(Paperhub Ir) 10 1016@j Corsci 2016 02 001

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

G Model

CS-6638; No. of Pages 15 ARTICLE IN PRESS


Corrosion Science xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Properties of oxide films formed on 316L SS and model alloys with


modified Ni, Cr and Si contents in high temperature water
Guangdong Han a , Zhanpeng Lu a,b,∗ , Xiangkun Ru a , Junjie Chen a , Jinlong Zhang a ,
Tetsuo Shoji c
a
Institute of Materials, School of Materials Science and Engineering, Shanghai University, 149 Yanchang Road, P.O. Box 269, Shanghai 200072, China
b
State Key Laboratory of Advanced Special Steels, Shanghai University, 149 Yanchang Road, P.O. Box 269, Shanghai 200072, China
c
New Industry Creation Hatchery Center, Tohoku University, 6-6-10 Aramaki Aoba, Aoba-ku Sendai 980-8579, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Oxide films formed on 316L stainless steel, 9Cr–26Ni–1.5Mo, and 9Cr–26Ni–1.5Mo–5Si in de-aerated high
Received 16 August 2015 temperature water have been characterized. High reactivity of Si and high solubility of silicon oxides in
Received in revised form 31 January 2016 high temperature water contribute to Si depletion in the outer oxide layer, Si enrichment in the inner
Accepted 1 February 2016
oxide layer near the metal/oxide interface, and the outer layer of large and continuously packed oxides
Available online xxx
with cavities. A higher Cr content in the alloys favors the formation of a thinner inner oxide layer. Si in
the model alloy affects the distribution of nickel oxide near the metal/oxide interface.
Keywords:
© 2016 Elsevier Ltd. All rights reserved.
A. Stainless steel
B. TEM
B. XPS
C. Oxidation
C. Reactor conditions

1. Introduction Neutron irradiation can cause depletion of Cr and Fe at the grain


boundaries, along with enrichment of Ni and Si [3–5,9,10]. Bruem-
Austenitic stainless steels, commonly used as in-core materials mer [10] has suggested that the Cr content at the grain boundaries
of nuclear power reactors, are susceptible to irradiation-assisted decreased to 8–10 wt.% and the Si and Ni contents increased to
stress corrosion cracking (IASCC), which is a critical issue for 6 wt.% and 30 wt.%, respectively, in several 300-series irradiated
the safe and economical operation of light water reactors. The in LWRs at 300–320 ◦ C up to 65 dpa. The low grain boundary Cr
elevated cracking susceptibility is attributed to both irradiation content may affect the formation of passive film, increasing the
damage in austenitic stainless steel and the aggressive environ- susceptibility to corrosion [5]. The increased susceptibility to IASCC
ment of the reactor coolants under irradiation. Irradiation can in highly irradiated stainless steels is closely related to the segre-
cause microstructural changes (dislocation loops, precipitate, and gation of Si at the grain boundaries. Was et al. [11] have reported
void formation) and microchemical changes (radiation-induced the high susceptibility to IASCC propagation in high Si alloys. Stain-
segregation of solute elements), and these micro-scale changes less steels with 3–5 wt.% Si exhibit high crack growth rates in high
enhance the susceptibility to intergranular stress corrosion crack- temperature water environments [12]. Li et al. [13] reported that Si
ing (IGSCC) [1,2]. Radiation-induced segregation (RIS) at the grain increases stress corrosion cracking (SCC) susceptibility, and they
boundaries in austenitic stainless steels is an important factor proposed three possible mechanisms to explain the Si effect on
that contributes to IASCC [1,3–5]. RIS results in local composition environmentally-assisted crack (EAC) behavior. Three mechanisms
changes at grain boundaries and thereby affects the corrosion resis- proposed by Li et al. are: (1) Si may lower the stacking fault energy,
tance of austenitic stainless steels [6–8]. and therefore facilitating the coplanar dislocation slip and ruptur-
ing the oxide film, (2) Si may decrease the strength of the oxide film,
(3) Si may increase the oxidation tendency of grain boundaries since
Si readily oxidizes. However, Jacobs [14] reported that grain bound-
ary Si enrichment has little impact on IASCC. Fukuya et al. [15]
∗ Corresponding author at: Institute of Materials, School of Materials Science and Chung et al. [16] found that commercial alloys with higher Si
and Engineering, Shanghai University, 149 Yanchang Road, P.O. Box 269, Shanghai contents are less susceptible to IG cracking than high-purity alloys
200072, China. Fax: +86 2156332475. irradiated under the similar conditions.
E-mail addresses: zplu@t.shu.edu.cn, zplu@t.shu.edu.cn (Z. Lu).

http://dx.doi.org/10.1016/j.corsci.2016.02.001
0010-938X/© 2016 Elsevier Ltd. All rights reserved.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
2 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Table 1
Chemical compositions of the test alloys (wt.%).

Material C Si Mn P S Cr Ni Mo Fe

316L SS 0.020 0.54 1.28 0.014 0.004 17.21 12.46 2.45 Bal.
9Cr–26Ni–1.5Mo 0.026 0.35 0.65 0.020 0.013 9.29 26.12 1.54 Bal.
9Cr–26Ni–1.5Mo–5Si 0.023 4.54 0.60 0.021 0.018 9.15 26.41 1.47 Bal.

