You are on page 1of 15

Journal of Nuclear Science and Technology

ISSN: 0022-3131 (Print) 1881-1248 (Online) Journal homepage: https://www.tandfonline.com/loi/tnst20

In situ measurement of corrosion of type 316L


stainless steel in 553 K pure water via the
electrical resistance of a thin wire

Kazushige Ishida & Derek Lister

To cite this article: Kazushige Ishida & Derek Lister (2012) In�situ measurement of corrosion of
type 316L stainless steel in 553 K pure water via the electrical resistance of a thin wire, Journal of
Nuclear Science and Technology, 49:11, 1078-1091, DOI: 10.1080/00223131.2012.730899

To link to this article: https://doi.org/10.1080/00223131.2012.730899

Published online: 23 Oct 2012.

Submit your article to this journal

Article views: 1142

View related articles

Citing articles: 2 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tnst20
Journal of Nuclear Science and Technology
Volume 49, No. 11, November (2012) pp. 1078–1091
http://www.tandfonline.com

ARTICLE

In situ measurement of corrosion of type 316L stainless steel in 553 K pure water via the electrical
resistance of a thin wire
Kazushige Ishidaa* and Derek Listerb
a
Hitachi Research Laboratory, Hitachi, Ltd., 7-2-1 Omika-cho, Hitachi-shi, Ibaraki-ken 319-1221, Japan; bDepartment of Chemical
Engineering, University of New Brunswick, P.O. Box 4400, Fredericton, New Brunswick E3B 5A3, Canada
(Received 3 April 2012; accepted final version for publication 4 September 2012)

A system for the in situ monitoring of corrosion depth via electrical resistance measurements was applied
to study the corrosion rate of type 316L stainless steel at 553 K in pure water. Corrosion depth was
measured using a 50 mm diameter wire probe mounted axially in the tube. Measurements were in good
agreement with literature data for both the hydrogen water chemistry (HWC) condition and the normal
water chemistry (NWC) condition. Oxide film analyses by scanning electron microscopy and laser
Raman spectroscopy on the wire probe and the tube showed no effects from shape of the test specimens
or the application of electric current. Corrosion kinetics was evaluated by fitting equations to the
measurements. Data for the HWC condition could be fitted by a two-step logarithmic–parabolic law. A
single-step logarithmic law fitted data for the NWC condition. Changes in corrosion rate by the water
chemistry changes were readily detected with the technique. Corrosion depth change could be observed
for the water chemistry change from the NWC condition to the HWC condition with electrochemical
corrosion potential (ECP) of 70.56 V vs. standard hydrogen electrode, which is lower than the ECP that
the phase of iron oxide changes from a-Fe2O3 to Fe3O4.
Keywords: type 316L stainless steel; high-temperature corrosion; hydrogen water chemistry; normal water
chemistry; in situ monitoring; corrosion kinetics

which circulates water in the RPV. Water at high


1. Introduction temperature (about 553 K) and of high purity
The reduction of radiation exposure during plant (electrical conductivity less than 30 mS m71 at room
inspection is an ongoing issue for nuclear power plants temperature, although with radiolytic oxygen, hydro-
[1,2]. One of the main sources of exposure is cobalt gen peroxide and hydrogen present) is circulated in the
activity accumulated on the piping surfaces. Radio- PLR piping. Hydrogen is also intentionally added in
active Co-60 is generated in the core by neutron some plants to mitigate stress corrosion cracking of
exposure of non-radioactive Co-59 and radioactive Co- structural materials. In the case of no hydrogen
58 is generated by neutron exposure of non-radioactive addition, oxygen at 100 mg kg71 and hydrogen
Ni-58. The radioactive cobalt is transported as ions in peroxide at 200 mg kg71 are typically contained in
water from the core region to the surfaces of piping the water (called the normal water chemistry (NWC)
connected to the reactor pressure vessel (RPV). Then, condition) [5,6]. In the case of hydrogen addition, the
the ions are incorporated into an oxide film which hydrogen concentration depends upon the plant
grows on the piping surfaces [3]. Knowing the specifications and the plant operating characteristics,
corrosion kinetics of the piping material is important but a concentration of 100–150 mg kg71 is typical
for evaluating the activity accumulation and the (called the hydrogen water chemistry (HWC)
radiation field build-up. condition).
Type 316L stainless steel (316L SS) is widely used The corrosion rate of stainless steel in high-
for piping in Japanese and other boiling water reactors temperature de-oxygenated water has been postulated
(BWRs) [4]. The piping of the primary loop recircula- as being controlled by the diffusion of metal ions across
tion system (PLR), which is the main source of the Cr2O3-based oxide layer, and it conforms to a
radiation exposure, is connected to the PLR pump parabolic rate law [7]. Other reports have expressed

*Corresponding author. Email: kazushige.ishida.eq@hitachi.com

ISSN 0022-3131 print/ISSN 1881-1248 online


Ó 2012 Atomic Energy Society of Japan. All rights reserved.
http://dx.doi.org/10.1080/00223131.2012.730899
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1079

