simulated PWR primary
12
ACCEPTED MANUSCRIPT

water [10]. The Cr content in the surface film on the as-treated MEP 316L SS
specimen was slightly higher than that on the as-treated EP 316L SS specimen. It has
been reported that EP was usually selected as a finishing treatment, which constituted
a particular surface film according to the selective dissolution of elements such as Fe
and Ni [10,16,21]. Electropolishing in the presence of a magnetic field would result

PT
in the accelerated selective leaching of these magnetic elements from the
nonmagnetic austenitic 316L SS [17,27,28]. According to magnetohydrodynamic

RI
(MHD) theory [22–26,50], a magnetic field can accelerate the interfacial mass
transport rate, which thus increases the limiting diffusion current of the metal

SC
dissolution. Aaboubi et al. [51] reported that the enhancement factor of the limiting
current density due to an applied magnetic field increases with the increase of the

U
limiting current density without a magnetic field. In the current case, applying a
AN
magnetic field during the electropolishing process is expected to enhance the
selective leaching of Fe and Ni that has been reported to occur in electropolishing
M

without a magnetic field.


Hryniewicz et al. [17,27,28] studied the surface composition of AISI 316L SS
D

film after various surface treatments using XPS. The atomic ratio of Cr/Fe in the bulk
TE

is 0.28 and the ratio after abrasive polishing is 0.48, whereas after standard EP it is
0.58 and after MEP is 1.41. The XPS results show that the changes in the
EP

composition of the surface film appear to have a significant influence on the


corrosion behavior of 316L SS [18,28,52]. It has been reported that the as-polished
nickel-base alloys can form oxide films with a double layer structure in simulated
C

PWR primary water [53]. The Cr/Fe atomic ratio in the alloy matrix is 1.61 for Alloy
AC

600 (taking Ni75Cr15Fe10, wt.%), 3.22 for Alloy 690 (taking Ni60Cr30Fe10) and
0.26 for 316L SS (taking Fe71Cr17Ni12). The Cr/(Fe+Ni) ratio in the alloy matrix is
0.20 for Alloy 600 (Ni75Cr15Fe10), 0.48 for Alloy 690 (Ni60Cr30Fe10) and 0.20 for
316L SS (Fe71Cr15Ni12). The Cr-rich inner layer plays the important role in
corrosion resistance. The bulk Alloys 600 and 690 have higher Cr/Fe ratios than that
of the MEP 316L SS surface layer but do not have better corrosion resistance in
simulated PWR primary water. The Cr/(Fe+Ni) ratio is thought to be more
13
ACCEPTED MANUSCRIPT

appropriate for evaluating the corrosion resistance of an alloy in an aqueous solution,


and it is significantly higher in the MEP 316L SS surface layer than in the bulk Alloys
600 and 690. Table 2 shows the calculated Cr/(Fe+Ni) ratio and the Cr content in the
Fe+Cr+Ni of the oxide films before and after exposure according to the XPS results.
The raw data were determined from the positions in the middle of each oxide layer.

PT
In simulated PWR primary water with 30 cc DH at 310 oC, Ni was located at the
potential where the elemental Ni was stable [54,55]. The candidate film formation

RI
reactions on Fe-Cr-Ni alloys in simulated PWR primary water environments are
described in Eqs. (1)-(6). The types of reactions and the resultant oxide film

SC
compositions are affected by the alloy chemical compositions and the electrochemical
potential. The potential-pH diagrams that were calculated at high temperatures can

