You are on page 1of 14

crystals

Article
Flow-Assisted Corrosion of Carbon Steel in Simulated Nuclear
Plant Steam Generator Conditions
Iva Betova 1 , Martin Bojinov 2, * and Vasil Karastoyanov 2

1 Institute of Electrochemistry and Energy Systems, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria;
i.betova@iees.bas.bg
2 Department of Physical Chemistry, University of Chemical Technology and Metallurgy, 1756 Sofia, Bulgaria;
vasko_kar@uctm.edu
* Correspondence: martin@uctm.edu

Abstract: Flow-assisted corrosion occurs via increased dissolution and/or mechanical degradation of
protective oxide formed on the surface of construction materials in direct contact with coolant liquids.
In the present paper, this phenomenon is studied on carbon steel in an ammonia-ethanolamine-
hydrazine electrolyte by in situ electrochemical impedance spectroscopy in conditions that closely
simulate those that prevail in nuclear plant steam generators. Based on the obtained results, a
quantitative kinetic model of the process is proposed and parameterized by nonlinear regression
of experimental data to the respective transfer function. On the basis of the experimental and
calculational results, it is concluded that flow-assisted corrosion of carbon steel is limited by oxide
dissolution and cation ejection processes and the protective layer–coolant interface. Expressions
for the film growth and corrosion release processes are proposed and successfully compared to
operational data.

Keywords: flow-assisted corrosion; carbon steel; steam generator; impedance spectroscopy; kinetic model

Citation: Betova, I.; Bojinov, M.;


Karastoyanov, V. Flow-Assisted
1. Introduction
Corrosion of Carbon Steel in Flow-accelerated or flow-assisted corrosion (FAC), also denoted as erosion–corrosion,
Simulated Nuclear Plant Steam of carbon or low-alloyed steel piping, occurs when the rate of dissolution of the protective
Generator Conditions. Crystals 2023, oxide film that forms on the internal piping surface into a stream of flowing water or wet
13, 1115. https://doi.org/10.3390/ steam is enhanced, leading to an increased wall thinning rate [1–9]. This phenomenon is
cryst13071115 encountered in both single and two-phase flow conditions and is the result of an increase
Academic Editors: Luntao Wang,
in the rate of corrosion or material dissolution, induced by the relative movement between
Baojie Dou, Xuejie Li and a corrosive fluid and a material surface; thus it does not involve erosion or cavitation
Zhongyu Cui damage. A thin layer of iron oxide (mostly magnetite) forms on the inside surface of
carbon steel exposed to deoxygenated water in the temperature range 100–300 ◦ C. This
Received: 25 June 2023 layer protects the underlying piping from the corrosive environment and limits further
Revised: 11 July 2023
corrosion. An increase in fluid velocity, temperature, and alteration of water chemistry
Accepted: 16 July 2023
can lead to magnetite dissolution at the oxide–coolant interface and its substitution by
Published: 17 July 2023
new oxide formed at the alloy–oxide interface, resulting in material removal and thinning
of the piping [2–8]. In general, the FAC process consists of two steps. The first step
is the production of soluble iron at the oxide–water interface, while the second step is
Copyright: © 2023 by the authors.
the transfer of the soluble and particulate corrosion products to the bulk flow across the
Licensee MDPI, Basel, Switzerland. diffusion boundary layer. Although FAC is characterized by a general reduction in pipe
This article is an open access article wall thickness for a given component, it frequently occurs over a limited area within this
distributed under the terms and component due to local high areas of turbulence [3,5,9].
conditions of the Creative Commons Several parameters influence the extent of degradation due to FAC, including the
Attribution (CC BY) license (https:// geometrical configuration of the components, piping orientation, position inside the pipe,
creativecommons.org/licenses/by/ flow Reynolds number, fluid chemistry, temperature, piping material, and flow turbulence
4.0/). structure, which affect the surface shear stress and mass transfer coefficients [9–12]. The

Crystals 2023, 13, 1115. https://doi.org/10.3390/cryst13071115 https://www.mdpi.com/journal/crystals


flow Reynolds number, fluid chemistry, temperature, piping material, and flow turbu-
lence structure, which affect the surface shear stress and mass transfer coefficients [9–12].
Crystals 2023, 13, 1115 The dissolution rate is, in fact, equal to the sum of the current densities of the dissolution 2 of 14
of the oxide and transfer of iron ions from the metal through the oxide layer, which em-
phasizes its role as one of the controlling factors of FAC [10,11]. Several works [13–17]
have been devoted
dissolution rate is,tointhe characterization
fact, equal to the sum of the thin
of the iron oxide
current layers
densities of formed on the car-
the dissolution of the
bonoxide and transfer of iron ions from the metal through the oxide layer, whichprevailing
steel components under the chemical and hydrodynamical conditions emphasizes
within theassecondary
its role one of thecircuit of PWR
controlling plants
factors of (de-aerated
FAC [10,11].and alkaline
Several worksturbulent
[13–17]water or
have been
wet steam).toHowever,
devoted the mechanism
the characterization of thin
of the the process
iron oxidehas layers
not been studied
formed in detail
on the carbon with
steel
electrochemical
components under methods. the Very
chemicalrecently, some new approaches,
and hydrodynamical including
conditions the wire
prevailing beamthe
within
electrode
secondary withcircuit
electrochemical
of PWR plants impedance spectroscopy,
(de-aerated and alkaline haveturbulent
shown promise
water orinwetprobing
steam).
theHowever,
initiation and propagation of erosion-corrosion [18,19] but, to
the mechanism of the process has not been studied in detail with electrochemical the best of our
knowledge,
methods. these methods some
Very recently, have not newbeen applied toincluding
approaches, steam-generator
the wirematerials and envi-
beam electrode with
electrochemical impedance spectroscopy, have shown promise in probing the initiation and
ronments.
propagation of erosion-corrosion
In that context, the aim of the present [18,19]paper
but, toisthe best of our the
to investigate knowledge,
corrosionthese methods
of a typical
have
alloy notinbeen
used steam applied
generatorsto steam-generator
(carbon steel 22K)materials
with aand environments.
specially designed setup to ensure
turbulent In that context, First,
conditions. the aim theofsetup
the present
and the paper is to investigate
associated hydrodynamicthe corrosion of a typical
calculations are
alloy used in steam generators (carbon steel 22K) with a specially
described. Second, in situ chrono-potentiometric and impedance spectroscopic measure- designed setup to ensure
turbulent
ments in the conditions.
temperatureFirst, rangethe setup °C
100–240 and inthe associated
a simulated hydrodynamic
coolant of a nuclear calculations
plant steamare
described.
generator are Second,
presented in and
situ discussed.
chrono-potentiometric
Quantitativeand impedanceofspectroscopic
interpretation the impedance measure-
data
ments in themodel
temperature ◦ C in a simulated coolant of a nuclear plant steam
with a kinetic allowsrange for its100–240
parameterization in terms of interfacial rate constants,
generator
diffusion are presented
coefficients of ionicanddefects,
discussed.
and Quantitative
field strengthinterpretation
in the forming ofoxide.
the impedance data
Finally, the
with a kinetic model allows for its parameterization in terms of
model is validated by comparison with operational data for FAC and directions for future interfacial rate constants,
diffusion
research arecoefficients
indicated. of ionic defects, and field strength in the forming oxide. Finally, the
model is validated by comparison with operational data for FAC and directions for future
research are
2. Materials andindicated.
Methods
The experiments
2. Materials described in the paper are performed in a recirculation loop
and Methods
equipped with a special insert for a FAC study, mounted immediately after the preheater
The experiments described in the paper are performed in a recirculation loop equipped
of the coolant (Figure 1). It is constructed from a tube made of AISI 316 stainless steel that
with a special insert for a FAC study, mounted immediately after the preheater of the coolant
plays the role of sample holder. The insert playing the role of flow accelerator is located
(Figure 1). It is constructed from a tube made of AISI 316 stainless steel that plays the role
in the middle of the tube. To simulate the coolant as a single-phase flow at large Reynolds
of sample holder. The insert playing the role of flow accelerator is located in the middle of
numbers, the k-ε model was implemented using a finite element method implemented in
the tube. To simulate the coolant as a single-phase flow at large Reynolds numbers, the
commercial software (Comsol Multiphysics, Burlington, MA, USA). Due to the axial sym-
k-ε model was implemented using a finite element method implemented in commercial
metry of the system, 2D calculations were performed. The fluid was modelled as pure
software (Comsol Multiphysics, Burlington, MA, USA). Due to the axial symmetry of
water at 100–240
the system, 2D°С and 90 barwere
calculations (density 0.95–0.82The
performed. g cm −3 and dynamic viscosity 10−3–10−4
fluid was modelled as pure water at
Pa100–240
s). ◦ C and 90 bar (density 0.95–0.82 g cm−3 and dynamic viscosity 10−3 –10−4 Pa s).