total corrosion rate of 316L SS in oxygenated water at


2. Experimental procedure
553–563 K as depending upon time raised to the 2/3
power [8] or being controlled by a logarithmic rate law 2.1. Measurements of corrosion depth and ECP
[9]. Corrosion depths at an early time of 200 h are The corrosion depth of 316L SS in 553 K pure
about 50–80% of those at 1000 h [9,10]. So, corrosion water has been reported to be of the order of a
rate at time less than 200 h is important for knowing micrometre or less [8–12]. The corrosion depths of
the corrosion kinetics. 316L SS wire specimens (diameter 50 mm) were
The corrosion rate of 316L SS in high-temperature evaluated using electrical resistance measurements
pure water depends upon the water chemistry, which applying a constant current (I) and measuring the
can include some amounts of dissolved oxygen, potential drop (V) intermittently to obtain the resis-
hydrogen and hydrogen peroxide, as well as the tance (R) via Ohm’s law (Equation (1)). The resistance
resulting electrochemical corrosion potential (ECP) R is also given by the resistivity (r), length (l) and
[10,11]. Corrosion rate was seen to increase, depending radius (r) of the wire as Equation (2), so that r can be
on the dissolved oxygen concentration between 55 mg obtained from Equation (3). Supposing uniform
kg71 and 1000 mg kg71, in corrosion tests of specimens (general) corrosion, corrosion depth (Dr) is calculated
immersed in 556 K pure water for up to 1000 h [10]. from the initial value (r0) and the value (rt) at the
However, the polarization curve in 553 K de-oxyge- immersion time (t) from Equation (4).
nated pure water was reported to have an active state
V
peak at about 70.36 V vs. standard hydrogen elec- R¼ ð1Þ
trode (V vs. SHE) and passive state at about 70.27 V I
vs. SHE [11].
l
The characteristics of oxide films formed in high- R¼r ð2Þ
temperature pure water were also reported to depend p  r2
on the water chemistry and were expected to affect rffiffiffiffiffiffiffiffiffiffi
corrosion [12]. What is more, the effects of hydrogen rl
r¼ ð3Þ
peroxide on the characteristics of oxide films may be pR
different from those of dissolved oxygen, even though
the ECP is the same [13]. It is clearly important to Dr ¼ r0  rt ð4Þ
expose specimens to high-temperature water contain-
ing hydrogen peroxide to study corrosion rate under The change of R by corrosion is represented by
the NWC condition. Equation (5) approximately.
The characteristics of oxide films formed in high-
temperature pure water were also reported to change
Rt  R0 2Dr
followed by the water chemistry change from NWC ffi ð5Þ
R0 r0
condition to HWC condition and vice versa [12].
Thickness of oxide film asymptotically decreases to
about 0.5 mm depending on the water chemistry Here, R0 is the initial resistance of the wire probe
condition [12]. However, the effects of water chemistry and Rt is the resistance of the wire probe at time t.
changes on the corrosion depth or the corrosion rate Equation (5) indicates that it is desirable for r to be
have not been clarified. small to increase the change of R.
In the work reported here, a system for monitoring A schematic drawing of the in situ monitoring
the corrosion depth in situ by measuring electrical system is shown in Figure 1. The 316L SS wire probe
resistance was applied to study the corrosion depth of (diameter, 50 + 5 mm; length, 18 cm; composition [%]:
316L SS at 553 K in pure water. In situ monitoring of Mn, 1.28; Ni, 10.55; Cr, 17.99; Mo, 2.99; Fe, balance)
corrosion depth is useful to obtain the corrosion rate of was mounted axially in the 316L SS tube (inner
the early stage or the effects of water chemistry change diameter, 7.04 mm; composition [wt%]: Mn, 1.69; Ni,
on corrosion rate, because it is not easy to remove a 12.15; Cr, 18.15; Mo, 2.58; Fe, balance). A photograph
specimen from high-temperature water exposed for of the 316L SS wire probe is also shown in Figure 1.
hour or less. Corrosion depth of 316L SS in 553 K Lead wires of 316L SS (diameter, 0.5 mm) were silver-
water was measured under the HWC condition with soldered to both ends of the wire probe and insulated
hydrogen at 150 mg kg71 and under the NWC with polytetrafluoroethylene (PTFE) heat-shrink to
conditions with hydrogen peroxide at 200 mg kg71 avoid contact with water. The 316L SS wire probe and
and the kinetic expressions for the corrosion were the 316L SS tube were used after degreasing by acetone
formulated for both conditions. The effects on corro- without polishing. No large scratches were seen on the
sion depth on water chemistry change from NWC 316L SS wire probe surfaces as shown in Figure 2 for
condition to HWC condition and vice versa were also scanning electron microscopy (SEM) images of the
studied. The ECP was also measured to consider the surface and cross section of as-received 316L SS wire
effect of ECP, as well as water chemistry. probe. Oxide film was not detected on the surfaces of
1080 K. Ishida and D. Lister

Figure 1. Schematic of in situ monitoring system. (a) Schematic of test section and measurement systems and (b) photo of wire
probe connected with lead wire.

Figure 2. SEM image of (a) cross section and (b) surface of as-received 316L SS wire probe.

the 316L SS wire probe by laser Raman spectroscopy


(LRS). Nu  kf
DT ¼ q ð6Þ
A platinum wire insulated with PTFE and a d
thermocouple were mounted close to the wire probe 
f
to measure the ECP and temperature, respectively. 8  ðRe  1000Þ  Pr
A current source (Keithley 2400) of 1 V with the Nu ¼ qffiffi   ð7Þ
resolution of 1079 V and a voltmeter (Keithley 2812A) 1 þ 12:7  8f  Pr2=3 1
of 10 mA with the resolution of 5 6 1076 mA were
connected to the 316L SS wire probe. The resistance of f ¼ ð0:79  lnðReÞ  1:64Þ2 ð8Þ
the 316L SS wire probe was measured every 2 min.
Current should be low to avoid pool boiling on the A potentiostat (Gamry PCI-4300) was connected to
316L SS wire probe surfaces. In this work, current of the platinum wire and the 316L SS tube and the ECP of
1 mA was applied. For this measuring condition, the the tube was measured every 5 min. Platinum wire was
amount of increasing temperature (DT ) was calculated worked as the hydrogen electrode during hydrogen
to be 0.0023 K from Fourier’s law (Equation (6)) using injection of 150 mg kg71. Potential of the platinum
Gnielinski equations for Nusselt number (Nu; Equa- wire (EPt,HWC ) at 533 K was calculated to be 70.52 V
tion (7)), pipe friction coefficient (f, Equation (8)), vs. SHE under the HWC condition with hydrogen at
Reynolds number (Re ), Prandtl number (Pr), thermal 150 mg kg71 from the Nernst equation (Equation (9))
conductivity (kf), equivalent diameter of flow path (d) using Henry’s constant of hydrogen (kH) reported in
and thermal flux (q) [14,15]. [16], molar fraction of dissolved hydrogen (CH2),
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1081

activity of hydrogen ion ([Hþ]), gas constant (RG) and


temperature (T). The ECP of 316L SS tube (E316LSS)
vs. SHE was calculated from Equation (10) using
measured potential of platinum wire (Emeas) against
316L SS tube.
!
RG  T kH  CH2
EPt;HWC ¼  ln 2
ð9Þ
2 ½Hþ 