U
provide the thermodynamic basis for the analysis of the candidate stable oxides [56–
AN
58]. According to the calculation potential for the Ni-NiO equilibrium line [54,55],
the open circuit potential for 316L SS was located in the Ni stable region, thereby
M

excluding the possibility of forming NiO as expressed by reaction (1). The atomic
percent of Fe in the surface films of the as-treated MP 316L SS are considerably
D

higher than that of Cr, thus favoring the preferential formation of an Fe-rich or Ni-
TE

rich outer layer oxide film, as shown in Eqs. (2) and (4) and Fig. 4a. For a higher
Cr/(Fe+Ni) ratio, reaction (3) is better than reaction (2), and reaction (4) should be
EP

less significant than reactions (5) and (6). The as-treated MEP surface has higher
Cr/(Fe+Ni) ratios than the MP and EP surfaces, thus meaning that reactions (3) or (6)
are expected to play important roles in the corrosion resistance in simulated PWR
C

primary water.
AC

Ni + H2O ⇋ NiO + H2(aq) (1)


3Fe + 4H2O ⇋ Fe3O4 + 4H2(aq) (2)
2Cr + 3H2O ⇋ Cr2O3 + 3H2(aq) (3)
Ni + 2Fe + 4H2O ⇋ NiFe2O4 + 4H2(aq) (4)
Ni + 2Cr + 4H2O ⇋ NiCr2O4 + 4H2(aq) (5)
Fe + 2Cr + 4H2O ⇋ FeCr2O4 + 4H2(aq) (6)
A thin oxide layer begins to form in less than a millisecond at low temperatures
14
ACCEPTED MANUSCRIPT

on the fresh metal after surface finishing [59]. The schematic diagrams of the thin
oxide films on the as-treated MP and MEP surfaces are shown in Fig. 14. The Cr
content of the thin oxide film on the as-treated MEP surface was ~45.4% and was
much higher than that for the as-treated MP surface, which was ~26.5%. The kinetic
rate laws of low temperature oxidation are logarithmic, either direct or inverse [60].

PT
The driving force for the logarithmic or inverse logarithmic growth of the film
thickness is provided by the electric field. The growth of an oxide requires ions to

RI
overcome an energy barrier in order to move into and through the oxide [59]. The
thermal energy that is available at room temperature is insufficient to overcome this

SC
barrier. The Cabrera-Mott model that was discussed by Fehlner [61] described the
proposed growth mechanism via the concept of electron tunneling. An electron can

U
penetrate an energy barrier without the requirement for thermal activation. Electrons
AN
tunneling through the oxide are captured by adsorbed oxygen on the oxide surface.
The charge separation that is established between the oxide surface (oxygen anionic)
M

and the metal (metal cationic) establishes an electric field across the oxide.
At high temperatures, the thermal energy is sufficient to overcome the energy
D

barrier, and the ion can move through the oxide. The parabolic kinetics for the oxide
TE

film growth are usually observed at high temperatures [62,63]. The diffusion of
cations and anions through the formed oxide film is the rate control step for metal
EP

oxidation [58]. Terachi et al. [35] suggested that the growth kinetics of the oxide film
on 316 stainless steel in the simulated PWR primary water at 320 oC followed a
parabolic rate law and formed an oxide film with a submicrometer thickness. The
C

growth kinetics of the oxide film on a Ni-16Cr-9Fe (wt.%) alloy in high temperature
AC

water at 325 oC followed a logarithmic rate law [47] and formed an oxide film with a
nanometer thickness. The oxide film grew relatively fast and a double layer oxide
with a thickness of 662 nm formed on the exposed MP surface. The growth of the
oxide film on the MP specimen in simulated PWR primary water followed a
parabolic kinetic rate, as shown in Fig. 14a. For the MEP specimen, a limited growth
of the prior-formed film was observed at a small nanometer scale during the exposure
to simulated PWR primary water. The growth of the oxide film on the MEP specimen
15
ACCEPTED MANUSCRIPT

in PWR primary water probably followed logarithmic kinetics, as shown in Fig. 14b.
The growth of the oxide film on the dark area of the EP specimen probably followed
parabolic kinetics and the growth of the oxide film on the light area of the EP
specimen probably followed logarithmic kinetics. It has been noted that the donor
density decreased with the increase of the Cr content, thereby resulting in a passive

PT
film on the specimen with a higher Cr content. This film has better stability and the
oxide is a better barrier against electron transfer [64,65]. The prior-formed passive