Figure
Figure 1. Scheme
1. Scheme of the
of the insert
insert used
used to investigate
to investigate flow-assisted
flow-assisted corrosion.
corrosion.

Samples for working electrodes were cut from 22K steel (AISI 1022) with a nominal
Samples for working electrodes were cut from 22K steel (AISI 1022) with a nominal
composition listed in Table 1. The actual composition estimated by glow-discharge optical
composition listed in Table 1. The actual composition estimated by glow-discharge optical
emission spectroscopy is also given in the Table.
emission spectroscopy is also given in the Table.
Table 1. Chemical composition of the steel used in the present investigation (wt.%, balance Fe).
Table 1. Chemical composition of the steel used in the present investigation (wt.%, balance Fe).
22K C Cr Cu Mn Mo Ni P S Si
22K C Cr Cu Mn Mo Ni P S Si
nominal
0.17–0.24 ≤0.25≤0.25
nominal 0.17–0.24 ≤0.30
≤0.30 0.7–1.0
0.7–1.0 ≤0.15
≤0.15 ≤0.25
≤0.25 ≤0.035
≤0.035 ≤0.035
≤0.035 0.17–0.37
0.17–0.37
analyzed
analyzed 0.22 0.22 0.23 0.23 0.040.04 0.91
0.91 0.12
0.12 0.40
0.40 0.012
0.012 0.02
0.02 0.310.31

Electrochemical measurements are carried out with a CompactStat.h10030 (Ivium Tech-


nologies, Eindhoven, The Netherlands) operating in a floating mode. In a three-electrode
configuration, a Pt plate (99.9%, Goodfellow) plays the role of a counter-electrode, whereas
the reference electrode is a Pd (99.9%, Goodfellow) cathodically polarized with 10 µA against
Crystals 2023, 13, 1115 3 of 14

another Pt to approximate the reversible hydrogen electrode (RHE). Before each measurement,
this electrode is calibrated vs. a 3 M KCl/AgCl/Ag reference (LL-type, Metrohm, Switzerland).
The potentials were recalculated to the standard hydrogen electrode scale using temperature
dependences of the RHE estimated with commercial software [20,21]. The water chemistry
used was ammonia-hydrazine-ethanolamine (AMETA, 3.0–5.0 mg kg−1 NH3 , 20–100 µg kg−1
N2 H4 , 0.3–0.5 mg kg−1 HOCH2 CH2 NH2 , pH25 ◦ C = 9.8). To prepare the electrolyte, p.a. N2 H4 ,
NH3 и HOCH2 CH2 NH2 (Sigma Aldrich) with an Fe content of less than 0.00005% were used.
Measurements were performed at 100, 130, 160, 180, 200, 220, and 240 ◦ C using the
following procedure. After filling the loop with coolant, the hot part, including the flow-
through cell with the flow-accelerating insert, is compressed to 90 bars, heated up to 80 ◦ C,
and purged with N2 (99.999%) for 16 h to reach a dissolved oxygen concentration <10 µg kg−1 .
After reaching this value, the temperature of the flow-through cell is increased gradually
so that the highest measurement temperature (240 ◦ C) is reached for ca. 2 h. Purging of
the feedwater in the reservoir with N2 continued till the end of the experiment duration
(typically 72–96 h). During the first 24 h, the inlet flow in the cell is 10 dm3 h−1 , from
24–48 h it is maintained at 4 dm3 h−1 , and from 48 to 72 h—increased again to 10 dm3 h−1 .
Impedance spectra were measured in a frequency range of 0.5 mHz to 11 kHz with an ac
signal of 40 mV (rms) or 1 µA (rms). Spectra measured by both methods were identical
within the reproducibility limit (±1% by impedance magnitude and ±2◦ by phase shift). The
interval of spectra registration is between 2 and 8 h. Linearity was checked by measuring
spectra with signal amplitudes between 10 and 40 mV, whereas causality was ensured by a
Kramers–Kronig compatibility test. For complex nonlinear least squares fitting of spectra to a
transfer function based on the proposed kinetic model, the Levenberg–Marquardt algorithm
implemented in an Origin Pro platform (Originlab, Northampton, MA, USA) was employed.
To estimate the oxide thickness, the exposed samples were galvanostatically re-
duced with a current density of −30 µA cm−2 at 22 ± 1 ◦ C in a borate buffer solution
(0.1 mol dm−3 Na2 B4 O7 + 0.005 mol dm−3 H3 BO3 , pH = 9.2) de-aerated with N2 . Exper-
iments were conducted in a three-electrode glass cell featuring a Pt (99.9%, Goodfellow)
plate counter electrode and a 3 M KCl/AgCl/Ag reference with an Autolab 302N poten-
tiostat/galvanostat (Metrohm Eco Chemie, Utrecht, The Netherlands). The concentration
of soluble iron generated during reduction was estimated with an ICP-OES apparatus
(Prodigy, Teledyne Leeman Labs, Mason, OH, USA).

3. Results
3.1. Hydrodynamic Calculations
The profiles of linear flow rate and Reynolds number in the flow-through cell calcu-
lated by the procedure described in the Experimental section are presented in Figure 2a,b
for a temperature of 240 ◦ C. The temperature dependence of maximum values of hydro-
dynamic parameters estimated in the vicinity of the working electrode are presented
Crystals 2023, 13, x FOR PEER REVIEW in
4 of 15
Figure 2c. Based on the performed calculations, it can be concluded that turbulent con-
ditions are ensured in the flow-accelerating insert in the studied temperature interval.