The changes of chemical composition at the grain boundary can Table 2


Test conditions employed for the oxidation of 316L SS and model alloy specimens
affect the oxidation behavior of austenitic stainless steel in high-
in simulated PWR primary water.
temperature water. The properties of the oxide film formed during
service, including the chemical composition, microstructure and Parameters Range
thickness, play important roles in the cracking process, especially Water chemistry 1200 ppm (by weight) B in boric acid + 2 ppm
at the initial stages [17,18]. The oxides at the crack tips are rele- (by weight) Li in lithium hydroxide
vant to those formed on the alloy surfaces in simulated pressurized Temperature 310 ± 0.5 ◦ C
Pressure 12 MPa
water reactor (PWR) primary water environments, suggesting that
Dissolved oxygen (DO) <5 ppb (by weight)
investigation of surface oxides could benefit further elucidating SCC Dissolved hydrogen (DH) <0.5 ppb (by weight)
mechanisms [19]. The objective of this work is to identify the effect Flow rate 5L/h
of the chemical composition changes, such as the depletion of Cr
and enrichment of Ni and Si as the simulation of RIS, on the oxi-
dation behavior of alloys in high-temperature water. The present SS, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si were about 54.5%,
work investigates the morphologies, cross-section characteristics 58.4%, and 46.3%, respectively.
and chemical composition of the oxide films grown on 316L SS
and model alloys exposed to high-temperature water for 114 h and 2.2. Immersion tests in high-temperature water
1194 h. Raman spectroscopy, X-ray photoelectron spectroscopy
(XPS), scanning electron microscopy (SEM), and transmission elec- The specimens were exposed to simulated PWR primary water
tron microscopy (TEM) are used as the characterization methods. with boric acid and lithium hydroxide for 114 h and 1194 h in a 5-L
(net volume) stainless steel autoclave equipped with a recircula-
tion loop fabricated by Toshin Kogyo, Co., Ltd. All the test specimens
2. Experimental were mounted vertically at the locations close to the center of the
autoclave. The test conditions are summarized in Table 2. The test
2.1. Material and specimen preparation solution in Table 2 is called de-aerated PWR primary water, which
is used for the immersion tests to simulate environments with rela-
The measured chemical compositions of 316L SS and model tively high oxidizing potentials under the irradiation and the crack
alloys are shown in Table 1. The model alloys were made by addi- tip environments. To reduce the metallic cations released from the
tions of pure Ni, Fe and Si in 316L SS, which had a lower Mo autoclave body during the immersion tests, the autoclave was at
content than that of 316L SS. The model alloy 9Cr–26Ni–1.5Mo first tested under similar conditions with the same water chem-
was used to simulate the depletion of Cr and enrichment of istry. After each test, the used PWR water was removed and new
Ni, and model alloy 9Cr–26Ni–1.5Mo–5Si was used to simulate PWR water was prepared for the new test.
the depletion of Cr and enrichment of Ni and Si at the grain
boundaries of irradiated austenitic stainless steel. The Mo con- 2.3. Surface characterization of the tested specimens
tents in the model alloys were lower than in 316L SS, which was
not considered critical in evaluating the difference in oxidation After the immersion tests in high-temperature water environ-
behavior in the present work. Model alloys 9Cr–26Ni–1.5Mo and ments, the surface morphologies of oxide films formed on the
9Cr–26Ni–1.5Mo–5Si were hot-forged and hot-rolled into sheets at immersed specimens were examined using FESEM operated at
1100 ◦ C in the same way. The alloy sheets were solution-annealed at 15 kV. The oxide films were analyzed using an INVIA laser Raman
1100 ◦ C for 20 min and were then quenched in water. Rectangular spectrometer with laser wavelength of 514.5 nm. XPS measure-
specimens (length × width × thickness = 10 mm × 5 mm × 2.5 mm) ments were performed on an ESCALAB 250Xi X-ray photoelectron
were cut from the 316L SS and model alloy sheets. The specimens spectrometer using a monochromatic Al K␣ source. Depth profil-
were mechanically ground with waterproof abrasive paper up to ing was performed over an area of 2.5 mm × 2.5 mm, and spectra
2000 grit successively. were collected within a 500-␮m spot. The sputtering rate was set
Field-emission SEM (FESEM; Apollo 300) along with electron at 0.25 nm/s with reference to the Ta2 O5 layer.
backscattering diffraction (EBSD) was used to characterize the grain Detailed cross-section characteristics were studied using a
boundary characteristics of the model alloys. To reduce the residual transmission electron microscope (TEM). Thin-foil specimens for
stress and strain on the surface, electropolishing was performed TEM observations were prepared using a 600i dual-beam focused
in 20% HClO4 + 80% CH3 COOH electrolyte (vol.%) following pol- ion beam (DB-FIB) with Ga ion sputtering after a protective Pt
ishing with 1 ␮m diamond paste. The acceleration voltage of the layer was deposited on the oxide scale. Large oxide particles were
FESEM beam for the EBSD measurement was 20 kV. The inverse avoided during the TEM specimen preparation by FIB. Hence, large
pole figures (IPF) are shown in Fig. 1. The measured grain size oxide particles were not examined and were absent from the TEM
took into account the twins. The average grain size of the 316L images. High-resolution TEM observations and selected area elec-
SS, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si was approximately tron diffraction (SAED) were performed with a JEM 2010F TEM
19.1, 9.7, and 19.1 ␮m with the present measurement parame- instrument operated at 200 kV and equipped with an energy-
ters, respectively. Grain boundaries with rotation angles between dispersive spectroscopy (EDS) system. The K-series was used for
15–180◦ were defined as random grain boundaries (RGB). The the TEM–EDS analysis.
grain boundaries shown
 in Fig. 2 indicated  that the grain bound- High angle annular dark field (HAADF) contrast is proportional
aries were mainly 3 and RGB, and the 3 fractions of 316L to the average atomic number, assuming constant thickness and

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 3

Fig. 1. IPFs of the 316L SS and model alloy specimens. (a) 316L SS, step size = 4 ␮m, (b) 9Cr–26Ni–1.5Mo, step size = 2 ␮m, and (c) 9Cr–26Ni–1.5Mo–5Si, step size = 4 ␮m.

Fig. 2. Grain boundary characteristics of the 316L SS and model alloys. (a) 316L SS, (b) 9Cr–26Ni–1.5Mo and (c) 9Cr–26Ni–1.5Mo–5Si.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
4 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Fig. 3. SEM morphologies of the oxide films formed on 316L SS and model alloys exposed to the simulated PWR primary water at 310 ◦ C for 114 h. (a) 316L SS, (b)
9Cr–26Ni–1.5Mo, and (c) 9Cr–26Ni–1.5Mo–5Si.

particles was inhomogeneous with the large ones dispersed on the


surface and the small ones developing from the subsurface layer. In
contrast, the 9Cr–26Ni–1.5Mo–5Si specimen surface was fully cov-
ered by large and continuously packed oxide particles with some
cavities.
Fig. 4 shows the Raman spectra of the oxide films formed on 316L
SS and model alloys exposed to the simulated PWR primary water
at 310 ◦ C for 114 h. Table 3 shows the Raman shifts of Raman spectra
peaks. The Raman peaks at 298, 320, 420, 472, 550, and 676 cm−1
are characteristics of Fe3 O4 , whereas the Raman peaks at 215, 490,
579, and 700 cm−1 are characteristics of NiFe2 O4 [21]. FeCr2 O4 is
responsible for the Raman peaks at 500 and 685 cm−1 , and NiCr2 O4
is responsible for the Raman peaks at 512 and 686 cm−1 [21]. The
Raman spectra of the oxide film formed on the 316L SS specimen
exhibited strong signals of Fe3 O4 , whereas the oxide film formed
on the 9Cr–26Ni–1.5Mo specimen exhibited the strong signals of
NiFe2 O4 . Both Fe3 O4 and NiFe2 O4 were detected in the oxide films
formed on the 9Cr–26Ni–1.5Mo–5Si specimen.
Fig. 5 shows the XPS depth profiles of the oxide films formed
on 316L SS, 9Cr–26Ni–1.5Mo, and 9Cr–26Ni–1.5Mo–5Si exposed
Fig. 4. Raman spectra of the oxide films formed on 316L SS and model alloys exposed
to the simulated PWR primary water at 310 ◦ C for 114 h. to the de-aerated, simulated PWR primary water at 310 ◦ C for
114 h. The oxide film formed on 9Cr–26Ni–1.5Mo was consider-
ably thicker than that formed on 316L SS, as shown in Fig. 5a and
density, the intensity in the HAADF images is proportional to the
b. For 316L SS and 9Cr–26Ni–1.5Mo, the signals generated at the
average Z [20]. In the present work, oxides are darker than the
initial stage of sputtering could come from both the outer and the
matrix in the STEM-HAADF images of the oxide film.
inner oxide films, as shown in Fig. 3. 9Cr–26Ni–1.5Mo–5Si speci-
men surface was fully covered by large and densely packed outer
3. Results layer oxide particles, and the signals generated at the initial stage
of sputtering were thought to be mainly come from the outer oxide
3.1. Immersion for 114 h film rich in Fe. With increasing the sputtering time, the Fe content
at first decreased and then gradually increased to values similar to
Fig. 3 shows the SEM morphologies of the 316L SS, that in the matrix. The oxygen content profiles were related to the
9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si surfaces after immer- difference of outer and inner layer oxides.
sion in the PWR primary water at 310 ◦ C for 114 h. Sparsely Fig. 6 shows the XPS spectra of O1s, Cr2p3/2, Fe2p3/2, Ni2p3/2
distributed polyhedral oxide particles were formed on both 316L and Si2p for the oxide film formed on the 9Cr–26Ni–1.5Mo–5Si
SS and 9Cr–26Ni–1.5Mo surfaces, the size of the outer-layer oxide