E316LSS ¼ Emeas  0:52 ð10Þ


Potential of platinum wire is constant under the
NWC condition [17]. Potential of platinum wire under
the NWC condition (EPt,NWC) was obtained empiri-
cally. Potential of 316L SS tube is controlled by
hydrogen peroxide concentration. So, ECP of 316L SS
tube under the NWC condition is the same as that
under overlapping term of the HWC condition and the
NWC condition. On the contrary, platinum wire works
as a quasi-hydrogen electrode under the overlapping
term of the HWC condition and the NWC condition
because the molar ratio of hydrogen to hydrogen
peroxide is more than 1 [17,18]. Potential of platinum
wire can be calculated from Equation (11) using Figure 3. Outline of high-temperature, high-pressure loop.
measured potential of platinum wire against 316L SS
tube under the NWC condition (E0 meas) and under the using a metering pump to control its concentration to
overlapping term of the HWC condition and the NWC 200 mg kg71 at the injection point. After traversing the
condition (E00 meas). The ECP of 316L SS tube (E316LSS) test section the water was cooled and then deionized by
vs. SHE was calculated from Equation (12) with passing through an ion exchange resin.
measured potential of platinum wire (Emeas) against Water chemistry conditions are summarized in
316L SS tube under the NWC condition. Table 2. Runs 1 and 2 demonstrated the in situ
measurement of corrosion depth via the resistance
EPt;NWC ¼ E0meas  E00meas  0:52 ð11Þ measurements of the wire probe and provided corro-
sion kinetics under typical BWR water chemistry
E316LSS ¼ Emeas þ EPt;NWC ð12Þ conditions. The effects of water chemistry change on
corrosion depth were studied in Runs 2 and 3. For the
case of Run 2, ECP under the HWC condition was
2.2. Apparatus and water chemistry conditions controlled to be lower than the ECP at which the phase
Corrosion depths were measured using the high- of iron oxide changes from a-Fe2O3 to Fe3O4. For the
temperature, high-pressure loop shown in Figure 3. case of Run 3, ECP under the HWC condition was
Operating conditions are listed in Table 1. The water in kept higher than the ECP at which the phase of iron
the pure water tank was circulated to the ion exchange oxide changes from a-Fe2O3 to Fe3O4.
resin column continuously to keep electrical conduc- In Run 1, specimens were immersed under the
tivity below 30 mS m71. For the NWC condition, the HWC condition for 168 h. In Run 2, the specimens
water was sparged with argon. For the HWC condi- were immersed for 118 h under the NWC condition,
tion, nitrogen mixed with 10% hydrogen gas was used 191 h under the HWC condition and then 51 h again
to control the dissolved hydrogen concentration to under the NWC condition. In Run 3, the specimens
150 mg kg71. The main water flow rate was kept at were immersed for 120 h under the NWC condition
10.8 kg h71 using a metering pump. Pressure at the test and then 90 h under the HWC condition.
section was set to 10 MPa to avoid boiling and the
water was heated to 553 K at the test section using a
heat exchanger and electrical heater. After reaching 2.3. Oxide film analysis
the steady state, temperature was controlled to with- Surface and cross section morphologies and che-
in +1K in the test section. Hydrogen peroxide (45 mg mical compositions of oxide films formed on the 316L
kg71) was injected just upstream from the test section SS wire probe and 316L SS tube were examined by
to avoid its decomposition before reaching the test SEM and LRS, respectively, to consider the effects of
section for getting the NWC condition. The flow rate shape of the test specimens and current application on
of hydrogen peroxide solution was set to 0.048 kg h71 the corrosion.
1082 K. Ishida and D. Lister