Figure 2. Cont. c 1×104


0.88
1×104
v / m s-1

0.86 v / m s-1 9×103


Re

Re
8×103
Crystals 2023, 13, 1115 4 of 14

c 1×104
0.88
1×104
c 1×104
0.88

v / m s-1
3
0.86 v / m s-1 1×1049×10

Re
Re
8×103

v / m s-1
0.86 v / m s-1 9×103

Re
0.84 Re 3
8×1037×10

0.84 3
7×1036×10
0.82
90 120 150 180 210 240
6×103
0.82 t / °C
90 120 150 180 210 240
t / °C
Figure 2. (a) Profiles of linear flow rate (m s−1) and (b) Reynolds number in the cell at 240 °С, (c
dependence
Figure of flow of
2. (a) Profiles rate and flow
linear Reynolds number
rate (m on(b)
s−1 ) and temperature.
Reynolds number in the cell at 240 ◦ C,
Figure 2. (a) Profiles of linear flow rate (m s−1) and (b) Reynolds number in the cell at 240 °С, (c)
(c) dependence of flow rate and Reynolds number on temperature.
dependence of flow rate and Reynolds number on temperature.
3.2. Influence of Temperature and Flow Rate on Corrosion Potential
3.2. Influence of Temperature and Flow Rate on Corrosion Potential
The corrosion-potential
3.2. Influence time Rate
of Temperature and Flow dependences in Potential
on Corrosion the interval 100–240 °С are presented in
The corrosion-potential time dependences in the interval 100–240 ◦ C are presented in
Figure 3.
The3.corrosion-potential time dependences in the interval 100–240 °С are presented in
Figure
Figure 3.
-0.50 100
-0.50 100
130
-0.55 130160
-0.55 160180
E / V vs. SHE

-0.60 180200
E / V vs. SHE

200220
-0.60
220240
-0.65 240
-0.65
-0.70
-0.70
-0.75
-0.75
-0.80
-0.80 0 10 20 30 40 50 60 70 80
0 10 20 30 40 50 60 70 80
t / ht / h
Figure 3. Corrosion potential of 22K in simulated secondary coolant at temperatures from 100 to 240 ◦ C.
Figure3.3.
Figure Corrosion
Corrosion potential
potential of 22К
of 22К in simulated
in simulated secondary
secondary coolant
coolant at temperatures
at temperatures from
from 100 100 to 240
to 240
°С. 3 −
°С. During the first 24 h of oxidation (at a volume flow rate 10 dm h ) potentials are 1

lower than −0.70 V and weakly dependent on time. Decreasing the volume flow rate
During
to 4 During
dm 3 h−the the
1 leads first
firstto24a 24 h oxidation
h of of oxidation
displacement thea(at
of(at a volume
volume
corrosion flow flow rate
rate 10
potential to dm103 hdm
more h−1) values
potentials
−1) 3potentials
positive are are
lowerthan
lower
regardless than of −0.70
−0.70
the V V
andand weakly
weakly
temperature dependent
ofdependent
oxidation. on time.
on time.
The Decreasing
Decreasing
subsequent thein
increase the
volume volume
the flow flow flow
raterate torate
leads 4 to 4
dm
to
dm 3ah
h leads
3return
−1 −1
leadsoftothe
a displacement
to acorrosion
displacement of the
potential corrosion
to
of the more
corrosionpotential
negative to more
values
potential in
to thepositive
more valuesvalues
temperature
positive regard-
range regard
130–220 ◦ C, whereas such a return is not observed at 100 and 240 ◦ C.
less
lessofofthe the temperature
temperature of of
oxidation. TheThe
oxidation. subsequent increase
subsequent in theinflow
increase the rate
flowleads
rate to a to a
leads
The interval of corrosion potentials at all temperatures is inserted in the corresponding
E-pH diagrams of the Fe–H2 O system calculated by commercial software [20] in Figure 4.
At temperatures up to 180 ◦ C, the thermodynamically stable corrosion product is predicted
to be the divalent iron ion FeOH+ , whereas at higher temperatures corrosion potentials
are located in the magnetite stability region. These calculations indicate that both aqueous
ions and magnetite are expected to be formed during the corrosion of 22K in the AMETA
secondary coolant.
The interval of corrosion potentials at all temperatures is inserted in the correspond-
ing E-pH diagrams of the Fe–H2O system calculated by commercial software [20] in Figure
4. At temperatures up to 180 °С, the thermodynamically stable corrosion product is pre-
dicted to be the divalent iron ion FeOH+, whereas at higher temperatures corrosion poten-
Crystals 2023, 13, 1115
tials are located in the magnetite stability region. These calculations indicate that 5both
of 14
aqueous ions and magnetite are expected to be formed during the corrosion of 22K in the
AMETA secondary coolant.

Figure 4. pH diagrams of an Fe–H22O O system


system at
at the studied temperatures. The
The rectangle shows the
interval of corrosion potentials at the operational pH of the coolant (marked with a vertical
vertical line).
line).

3.3. Influence of Temperature and Flow Rate on Impedance Spectra


The impedance spectra of 22K in AMETA electrolyte at the investigated temperatures
(100–240 ◦ C) and oxidation times (1–72 h) are collected in Figures 5–10 in Bode coordinates.
To facilitate the identification of processes in the high-frequency region, 95% of the elec-
trolyte resistance is subtracted from the spectra. The impedance magnitude at frequencies
approaching zero (Zf→0 ) that can be interpreted as the polarization resistance (i.e., the
inverse of corrosion/oxidation rate) decreases with temperature in the interval 100–180 ◦ C
and increases at higher temperatures, i.e., the temperature dependence of the corrosion rate
is non-monotonous.
(100–240 °С)
(100–240 °С) and
and oxidation
oxidation times
times (1–72(1–72 h)
h) are
are collected
collected inin Figures
Figures 5‒10
5‒10 in
in Bode
Bode coordi-
coordi-
nates. To facilitate the identification of processes in the high-frequency region,
nates. To facilitate the identification of processes in the high-frequency region, 95% of 95% of the
the
electrolyte resistance is subtracted from the spectra. The impedance
electrolyte resistance is subtracted from the spectra. The impedance magnitude at fre- magnitude at fre-
quencies approaching
quencies approaching zerozero (Z
(Zff→00)) that
that can
can be
be interpreted
interpreted as as the
the polarization
polarization resistance
resistance (i.e.,
(i.e.,

the inverse
the inverse of of corrosion/oxidation
corrosion/oxidation rate) rate) decreases
decreases with with temperature
temperature inin the
the interval
interval 100–
100–
Crystals 2023, 13, 1115 180 °C and increases at higher temperatures, i.e., the temperature dependence of the6 cor-
of 14
180 °C and increases at higher temperatures, i.e., the temperature dependence of the cor-
rosion rate
rosion rate is
is non-monotonous.
non-monotonous.