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 5

Table 3
The Raman shifts of Raman spectra peaks of the oxide films formed on 316L SS and model alloys exposed to the simulated PWR primary water at 310 ◦ C for 114 h.

Alloys Raman shift (cm−1 )

316L SS 319 466 547 680


9Cr–26Ni–1.5Mo 332 487 580 696
9Cr–26Ni–1.5Mo-5Si 324 478 566 685

100 sputtering [26,27]. The signals of Cr0 and Fe0 were not detected

80
a O1s
Fe2p
after 3000 s of sputtering, and only weak signals were detected
after 3600 s of sputtering. The signals generated before 3600 s of
Atomic percent (at%)

Cr2p sputtering mainly came from the oxide film. The Si2p characteristic
60 Ni2p peak signals were not detected before 3600 s of sputtering, indicat-
ing that the Si in the oxides formed on the 9Cr–26Ni–1.5Mo–5Si
40 specimen was too low to be detected. No characteristic peak sig-
nals corresponding to SiO2 (103.6 eV) were identified during XPS
20 profiling [28,29].
Peak shape and peak binding energy is useful for the identifi-
0 cation of relatively pure iron oxides. Biesinger et al. [30] reported
0 200 400 600 800 1000 1200 1400 1600
Sputtering time (s) the fitting parameters for Fe 2p3/2 spectral binding energy: Fe3 O4
100 (Fe2+ ), 708.4 and 709.2 9 eV; Fe3 O4 (Fe3+ ), 710.2, 711.2, 712.3, 713.4

80
b O1s
Fe2p
and 714.5 eV; FeCr2 O4 : 709.0, 710.3, 711.2, 713.0 and 713.8 eV;
NiCr2 O4 , 709.5, 710.7, 712.2 and 713.7 eV. McIntyre and Zetaruk
Atomic percent (at%)

Cr2p [31] reported that the binding energies of major core lines in Fe3 O4
60 Ni2p (Fe2+ ) and NiFe2 O4 were 708.3 ± 0.15 and 710.6 ± 0.2 eV. The oxide
film was inhomogeneous thus the signals generated during sput-
40 tering could come from both the outer and the inner oxide films. In
the present work, the oxides formed on 316L SS and model alloys
20 were complex and mixed metal oxides, which were mainly spinel-
type oxides Fex Cry Niz O4 . The exactly chemical states of Fe2p3/2
0 were complicated. The changes of Fe2p spectra were emphasized,
0 600 1200 1800 2400 3000 3600 4200 and 711 eV related to Fe+3 as well as 709 eV related to Fe2+ was
Sputtering time (s)
marked as reference.
100

80
c O1s
3.2. Immersion for 1194 h
Atomic percent (at%)

Fe2p
Cr2p
60 Ni2p Fig. 7 shows the morphologies of the oxide films formed
on 316L SS, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si after
40 immersion in the simulated PWR primary water for 1194 h. The
oxide films formed on 316L SS and 9Cr–26Ni–1.5Mo specimens
20 were composed of polyhedral oxides, and the specimen surfaces
were not fully covered by the outer layer oxide particles. The
0 9Cr–26Ni–1.5Mo–5Si specimen surface was fully-covered by large
0 900 1800 2700 3600 4500 5400 6300
and continuously packed oxide particles with some cavities. The
Sputtering time (s)
average size of the outer oxide particles formed on 316L SS was
Fig. 5. Composition profiles in the depth of the oxide films formed on 316L SS and smaller than that of 9Cr–26Ni–1.5Mo, as shown in Figs. 3 and 7,
model alloys exposed to the simulated PWR primary water at 310 ◦ C for 114 h. (a) which was consistent with the observation on the surfaces of 316
316L SS, (b) 9Cr–26Ni–1.5Mo, (c) 9Cr–26Ni–1.5Mo–5Si. SS and tailored alloys with various Cr contents from 5 to 20% after
immersion in high-temperature hydrogenated water [19].
Fig. 8 shows the Raman spectra of the surface films formed on
specimen after 114 h immersion. In the Cr2p3/2 spectra (Fig. 6b), 316L SS, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si after immer-
the peak at approximately 576.7 eV was assigned to Cr3+ [22,23]. sion for 1194 h. Table 4 shows the Raman shifts of Raman
In the Fe2p3/2 spectra (Fig. 6c), the main peak at approximately spectra peaks. Similar to the results of immersion for 114 h, the
711 eV was assigned to Fe3+ , and the peak at approximately 709 eV Raman spectra of the oxide film formed on the 316L SS and
was from Fe2+ [24]. Two different chemical states were detected in 9Cr–26Ni–1.5Mo–5Si specimens exhibited the strong signals of
the Ni2p3/2 spectra: the peak at 853 eV was assigned to metallic Ni0 Fe3 O4 and the oxide film formed on the 9Cr–26Ni–1.5Mo specimen
and the peak at 855.3 eV was assigned to Ni2+ [22,23]. After approx- exhibited the strong signals of NiFe2 O4 .
imately 600 s sputtering, no signals corresponding to Cr2p were Figs. 9–12 show the TEM results of the oxide films formed on
generated due to the lower Cr content in the outer layer. However, 316L SS exposed to the simulated PWR primary water for 1194 h.
strong signals corresponding to Ni0 were detected after 600 s sput- As mentioned previously, the high angle annular dark field (HAADF)
tering, and the peak decomposition of Ni2p3/2 was similar to that contrast is proportional to the average atomic number, and the
after 3600 s. Ni-containing oxides are stable in the de-aerated, sim- oxides are darker than the matrix in the STEM-HAADF images. Fig. 9
ulated PWR primary water according to the potential-pH diagram shows the duplex structure of the oxide film formed on 316L SS.
of Ni–H2 O at 300 ◦ C [25]. However, Ni-containing oxides is very sus- The outer layer was composed of discontinuous and large oxide
ceptible to reduction by ion bombardment during XPS profiling, so grains, whereas the inner layer was composed of dense and fine
the detected Ni0 might be reduced from Ni2+ during the argon ionic oxide grains. The thickness of the inner layer determined by TEM

