논문자료 9cr corrosion 부식 pitting

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/276135918

The Corrosion Behavior of 9Cr Ferritic–


Martensitic Heat-Resistant Steel in Water and
Chloride Environment

Article in Journal of Engineering Materials and Technology · July 2015


DOI: 10.1115/1.4030430

CITATIONS READS

0 68

3 authors:

Zhen Zhang Preet M. Singh


Nanjing Institute of Technology Georgia Institute of Technology
11 PUBLICATIONS 15 CITATIONS 162 PUBLICATIONS 1,206 CITATIONS

SEE PROFILE SEE PROFILE

Zhengfei Hu
Tongji University
46 PUBLICATIONS 322 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The corrosion behavior of martensitic steel in chloride and supercritical water environment View
project

All content following this page was uploaded by Zhen Zhang on 16 March 2016.

The user has requested enhancement of the downloaded file.


The Corrosion Behavior
of 9Cr Ferritic–Martensitic
Z. Zhang
School of Materials Science and Engineering, Heat-Resistant Steel in Water
Tongji University,
Shanghai 201804, China and Chloride Environment
P. M. Singh The corrosion behavior of 9Cr ferritic–martensitic heat-resistant steel was investigated
School of Materials Science and Engineering,
Georgia Institute of Technology,
in water and chloride environment at room temperature (RT). The results of linear polar-
Atlanta, GA 30332-0245
ization, electrochemical impedance spectroscopy (EIS), and potentiodynamics (PD)
polarization tests on long-term exposure show that 9Cr ferritic–martensitic steel has
weaker corrosion resistance and greater pitting corrosion tendency in higher chloride
Z. F. Hu concentrations. Corresponding scanning electron microscopy (SEM) observation dis-
School of Materials Science and Engineering,
plays that higher concentration chloride promotes the pitting initiation. During long-
Tongji University,
term exposure, pitting susceptibility decreases, the average pit size increases, and the
Shanghai 201804, China
density declines in higher chloride concentrations. Pits in the grains and along the grain
boundaries are observed by optical microscope (OM), and it indicates that inclusions in
grains and carbide particles at grain boundaries are the sites susceptible to pitting
initiation. [DOI: 10.1115/1.4030430]