Table 1. Operating conditions of high-temperature, high- which is much smaller than the inner radius of 316L SS
pressure loop. tube (3.52 mm). Hydrogen peroxide or hydrogen will
Parameters Values diffuse two-dimensionally to the 316L SS wire probe
surfaces, but will diffuse one-dimensionally to the 316L
Electrical conductivity at pure water 530 SS tube surfaces. Diffusion amount of hydrogen
tank (mS m71)
Flow rate of high-pressure metering 10.8 peroxide or hydrogen to the metal surfaces for the
pump (kg h71) 316L SS wire probe will be about eight times as large as
Test section that for the 316L SS tube by the calculation using
Temperature (K) 553 Equation (17).
Pressure (MPa) 10
Hydrogen peroxide concentration at 45
hydrogen peroxide solution tank (mg kg71)
2p  ðr þ dÞ  l
ðratioÞ ¼ ð17Þ
Flow rate of hydrogen peroxide injection 0.048 2p  r  l
pump (kg h71)
On the contrary, metal ion concentration dissolved
form the 316L SS wire probe surfaces will be eight
times as dilute as that dissolved from the 316L SS tube
Table 2. Water chemistry conditions. surfaces. Corrosion of the 316L SS wire probe may be
Chemicals larger than that of the 316L SS tube due to these
(mg kg71) phenomena.
Immersion Water chemistry The shape of specimens may also affect the surface
Run no. time (h) condition H2O2 H2
morphologies or chemical composition of oxide film.
1 168 HWC – 150 Corrosion increases the volume of 316L SS because the
2 118 NWC 1 200 – Pilling–Bedworth ratios of iron, nickel and chromium,
191 HWC – 150 major elements included in 316L SS, are 2.06, 3.92 and
51 NWC 2 200 –
3 120 HWC 200 – 1.68, respectively [21]. Oxide film thickness is 1 mm or
90 NWC – 150 less, which is a significant thickness relative to the 316L
SS wire probe radius but is a negligible thickness
relative to the 316L SS tube radius. So oxide film on
The shape of test specimens will affect the diffusion the 316L SS tube surface may grow one-dimensionally.
amount of hydrogen peroxide or hydrogen to the metal On the contrary, oxide film on the 316L SS wire probe
surfaces and the diffusion amount of metal ion from may grow two-dimensionally.
metal surfaces. Diffusion layer thickness (d) was Current application may function as cathodic
calculated to be 169 mm from Equation (13) using the protection [22,23], which may affect surface and cross
mass transfer coefficient of straight pipe (Km, Equation section morphologies or chemical composition of oxide
(14)), diffusion coefficient of hydrogen peroxide (D), film. However, the increased temperature of the 316L
equivalent diameter of flow path (d), Reynolds number SS wire probe by supplying a current is negligible as
(Re) and Schmidt number (Sc) [15,19,20]. Diffusion described in Section 2.1.
coefficient of hydrogen peroxide (D) at 553 K was
obtained from the Einstein–Stokes equation (Equation
3. Results and discussion
(15)) using viscosity of water (Z), Boltzmann constant
(kb) and radius of chemical species (rC) [15]. Equivalent 3.1. Effects of temperature elevation
diameter of flow path (d) was calculated from Equation Measurements of resistance during temperature
(16) using volume (M) and surface area (S) of test elevation in Run 1 are shown in Figure 4. Data
section. considered to be for the steady state at each
temperature are plotted as open circles. Set tempera-
D
d¼ ð13Þ tures are also shown in Figure 4. Based on the
Km measurements during temperature elevation, the tem-
perature dependence of specific resistivity of the 316L
D
K ¼ 0:0165  Re0:86  Sc0:33  ð14Þ SS wire probe was obtained and is shown in Figure 5,
d along with literature data for 316L SS [14]. Resistivity
was calculated supposing that the diameter of 316L SS
kT wire probe was 50 mm. The result was empirically
D¼ ð15Þ
6p  Z  rC expressed as Equation (18).
4M
d¼ ð16Þ
S r ¼ c1  T2 þ c2  T þ c3 ð18Þ

Diffusion layer thickness is about three times as Here, c1 ¼ 3.434 6 10712 O m K72, c2 ¼ 71.502 6
large as the radius of 316L SS wire probe (25 mm), but 1079 O m K71, c3 ¼ 9.843 6 1077 O m.
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1083

The correlation coefficient is 0.998, which is better AISI 660 SS, resistivity changed less than 5% [24]. So,
than that for first order fitting (correlation coefficient: the effect of heat treatment may be small for 316L SS,
0.967). These results are about 25% higher than the because 316L SS is one type of austenitic stainless steel.
literature data at 553 K. Diametrical variation From these considerations, the compositional variation
(+5 mm) of 316L SS wire probe is considered to be is considered to be the most probable cause.
one reason for the deviation of +20%. Some data for
annealed specimens have been summarized in the
literature [24]. Data scatter of +6% (grey area of 3.2. Fluctuation of resistance measurements
Figure 5) is attributed to compositional variation Measurements of resistance at 553 K in Run 1 are
within the specifications of 316L SS and this is a shown in Figure 6. After reaching the set temperature
second reason for the resistivity difference. Regarding at the test section, the corrosion depth measurement
the effect of cold working on resistivity of 316L SS, it was started and the immersion time was defined as 0 h.
was reported that resistivity of specimens strained 40% Fluctuations in the resistance measurements are with-
in tension is about 1% higher than that of annealed in +0.05 O , which corresponds to a corrosion depth
specimens [25]. The effect of heat treatment is of +0.007 mm. Since corrosion depths are about
considered as a third reason. Resistivity was reported 0.5 mm, the fluctuations are about +1.4%.
to change about 10–100% in precipitation-hardened
type AISI 631 SS [24]. However, for austenitic type
3.3. Corrosion depth
The measurements of corrosion depth for Run 1 are
shown in Figure 7. Measured ECP at the steady state
is 70.56 V vs. SHE (listed in Table 3). Literature data
for 561 K water containing small amounts of dissolved
oxygen (55 mg kg71) with no hydrogen (no ECP
reported) [10] and for 561 K water containing dis-
solved oxygen (15 mg kg71) and dissolved hydrogen
(150 mg kg71; ECP of 70.46 V vs. SHE) [12] are also
plotted. The results of this work are in good agreement
with the data of [12], although the data of [10] are
about one order of magnitude lower. Since corrosion at
or near the activate state is known to be sensitive to
ECP, and since a slight amount of oxygen will change
the ECP, the deviation from [10] results may be due to
differences in ECP. Hydrogen enhances the anodic
current, which would cause of lowering the ECP even
though oxygen concentration is the same [26]. Hydro-
Figure 4. Resistance change during temperature elevation gen is the reducing agent to dissolve the protective
for Run 1. oxide film and that might enhance the corrosion.

Figure 5. Temperature dependence of resistivity of 316L SS Figure 6. Immersion time dependence of measured
wire probe. resistance for Run 1.
1084 K. Ishida and D. Lister

Figure 7. Comparison of corrosion depth of 316L SS under


HWC condition among literature data and results of this
work for Run 1.
Figure 8. Comparison of corrosion depth of 316L SS under
NWC condition among literature data and results of this
Table 3. Measured ECP of steady state. work for Run 2.