RI
film with a higher Cr content of ~45.5% (at.%) on the as-treated MEP surface acted
as a barrier against the diffusion of metal and oxygen ions. Once the ion can hardly

SC
move through the oxide barrier through diffusion, the electric field provides the
driving force and the growth of the oxide film follows logarithmic or inverse

U
logarithmic kinetics. A high electrical field can be established with a more
AN
complicated layer structure that serves as a barrier for metal dissolution, which, in
turn, leads to logarithmic or inverse logarithmic growth of the film thickness [66].
M

The growth rate of a chromia film is considerably lower than that of a Fe-based spinel
film because the diffusion coefficients of the metal ions are much lower in chromia
D

[44]. The growth rate of the oxide film on the MEP surface in high temperature water
TE

was much lower than on the MP surface. The exposed EP specimens nearly retained
their original brightness while the exposed EP specimens had uneven surfaces. This
EP

may be due to the agitation of the solution by the magnetic field, which makes the
surface composition of the specimen more uniform.
It has been noted that there are two types of oxide films that form on specimen
C

surfaces after exposure to simulated PWR primary water [10]. The Type I film was a
AC

relatively thick and visible oxide that is normally comprised of double layers. The
Type II film was a thin and almost invisible film that is normally composed of a
compact single layer film. The oxide films that formed on the MP specimens after
1050 h of exposure to simulated PWR primary water at 310 oC are considered to be
Type I films. The film that formed on the MEP specimen after exposure was
considered to be a Type II film. The higher Cr content on the MEP surface is expected
to facilitate Type II passivity and retard the transition from Type II passivity to a Type
16
ACCEPTED MANUSCRIPT

I oxide. Thus, the improved corrosion resistance of the MEP surface was derived
from the higher Cr content in the prior-formed surface film. It has been reported that
Type I oxide films could form on the 316L SS ground specimens and MP specimens
after exposure periods shorter than 1050 h [10,19,20,33,34]. It is supposed that the
MEP surface would remain as a Type II film after some longer exposure periods

PT
under ideal conditions. The surface films on the PWR components throughout their
service life could be affected by many variables during plant operations and shut

RI
down periods. More systematic results are required to show the long term
sustainability of the MEP surface in PWR primary water.

SC
Stainless steels owe their superior corrosion resistance in low temperature
aqueous solutions to the formation of chromia films, which are also expected to form

U
during high temperature aqueous corrosion. A relatively high Cr content and more
AN
protective passive film can be achieved by an MEP treatment to accelerate the
selective dissolution of magnetic elements, such as Fe and Ni during electropolishing
M

in the presence of a magnetic field. This film prevents the metals from forming thick
spinel oxides in high temperature water, thereby improving the corrosion and
D

oxidation resistance in high temperature water.


TE

5. Conclusions
EP

The morphologies and chemical compositions of oxide films on as-treated MP


and MEP 316L stainless steel surfaces before and after exposure to simulated PWR
primary water were characterized.
C

1. Thin oxide films formed on the as-treated MP and MEP surfaces. The surface
AC

film thickness was 2.1 nm on the as-treated MP specimen and 2.4 nm on the as-
treated MEP specimen. The Cr/(Fe+Ni) ratio of the as-treated MEP surface film
was higher than that of the as-treated MP surface film.
2. After exposure to simulated PWR primary water at 310 oC for 1050 h, a double
layer submicron-meter thickness oxide film was found on the MP surface,
whereas a single layer nanometer thickness oxide film was formed on the MEP
surface. The oxide film on the MP specimen was a 662 nm thick spinel oxide
17
ACCEPTED MANUSCRIPT

consisting of a Fe-rich outer layer and a Cr-rich inner layer. There were two
sublayers in the inner layer with a higher Cr content and a lower Fe content in the
sub-inner layer near the outer layer rather than near the alloy matrix. The MEP
surface retained its original brightness and the surface film was a Cr-rich oxide
with a thickness of approximately 4 nm.