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
6 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

3+
a O
2-
c Fe Fe2+

Intensity (arbitrary units)


Intensity (arbitrary units)

0
Fe

600 s
600 s
3600 s
3600 s
6000 s
6000 s

536 534 532 530 528 526 730 725 720 715 710 705 700
Binding energy (eV) Binding energy (eV)
0
b Cr
3+ d Ni

Intensity (arbitrary units)


Intensity (arbitrary units)

2+
0
Ni
Cr 600 s

600 s
3600 s
3600 s
6000 s
6000 s

595 590 585 580 575 570 875 870 865 860 855 850 845
Binding energy (eV) Binding energy (eV)
0
e Si
Intensity (arbitrary units)

6000 s
5400 s

4800 s
4200 s
3600 s
2400 s
600 s

104 102 100 98 96


Binding energy (eV)
Fig. 6. (a) O1s core level spectra, (b) Cr2p3/2 core level spectra, (c) Fe2p3/2 core level spectra, (d) Ni2p3/2 and (e) Si2p core level spectra after different sputtering periods
of oxide film formed on 9Cr–26Ni–1.5Mo–5Si exposed to the simulated PWR primary water at 310 ◦ C for 114 h.

Table 4
The Raman shifts of Raman spectra peaks of the oxide films formed on 316L SS and model alloys exposed to the simulated PWR primary water at 310 ◦ C for 1194 h.

Alloys Raman shift (cm−1 )

316L SS 310 468 551 681


9Cr–26Ni–1.5Mo 329 473 570 689
9Cr–26Ni–1.5Mo–5Si 322 470 563 681

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 7

Fig. 7. SEM morphologies of the oxide films formed on 316L SS and model alloys exposed to the simulated PWR primary water at 310 ◦ C for 1194 h. (a) 316L SS, (b)
9Cr–26Ni–1.5Mo, (c) 9Cr–26Ni–1.5Mo–5Si.

Table 5
Chemical compositions of the outer layer oxide particles shown in Figs. 9, 13 and 20 by TEM–EDS. The scatter band is the standard deviation of measurement over five
different positions of outer oxides.

Fe (at%) Cr (at%) Ni (at%) O (at%) Si (at%)

316L SS 71.55 ± 4.0 1.63 ± 0.64 3.33 ± 0.93 22.38 ± 5.19 1.12 ± 0.36
9Cr–26Ni–1.5Mo 54.52 ± 2.66 0.60 ± 0.28 15.29 ± 1.48 28.75 ± 2.61 0.85 ± 0.21
9Cr–26Ni–1.5Mo–5Si 45.85 ± 3.81 0.66 ± 0.21 18.43 ± 0.15 30.32 ± 3.52 2.04 ± 0.24

Fig. 9. STEM-HAADF image of the oxide film formed on 316L SS exposed to the
simulated PWR primary water at 310 ◦ C for 1194 h.

was approximately 100 nm. The EDS area mapping of the oxide
film shown in Fig. 10 indicated that the outer layer was Fe-rich and
the inner layer was Cr-rich, which was consistent with the results
of the EDS line scan profiles of the oxide film shown in Fig. 11b.
Fig. 8. Raman spectra of the oxide films formed on 316L SS and model alloys exposed The TEM–EDS results of the outer oxide particles are shown in
to the simulated PWR primary water for 1194 h at 310 ◦ C. Table 5. The diffraction pattern of the outer layer oxides shown
in Fig. 12b indicated that these oxide particles had the spinel-
type crystal structure, and were likely (Ni, Fe) Fe2 O4 , which was

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
8 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Fig. 10. STEM-HAADF image and corresponding EDS mappings (acquisition time = 125.9 s) of the rectangular region for the Cr, Ni, Fe, O, and Si of the oxide film formed on
316L SS after exposure to the simulated PWR primary water at 310 ◦ C for 1194 h.

Fig. 11. (a) STEM-HAADF image of the oxide film formed on 316L SS after 1194 h immersion in the simulated PWR primary water, (b) EDS line scan profiles (acquisition
time = 30 s) shown in (a) along the white line.

Fig. 12. (a) TEM cross-sectional micrograph of the oxide film formed on 316L SS exposed to the simulated PWR primary water at 310 ◦ C for 1194 h, (b) the electron diffraction
patterns of region I shown in (a).