Keywords: 9Cr ferritic–martensitic steel, polarization, pitting corrosion

Introduction composition specified by standard ASTM and used in the tests is


given in Table 1 and basic mechanical properties are given in
Tempered ferritic–martensitic steels of 9Cr (wt. %) are most
Table 2. The initial microstructures of P92 steels are shown in
popular materials used to fabricate components in power plants
Fig. 1. The normalizing and tempering treatment of steel resulted
for their excellent mechanical properties at elevated temperature
in a tempered martensite. The structure composed of block mar-
[1,2]. Their prominent creep strength with suitable toughness is
tensites, which consist of lath martensites decorated with stringers
achieved by martensitic microstructure coupled with dispersion of
of M23C6 carbides that were Cr enriched at the lath boundaries,
fine precipitates. In advanced 9Cr1Mo steels, P92 is a renewed
prior austenite grain boundaries, and subgrain boundaries. In the
martensitic heat-resistant steel, high creep strength is attained by
intralath regions, fine MC type carbides rich in V and Nb were
the addition of V, Nb, and N which form fine precipitates, and
homogeneously dispersed.
creep strength of these steels has been improved further by replac-
Flat round coupons (with exposed area 1.0 cm2) were machined
ing part of Mo with W and adding B. Furthermore, in order to
for electrochemical tests and long-term exposure tests, respec-
achieve oxidation resistance at high temperature, the minor addi-
tively. All samples were machined along the pipe in longitudinal
tions of silicon and sulphur are required [3,4].
direction. All specimens were dry-abraded up to 2000 grit,
9Cr ferritic–martenistic steel products, such as tube, pipe, and
degreased with ethanol, and dried in forced air. The initial dimen-
components, are generally transported and stored for months or
sions of these samples after polishing were measured for corrosion
even years before processing or putting into operation. During
rate calculations. All tests were started immediately after the pol-
these days, corrosion attack might have a significant effect on the
ishing and cleaning. The composition of the solution used in these
long-term stability of the components at elevated temperature.
tests was shown in Table 3. Each solution was prepared so that the
The high temperature properties of these steels have been widely
solution composition for a given type of tests was the same.
studied [5–7], however the corrosion characteristic of this kind of
steel at RT in certain aqueous environments is not well under-
stood. This work investigated the phenomena of corrosion
Testing Procedures. All tests were performed at RT using a
characteristics in 9Cr ferritic–martensitic steel under special envi-
Gamry PC4 electrochemical measurement system in a sealed
ronment (water, water þ chloride ion, and exposure time) at RT.
three-electrode electrochemical testing cell. The saturated calomel
Conventional electrochemical techniques are used along with EIS,
electrode (SCE) was used as reference electrode. A single plati-
OM, and SEM to understand the corrosion mechanism and iden-
num foil flag was employed as a counter electrode. Corrosion
tify the steel surface morphology after immersion.
potential (Ecorr) measurements were performed at different inter-
vals during the test where Ecorr was measured for a few minutes
period. An average Ecorr value was calculated from each period
Material and Experimental and used in the Ecorr versus time. Linear polarization resistance
(LPR) data are recorded during about 18 days immersion in water
Materials and Testing Environments. The test specimens
and chloride solutions for each specimen. The scan rate was
were machined from a commercial 9Cr ferritic–martensitic steel
0.020 mV/s from starting at 20 mV versus Ecorr and scanning to
P92 pipe sectioned in the longitudinal direction. Its chemical
þ20 mV versus Ecorr. The Ecorr value was measured by the poten-
tiostat just prior to the LPR routine. The EIS of P92 steel was
Contributed by the Materials Division of ASME for publication in the JOURNAL OF
ENGINEERING MATERIALS AND TECHNOLOGY. Manuscript received July 31, 2014; final
traced during 296 hr immersion in chloride solutions of various
manuscript received April 17, 2015; published online May 8, 2015. Assoc. Editor: concentrations. The impedance diagrams were recorded at the free
Georges Cailletaud. corrosion potential (Ecorr) after reading a steady state value by

Journal of Engineering Materials and Technology JULY 2015, Vol. 137 / 031009-1
C 2015 by ASME
Copyright V

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Chemical composition of P92 steel (mass percent, %)

C Mn P S Si Cr W Mo V Nb N B Al Ni

ASTM A335-2003 0.07–0.13 0.3–0.6 0.02 0.01 0.5 8.5–9.5 1.5–2.0 0.3–0.6 0.15–0.25 0.04–0.09 0.03 0.07 0.001–0.006 0.04 0.4
Test 0.1 0.45 0.015 0.008 0.3 8.82 1.57 0.35 0.2 0.078 0.037 0.0027 0.006 0.11

Table 2 Basic mechanical properties of P92 steel

Temperature/  C 0.2% proof strength/ (MPa) Tensile strength/ (MPa) Elongation at fracture/ (%) Reduction of area/ (%)