104 60 60
104 60 60
100 °C 104 100 °C
100 °C 104 100 °C
50 50
50 50

40 40
cm2 2

40

cm2 2
40

phase/ /° °

phase/ /° °
|Z|/ /ΩΩcm

1033

|Z|/ /ΩΩcm
10 30 103 30

phase
103

phase
30 30
|Z|

20

|Z|
20 20
27.2
37.1
37.1 20
27.2 49.3
37.1 49.3
37.1 10 57.9
10
102 49.3 10 57.9
59.8 10
102 49.3 102 59.8
57.9
57.9 102
0 0
10-3 10-2 10-1 100 101 102 0
10-3 10-2 10-1 100 101 102 10-3 10-2 10-1 100 101 1022 0
10-3 10-2 10-1 100 101 10
ff // Hz
Hz
ff // Hz
Hz
Figure 5. Impedance spectra at 100 °C as a function of time. Points—experimental values, solid
◦ C as
Figure
Figure 5.
5. Impedance
Impedance spectra
spectra at
at 100
100 °C as aa function
function of time. Points—experimental
Points—experimental values,
values, solid
solid
lines—best-fit calculation according to the proposed model. Parameter is the exposure time in h.
lines—best-fit
lines—best-fit calculation
calculation according
according toto the
the proposed
proposed model.
model. Parameter
Parameter is
is the
the exposure
exposure time
time in
in h.
h.

104 60 104 60
104 60 104 60
130 °C 130 °C
130 °C 130 °C
50 50
50 50
40 40
cm2 2

40
cm2 2

40
phase/ /° °

phase/ /° °
|Z|/ /ΩΩcm

1033
|Z|/ /ΩΩcm

10 1033
30 10
phase

30 30

phase
30
|Z|

29.8 20
|Z|

29.8 20 29.8 20
40.0
40.0
29.8
40.0
20
48.6 40.0
48.6 10 48.6
64.5 10 48.6
102 64.5 64.5 10
102 74.0 102 64.5 10
74.0 102 74.0
74.0
0
10-3 10-2 10-1 100 101 102 0 0
10-3 10-2 10-1 100 101 102 10-3 10-2 10-1 100 101 102 0
10-3 10-2 10-1 100 101 102
ff // Hz
Hz
Crystals 2023, 13, x FOR PEER REVIEW ff // Hz
Hz 7 of 15
Figure 6.
Figure 6. Impedance spectra
Impedance spectra
6. Impedance at
spectra at 130
at 130 ◦ C as
°C
130 °C a function
as aa function of
of time.
time. Points—experimental values,
Points—experimental values, solid
values, solid
solid
Figure as function of time. Points—experimental
lines—best-fit calculation
lines—best-fit calculation according
calculation according
according toto the
to the proposed
the proposed model. Parameter
proposed model. Parameter is
Parameter is the
is the exposure
the exposure time in h.
exposure time in h.
lines—best-fit
50 50
160 °C

40 160 °C
40
|Z| / Ω cm2

|Z| / Ω cm2

103 103
phase / °

30
phase / °

30

20 20
2.1
6.4 10 102 50.3 10
2 13.0 60.5
10
29.6 69.3
34.7
0 0
10-3 10-2 10-1 100 101 102 10-3 10-2 10-1 100 101 102
f / Hz f / Hz
Impedance spectra
spectra at ◦ C as a function of time. Points—experimental values, solid
Figure
Figure 7.
7. Impedance at 160
160 °C as a function of time. Points—experimental values, solid
lines—best-fit calculation according
according to
to the
the proposed
proposed model.
model. Parameter
Parameter is
is the
the exposure
exposure time
time in
in h.
h.

180 °C 50 180 °C 50

40 40
10 29.6 69.3
29.6 69.3
34.7
34.7 00
-3 -2 -1 0 1 2
00
10
10-3 10
10-2 10
10-1 10
100 10
101 10
102 10-3
10-2
10 -1
100
10 1
10 2
10-3 10-2 10-1 100 101 102
ff // Hz
Hz ff // Hz
Hz
Crystals 2023, 13, 1115 Figure 7 of 14
Figure 7.
7. Impedance
Impedance spectra
spectra at
at 160
160 °C
°C as
as aa function
function of
of time.
time. Points—experimental
Points—experimental values,
values, solid
solid
lines—best-fit
lines—best-fitcalculation
calculationaccording
accordingto tothe
theproposed
proposedmodel.
model. Parameter
Parameterisisthe
theexposure
exposuretime
timein
inh.
h.

180
180°C
°C 50
50 180
180°C
°C 50
50

40
40 40
40
cm22

cm22
phase/ /°°

phase/ /°°
3 3
10
|Z|/ /ΩΩcm

10

|Z|/ /ΩΩcm
103 30 103 30
30 30

phase

phase
20 20
|Z|

|Z|
20 20
1.3
1.3 45.2
8.0
8.0 10
10 10 2 45.2 10
10
10 2
17.6 102 50.6
102 17.6 50.6
68.7
24.3 68.7
24.3
00 00
-3 -2 -1 0 1 2 -3 -2 -1 0 1 2
10 10 10 10 10 10 10
10-3 10
10-2 10
10-1 10
100 10
101 10
102
10-3 10-2 10-1 100 101 102
ff // Hz
Hz ff // Hz
Hz
Figure 8.
Figure8.
Figure Impedance
8. Impedance spectra
Impedance spectra at
spectra at 180
180◦°C
at180 °C as
C asasaaafunction
function of
functionof time.
oftime. Points—experimental
time.Points—experimental values,
Points—experimentalvalues, solid
values,solid
solid
lines—best-fit calculation
lines—best-fitcalculation according
calculationaccording to the
accordingtotothe proposed
theproposed model.
proposedmodel. Parameter
model.Parameter is the
Parameterisisthe exposure
theexposure time
exposuretime in h.
timeininh.
h.
lines—best-fit

70
70 10 4 60
60
104 200 °C
200 °C 200 °C
200 °C
60
60 50
50
50
50 40
40
cm22

cm22
phase/ /°°

phase/ /°°
|Z|/ /ΩΩcm

|Z|/ /ΩΩcm

1033 10 40
40 103310
30
phase

phase
30
30
30
|Z|

|Z|

1.0
1.0 27.6
27.6 20
20
4.7
4.7
20
20 40.6
40.6
8.1 48.4
8.1 48.4
2 11.4 10 10 2 51.7 10
10
10
102
11.4
21.0 10 102 51.7
69.8
21.0 69.8
00 00
-3 -2 -1 0 1 2 -3 -2 -1 0 1 2
10
10-3 10
10-2 10
10-1 10
100 10
101 10
102 10 10 10 10 10 10
10-3 10-2 10-1 100 101 102
ff // Hz
Hz ff // Hz
Hz
Crystals 2023, 13, x FOR PEER REVIEW 8 of 15
Figure
Figure 9. Impedance spectra at 200 ◦°CC as aa function of time. Points—experimental values, solid
Figure 9.
9. Impedance spectra at at 200
200 °C as as a function
function of
of time.
time. Points—experimental
Points—experimental values,values, solid
solid
lines—best-fit
lines—best-fitcalculation
lines—best-fit calculationaccording
calculation accordingto
according tothe
to theproposed
the proposedmodel.
proposed model.Parameter
model. Parameterisisisthe
Parameter theexposure
the exposuretime
exposure timein
time inh.
in h.
h.
60 70
240 °C 240 °C

104 50 60
104
50
40
-phase / °

phase / °
2

40
|Z| / Ω cm2

|Z| / kΩ cm

30
103 30
2.1 20 103 51.3
70.7
6.3
74 20
7.9
83.5
12.9 10
17.5 95.6 10
24.4
2 0
10
10-3 10-2 10-1 100 101 102 102 0
10-3 10-2 10-1 100 101 102
f / Hz
f / Hz
Figure
Figure10.
10. Impedance
Impedancespectra
spectraatat240
240°C◦ Cas
as aa function
function of
of time.
time. Points—experimental
Points—experimentalvalues,
values,solid
solid
lines—best-fit
lines—best-fit calculation according to the proposed model. Parameter is the exposure time in
calculation according to the proposed model. Parameter is the exposure time in h.
h.