PT
3. The relatively high Cr/(Fe+Ni) ratio of the as-treated MEP surface as the result of
the accelerated selective dissolution of the ferromagnetic elements, such as Fe

RI
and Ni during electropolishing in the presence of a magnetic field, contributed to
the excellent oxidation resistance in simulated PWR primary water.

SC
Acknowledgements

U
This work was jointly supported by National Natural Science Foundation of
AN
China (NSFC No. 51571138), National Key R&D Program of China No.
2017YFB0703002, National Science and Technology Major Project No.
M

2015ZX06002005, and the “111” Project from State Administration of Foreign


Expert Affairs. The support from the Instrument Analytical and Research Center,
D

Shanghai University is acknowledged.


TE

References
EP

[1] T. Terachi, K. Fujii, K. Arioka, J. Nucl. Sci. Technol. 42 (2005) 225–232.


[2] P.L. Andresen, M.M. Morra, Corrosion. 64 (2008) 15–29.
[3] P.L. Andresen, M.M. Morra, J. Nucl. Mater. 383 (2008) 97–111.
C

[4] Z. Lu, T. Shoji, T. Dan, Y. Qiu, T. Yonezawa, Corros. Sci. 52 (2010) 2547–
AC

2555.
[5] L. Tribouilloy, F. Vaillant, J.-M. Olive, M. Puiggali, Adv. Mater. Sci. 7 (2007)
61–69.
[6] J. Chen, Z. Lu, Q. Xiao, X. Ru, G. Han, Z. Chen, B. Zhou, T. Shoji, J. Nucl.
Mater. 472 (2016) 1–12.
[7] J. Chen, Z. Lu, Q. Xiao, X. Ru, J. Ma, T. Shoji, J. Nucl. Mater. 496 (2017) 313–
324.
18
ACCEPTED MANUSCRIPT

[8] R. Obata, M. Koshiishi, H. Anzai, Correlation between oxide film and stress
corrosion cracking susceptibility of surface cold worked L-grade stainless steels,
in: 14th Int. Conf. Environ. Degrad. Mater. Nucl. Power Syst. Virginia Beach,
VA, August 23-27, 2009, pp. 622–634.
[9] F. Scenini, R.C. Newman, R.A. Cottis, R.J. Jacko, Corrosion. 64 (2008) 824–

PT
835.
[10] G. Han, Z. Lu, X. Ru, J. Chen, Q. Xiao, Y. Tian, J. Nucl. Mater. 467 (2015)

RI
194–204.
[11] A. Baron, W. Simka, G. Nawrat, D. Szewieczek, A. Krzyżak, J. Achiev. Mater.

SC
Manuf. Eng. 18 (2006) 55–58.
[12] S. Cissé, L. Laffont, B. Tanguy, M.C. Lafont, E. Andrieu, Corros. Sci. 56 (2012)

U
209–216.
AN
[13] J. Kaneda, H. Tamako, R. Ishibashi, Effects of surface treatments on
microstructure, hardness and residual sress in type 316l stainless steel, in: 14th
M

Int. Conf. Environ. Degrad. Mater. Nucl. Power Syst. Virginia Beach, VA,
August 23-27, 2009, pp. 791–802.
D

[14] E.-S. Lee, Int. J. Adv. Manuf. Technol. 16 (2000) 591–599.


TE

[15] S.E. Ziemniak, M. Hanson, P.C. Sander, Corros. Sci. 50 (2008) 2465–2477.
[16] R. Rokicki, T. Hryniewicz, Trans. Inst. Met. Finish. 90 (2012) 188–196.
EP

[17] T. Hryniewicz, R. Rokicki, K. Rokosz, Surf. Coat. Technol. 202 (2008) 1668–
1673.
[18] K. Rokosz, J. Lahtinen, T. Hryniewicz, S. Rzadkiewicz, Surf. Coatings Technol.
C

276 (2015) 516–520.