20 465 660 20 15

Table 3 Composition of the solution used in tests

Solution A Solution B Solution C

Distilled water 1.0% NaCl 3.5% NaCl

Fig. 2 OCP evolution with time under various NaCl


concentrations

concentration, are shown in Fig. 2. Specimen in water exhibits an


increase at the beginning of the OCP evolution. This increase in
the OCP can be explained as the growth of surface film, although
such film may not be fully passive [10,11]. Then, the OCP level
reaches steady state potential at around 30 mV. Under these con-
ditions, some areas on the metal surface generate small and iso-
Fig. 1 Metallurgical structure and schematic diagram of P92
lated pits, while other areas remain passive, which exhibits a
steel-tempered martensite [8,9]
competition between the film growth and breakdown [12]. The
OCPs in chloride solutions shift toward a negative direction in the
applying a sinusoidal potential perturbation signal of 10 mV beginning until approach the steady state potential, which prob-
amplitudes through a frequency domain from 10 kHz down to ably due to the greater pitting corrosion rate than the film growth
10 MHz. The absolute impedance (|z|) of the metal/solution inter- rate. It can also reveal that the steady state potential value
face and the phase shift (h) were measured as a function of the fre- becomes significantly more negative the higher the chloride
quency (f). Long-term exposure tests were carried out in 250 ml amount in solution, which indicates that the presence of chloride
glass beakers for 22 days at RT. PD tests were performed immedi- promotes pitting corrosion on P92 steel.
ately on specimens from each corrosion solution after exposure. Changes in LPR with time for P92 steel, as a function of chlo-
The scan was performed at 0.15 mV/s from 0.2 V of the meas- ride concentration, are shown in Fig. 3. The LPR values in water
ured Ecorr to 0.8 V versus SCE and returned to the starting poten- are clearly higher than that in NaCl solutions, which show the
tial. The corrosion potentials (Ecorr), breakdown potentials (Ebd) weaker corrosion resistant ability of P92 steel in chloride solu-
and repassivation (Erp) were obtained from the potentiodynamitic tions. The first peak in the LPR evolution of water can be
polarization curves. Surface of samples was characterized using explained as the growth and breakdown of surface film. Then the
the OM and SEM. LPR level stays in a narrow range, which confirms the stable pas-
sive layer on the surface of steels. The LPRs of specimens in chlo-
Results and Discussion ride solutions exhibit a similar tendency as that in Fig. 2, and it
can reflect the change of corrosion rate of specimen during expo-
Polarization Measurements. Changes in open circuit potential sure. It can be seen that the corrosion rate increases with increas-
(OCP) with time for P92 steel, as a function of NaCl ing chloride concentrations.

031009-2 / Vol. 137, JULY 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


impedance starts to decrease only after 0.5 hr of immersion, which
reveals that the surfaces are more susceptible to pitting attack.
The decrease of the overall impedance values in 3.5% NaCl solu-
tions also reveals a weakness of the corrosion resistance, in good
agreement with the polarization curves.
The Bode plots of P92 steel immersed in different NaCl con-
centrations are shown in Fig. 5. As it can be seen in Fig. 5(a),
there are two time constant in Bode phase angle plots. According
to some researchers [16,17], the impedance data were analyzed by
using the model where Rs is in series with two RQ elements,
which consists of the resistance R and the constant phase angle
element Q. The parameters calculated for P92 steel are given in
Table 4. The low-frequency part R2Q2 may represent a diffusion
phenomenon [18], and the high-frequency part R1Q1 represent the
capacitive behavior of the passive film formed, coupled with a
resistance due to the ionic paths through the oxide film [19]. The
Fig. 3 LPR evolution with time under water and different NaCl Rp was calculated using Rp ¼ R1 þ R2, where R1 is the resistance
concentrations that due to the ionic paths through the oxide film, and R2 repre-
sents the corresponding charge-transfer resistance. The effects of
chloride to the electrochemical properties of the system, as evi-
Electrochemical Impedance Measurements. The Nyquist denced by changes in the values of Rp in the EIS spectra, are
plots of P92 steel in water and chloride solutions after 296 hr shown in Fig. 6. It can be observed that Rp values in 1.0% NaCl
immersions are shown in Fig. 4. It can be seen that the Nyquist increase significantly with immersion time up to 24 hr and then
plots consist a capacitive loop, with the center under the real im- the value decreases to a more-or-less constant value. Similar
pedance axis, which indicates passive films formed at the surface changes happen on samples in 3.5% NaCl, but the maximum
and is a typical behavior for the solid metal electrodes that show value of Rct occurs at 1 hr. This confirmed the stable passive layer
frequency dispersion of the impedance data [13]. Dispersion has on steels in chloride solutions and can be concluded that the sam-
been attributed to roughness and other inhomogeneities of the ple in 3.5% NaCl has a weaker corrosion resistance than that in
solid surface [14]. As it can be seen in Fig. 4(a), the radius of the 1.0% NaCl, in accord with the polarization curves and the Nyquist
capacitive arc increases with time until it reaches maximum at plots. The decrease of Rs may probably because the cation con-
121 hr, then it decreases with further increase in time. Similar centrations increase with immersion time. The factor n, defined as
changes can be seen in 1.0% and 3.5% NaCl solutions, the maxi- a Q power, for 0.5 < n < 1, it describes the distribution of the
mum radius occurs at 24 hr and 0.5 hr, respectively. In the imped- dielectric relaxation times in the frequency space [16].
ance measurement, the radius of the semicircular arc is related to
the polarization resistance of the passive film. The increase of it
reveals an enhancement of the passive film and its protective Long-Term Exposure Measurements. PD curves of P92
behavior in the test solutions. Otherwise, the decrease means the steels exposed in water and chloride solutions are shown in Fig. 7
protective ability is lost [15]. In 3.5% NaCl solutions, the with the corrosion potentials (Ecorr), breakdown potentials (Ebd),