Deconvolution of the impedance spectra indicates three processes with different


characteristic frequencies in the phase angle curves. According to the published interpre-
tation of the impedance spectra of construction materials in high-temperature water elec-
trolytes [22–27], the process with the highest characteristic frequency is related to the elec-
trical properties of the protective layer of oxide; that at intermediate frequencies, with ion
102 0
10-3 10-2 10-1 100 101 102 102 0
10-3 10-2 10-1 100 101 102
f / Hz
f / Hz
Figure 10. Impedance spectra at 240 °C as a function of time. Points—experimental values, solid
Crystals 2023, 13, 1115 8 of
lines—best-fit calculation according to the proposed model. Parameter is the exposure time in h. 14

Deconvolution of the impedance spectra indicates three processes with different


characteristic frequencies
Deconvolution of theinimpedance
the phase angle curves.
spectra According
indicates three to the published
processes interpre-
with different
tation of the impedance
characteristic frequenciesspectra
in theof construction
phase materials
angle curves. in high-temperature
According waterinter-
to the published elec-
trolytes [22–27],
pretation the process spectra
of the impedance with theofhighest characteristic
construction frequency
materials is related to the
in high-temperature elec-
water
trical properties
electrolytes of the
[22–27], theprotective
process withlayer ofhighest
the oxide; that at intermediate
characteristic frequencies,
frequency is relatedwith ion
to the
and electron
electrical transferofatthe
properties theprotective
oxide–coolant
layerinterface,
of oxide; and
thatthe low-frequency
at intermediate process is iden-
frequencies, with
tified
ion aselectron
and the transport
transferof at
point
the defects in the oxide.
oxide–coolant Spectra
interface, and calculated using theprocess
the low-frequency model
described
is in as
identified thethe
Discussion
transportsection aredefects
of point shownin with
thesolid lines
oxide. and demonstrate
Spectra its ability
calculated using the
model described
to reproduce in the Discussion
quantitatively section
both the are shown
magnitude and with solid linesdependence
the frequency and demonstrate
of exper-its
ability
imental tospectra.
reproduce
Thus,quantitatively
the data canboth the magnitude
be used and the
to parameterize thefrequency
model anddependence
estimate rates of
experimental
of productionspectra. Thus,
of soluble ironthe data
and can be used
magnetite to parameterize
particles, the model
which is attempted inand estimate
the next part
rates
of theofpaper.
production of soluble iron and magnetite particles, which is attempted in the next
part of the paper.
3.4. Estimation of Oxide Film Thickness
3.4. Estimation of Oxide Film Thickness
The potential quantity of charge curves (chrono-potentiometric curves at constant
Thedensity)
current potentialinquantity
de-aerated of charge curvessolution
borate buffer (chrono-potentiometric
(pH 8.4) of a range curves at constant
of samples oxi-
current density) in de-aerated borate buffer solution (pH 8.4) of a range
dized for 72 h at different temperatures in AMETA coolant are collected in Figure of samples oxidized
11.
for 72 h at different temperatures in AMETA coolant are collected in Figure 11.
-0.2
a 400 b
100 °C
350
-0.3 130 °C
E / V vs. SHE

160 °C
240 °C
300
L / nm

-0.4 250

200

-0.5 150
method of tangents
100 divalent Fe concentration
-0.6 50
0 50 100 150 200 250 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7
-2
Q / mC cm 103 T-1 / K-1

Figure11.
Figure 11. (a)
(a) Potential
Potential quantity
quantity of
of charge
charge curves
curves (chrono-potentiometric
(chrono-potentiometriccurves
curvesat
atconstant
constant current
current
density) in
density) in de-aerated
de-aerated borate
borate buffer
buffer solution
solution (pH
(pH 8.4)
8.4) of
of aa range
range of
of samples
samples oxidized
oxidized forfor 72
72 hh at
at
different temperatures.
different temperatures. TheThe dashed
dashed lines
lines show
show the
the used
used method
method ofof estimation
estimation of
of the
the quantity
quantity ofof
charge equivalent to the oxide layer thickness; (b) Calculated thickness vs. temperature dependence
using the two methods outlined in the text.

The method to estimate oxide thickness is based on the assumption that during
constant current polarization in borate buffer solution, reductive dissolution of the oxide
proceeds according to the overall reaction

Fe3 O4 +8H+ +2e− → 3Fe2+ +4H2 O (1)

The oxide thickness is calculated using two methods—the method of tangents in


potential-charge curves, illustrated in Figure 11 via Faraday’s law with 100% coulombic
efficiency, and from the concentration of divalent iron measured by ICP-OES in the solution
after completion of polarization (once again, using Faraday’s law). The estimates obtained
via the two methods coincide with the reproducibility of the results, as shown in Figure 11
(right). These values are used to calibrate the calculations described in the following section.

4. Discussion
4.1. Description of the Model
The present model is based on the mixed-conduction model for oxide films (MCM),
in analogy to previous estimates of the rates of oxidation and corrosion of low-alloyed and
Crystals 2023, 13, 1115 9 of 14

stainless steels in simulated primary and secondary coolants of nuclear power plants [22–27].
Assuming that the protective oxide on the studied steel is magnetite, the following processes
take place at the alloy–oxide (A/O) and oxide–coolant (O/C) interfaces

k
Fe
(A/O) Fem → Fei•• +2e0
k ••
O
(A/O) 3Fem → 3FeFe +4VO +8e0
k y+
(O/C) Fei•• →
2Fe
Feaq + (y − 2)e0
k (2)
•• 2O
(O/C) 4H2 O + 4VO → 4OO +8H+
kd
(O/C) Fe3 O4 +2H+ +2e− +2H2 O  3Fe(OH)2,aq
(O/C) H2 O + 2e− → H2 +2OH−

Thus, the reactions of the formation of soluble iron and hydroxide particles proceed
in parallel at the protective oxide–coolant interface. The processes at the inner and outer
interfaces are coupled by oxygen transport via vacancies and iron transport via interstitial
cations. Accordingly, the transfer function that describes the impedance of the system has
the form [27]

1   −1
Z = Rel + Zox + ZO/C , ZO/C = −1
, Zox = Ze−1 + Zion,O
−1 −1
+ + Zion,Fe (3)
jωCO/C + RO/C