AC

[19] S. Wang, Y. Hu, K. Fang, W. Zhang, X. Wang, Corros. Sci. 126 (2017) 104–120.
[20] Y. Guo, E.H. Han, J. Wang, Mater. Corros. 66 (2015) 670–680.
[21] T. Hryniewicz, K. Rokosz, Mater. Chem. Phys. 123 (2010) 47–55.
[22] A. Bund, S. Koehler, H.H. Kuehnlein, W. Plieth, Electrochim. Acta. 49 (2003)
147–152.
[23] R.A. Tacken, L.J.J. Janssen, J. Appl. Electrochem. 25 (1995) 1–5.
[24] Z. Lu, D. Huang, W. Yang, Corros. Sci. 47 (2005) 1471–1492.
19
ACCEPTED MANUSCRIPT

[25] Z. Lu, D. Huang, W. Yang, J. Congleton, Corros. Sci. 45 (2003) 2233–2249.


[26] L.M.A. Monzon, J.M.D. Coey, Electrochem. Commun. 42 (2014) 38–41.
[27] T. Hryniewicz, K. Rokosz, R. Rokicki, , Corros. Sci. 50(2008) 2676–2681.
[28] T. Hryniewicz, K. Rokosz, Surf. Coat. Technol. 204 (2010) 2583–2592.
[29] T. Hryniewicz, P. Konarski, K. Rokosz, R. Rokicki, Surf. Coat. Technol. 205

PT
(2011) 4228–4236.
[30] T. Hryniewicz, K. Rokosz, Front. Mater. 1 (2014) 1–7.

RI
[31] T. Hryniewicz, K. Rokosz, Mater. Chem. Phys. 122 (2010) 169–174.
[32] T. Hryniewicz, R. Rokicki, K. Rokosz, Corrosion. 64 (2008) 660–665.

SC
[33] J. Chen, Q. Xiao, Z. Lu, X. Ru, H. Peng, Q. Xiong, H. Li, J. Nucl. Mater. 489
(2017) 137–149.

U
[34] Y.J. Kim, Corrosion. 51 (1995) 849–860.
AN
[35] T. Terachi, T. Yamada, T. Miyamoto, K. Arioka, K. Fukuya, J. Nucl. Sci.
Technol. 45 (2008) 975–984.
M

[36] T. Shoji, Z. Lu, H. Murakami, Corros. Sci. 52 (2010) 769–779.


[37] S. Ashokkumar, J. Adler-Nissen, J. Food Eng. 105 (2011) 537–544.
D

[38] A. Machet, A. Galtayries, P. Marcus, P. Combrade, P. Jolivet, P. Scott, Surf.


TE

Interface Anal. 34 (2002) 197–200.


[39] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, Handbook of X ray
EP

photoelectron spectroscopy, 1995.


[40] J. Wang, X. Li, F. Huang, Z. Zhang, J. Wang, R.W. Staehle, Corrosion. 70
(2014) 598–614.
C

[41] Z. Zhang, J. Wang, E.H. Han, W. Ke, Corros. Sci. 53 (2011) 3623–3635.
AC

[42] D. Ramachandran, R. Egoavil, A. Crabbe, T. Hauffman, A. Abakumov, J.


Verbeeck, I. Vandendael, H. Terryn, D. Schryvers, J. Microsc. 264 (2016) 207–
214.
[43] B.D. Hosterman, Raman spectroscopic study of solid solution spinel oxides,
2011, p. 101.
[44] J. Robertson, Corros. Sci. 32 (1991) 443–465.

20
ACCEPTED MANUSCRIPT

[45] R. Soulas, M. Cheynet, E. Rauch, T. Neisius, L. Legras, C. Domain, Y. Brechet,


J. Mater. Sci. 48 (2013) 2861–2871.
[46] B. Stellwag, Corros. Sci. 40 (1998) 337–370.
[47] D. Féron, J.M. Olive, Corrosion issues in light water reactors: Stress corrosion
cracking (EFC 51), CRC Press, 2007, p. 53.