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 9

indicating that both inner and outer oxides had the spinel-type
crystal structure Fex Niy Crz O4 .
Figs. 20–24 show the TEM results of the oxide film formed
on 9Cr–26Ni–1.5Mo–5Si exposed to the simulated PWR pri-
mary water at 310 ◦ C for 1194 h. The cross-sectional morphology
shown in Fig. 20 indicated that the oxide film formed on
9Cr–26Ni–1.5Mo–5Si had a duplex layer structure: outer-layer,
large oxide particles covered the specimen surface and the inner
layer was composed of fine oxides. The outer-layer oxides were
inhomogeneous, and the thickness of the inner layer was thinner
beneath the larger oxide particles. The thickness of the inner layer
determined by TEM ranged from 100 to 420 nm. The inner oxide
film close to oxide/metal interface was darker than the common
inner oxide film in the STEM-HAADF image, as shown in Figs. 20–22.
Fig. 13. STEM-HAADF image of the oxide film formed on 9Cr–26Ni–1.5Mo after The EDS mappings shown in Figs. 21 and 22 indicated that Fe and Ni
exposure to the simulated PWR primary water at 310 ◦ C for 1194 h.
were depleted in the inner oxide film close to oxide/metal interface,
which was confirmed by the EDS line scan profiles of the oxide film
consistent with the Raman result shown in Fig. 8. The EDS line scan shown in Fig. 23b. The TEM–EDS microanalysis, as shown in Table 6,
profiles of the oxide film indicated that the content of Cr was much indicated that the inner oxide film close to oxide/metal interface
higher than Fe and Ni, as shown in Fig. 11b. Ni was enriched near (region I) had higher contents of Si and O and lower contents of
the oxide/metal interface, which was confirmed in the EDS line scan Ni and Fe than the common oxide film, like region II. The electron
profiles of the oxide film. diffraction patterns shown in Fig. 24b and c indicated that both the
Figs. 13–19 show the TEM results of the oxide film formed on outer layer oxides and the inner layer oxides had the spinel-type
9Cr–26Ni–1.5Mo exposed to the simulated PWR primary water for Fex Niy Crz O4 crystal structure.
1194 h. A duplex oxide film was formed on the 9Cr–26Ni–1.5Mo
specimen surface after immersion for 1194 h, as shown in Fig. 13.
4. Discussion
Table 5 shows that the outer oxide formed on the 9Cr–26Ni–1.5Mo
specimen surface had a higher content of Ni and a lower content of
The oxide films formed on 316L SS, 9Cr–26Ni–1.5Mo and
Cr than that on the 316L SS specimen. The inner oxide film was inho-
9Cr–26Ni–1.5Mo–5Si after exposure to the simulated PWR primary
mogeneous, and its thickness, as determined by TEM, ranged from
water at 310 ◦ C for 1194 h have a duplex structure, as shown in
160 to 700 nm. There were some darker areas (such as area I shown
Figs. 9, 13 and 20. TEM analysis shows that the outer layer oxides
in Fig. 13) in the inner oxide film. Fig. 16a shows that the darker
formed on 316L S, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si are
areas in the oxide film shown in the STEM-HAADF images were
(Ni, Fe) Fe2 O4 .
composed of porous granular oxide. High resolution TEM images
Si in the model alloy affects both the chemical composition
of the dark (granular) zones exhibited lattice fringes, as shown in
profile and the surface morphology of the oxide film. After expo-
Fig. 16c. Images of selected area electron diffraction exhibited poly-
sure to the simulated PWR primary water at 310 ◦ C for 114 h
crystalline characteristics, as shown in Fig. 16b, and the dark field
and 1194 h, the 316L SS and 9Cr–26Ni–1.5Mo specimen surfaces
images also indicated the presence of crystals.
are covered by dispersed faceted oxide particles, whereas the
There were several factors that could cause the occurrence of
9Cr–26Ni–1.5Mo–5Si surface is covered by large, continuously
the dark zones in the STEM-HAADF images, such as the chemical
packed oxide particles with cavities. The large oxide particles
compositions and the quantity of the existing oxides. TEM obser-
on 9Cr–26Ni–1.5Mo–5Si specimen may not mainly deposit from
vations show that the dark zones were typical granular regions, as
the metallic cations released from autoclave body corrosion
shown in Figs. 16 a and 19 a. Results of TEM–EDS line scan profiles
because they are not observed on the surfaces of 316L SS and
show that the Fe contents in the dark zones were lower than those
9Cr–26Ni–1.5Mo in the same immersion experiments. According
in the relatively bright zones, as shown in Fig. 15a. The contrast of
to the thermodynamic data from HSC software [32], the change
the TEM–EDS Cr mapping did not show any significant decrease in
of standard Gibbs free energy for the reaction of Si with water at
the dark zones, as shown in Fig. 14. These results indicated that the
310 ◦ C is calculated to be −415.69 kJ/mol (equivalent to standard
local chemical composition would contribute to the observed dark
electrode potential of −1.08 V(SHE)), taking G0 H2O (310 ◦ C) as
zones.
−193.99 kJ/mol and G0 SiO2 (310 ◦ C) as −803.67 kJ/mol, according
The EDS area mappings shown in Fig. 14 indicated that there was
to Eq. (1). Si in the model alloy tends to react with high tempera-
a Ni-enriched thin layer, which also was confirmed by the EDS line
ture water to form oxides and at the same time affects the oxidation
scan profiles of the oxide film cross-section shown in Fig. 15a. How-
behavior of other metals. The solubility of silicon oxides in water
ever, the Ni enrichment layer was not continuous. Fig. 13 shows that
increases with increasing temperature from 0 ◦ C to about 310 ◦ C,
the inner oxide film just beneath the Ni enrichment regions was
which is strongly affected by the forms of silicon oxides such as
thinner than that beneath the regions without Ni enrichment. The
amorphous silica quartz [33,34]. The solubility for amorphous sil-
EDS area mappings shown in Fig. 17 showed that Ni enriched adja-
ica in 310 ◦ C water varied in the range of about 1400 to 1800 mg/kg
cent to some oxide/metal interfaces, which was consistent with the
[33,34], which is about a half or lower for quartz. Even though the
results of the EDS line scan profiles of the oxide film cross-section
solubility of silicon oxide in PWR primary water could be different
shown in Fig. 18b.
from that in pure water, it is expected to be still high by referring the
Fig. 19 shows the micrograph and electron diffraction patterns
effect of composition of high temperature solutions on solubility of
of region II shown in Fig. 13. For the inner layer of the oxide film,
silicon oxide [33].
the atomic percents of Fe, Ni and Cr were not stoichiometric and
it was difficult to deduce the exact chemical formula of the oxide. Si + 2H2 O = SiO2 + 2H2 (1)
The Cr content in the inner layer formed on 9Cr–26Ni–1.5Mo was
much lower than that of 316L SS. Fig. 19b and c shows the electron Due to the high solubility of silicon oxides in high temperature
diffraction patterns of the designated area in TEM image (Fig. 19a), water, the oxidized silicon on the alloy surface can be dissolved

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
10 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Table 6
Chemical compositions of regions I and II shown in Fig. 21 by TEM–EDS. The scatter band is the standard deviation of measurement over five different positions of regions I
and II.

Fe (at%) Cr (at%) Ni (at%) O (at%) Si (at%)

Region I 17.35 ± 2.06 18.17 ± 0.99 8.74 ± 1.96 41.51 ± 2.62 14.22 ± 1.17
Region II 23.18 ± 3.27 17.18 ± 0.93 25.62 ± 1.63 31.77 ± 1.93 2.33 ± 0.41

in the solution during the immersion in PWR primary water. This The HAADF images (Fig. 20) show that the inner oxide film close
explains why silicon is depleted in the outer layer of the oxide film, to the interface between the oxide/matrix on 9Cr–26Ni–1.5Mo–5Si
which has been shown by XPS results in Fig. 6 and TEM results in is characterized by a lower average atomic number compared with
Tables 5 and 6. the common inner oxide film. The inner oxide beneath the interface
of the oxide/metal exhibits higher contents of O and Si and lower

Fig. 14. STEM-HAADF image and EDS mappings (acquisition time = 449 s) for Cr, Ni, Fe, O, and Si in the oxide film formed on 9Cr–26Ni–1.5Mo after 1194 h immersion in
simulated PWR primary water at 310 ◦ C.

a matrix b matrix
inner oxide inner oxide
outer O
oxide Fe O
Counts

Counts

Cr Fe
Ni Cr
Si Ni
Si

0 100 200 300 400 500 600 0 100 200 300 400 500 600 700 800 900
Position / nm Position / nm

Fig. 15. (a) EDS line scan profiles (acquisition time = 56 s) of the oxide film shown in Fig. 14 along the white line a, (b) EDS line scan profiles (acquisition time = 97 s) of the
oxide film shown in Fig. 14 along the white line b.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 11

Fig. 16. (a) TEM image and (b) electron diffraction pattern of region I shown in Fig. 13, (c) TEM image of circular region in (a).