Run no. Water chemistry condition ECP (V vs. SHE)


1 HWC 70.56
2 NWC 1 þ0.15 (Run 1) are shown in Figure 10. Raman spectra were
HWC 70.56
NWC 2 0.00 measured at three different points for each specimen
3 NWC þ0.15 because NiFe2O4 and a-Fe2O3 are localized as the
HWC 70.42 outer layer [12]. For both specimens, a peak at 680–
685 cm71 can be detected, which is attributed to
FeCr2O4 or Fe3O4. In the case of the wire probe, a
peak is also detected at 811 cm71, and since this peak
The results of corrosion depth measurements in is also seen on the specimen before exposure in the
pure water at 553 K under the NWC condition for loop, it cannot be attributed to a compound formed by
118 h (Run 2) are shown in Figure 8 along with immersion in high-temperature water.
literature data and the experimental conditions at Kim [12] describes the characteristics of the oxide
which they were obtained [8–10,12,27]. Measured ECP film formed in 561 K water containing 150 mg kg71 of
of the NWC condition (NWC 1) was þ0.15 V vs. SHE. hydrogen and 15 mg kg71 of oxygen as analysed by
The results of this work for water containing hydrogen SEM and transmission electron microscopy (TEM).
peroxide are in good agreement with the rather Large crystallites of g-Fe2O3 and a-Fe2O3 as the outer
scattered literature data, which were obtained for layer and fine grains containing FexCr37xO4 as the
water containing oxygen. inner layer were observed. g-Fe2O3 and a-Fe2O3 were
not detected in the present work. The differences
might be due to a slight difference in water chem-
3.4. Oxide film analysis istry. Only hydrogen (150 mg kg71) was added and
The SEM images of the surface and cross section of ECP was about 70.56 V vs. SHE in the present
the oxide film formed on the wire probe and on the work, whereas in [12] oxygen was added to the
tube under the HWC condition as Run 1 are shown in amount of 15 mg kg71 and ECP was about 70.46 V
Figure 9. For both specimens, octahedral crystallites vs. SHE. According to thermodynamic calculations
(constituting the outer layer) are scattered on the [28], the phase of iron oxide changes from a-Fe2O3 to
surface and along the interface at the inner layer of the Fe3O4 at an ECP of 70.47 V vs. SHE. Oxygen of
oxide film. There is no apparent difference in 15 mg kg71 would make the water chemistry of [12]
morphology. study a little more oxidative than that in the present
Raman spectra of the oxide films formed on the work, possibly leading to the formation of the higher-
wire probe and on the tube under the HWC condition order iron oxide.
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1085

Figure 9. SEM image of surface and cross section of oxide film formed under HWC condition for Run 1.

and then 51 h again under the NWC condition


(measured ECP þ 0.00 V vs. SHE). For both speci-
mens, octahedral crystallites form a densely packed
outer layer but no crystallites can be seen for the
inner layer. There is no apparent difference in
morphology.
Raman spectra of the oxide films formed on the
wire probe and on the tube after Run 2 are shown in
Figure 12. Raman spectra were also measured at three
different points for each specimen. For both specimens,
peaks due to a-Fe2O3 (293, 412, 612 cm71), Fe3O4
(660 cm71) and NiFe2O4 (497, 699 cm71) are detected.
However, relative peak intensity of a-Fe2O3 for 316L
SS tube is smaller than that for 316L SS wire probe.
The Raman scattering power of a-Fe2O3 is well known
to be much larger than that of Fe3O4 [29]. So, a small
difference in the a-Fe2O3 formation under the NWC
condition may cause the difference in relative peak
intensity of a-Fe2O3.
Again, in the case of the wire probe, the peak at
811 cm71 is attributed to contamination before exposure
in the loop. Kim [12] reported on the SEM and TEM
Figure 10. Raman spectra of oxide film formed under HWC characterization of oxide films formed in 561 K water
condition for Run 1. containing nominally 15 mg kg71 of hydrogen and 200 mg
kg71 of oxygen with some changes in conditions. Speci-
The SEM images of the surface and cross section of mens were immersed under the NWC condition for 2
oxide film formed on the wire probe and on the tube weeks, the HWC condition for 2 weeks and then the NWC
after Run 2 are shown in Figure 11. The specimens condition again for 2 weeks. As in the experiments
were immersed for 118 h under the NWC condition reported here, large crystallites of g-Fe2O3, a-Fe2O3 and
(measured ECP þ 0.15 V vs. SHE), 191 h under the NiFe2O4 as the outer layer and fine grains containing
HWC condition (measured ECP 7 0.56 V vs. SHE) FexCr37xO4 as the inner layer were detected.
1086 K. Ishida and D. Lister

Figure 11. SEM image of surface and cross section of oxide film formed under NWC condition for Run 2.

measurements of the resistance of a thin wire is


therefore judged to be a useful and valid technique.

3.5. Corrosion kinetics


Corrosion of stainless steel in pure water at 553 K
is thought to be controlled by a parabolic rate law or a
logarithmic rate law described as Equation (19) or
Equation (20), respectively [7,9]. Consequently, the
corrosion depth data were fitted with the equations:
dW a pffiffi
¼ pffiffi or W ¼ a0  t ð19Þ
dt 2 t

dW bk
¼ or W ¼ k  lnðb0 t þ 1Þ ð20Þ
dt btþ1

where W is the corrosion depth and a, a0 , b, b0 and k


are constants.
The fitting results of the corrosion depth measured
under the HWC condition are shown in Figure 13.
Neither of the two equations can be fitted individually.
Figure 12. Raman spectra of oxide film formed under NWC Corrosion is therefore postulated to be controlled by a
condition for Run 2.
logarithmic rate law for the initial 1.4 h and by a
parabolic rate law thereafter. Thus, data were fitted by
The results of corrosion measurements and oxide Equation (20) until 1.4 h and data after 1.4 h were
film analyses on the wire probe and the tube show no fitted by Equation (21).
effects from shape of the test specimens or the
application of electric current. Evaluating corrosion pffiffiffiffiffiffiffiffiffiffi
depth under high-temperature pure water by in situ W ¼ a0  t  t 0 þ W0 ð21Þ
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1087

Figure 14. Fitting results of corrosion depth under NWC


condition for Run 2.