PT
[48] X. Ren, K. Sridharan, T.R. Allen, J. Nucl. Mater. 358 (2006) 227–234.
[49] S.J. Rothman, L.J. Nowicki, G.E. Murch, J. Phys. F Met. Phys. 10 (1980) 383–

RI
398.
[50] Z. Lu, T. Shoji, W. Yang, Corros. Sci. 52 (2010) 2680–2686.

SC
[51] O. Aaboubi, J. Electrochem. Soc. 137 (1990) 1796.
[52] K. Rokosz, T. Hryniewicz, Adv. Mater. Sci. 12 (2012) 19–21.

U
[53] X. Ru, Z. Lu, J. Chen, G. Han, J. Zhang, P. Hu, X. Liang, J. Nucl. Mater. 497
AN
(2017) 37–53.
[54] D.S. Morton, S.A. Attanasio, G.A. Young, Primary water SCC understanding
M

and characterization through fundamental testing in the vicinity of the


nickel/nickel oxide phase transition, in: 10th Int. Conf. Environ. Degrad. Mater.
D

Nucl. Power Syst. – Water React., 2001.


TE

[55] S.A. Attanasio, D.S. Morton, Measurement of the nickel/nickel oxide transition
in Ni-Cr-Fe Alloys and updated data and correlations to quantify the effect of
EP

aqueous hydrogen on primary water SCC, in: 11th Int. Conf. Environ. Degrad.
Mater. Nucl. Power Syst. – Water React., 2003, pp. 143–154.
[56] B. Beverskog, I. Puigdomenech, Corros. Sci. 39 (1997) 43–57.
C

[57] B. Beverskog, I. Puigdomenech, Corros. Sci. 39 (1997) 969–980.


AC

[58] B. Beverskog, I. Puigdomenech, Corros. Sci. 55 (1999) 1077–1087.


[59] F.P. Fehlner, M.J. Graham, Thin oxide film formation on metals, in: P. Marcus
(Ed.), Corrosion mechanisms in theory and practice, Third, CRC press, 2012, pp.
217–220.
[60] F.P. Fehlner, N.F. Mott, Oxid. Met. 2 (1970) 59–99.
[61] F.P. Fehlner, Low-temperature oxidation: the role of vitreous oxides, John
Wiley & Sons New York Ny, 1986, p.148.
21
ACCEPTED MANUSCRIPT

[62] M. Schutze, High-temperature corrosion, in: P. Marcus (Ed.), Corrosion


mechanisms in theory and practice, Third, CRC press, 2012, pp. 578–584.
[63] L. Marchetti, S. Perrin, F. Jambon, M. Pijolat, Corros. Sci. 102 (2016) 24–35.
[64] M. Liu, X. Cheng, X. Li, Y. Pan, J. Li, Appl. Surf. Sci. 389 (2016) 1182–1191.
[65] T. Ohtsuka, M. Abe, T. Ishii, J. Electrochem. Soc. 162 (2015) C528–C535.

PT
[66] H.H. Strehblow, P. Marcus, Mechanisms of pitting corrosion, in: P. Marcus
(Ed.), Corrosion mechanisms in theory and practice, Third, CRC press, 2012, p.

RI
351.

U SC
AN
M
D
TE
C EP
AC

22
ACCEPTED MANUSCRIPT

Figure captions
Fig. 1. Schematic diagram of magnetoelectropolishing system, and the Lorentz force
generated by the magnetic field and electric filed.
Fig. 2. SEM morphologies of the 316L SS surface with (a) MP, (b) EP and (c) MEP
surface treatments before exposure to the PWR primary water.

PT
Fig. 3. Surface roughness and micro-hardness of the 316L SS specimens with MP, EP
and MEP treatments before exposure to the PWR primary water.