In the above equations, the impedance of the oxide–coolant interface ZO/C is repre-
sented as a parallel combination of an interfacial capacitance CO/C and a charge transfer
resistance of a single-step reaction (water reduction with hydrogen evolution) RF/S . On
the other hand, the impedance of the protective oxide Zox is a parallel combination of the
impedances of its electric properties, Ze , and two ionic transport impedances correspond-
ing to oxygen (Zion,O ) and iron (Zion,Fe ), respectively. The detailed expressions of these
impedances are given below [27]

RT [1 + jωρd εε 0 exp(2KL)] F RT k2O + k2Fe


Ze ≈ ln , K= E, ρd = 2 (4)
2jωFELCsc 1 + jωρd εε 0 RT F De kO + k Fe

RT RT
Zion,O ≈  q  , Zion,M ≈  q  (5)
4jω 4jω
4F2 k O ( 1 − α ) 1 + 1+ DO K2
4F2 k Fe ( 1 − α ) 1 + 1+ DFe K2

In these expressions, De , DO , and DFe are the diffusion coefficients of electrons, oxygen
vacancies, and iron cation interstitials, L is the protective film thickness, E is the electric
field strength in the oxide, ε its dielectric constant, α is the part of the potential consumed
as a potential drop at the oxide–coolant interface.
According to the model, the film thickness increases with time following the equation

1 h i 3α FE
L ( t ) = L0 + ln 1 + Vm,ox kO bO e−bO L0 t , bO = O (6)
bO RT

Here, LO is the initial film thickness, Vm,ox —its molar volume (44.5 cm3 mol−1 for
Fe3 O4 ), and αO —the transfer coefficient of the oxidation reaction at the alloy–oxide interface.
The rate of dissolution of Fe through the oxide is given by:
 
dn Fe2+ (t) 2α Fe FE
≈ k Fe exp − L(t) (7)
dt RT
Here, LO is the initial film thickness, Vm,ox—its molar volume (44.5 cm3 mol−1 for Fe3O4),
and αО—the transfer coefficient of the oxidation reaction at the alloy–oxide interface. The
rate of dissolution of Fe through the oxide is given by:
dnFe2+ (t )  2α FE 
≈ kFe exp  − Fe L(t )  (7)
Crystals 2023, 13, 1115 dt  RT  10 of 14

The integration of the last equation gives the possibility to calculate the corrosion
release rate (i.e., the rate of generation of soluble iron ions) as the alloy thinning rate in
The integration of the last equation gives the possibility to calculate the corrosion
mm/y.
release rate (i.e., the rate of generation of soluble iron ions) as the alloy thinning rate in
mm/y.  2α FE 
vcorr = 3.153 × 108Vm, Fe k Fe exp  − 2αFe FE L(t )  (8)
vcorr = 3.153 × 108 Vm,Fe k Fe exp  − RT L(t)
Fe
(8)
RT
Vm,Fe being the atomic volume of Fe (7.2 cm3 mol−1).
Vm,Fe being the atomic volume of Fe (7.2 cm3 mol−1 ).
4.2. Estimation of Kinetic and Transport Parameters
4.2. Estimation of Kinetic and Transport Parameters
Using a complex nonlinear regression of the experimental impedance spectra with
Using a complex nonlinear regression of the experimental impedance spectra with
respect to the model equations by the Levenberg–Marquardt algorithm, estimates of the
respect to the model equations by the Levenberg–Marquardt algorithm, estimates of the
kinetic and transport parameters are obtained. As mentioned already above, the calcula-
kinetic and transport parameters are obtained. As mentioned already above, the calcu-
tions were calibrated with the oxide thicknesses at the end of exposure obtained by ca-
lations were calibrated with the oxide thicknesses at the end of exposure obtained by
thodic reduction in borate buffer at room temperature. The dependences of the main pa-
cathodic reduction in borate buffer at room temperature. The dependences of the main
rameters, layer thickness L, field strength Е, interfacial capacitance CO/C, and charge trans-
parameters, layer thickness L, field strength E, interfacial capacitance CO/C , and charge
fer resistance
transfer RO/C Rat the
resistance oxide/coolant interface at the time of exposure, are collected in
O/C at the oxide/coolant interface at the time of exposure, are collected
Figure 12. The dependences
in Figure 12. The dependences of of
thethe
remaining parameters
remaining parametersononoxidation time
oxidation were
time found
were foundto
be comparatively insignificant.
to be comparatively insignificant.
450 60
100
100
400 130
130
160 50
350 160
180
180
200
300 220 40 200
E / kV cm-1

220
240
L / nm

250 240
30
200
150 20
100
10
50
Crystals 2023, 13, x FOR PEER REVIEW 0 11 of 15
1 10 100 0 10 20 30 40 50 60 70 80
time / h time / h

100
130
160
2 180
200
CO/C / mF cm-2

RO/C / kΩ cm2

220
240
10
100
130
1 160
180
200
220
240

0 1
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
time / h time / h

Figure12.
Figure Filmthickness
12.Film thickness(L),
(L),field
fieldstrength
strength(E),
(E),interfacial
interfacialcapacitance
capacitance(C
(CO/C ), ),and
O/C andcharge
chargetransfer
transfer
resistance
resistance(R
(RO/C) depending
) depending ononexposure
exposuretime at
time different
at temperatures
different temperatures ◦ C).
(°С).
(
O/C

Thegrowth
The growthof
of the
the oxide
oxide follows a direct
direct logarithmic
logarithmiclaw lawatatall
alltemperatures
temperaturesininaccor-
ac-
dance with model predictions. The field strength in the oxide decreases with
cordance with model predictions. The field strength in the oxide decreases with time due time due
to the accumulation of point defects forming a space charge. It is worth noting
to the accumulation of point defects forming a space charge. It is worth noting that the that the
decrease of the volume flow rate at 24 h (from 10 to 4 dm −1 ) has some influence on the
3 h−1
decrease of the volume flow rate at 24 h (from 10 to 4 dm h ) has some influence on the
3

field strength and the charge transfer resistance, whereas no influence of the subsequent
increase of flow rate at 48 h is observed. This means that the oxide formed in turbulent
conditions is weakly influenced by the hydrodynamic parameters, as inferred already
from the analysis of the experimental impedance spectra.
Figure 12. Film thickness (L), field strength (E), interfacial capacitance (CO/C), and charge transfer
resistance (RO/C) depending on exposure time at different temperatures (°С).