Figure 13. Fitting results of corrosion depth under HWC


condition for Run 1.

Here, W0 is the corrosion depth at 1.4 h and t0 is


the period of data fitted by the logarithmic rate law (1.4
h). This multi-step regression curve fits the measured
data well; the logarithmic rate law until 1.4 h with
b0 ¼ 115.7 h71 and k ¼ 0.02665 mm h71 and the
parabolic rate law after 1.4 h with a0 ¼ 0.02651 mm
h70.5. Such multi-step control of corrosion may reflect
the change in kinetics as formation of the oxide film on
the bare surface gives way to metal ion diffusion
through the film.
The fitting results of the corrosion depth measured
under the NWC condition (NWC 1) in Run 2 are
shown in Figure 14. In this case, the data are fitted Figure 15. Corrosion rate under HWC condition, NWC
condition and Run 2 calculated using fitting equations.
nicely by the logarithmic rate law (b0 ¼ 3191 h71 and
k ¼ 0.02834 mm h71).
Corrosion rates were calculated using the fitting 6.7 6 1074 mm h71, with which the result of the
equation for the HWC condition (Run 1) and the present work is in good agreement (3.7 6 1074 mm
NWC condition (Run 2), and are shown in Figure 15. h71 at 100 h for the NWC condition). However,
Corrosion rate of the HWC condition is larger than transpassivation potential was 0.00 V vs. SHE,
that of the NWC condition excluding the initial period which is lower than the ECP for the NWC condition
of less than 0.1 h. The active state at low ECP in the present work. Corrosion rate increases with
(70.34 V vs. SHE) was observed by anodic polariza- increasing ECP above transpassivation potential. This
tion curve measurement and the critical current of may imply that not only ECP but also hydrogen
passivation was reported to be 28 mA m72 [11]. peroxide itself affect the corrosion rate. Hydrogen
Corrosion rate was calculated from the critical current peroxide may assist the oxide formation to control
of passivation as about 3.7 6 1073 mm h71 supposing corrosion.
corrosion by divalent metal ions and density of 79.8 t Corrosion kinetics and corrosion rate clearly
m73 and the present result of the present work depend on water chemistry such as ECP and hydrogen
(1.5 6 1073 mm h71 at 100 h for the HWC condition) peroxide. Differences in corrosion kinetics and corro-
is in good agreement. Passive current density was also sion rate may contribute to the difference in the oxide
reported to be 5 mA m72 [11]. Corrosion rate film. Base metal is covered by the oxide film under the
calculated from passive current density was NWC condition, which controls the corrosion rate.
1088 K. Ishida and D. Lister

However, base metal is covered insufficiently by the change, 54 h was needed to get the steady state of ECP.
oxide film under the HWC condition. The ECP of the HWC condition for Run 3 is higher
than that for Run 2. The reason for the ECP difference
itself and for the difference in the decrease of ECP
3.6. Effects of water chemistry change on corrosion between Run 2 and Run 3 under HWC condition may
Corrosion depths for Run 2 and Run 3 are shown be because of a small amount of oxygen ingress into
in Figures 16 and 17, respectively. For the case of Run the pure water tank in Run 3. Deionized water sparged
2, ECPs of the first NWC condition (NWC 1), the by argon was supplied continuously during Run 3.
HWC condition and the second NWC condition Oxygen might not have been sufficiently degassed from
(NWC 2) were þ0.15, 70.56 and þ0.00 V vs. SHE the water. The change from the NWC condition to the
at the steady state, respectively. Water chemistry was HWC condition did not increase the corrosion depth
changed from the NWC condition to the HWC for Run 3.
condition at 118 h. After the water chemistry change, The difference of corrosion behaviour by the water
14 h was needed to get the steady state of ECP. Water chemistry change might be attributable to ECP.
chemistry was changed from the HWC condition to According to thermodynamic calculations [28], the
the NWC condition at 309 h. After the water chemistry phase of iron oxide changes from a-Fe2O3 to Fe3O4 at
change, 12 h was needed to get the steady state of ECP. an ECP of 70.47 V vs. SHE. Since in Run 2 the ECP
The change from the NWC condition to the HWC decreases below 70.47 V vs. SHE under the HWC
condition increases the corrosion depth above the condition, the oxide film can undergo reconstruction
extrapolated line to continue corrosion under the NWC and promote an increase in corrosion.
condition at the same time as ECP is decreased Corrosion depth was fitted by the Equation (22)
to 70.56 V vs. SHE. By contrast, the change from the and corrosion rates were calculated.
HWC condition to the NWC condition at 309 h imme-
diately increases the corrosion depth above the extra- n
W ¼ a00  ðt  t0 Þ þ W0 ð22Þ
polated line for corrosion under the HWC condition
after the chemistry change.
For the case of Run 3, ECPs of NWC and HWC Here, a0 ¼ 0.008 mm h0.05, t0 ¼ 130 h, W0 ¼ 0.364
were þ0.15 and 70.42 V vs. SHE, respectively. Water mm and n ¼ 0.95 from 130 h to 160 h; a0 ¼ 0.0012 mm
chemistry was changed from the NWC condition to the h0.32, t0 ¼ 160 h, W0 ¼ 0.384 mm and n ¼ 0.68 from
HWC condition at 118 h. After the water chemistry 160 h to 309 h; and a0 ¼ 0.0019 mm h0.45, t0 ¼ 309 h,

Figure 16. Effects of water chemistry change on corrosion Figure 17. Effects of water chemistry change on corrosion
depth for Run 2. depth for Run 3.
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1089

W0 ¼ 0.423 mm and n ¼ 0.55 from 309 h to 360 h.