Fig. 17. STEM-HAADF image and EDS mappings (acquisition time = 210 s) for Cr, Ni, Fe, O, and Si of the oxide film formed on 9Cr–26Ni–1.5Mo after 1194 h immersion in the
simulated PWR primary water at 310 ◦ C.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
12 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Fig. 18. (a) STEM-HAADF image of the oxide film formed on 9Cr–26Ni–1.5Mo after 1194 h immersion in the simulated PWR primary water at 310 ◦ C and (b) EDS line scan
profiles (acquisition time = 81 s) of oxide film shown in (a) along the white line.

Fig. 19. TEM micrograph with electron diffraction patterns of the oxides formed on 9Cr–26Ni–1.5Mo exposed to the simulated PWR primary water at 310 ◦ C for 1194 h. (a)
TEM image, (b) the diffraction pattern of the oxides of region I shown in (a), and (c) the diffraction pattern of the oxides of region II shown in (a). The white dotted line in (a)
indicates the interface between the inner oxide film and the matrix.

Table 7
Experimental results of oxide film formed on 316L SS and model alloys exposed to the simulated PWR primary water at 310 ◦ C.

Alloys Outer oxide film Inner oxide film Near the interface of the oxide/metal

316L SS
Oxide shape: with straight edges and planar Thickness: about 100 nm Local Ni enrichment
faces Fe:Cr:Ni = 71.55:1.63:3.33 (at%)
9Cr–26Ni–1.5Mo Oxide shape: with straight edges and planar Thickness: 160–700 nm Local Ni enrichment
faces Fe:Cr:Ni = 54.52:0.60:15.29 (at%) Exist porous granular areas
9Cr–26Ni–1.5Mo–5Si Oxide shape: with curving edges, blunt angles Thickness: 100 −420 nm Si enrichment beneath the interface of
and some cavities Fe:Cr:Ni = 45.85:0.66:18.43 the oxide/metal
(at%)

contents of Fe and Ni than its surrounding areas, which is different


from those on 316L SS and 9Cr–26Ni–1.5Mo. Silicon oxides near
the metal/oxide interface can sustain due to the protectiveness of
the oxide film on it, as shown in Table 6.
The Cr content in the inner oxide is similar to the level of alloy
matrix, and the Cr content in the inner oxide formed on 316L SS
was higher than that on 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si
specimens, as shown in Figs. 11, 15 and 23. The inner layer formed
on stainless steels in high temperature aqueous solutions are
related to the varying diffusion rates of the metal ions in the
oxide, and the diffusion rates follow the order: Fe2+ > Ni2+  Cr3+
[35–38]. The inner oxide film is crucial for the dissolution of metal
atoms from the metal to the oxide film or oxygen diffusion from
the oxide film to the metal matrix [18,35,39–41]. The cation dif-
fusion rate in the Cr-rich oxide layer is slower than that in Fe
or Ni oxides [18,19,41]. The Cr-rich inner layer formed on stain-
Fig. 20. STEM-HAADF image of the oxide film formed on 9Cr–26Ni–1.5Mo–5Si less steels favors the decreasing of the growth rate of the oxide
exposed to the simulated PWR primary water at 310 ◦ C for 1194 h. The black dotted
scale, acting as a barrier against the diffusion of metal and oxygen
line indicates the interface of the outer- and inner-layer oxide film.
ions [19,41,42]. Therefore, the oxide film formed on 316L SS speci-

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 13

Fig. 21. STEM-HAADF image of region I shown in Fig. 20 and the corresponding EDS mappings (acquisition time = 1011 s) for Cr, Ni, Fe, O, and Si. The black dotted line
indicates the interface of the outer- and inner-layer oxide film.

Fig. 22. STEM-HAADF image of the oxide film formed on 9Cr–26Ni–1.5Mo–5Si after immersion for 1194 h and corresponding EDS mappings (acquisition time = 1361 s) for
Cr, Ni, Fe, O, and Si.

men has a thinner inner layer than those on 9Cr–26Ni–1.5Mo and It has been reported that adding a certain amount of Si in
9Cr–26Ni–1.5Mo–5Si specimens, as shown in Figs. 9, 13 and 20. The austenitic alloys could improve the oxidation resistance in air by
effect of Si on the distribution of chromium in the inner layer is not forming protective oxide scales [43]. It has been found that addi-
significant due to the strong tendency of Cr to form oxides. tion of silicon had the potential of improving the formation of scales
Results in Figs. 3 and 7 show that the presence of Si in the model of Ni–Al and Ni–Cr–Al alloys in high temperature atmosphere envi-
alloy would facilitate the formation of other metallic oxides thus ronments [44–47]. The protectiveness of oxides in air and in high
favors the formation of a continuous iron-rich spinel outer layer temperature water could be significantly different due to different
even after a short term immersion in high temperature water. The oxidation mechanisms. Silicon oxides can act as a protective bar-
cavities in the outer-layer oxide particles on 9Cr–26Ni–1.5Mo–5Si rier mitigating further oxidation in high temperature atmospheric
specimen can be related to the oxidation of Si and the subsequent environments while silicon oxides would not provide adequate
dissolution of the Si-bearing oxides in the solution.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
14 G. Han et al. / Corrosion Science xxx (2016) xxx–xxx

Fig. 23. (a) STEM-HAADF image of the oxide film formed on 9Cr–26Ni–1.5Mo–5Si after immersion in simulated PWR primary water at 310 ◦ C for 1194 h. The black dotted
line indicates the interface of the outer- and inner-layer oxide film (b) EDS line scan profiles (acquisition time = 106 s) of the oxide film shown in (a) by the white line.

Fig. 24. (a) TEM micrograph with SAED patterns of 9Cr–26Ni–1.5Mo–5Si after exposure to the simulated PWR primary water at 310 ◦ C for 1194 h. The white dotted line
indicates the interface of the oxide/metal. (b) and (c) the SAED pattern corresponding to the designated areas I and II in (a), respectively.