The growth of the oxide follows a direct logarithmic law at all temperatures in ac-
cordance with model predictions. The field strength in the oxide decreases with time due
Crystals 2023, 13, 1115 11 of 14
to the accumulation of point defects forming a space charge. It is worth noting that the
decrease of the volume flow rate at 24 h (from 10 to 4 dm3 h−1) has some influence on the
field strength and the charge transfer resistance, whereas no influence of the subsequent
field strength
increase andrate
of flow the charge
at 48 h transfer resistance,
is observed. whereas
This means thatnothe
influence of the subsequent
oxide formed in turbulent
increase of flow rate at 48 h is observed. This means that the oxide
conditions is weakly influenced by the hydrodynamic parameters, as inferred formed in turbulent
already
conditions is weakly influenced by the hydrodynamic
from the analysis of the experimental impedance spectra. parameters, as inferred already from
the analysis of the experimental impedance spectra.
The temperature dependences of kinetic and transport parameters at the end of ex-
The(ca.
posure temperature dependences
72 h) are shown of kinetic
in Arrhenius and transport
coordinates in Figureparameters
13. at the end of
exposure (ca. 72 h) are shown in Arrhenius coordinates in Figure 13.
10-11 10-9
rate constant / mol cm-2 s-1

rate constant / cm s-1


k2Fe
k2O

10-12 kFe
kO

10-10
Crystals 2023, 13, x FOR PEER REVIEW 12 of 15
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7
103 T / K-1 103 T / K-1

400
30
350

300

E / kV cm-1
D / cm2 / s-1

L / nm

10-17 250
20
L / nm
200
E / kV cm-1
DFe 150
DO 10
100

10-18 50
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7
3 -1
10 T / K 103 T / K-1
0.08
80 2.0
0.92
RO/C-1 / kΩ-1 cm-2

0.06
CF/S / mF cm-2
Csc / μF cm-2

60 1.5 0.88

40 1.0 0.04 0.84

Csc / μF cm-2
20 0.5 0.80
CO/C / mF cm-2 RO/C-1 / kΩ-1 cm-2
0.02
α 0.76
0 0.0
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7
3 -1
10 T / K 103 T / K-1

Temperature dependences of kinetic and transport parameters at


Figure 13. Temperature at the
the end
end of
of exposure.
exposure.

The following conclusions can be drawn on the basis of their values:


• The rate constants of electrochemical processes at the alloy–oxide and oxide–coolant
interfaces increase exponentially with temperature, as expected. The values of the
apparent activation energies are 33 ± 2 and 23 ± 2 kJ mol−1 for the reactions at the inner
and outer interfaces, values that are typical for electrochemical processes;
Crystals 2023, 13, 1115 12 of 14

The following conclusions can be drawn on the basis of their values:


• The rate constants of electrochemical processes at the alloy–oxide and oxide–coolant
interfaces increase exponentially with temperature, as expected. The values of the
apparent activation energies are 33 ± 2 and 23 ± 2 kJ mol−1 for the reactions at the
inner and outer interfaces, values that are typical for electrochemical processes;
• The diffusion coefficients of oxygen and iron ions also follow an Arrhenius dependence
as a first approximation, the activation energy being on the order of 45 kJ mol−1 . That
value is ca. two times lower than the activation energy for magnetite formation in
steam via solid-state diffusion, i.e., the transport of cations and anions most probably
proceeds via grain boundaries in the nanocrystalline oxide;
• The field strength decreases with temperature in accordance with previous calculations
for stainless steel in a borate buffer solution at 150–300 ◦ C [28]. This is due to an increase
in the number of defects in the film with the temperature. It is important to mention
that the field strength in the oxide on carbon steel is much lower than that in the oxide
on stainless steel, which can be due to a different composition of the protective layer
(magnetite on carbon steel and chromite FeCr2 O4 on type 316 stainless steel);
• Both the capacitance of the space charge layer in the oxide and the capacitance at
the oxide–coolant interface in general decrease with temperature. The space charge
capacitance has relatively high values, which is in accordance with the fact that
magnetite is a degenerated semiconductor with a small depletion layer width. The
interfacial capacitance values (0.5–2.0 mF cm−2 ) point out to its nature as a pseudo-
capacitance of an intermediate of the film dissolution or water reduction reaction.
From the point of view of recent reviews of flow-assisted corrosion [3–8], the most
interesting dependence is that of the inverse of the charge transfer resistance at the oxide–
coolant interface (RO/C −1 ). This parameter passes through a maximum at ca. 160 ◦ C in
accordance with operational data for the flow-assisted corrosion rate in steam generators of
both conventional and nuclear power plants. It can be argued that in the studied conditions,
the rate of flow-assisted corrosion is determined by the processes at the interface of oxide–
solution and the maximum is due to the opposing influence of two factors—magnetite
solubility, which decreases with increasing temperature [7], and the dissolution rate of the
oxide, which increases with temperature.

5. Conclusions
The flow-assisted corrosion of carbon steel type 22K in an ammonia-ethanolamine
secondary coolant of a nuclear power plant is studied by in situ EIS in the temperature
interval 100–240 ◦ C at 90 bar and inlet volume flow rates 4–10 dm3 h−1 . Hydrodynamic
calculations indicate that the developed flow-through cell with an accelerating insert
ensures linear flow rates up to 1 m s−1 and Reynolds numbers above 104 , i.e., turbulent
conditions are reached. The influence of flow rate on corrosion potential is explained by a
change of the main thermodynamically stable product with increasing temperature (from
FeOH+ aq in the interval 100–160 ◦ C to magnetite at higher temperatures). The effect of
changing the flow rate on electrochemical impedance spectra is relatively insignificant,
which means that in the present experimental conditions, corrosion rates are limited by
interfacial processes and solid-state transport and not by the convective transfer of products
in the coolant.
The results are interpreted with a quantitative kinetic model featuring two parallel
processes—growth and dissolution of oxide with soluble Fe(OH)2 as the end product
and dissolution of Fe through the protective film resulting in soluble iron ions (probably
FeOH+ ) as the end product. Equations for the fluxes of iron hydroxide are limited by the
dissolution rate of oxide and divalent iron ions are limited by the transport and ejection
of interstitials from the oxide. The model is calibrated by oxide thicknesses at the end of
exposure estimated via galvanostatic cathodic reduction.
The model is fully parameterized by complex nonlinear regression of experimental
impedance spectra to the respective transfer function, allowing for the rate constants of film
Crystals 2023, 13, 1115 13 of 14

formation and metal oxidation at the alloy–oxide interface, diffusion coefficients of oxygen
vacancies, and interstitial iron cations, as well as field strength in the growing oxide to be
estimated as depending on temperature (100–240 ◦ C). On the basis of these calculations,
equations for oxide growth and release rates with time were derived. As a main conclusion,
the soluble iron release rate is proposed to obey the equation
 
2α FE
vcorr (t) = 2.271 × 109 k Fe exp − Fe L(t)
RT

i.e., it is determined by both the rate of iron oxidation at the alloy–oxide interface, the field
strength in the oxide, and the thickness of the protective layer. The temperature dependence
of the flow-assisted corrosion rate in the present experimental conditions indicates that it
is expressed by the reciprocal value of the charge transfer resistance at the oxide–coolant
interface. Further research is needed in order to establish quantitative correlations between
iron release rates and hydrodynamic parameters in a range that covers the operational
conditions of steam generators.