RI
Fig. 4. XPS depth profiles of the surface films on the as-treated (a) MP, (b) EP and (c)
MEP specimens.

SC
Fig. 5. (a) Cr/Fe and (b) Cr/( Fe+Ni) atomic percent ratios through the depth profiles
of the surface films on the as-treated MP, EP and MEP specimens.

U
Fig. 6. High-resolution XPS spectra for the as-treated MP and MEP specimens at a
AN
sputter depth of 1 nm.
Fig. 7. Optical morphologies of the 316L SS specimens with various surface
M

treatments after exposure to simulated PWR primary water at 310 oC for 1050 h.
Fig. 8. SEM morphologies of the surface films formed on 316L SS specimens after
D

exposure to simulated PWR primary water at 310 oC for 1050 h, (a)(b) the MP
TE

specimen, (c)(d) the dark area on the EP specimen, (e)(f) the bright area on the EP
specimen, and (g)(h) the bright area on the MEP specimen.
EP

Fig. 9. Raman spectra of the oxide films formed on 316L SS with various surface
treatments after exposure to simulated PWR primary water at 310 oC for 1050 h.
Fig. 10. XPS depth profiles of the surface films after exposure to simulated PWR
C

primary water at 310 oC for 1050 h, (a) MP, (b) EP dark area, (c) EP bright area and
AC

(d) MEP.
Fig. 11. (a) STEM-HAADF image of the oxide film formed on MP 316L SS after
exposure to simulated PWR primary water at 310 oC for 1050h, (b) detail area of (a)
and corresponding EDS maps of the rectangular region for Fe, Cr, Ni, O and (c) EDS
line scan profile shown in (a).
Fig. 12. TEM images and SAED patterns of the oxide film formed on MP 316L SS
after exposure to simulated PWR primary water at 310 oC for 1050h: (a) TEM (bright
23
ACCEPTED MANUSCRIPT

field) image, and the SAED patterns of I, II, and III, (b) high resolution TEM (bright
field) image of the boundary for III and II and (c) high resolution TEM (bright field)
image of the boundary of oxide layer II and the outer layer particles of I.
Fig. 13. High-resolution XPS spectra for the MEP specimens after exposure to
simulated PWR primary water at 310 oC for 1050 h at a sputter depth of 2 nm.

PT
Fig. 14. Schematic diagrams of the oxide films on the MP and MEP 316L SS surfaces
before and after exposure to high temperature water as well as the oxide film growth

RI
rate law, (a) parabolic rate law and (b) logarithmic rate law.

U SC
AN
M
D
TE
C EP
AC

24
ACCEPTED MANUSCRIPT

Table captions

Table 1 Chemical compositions of 316L SS (wt.%).


Table 2 Ratios of Cr/(Fe+Ni) and Cr content (at.%) in the oxide films on the 316L SS
specimens before and after exposure from the XPS data.

PT
RI
U SC
AN
M
D
TE
C EP
AC

25
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 1. Schematic diagram of magnetoelectropolishing system, and the Lorentz force
generated by the magnetic field and electric filed.

U
AN
M
D
TE
C EP
AC

Fig. 2. SEM morphologies of the 316L SS surface with (a) MP, (b) EP and (c) MEP
surface treatments before exposure to the PWR primary water.

26
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 3. Surface roughness and micro-hardness of the 316L SS specimens with MP, EP
and MEP treatments before exposure to the PWR primary water.

U
AN
M
D
TE
C EP
AC

Fig. 4. XPS depth profiles of the surface films on the as-treated (a) MP, (b) EP and (c)
MEP specimens.

27
ACCEPTED MANUSCRIPT

PT
RI
Fig. 5. (a) Cr/Fe and (b) Cr/( Fe+Ni) atomic percent ratios through the depth profiles
of the surface films on the as-treated MP, EP and MEP specimens.