protectiveness due to the high solubility in high temperature water of other alloying elements in the model alloy has been proved by
environments [12,33,34]. the results of the outer surface films, as shown in Figs. 3 and 7.
Local Ni-enrichment near the oxide/metal interface is observed Table 7 summarizes the experimental results of oxide films
for 316L SS and 9Cr–26Ni–1.5Mo specimens, as shown in formed on 316L SS and model alloys exposed to the simulated
Figs. 10, 14 and 17. Ni enrichment near the oxide/metal interface or PWR primary water at 310 ◦ C for 1194 h. In the present work,
ahead of the crack tip has been previously observed for austenitic the thickness of the inner layer determined by TEM of 316L
stainless steels in simulated PWR primary water [48–50]. Ni enrich- SS, 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo–5Si is about 100 nm,
ment near the oxide/metal interface could be related to its lower 160–700 nm and 100–420 nm after 1194 h immersion, respec-
oxidation rates compared with Fe and Cr during the formation of tively. The oxidation rate constants are not calculated because
the oxide film. Local depletion of iron in region I and almost uniform more data points at different immersion time periods are neces-
distribution of oxygen in Fig. 17 indicates that the inward diffusion sary and the oxidation rate law for the present test conditions is not
of oxygen would be more difficult than the outward diffusion of defined. Ziemniak and Hanson [52] studied the corrosion behavior
ferrous/ferric ions in the inner oxide film. Ni is the least likely to of 304 SS in a 10,000 h test conducted in hydrogenated water at
be oxidized due to the lower affinity for oxygen than for Cr and 260 ◦ C, and suggested that the corrosion kinetics were parabolic.
Fe; therefore the rate of incorporation of Ni in the oxide scale is Generally, the oxidation of austenitic alloys in high temperature fol-
lower than that of Cr and Fe, resulting in a high Ni concentration lows the parabolic law [52–54], where the increasing rate of oxide
between oxide and metal [41,51]. Das et al. [18] suggested that Fe film thickness decreases with increasing immersion time. Although
and Cr atoms segregated faster than Ni atoms at the surface, and Ni the oxidation rate law for the alloy/environment combinations in
enrichment occurred at the inner oxide layer on Fe–Cr–Ni ternary the present research is not determined, it is expected that fur-
alloys. Ni enrichment is not continuous, and no Ni enrichment is ther increasing the oxidation duration time after 1194 h immersion
observed beneath the porous oxides, as shown in Fig. 14, indicating would produce only a slight increase of the film thickness.
that the porous oxides have a negative influence on the Ni enrich- It has been reported that austenitic stainless steels with high
ment near the interface. The micro pores in the porous oxides are Si contents exhibit high SCC growth rates in high temperature
expected to provide more shortcuts for the inward diffusing oxygen water environments [12,13]. Andresen and Morra [12] suggested
and outward diffusing metal ions. The enhanced inward diffus- that crack growth rates of stainless steels with elevated Si con-
ing oxygen favors the formation of the Ni-oxides. Moreover, these centrations were insensitive to changes in both corrosion potential
shortcuts would contribute to the outward diffusion of nickel ions. and stress intensity. Addition of Si in austenitic stainless steels
Ni-enrichment near the oxide/metal interface was not observed on would affect the oxidation process of the model austenitic alloys in
9Cr–26Ni–1.5Mo–5Si specimen, which is likely due to the effect of high-temperature water and therefore the stress corrosion cracking
Si on the oxidation behavior of Ni. The effect of Si on the oxidation behavior. The initial oxidation behavior of the crack tip is impor-
tant for SCC susceptibility and the crack growth rate [17,18]. The

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001
G Model
CS-6638; No. of Pages 15 ARTICLE IN PRESS
G. Han et al. / Corrosion Science xxx (2016) xxx–xxx 15