Author Contributions: Conceptualization, M.B. and I.B.; methodology, M.B.; validation, I.B. and
V.K.; formal analysis, I.B.; investigation, V.K.; resources, M.B.; data curation, V.K.; writing—original
draft preparation, I.B.; writing—review and editing, M.B.; visualization, I.B. All authors have read
and agreed to the published version of the manuscript.
Funding: This research was funded by The National Scientific Fund of Bulgaria, grant number
KTI-06-H59/4-2021 “Deterministic modeling of degradation of structural materials for energy systems
in high-temperature electrolytes”. Measurements were performed with equipment of the National
Scientific Infrastructure “Energy Storage and Hydrogen Energy” (ESHER, contract DO1-160/28.08.18).
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ahmed, W.H. Evaluation of the proximity effect on flow-accelerated corrosion. Ann. Nucl. Energy 2010, 37, 598–605. [CrossRef]
2. Betova, I.; Bojinov, M.; Saario, T. Predictive Modelling of Flow-Accelerated Corrosion—Unresolved Problems and Issues; VTT-R-08125-10;
VTT Technical Research Centre of Finland: Espoo, Finland, 2010; p. 55.
3. Poulson, B. Predicting and Preventing Flow Accelerated Corrosion in Nuclear Power Plant. Int. J. Nuclear Energy 2014, 2014, 423295.
[CrossRef]
4. Dooley, R.B. Flow-Accelerated Corrosion in Fossil and Combined Cycle/HRSG Plants. Power Plant Chem. 2008, 10, 68–89.
5. Trevin, S. Flow accelerated corrosion (FAC) in nuclear power plant components. In Nuclear Corrosion Science and Engineering;
Feron, D., Ed.; Woodhead Publishing Limited: Sawston, UK, 2012; pp. 186–229.
6. Tomarov, G.V.; Shipkov, A.A. Flow-Accelerated Corrosion Wear of Power-Generating Equipment: Investigations, Prediction, and
Prevention. Therm. Eng. 2018, 65, 493–503. [CrossRef]
7. Dooley, R.B.; Lister, D. Flow-Accelerated Corrosion in Steam-Generating Plants. Power Plant Chem. 2018, 20, 194–244.
8. Gipon, E.; Trevin, S. Flow-accelerated corrosion in nuclear power plants. In Nuclear Corrosion—Research, Progress and Challenges,
European Federation of Corrosion (EFC) Series; Elsevier: Amsterdam, The Netherlands, 2020; pp. 213–250.
9. Schmitt, G.; Bakalli, M. Flow-Assisted Corrosion. In Shreir’s Corrosion, 4th ed.; Richardson, T., Ed.; Elsevier: Amsterdam,
The Netherlands, 2007; Volume 2, pp. 954–987.
10. Vermorel, F.; Romand, M.; Charbonnier, M.; Bouchacourt, M. Caractérisation par émission et diffraction des rayons X des
couches d’oxydes formées par corrosion-érosion des aciers dans les circuits secondaires des centrales REP. J. Phys. IV 2000, 10,
Pr10-313–Pr10-322. [CrossRef]
11. Nazradani, S.; Nakka, R.K.; Hopkins, D.; Stevens, J. Characterization of oxides on FAC susceptible small-bore carbon steel piping
of a power plant. Int. J. Press. V. Pip. 2009, 86, 845–852. [CrossRef]
12. Laleh, M.; Xu, Y.; Tan, M.Y.J. A three-dimensional electrode array probe designed for visualising complex and dynamically
changing internal pipeline corrosion. Corros. Sci. 2023, 211, 110924. [CrossRef]
13. Moed, D.H.; Weerakul, S.; Lister, D.H.; Leaukosol, N.; Rietveld, L.C.; Verliefde, A.R.D. Effect of Ethanolamine, Ammonia, Acetic
Acid, and Formic Acid on Two-Phase Flow Accelerated Corrosion in Steam–Water Cycles. Ind. Eng. Chem. Res. 2015, 54,
8963–8970. [CrossRef]
Crystals 2023, 13, 1115 14 of 14

14. Olmedo, A.M.; Bordoni, R. Studies of Oxide Layers Grown at 260 ◦ C on A106 B Carbon Steel in Aqueous Medium with
Ethanolamine or Morpholine. Mater. Sci. Appl. 2015, 6, 783–791. [CrossRef]
15. Kim, S.; Kim, T.; Lee, Y.; Kim, J.H. Soft X-ray study on electronic structure of passive oxide layer of carbon steel and low alloy
steel in flow-accelerated corrosion environments of pressurized water reactors. Corros. Sci. 2019, 159, 108143. [CrossRef]
16. Zhang, T.; Li, T.; Lu, J.; Guo, Q.; Xu, J. Microstructural Characterization of the Corrosion Product Deposit in the Flow-Accelerated
Region in High-Temperature Water. Crystals 2022, 12, 749. [CrossRef]
17. Hu, Y.; Xin, L.; Hong, C.; Han, Y.; Lu, Y. Microstructural Understanding of Flow Accelerated Corrosion of SA106B Carbon Steel in
High-Temperature Water with Different Flow Velocities. Materials 2023, 16, 3981. [CrossRef] [PubMed]
18. Huo, Y.; Tan, M.Y.-T.; Forsyth, M. Visualising dynamic passivation and localised corrosion processes occurring on buried steel
surfaces under the effect of anodic transients. Electrochem. Commun. 2016, 66, 21–24. [CrossRef]
19. Xu, Y.; Tan, M.Y.J. Probing the initiation and propagation processes of flow accelerated corrosion and erosion corrosion under
simulated turbulent flow conditions. Corros. Sci. 2019, 151, 163–174. [CrossRef]
20. HSC Software, version 6.0; Outokumpu Research Oy: Outokumpu, Finland, 2006.
21. OLI Analyzer Studio, version 3.1; OLI Systems, Inc.: Parsippany, NJ, USA, 2010.
22. Sipilä, K.; Bojinov, M.; Mayinger, W.; Saario, T.; Stanislowski, M. Effect of chloride and sulfate additions on corrosion of low alloy
steel in high-temperature water. Electrochim. Acta 2015, 173, 757–770. [CrossRef]
23. Järvimäki, S.; Saario, T.; Sipilä, K.; Bojinov, M. Effect of hydrazine on general corrosion of carbon and low-alloyed steels in
pressurized water reactor secondary side water. Nucl. Eng. Des. 2015, 295, 106–115. [CrossRef]
24. Sipilä, K.; Bojinov, M.; Jäppinen, E.; Mayinger, W.; Saario, T.; Selektor, M. Localized corrosion of pressure vessel steel in a boiling
water reactor cladding flaw—Modeling of electrochemical conditions and dedicated experiments. Electrochim. Acta 2017, 241,
10–27. [CrossRef]
25. Alami, A.; Bojinov, M. Effect of sulfate and dissolved hydrogen on oxide films on stainless steel in high-temperature water. J. Solid
State Electrochem. 2017, 21, 3505–3518. [CrossRef]
26. Jäppinen, E.; Ikäläinen, T.; Järvimäki, S.; Saario, T.; Sipilä, K.; Bojinov, M. Corrosion behavior of carbon steel coated with
octadecylamine in the secondary circuit of a pressurized water reactor. J. Mater. Eng. Perform. 2017, 26, 6037–6046. [CrossRef]
27. Bojinov, M.; Jäppinen, E.; Saario, T.; Sipilä, K.; Toivonen, A. Effect of lead and applied potential on corrosion of carbon steel in
steam generator crevice solutions. Corros. Sci. 2019, 159, 108117. [CrossRef]
28. Bojinov, M.; Kinnunen, P.; Lundgren, K.; Wikmark, G. A Mixed-Conduction Model for the Oxidation of Stainless Steel in a
High-Temperature Electrolyte—Estimation of Kinetic Parameters of Oxide Layer Growth and Restructuring. J. Electrochem. Soc.
2005, 152, B250–B261. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like