The results are shown in Figure 15. Corrosion rates Abbreviations and nomenclature
under the HWC condition in Run 2 (from 130 h to 309 304 SS: type 304 L stainless steel
h) are lower than that for the HWC condition 316L SS: type 316L stainless steel
measured in Run 1. This means that oxide film BWR: boiling water reactor
formed under NWC condition until 118 h provided ECP: electrochemical corrosion
corrosion protection. However, oxide film for corro- potential
sion protection may be partially damaged, according to HWC: hydrogen water chemistry
the results that the increase of corrosion rate only LRS: laser Raman spectroscopy
occurred when ECP is below 70.47 V vs. SHE at NPP: nuclear power plant
which the phase of iron oxide changes from a-Fe2O3 to NWC: normal water chemistry
Fe3O4. PLR: primary loop recirculation system
PTFE: polytetrafluoroethylene
V vs. SHE: volt versus standard hydrogen
4. Conclusion electrode
The in situ system for monitoring corrosion via RPV: reactor pressure vessel
electrical resistance measurements of a thin wire was SEM: scanning electron microscopy
applied to study the corrosion of 316L SS in pure water TEM: transmission electron microscopy
at 553 K. Corrosion depths were measured under the a: constant
HWC condition (150 mg kg71 of hydrogen) and the a0 : constant
NWC condition (200 mg kg71 of hydrogen peroxide) a00 : constant
and the kinetic expressions for the corrosion were b: constant
formulated for both conditions. The effects on corro- b0 : constant
sion depth from water chemistry change from the c 1: constant
NWC condition to the HWC condition and vice versa c 2: constant
were also studied. c 3: constant
Corrosion depth could be measured using a 50 mm CH2: molar fraction of dissolved
diameter wire probe mounted axially in the 316L SS hydrogen
tube. Measurements were in good agreement with d: equivalent diameter of flow path
literature data for both the HWC and the NWC D: diffusion coefficient of hydrogen
conditions. Differences in oxide morphology could not peroxide
be observed between the 316L SS wire and a 316L SS EPt,NWC: potential of platinum wire under
tube by SEM analysis or in oxide chemical composi- NWC condition
tion between the wire and tube by LRS analysis. The EPt,HWC: potential of platinum wire under
results of corrosion measurements and oxide film HWC condition
analyses on the wire probe and the tube showed no E316LSS: ECP of 316L SS tube
effects from shape of the test specimens or the Emeas: measured potential of platinum
application of electric current. wire
Corrosion kinetics was evaluated by fitting equa- E0 meas: measured potential of platinum
tions to the measurements of corrosion depth. The data wire against 316L SS tube under
for the HWC condition could not be fitted individually NWC condition
by either a parabolic or a logarithmic rate law E00 meas: measured potential of platinum
equation, but a two-step logarithmic–parabolic law wire under overlapping term of
fitted them very well. A single-step logarithmic law HWC and NWC conditions
fitted the data for the NWC condition very well. f: pipe friction coefficient
Changes in corrosion rate by the water chemistry [Hþ]: activity of hydrogen ion
changes were readily detected with the resistance I: current
technique. Corrosion amount change could be clearly k: constant
observed for the water chemistry change from the kb: Boltzmann constant
NWC condition to the HWC condition with ECP kH: Henry’s constant
of 70.56 V vs. SHE, which was lower than the ECP at Km: mass transfer coefficient of
which the phase of iron oxide changes from a-Fe2O3 to straight pipe
Fe3O4 (70.47 V vs. SHE). However, the corrosion l: length of 316L SS wire probe
amount change could not be observed for the water M: volume of test section
chemistry change from the NWC condition to the n: constant
HWC condition with ECP of 70.42 V vs. SHE, which Nu: Nusselt number
was higher than the phase of iron oxide changes Pr: Prandtl number
from a-Fe2O3 to Fe3O4. R: resistance
1090 K. Ishida and D. Lister