results in the present work imply that Si would significantly affect [5] Z. Jiao, G.S. Was, Acta Mater. 59 (2011) 1220–1238.
the stress corrosion behavior through its strong effect on the initial [6] G.S. Was, S.M. Bruemmer, J. Nucl. Mater. 216 (1994) 326–347.
[7] K. Fukuya, M. Nakano, K. Fujii, T. Torimaru, J. Nucl. Sci. Technol. 41 (2004)
oxidation behavior. The results would be useful for understanding 594–600.
irradiation assisted stress corrosion cracking where Si-enrichment [8] S.J. Zinkle, P.J. Maziasz, R.E. Stoller, J. Nucl. Mater. 206 (1993) 266–286.
at the grain boundaries of reactor core internal alloys due to irradi- [9] T. Toyama, Y. Nozawa, W. Van Renterghem, Y. Matsukawa, M. Hatakeyama, Y.
Nagai, A. Al Mazouzi, S. Van Dyck, J. Nucl. Mater. 425 (2012) 71–75.
ation is a great concern. It is noted that SCC growth is affected by the [10] S.M. Bruemmer, Proceedings of the 10th International Conference on
oxidation behavior and the oxide film toughness in terms of film Environmental Degradation of Materials in Nuclear Power Systems—Water
rupture strain or film degradation strain, and the crack tip mechan- Reactors, Paper No. 0008V, NACE, Houston, TX, 2001, CRDOM.
[11] G.S. Was, Y. Ashida, P.L. Andresen, Corros. Rev. 29 (2011) 7–49.
ics [12,53,54]. Care should be taken when correlating oxidation
[12] P.L. Andresen, M.M. Morra, Effects of Si on SCC of irradiated and unirradiated
behavior directly to SCC growth behavior. stainless steels and nickel alloys, Proceedings of the 13th International
Conference on Environmental Degradation of Materials in Nuclear Power
Systems—Water Reactors (2005) 87–106, TMS.
5. Conclusions
[13] G.F. Li, Y. Kaneshima, T. Shoji, Corrosion 56 (2000) 460–469.
[14] A.J. Jacobs, in: A.S. Kumar, D.S., Gelles, R.K. Nanstad, E.A. Little (Eds.), 16th
The morphology, chemical composition and microstructure International Symposium on Radiation on Materials, ASTM-STP 1175, ASTM
Philadelphia, 1993, 902.
of the oxide film formed on 9Cr–26Ni–1.5Mo–5Si specimen are
[15] K. Fukuya, K., Nakata, A. Horie, An IASCC study using higher energy ion
significantly different from those on 316L and 9Cr–26Ni–1.5Mo irradiation, in: D. Cubicciotti (Ed.), Proceedings of the 5th International
specimens exposed to a de-aerated, simulated PWR primary water Conference on Environmental Degradation of Materials in Nuclear Power
at 310 ◦ C. Systems—Water Reactors, ANS, Monterey, CA, 1992, 814.
[16] H.M. Chung, W.E., Ruther, J.E., Sanecki, T.F. Kassner, Irradiation-induced
sensitization and stress corrosion cracking of Type 304 stainless steel core
(1) High reactivity of Si and high solubility of silicon oxides in internal components, in: D. Cubicciotti (Ed.), Proceedings of 5th International
high temperature water contribute to the observed Si-depletion Conference on Environmental Degradation of Materials in Nuclear Power
Systems—Water Reactors, ANS, Monterey, CA, 1992, 795.
in the outer oxide layer and the local Si-enrichment near the [17] K. Fujiwara, H. Tomari, K. Shimogori, T. Fukuzuka, Corrosion 38 (1982) 69–76.
metal/oxide interface for 9Cr–26Ni–1.5Mo–5Si alloy. [18] N.K. Das, K. Suzuki, K. Ogawa, T. Shoji, Corros. Sci. 51 (2009) 908–913.
(2) High reactivity of Si in 9Cr–26Ni–1.5Mo–5Si with water [19] T. Terachi, T. Yamada, T. Miyamoto, K. Arioka, K. Fukuya, J. Nucl. Sci. Technol.
45 (2008) 975–984.
enhances the oxidation of other alloy elements to form a con- [20] S. Lozano-Perez, T. Yamada, T. Terachi, M. Schröder, C.A. English, G.D.W.
tinuous iron-rich spinel outer layer even after a short term Smith, C.R.M. Grovenor, B.L. Eyre, Acta Mater. 57 (2009) 5361–5381.
immersion in high temperature water. The prior-oxidation of [21] M.D.A. Cunha Belo, M. Walls, N.E. Hakiki, J. Corset, E. Picquenard, G. Sagon, D.
Noël, Corros. Sci. 40 (1998) 447–463.
Si and the subsequent dissolution of the Si-bearing oxides in
[22] J.C. Langevoort, I. Sutherland, L.J. Hanekamp, P.J. Gellings, Appl. Surf. Sci. 28
high temperature water contribute to the observed cavities in (1987) 167–179.
the outer oxide layer. [23] V.A. Maroni, C.A. Melendres, T.F. Kassner, R. Kumar, S. Siegel, J. Nucl. Mater.
172 (1990) 13–18.
(3) The inner layer formed on 316L SS with about 17 wt.%
[24] T. Yamashita, P. Hayes, Appl. Surf. Sci. 254 (2008) 2441–2449.
Cr is thinner than those formed on 9Cr–26Ni–1.5Mo and [25] R.W. Staehle, J.A. Gorman, Corrosion 59 (2003) 931–994.
9Cr–26Ni–1.5Mo–5Si model alloys with about 9 wt.% Cr. The [26] A. Machet, A. Galtayries, P. Marcus, P. Combrade, P. Jolivet, P. Scott, Surf.
Cr content in the inner oxide formed on 316L SS is higher than Interf. Anal. 34 (2002) 197–200.
[27] W.J. Kuang, X.Q. Wu, E.-H. Han, Corros. Sci. 63 (2012) 259–266.
that on 9Cr–26Ni–1.5Mo and 9Cr–26Ni–1.5Mo-5Si specimens. [28] S.P. Chenakin, G. Melaet, R. Szukiewicz, N. Kruse, J. Catal. 312 (2014) 1–11.
A high Cr-content in the alloys favors the formation of a protec- [29] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, G.E. Mullenberg (Eds.),
tive thin Cr-rich inner layer that hinders the further growth of Hand-book of X-ray Photoelectron Spectroscopy, PerkinElmer Corporation,
Physical Electronics Division, Eden Prairie, MN, USA, 1979.
the inner layer. The effect of Si on the distribution of chromium [30] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.St.C.
in the inner layer is not significant due to the strong tendency Smart, Appl. Surf. Sci. 257 (2011) 2717–2730.
of Cr to form oxides. [31] N.S. McIntyre, D.G. Zetaruk, Anal. Chem. 49 (1977) 1521–1529.
[32] HSC Chemistry 5.11., Outokumpu Research Oy, Pori, Finland.
[33] R.O. Fournier, J.J. Rowe, Am. Mineral. 62 (1977) 1052–1056.
Ni enrichment near the oxide/metal interfaces is observed for [34] A.G. Carr, J.W. Tester, Fluid Phase Equilib. 337 (2013) 288–297.
316L SS and 9Cr–26Ni–1.5Mo specimens but not observed for [35] J. Robertson, Corros. Sci. 32 (1991) 443–465.
[36] R. Dieckmann, Solid State Ion. 12 (1984) 1–22.
9Cr–26Ni–1.5Mo–5Si specimen, indicating the effect of Si on the
[37] R. Dieckmann, T.O. Mason, J.D. Hodge, H. Schmalzried, Ber. Bunsen. Phys.
oxidation kinetics of Ni. Chem. 82 (1978) 778–783.
[38] M.G.C. Cox, B. Mcenaney, V.D. Scott, Philos. Mag. 26 (1972) 839–851.
[39] B. Stellwag, Corros. Sci. 40 (1998) 337–370.
Acknowledgments
[40] G.D. Han, Z.P. Lu, X.K. Ru, J.J. Chen, Q. Xiao, Y.W. Tian, J. Nucl. Mater. 467
(2015) 194–204.
This work has been supported by the International Cooperative [41] M.C. Sun, X.Q. Wu, Z.E. Zhang, E.-H. Han, Corros. Sci. 51 (2009) 1069–1072.
[42] D.H. Lister, R.D. Davidson, E. McAlpine, Corros. Sci. 27 (1987) 113–140.
Project, sponsored by the Science and Technology Commission of
[43] X.Q. Xu, X.F. Zhang, X.Y. Sun, Z.P. Lu, Corros. Sci. 65 (2012) 317–321.
Shanghai Municipality No. 13520721200, the Ph.D. Programs Foun- [44] Y. Wu, Y. Niu, Corros. Sci. 49 (2007) 1656–1672.
dation of the Ministry of Education of China No. 20123108110021, [45] Y. Wu, F. Gesmundo, Y. Niu, Oxid. Met. 65 (2005) 53–74.
and Shanghai Pujiang Project No. 12PJ1403600. [46] S. Wang, Y. Wu, C.S. Ni, Y. Niu, Corros. Sci. 51 (2009) 511–517.
[47] S. Wang, Y. Wu, F. Gesmundo, Y. Niu, Oxid. Met. 69 (2008) 299–315.
[48] K. Kruska, S. Lozano-Perez, D.W. Saxey, T. Terachi, T. Yamada, G.D.W. Smith,
References Corros. Sci. 63 (2012) 225–233.
[49] T. Terachi, K. Fujii, K. Arioka, J. Nucl. Sci. Technol. 42 (2005) 225–232.
[1] K.J. Stephenson, G.S. Was, J. Nucl. Mater. 444 (2014) 331–341. [50] M. Nezakat, H. Akhiani, S. Penttilä, S.M. Sabet, J. Szpunar, Corros. Sci. 94
[2] S.M. Bruemmer, E.P. Simonen, P.M. Scott, P.L. Andresen, G.S. Was, J.L. Nelson, J. (2015) 197–206.
Nucl. Mater. 274 (1999) 299–314. [51] K.I. Choudhry, S. Mahboubi, G.A. Botton, J.R. Kish, I.M. Svishchev, Corros. Sci.
[3] E.A. Kenik, J.T. Busby, Mater. Sci. Eng., R. 73 (2012) 67–83. 100 (2015) 222–230.
[4] O.K. Chopra, A.S. Rao, J. Nucl. Mater. 409 (2011) 235–256. [52] S.E. Ziemniak, M. Hanson, Corros. Sci. 44 (2002) 2209–2230.
[53] T. Shoji, Z.P. Lu, H. Murakami, Corros. Sci. 52 (2010) 769–779.
[54] Z.P. Lu, T. Shoji, H. Xue, C.Y. Fu, J. Press. Vessel Technol. 135 (2013)
021402-1–021402-9.

Please cite this article in press as: G. Han, et al., Properties of oxide films formed on 316L SS and model alloys with modified Ni, Cr and
Si contents in high temperature water, Corros. Sci. (2016), http://dx.doi.org/10.1016/j.corsci.2016.02.001

You might also like