U SC
AN
M
D
TE

Fig. 6. High-resolution XPS spectra for the as-treated MP and MEP specimens at a
sputter depth of 1 nm.
C EP
AC

Fig. 7. Optical morphologies of the 316L SS specimens with various surface


treatments after exposure to simulated PWR primary water at 310 oC for 1050 h.
28
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 8. SEM morphologies of the surface films formed on 316L SS specimens after
exposure to simulated PWR primary water at 310 oC for 1050 h, (a)(b) the MP
specimen, (c)(d) the dark area on the EP specimen, (e)(f) the bright area on the EP
29
ACCEPTED MANUSCRIPT

specimen, and (g)(h) the bright area on the MEP specimen.

PT
RI
SC
Fig. 9. Raman spectra of the oxide films formed on 316L SS with various surface
treatments after exposure to simulated PWR primary water at 310 oC for 1050 h.

U
AN
M
D
TE
C EP
AC

Fig. 10. XPS depth profiles of the surface films after exposure to simulated PWR
primary water at 310 oC for 1050 h, (a) MP, (b) EP dark area, (c) EP bright area and
(d) MEP.
30
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

Fig. 11. (a) STEM-HAADF image of the oxide film formed on MP 316L SS after
D

exposure to simulated PWR primary water at 310 oC for 1050h, (b) detail area of (a)
TE

and corresponding EDS maps of the rectangular region for Fe, Cr, Ni, O and (c) EDS
line scan profile shown in (a).
C EP
AC

31
ACCEPTED MANUSCRIPT

PT
RI
U SC
Fig. 12. TEM images and SAED patterns of the oxide film formed on MP 316L SS
AN
after exposure to simulated PWR primary water at 310 oC for 1050h: (a) TEM (bright
field) image, and the SAED patterns of I, II, and III, (b) high resolution TEM (bright
M

field) image of the boundary for III and II and (c) high resolution TEM (bright field)
image of the boundary of oxide layer II and the outer layer particles of I.
D
TE
C EP

Fig. 13. High-resolution XPS spectra for the MEP specimens after exposure to
AC

simulated PWR primary water at 310 oC for 1050 h at a sputter depth of 2 nm.

32
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 14. Schematic diagrams of the oxide films on the MP and MEP 316L SS surfaces
before and after exposure to high temperature water as well as the oxide film growth
rate law, (a) parabolic rate law and (b) logarithmic rate law.

33
ACCEPTED MANUSCRIPT

Table 1 Chemical compositions of 316L SS (wt.%).


Alloy C Si Mn P S Cr Ni Mo Fe
316L SS 0.023 0.16 1.11 0.015 0.002 17.00 12.27 2.32 Bal.

Table 2 Ratios of Cr/(Fe+Ni) and Cr content (at.%) in the oxide films on the 316L SS
specimens before and after exposure from the XPS data.

PT
State Position in XPS Cr/(Fe+Ni) Cr content
Matrix 10 nm (As-treated MP) 0.18 15.3%

RI
As-treated MP 1 nm 0.36 26.5%
As-treated EP 1 nm 0.83 45.3%

SC
As-treated MEP 1 nm 0.84 45.4%
300 nm (sub-layer M) 0.48 32.4%
Exposed MP

U
500 nm (sub-layer O) 0.41 29.0%
310 nm (dark area) 0.56 35.9%
AN
Exposed EP
2.5 nm (bright area) 1.27 55.9%
Exposed MEP 2 nm 0.74 42.5%
M
D
TE
C EP
AC

34
ACCEPTED MANUSCRIPT

Highlights

316L SS treated by magneto-electropolishing (MEP) under a magnetic field.


About 2 nm thick oxide film on as-treated mechanically polished (MP) and MEP
surfaces.
A double layer sub-micrometer thick oxide on exposed MP surface in PWR

PT
water.
A nanometer-thick oxide of high Cr content on exposed MEP surface in PWR

RI
water.
High Cr/(Fe+Ni) ratio of MEP surface contributed to its excellent oxidation

SC
resistance.

U
AN
M
D
TE
C EP
AC

You might also like