RG: gas constant [6] R.L. Cowan, M.E. Indig, J.N. Kass, R.J. Law, and L.L.
R: radius of 316L SS wire probe Sundberg, Experience with Hydrogen Water Chemistry in
Boiling Water Reactors, Water Chemistry of Nuclear
r t: initial radius of 316L SS wire Reactor Systems 4 Vol. 1, Bournemouth, UK, October
probe 13–17, 1986, pp. 29–36.
rC: radius of chemical species [7] J. Robertson, The mechanism of high temperature
Re: Reynolds number aqueous corrosion of stainless steels, Corros. Sci. 32
R0: initial resistance of 316L SS wire (1991), pp. 443–465.
[8] C. Degueldre, S. O’Prey, and W. Francioni, An in-line
probe diffuse reflection spectroscopy study of the oxidation of
Rt: resistance of 316L SS wire probe stainless steel under boiling water reactor conditions,
at t Corros. Sci. 38 (1996), pp. 1763–1782.
S: surface area of test section [9] H. Inagaki, A. Nishikawa, Y. Sugita, and T. Tsuji,
Sc: Schmidt number Synergy effect of simultaneous zinc and nickel addition
on cobalt deposition onto stainless steel in oxygenated
t: immersion time high temperature water, J. Nucl. Sci. Technol. 40 (2003),
t0 : immersion time from corrosion pp. 143–152.
rate change [10] K. Ohashi, T. Honda, E. Kashimura, and Y. Furutani,
T: temperature Effect of dissolved oxygen on corrosion of ferrous
t0: period of data fitted by materials in high temperature and high pure water,
Boshoku Gijyutsu (Corros. Eng.) 37 (1988), pp. 198–204
logarithmic rate law (1.4 h) [in Japanese].
V: potential drop [11] M. Tachibana, K. Ishida, Y. Wada, R. Shimizu, N. Ota,
W: corrosion depth and N. Hara, Determining factor for anodic polariza-
W0 : corrosion depth just before tion curve of typical structural materials of boiling water
corrosion rate change reactors in high temperature–high purity water, J. Nucl.
Sci. Technol. 49 (2012), pp. 253–262.
W 0: corrosion depth at 1.4 h [12] Y.J. Kim, Characterization of the oxide film formed on
Dr: corrosion depth of 316L SS wire type 316 stainless steel in 2888C in cyclic normal and
probe hydrogen water chemistry, Corrosion 51 (1995), pp. 849–
d: diffusion layer thickness 860.
Z: viscosity of water [13] T. Miyazawa, S. Uchida, T. Satoh, Y. Morishima, T.
Hirose, Y. Satoh, K. Iimura, Y. Wada, H. Hosokawa,
r: resistivity of 316L SS wire probe and N. Usui, Effects of hydrogen peroxide on corrosion
of stainless steel, (IV) Determination of oxide film
properties with multilateral surface analyses, J. Nucl.
Acknowledgements Sci. Technol. 42 (2005), pp. 233–241.
[14] The Japan Society of Mechanical Engineers, JSME
The authors are grateful to A. Feicht, L. Liu and P.
Data Book: Heat Transfer, 4th ed., Maruzen Co., Ltd.,
Srisukvatananan, research staff in the Nuclear Engineering
Tokyo, 1986, p. 55 [in Japanese].
Group at the University of New Brunswick, for their help
[15] The Japan Society of Mechanical Engineers, JSME
with the experiments and the surface analyses. The Natural
Steam Table, The Japan Society of Mechanical En-
Sciences and Engineering Research Council of Canada and
gineers, Tokyo, 1999, p. 29 [in Japanese].
the CANDU Owners Group are thanked for financial
[16] H.A. Pray, C.E. Schweickert, and B.H. Minnich,
support.
Solubility of hydrogen, oxygen, nitrogen, and helium
in water, Ind. Eng. Chem. 44 (1952), pp. 1146–1151.
[17] Y.J. Kim, Effect of variations in simulated BWR water
References chemistry on high temperature electrochemistry of
[1] R.A. Shaw, Getting at the source: reducing radiation stainless steel, Corrosion 2005, Houston, TX, April 3–
fields, Nucl. Technol. 44 (1979), pp. 97–103. 7, 2005, PN05579.
[2] C.C. Lin, A review of corrosion product transport and [18] Y.J. Kim and P.L. Andresen, Data Quality, Issues and
radiation field buildup in boiling water reactors, Prog. Guidelines for ECP Measurement in High Temperature
Nucl. Energy 51 (2009), pp. 207–224. Water, Corrosion 2001, Houston, TX, March 11–16,
[3] D.H. Lister, The transport of radioactive corrosion 2001, PN01137.
products in high-temperature water, Nucl. Sci. Eng. 59 [19] F.P. Berger and K.-F.F.-L. Hau, Mass transfer in
(1976), pp. 406–426. turbulent pipe flow measured by the electrochemical
[4] Y. Okamura, A. Sakashita, T. Fukuda, H. Yamashita, method, Int. J. Heat Mass Transf. 20 (1997), p. 1185.
and T. Futami, Latest SCC Issue of Core Shroud and [20] A.J. Elliot, D.R. McCracken, G.V. Buxton, and N.D.
Recirculation Piping in Japanese BWRs, Transactions of Wood, Estimation of rate constant for near-diffusion-
the 17th International Conference on Structural Me- controlled reactions in water at high temperatures, J.
chanics in Reactor Technology, Prague, Czech Republic, Chem. Soc. Faraday Trans. 86 (1990), pp. 1539–1547.
August 17–22, 2003, Paper Number WG01-1. [21] N.B. Pilling and R.E. Bedworth, The oxidation of
[5] Y. Wada, N. Shigeneka, N. Uetake, and S. Uchida, metals at high temperature, J. Inst. Met. 29 (1923), pp.
Numerical Simulation of SCC Environment in a BWR 529–591.
Primary Coolant System, Proceedings of the Eighth [22] R.B. Mears and R.H. Brown, A theory of cathodic
International Symposium on Environmental Degrada- protection, Trans. Electrochem. Soc. 74 (1938), pp. 519–
tion of Materials in Nuclear Power Systems – Water 531.
Reactor Vol. 1, Amelia Island, Florida, August 10–14, [23] R.H. Brown and R.B. Mears, Cathodic protection,
1997, pp. 574–581. Trans. Electrochem. Soc. 81 (1942), pp. 455–483.
Journal of Nuclear Science and Technology, Volume 49, No. 11, November 2012 1091

[24] R.H. Bogaard, P.D. Desai, H.H. Li, and C.Y. Ho, [27] D.H. Lister and G. Venkateswaran, Effects of magne-
Thermophysical properties of stainless steels, Thermo- sium and zinc additives on corrosion and cobalt
chim. Acta 218 (1993), pp. 373–393. contamination of stainless steels in simulated BWR
[25] S.W. Yang and J.E. Spruiell, Cold-worked state and coolant, Nucl. Technol. 125 (1999), pp. 316–331.
annealing behavior of austenitic stainless steel, J. Mater. [28] D. Cubicciotti, Equilibrium chemistry of nitrogen and
Sci. 17 (1982), pp. 677–690. potential – pH diagrams for the Fe–Cr–H2O system in
[26] Y.J. Kim, Electrochemical Interactions of Hydrogen, BWR water, J. Nucl. Mater. 167 (1989), pp. 241–248.
Oxygen and Hydrogen Peroxide on Metal Surfaces in [29] O.N. Shebanova and P. Lazor, Raman spectroscopic
High Temperature, High Purity Water, Proceedings of study of magnetite (FeFe2O4): a new assignment for the
the Eighth International Symposium on Environmental vibrational spectrum, J. Solid State Chem. 174 (2003),
Degradation of Materials in Nuclear Power Systems – pp. 424–430.
Water Reactors Vol. 2, August 10–14, 1997, Amelia
Island, FL, pp. 641–648.

You might also like