Fig. 4 The Nyquist plots of P92 steel during 296 hr immersion in water and various NaCl con-
centrations: (a) distilled water, (b) 1.0% NaCl, and (c) 3.5% NaCl

Journal of Engineering Materials and Technology JULY 2015, Vol. 137 / 031009-3

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 6 Rp as a function of time. Rp resistance was calculated
using Rp 5 R1 1 R2, where R1 and R2 were parameters from the
fitting procedure of P92 steel chloride solutions, as indicated.

film repassivation kinetics, which may lead to possible localized


corrosion attack. After 22 days exposure, the specimen exposed in
water had a larger hysteresis loop and the difference between Erp
and Ebd value became greater than the original specimen, which
shows that the pitting resistance became weaker after long-term
exposure in water. However, the specimen exposed in chloride
solutions had a more negative Ebd value than original conditions,
but the Ebd value was closer to Erp value than the original speci-
men. It showed that the specimen had a decreased pitting suscepti-
bility after 22 days exposure in chloride solutions. As can be seen
in Fig. 8, the Ebd value decreases with increase of chloride con-
Fig. 5 The Bode plots of P92 steel immersed in different NaCl tent, which shows the instability of surface film at higher chloride
concentrations: (a) 1.0% NaCl and (b) 3.5% NaCl content. We can also see that the value of Ebd goes downhill very
quickly when the content of chloride is lower than 1.0%, but it
and repassivation (Erp) values provided in Table 5. The specimen falls slowly when the value is higher than that. It can be proved
exposed in 3.5% NaCl solution had a more negative Ebd and Erp that there exists a critical value of chloride content. The P92 steel
value than specimen exposed in water and 1.0% NaCl solution, exhibits good corrosion resistance in environments where the
indicating a greater tendency to pitting and a reduced tendency to content of chloride is lower than this value.
repassivate at conditions of original and after 22 days exposure.
Both Ecorr and Erp decrease with an increase in the chloride con- Surface Morphology Analysis. The macroscopic of surface of
centration. Since the surface reaction should not change with the specimen after 22 days immersion in water and chloride solutions
same water and acidity conditions, such the decrease in the Ecorr are shown in Fig. 9. As we can see in Fig. 9(a), there are almost
and Erp suggests an increase in the corrosion activity. Before 22 no corrosion products on the surface of specimen after 22 days
days exposure, the original specimen exposed in water had a more immersion in water. However, more corrosion products are pro-
negative Ecorr value than Erp, oscillations due to a very low ionic duced especially at higher concentration of chloride. The long-
strength and no distinct current hysteresis was shown by the term exposure tests in chloride solution show that the generated
reverse potential scan, indicating that it did not exhibit clear passi- corrosion products are easy to drop out from the surface of speci-
vation behavior and pitting potential with anodic polarization in men, and it indicates that the corrosion behavior of P92 steel in
water. However, the reverse loop produced a hysteresis loop and chloride solution is weight loss. The amount of weight loss can
Ecorr exhibited higher values than Erp in chloride solution. Clearly, represent the corrosion rate and it shows that chloride has signifi-
chloride destabilized the surface film formation and decreased the cant effects on the corrosion performance of P92 steel.

Table 4 Effect of immersion time on the impedance fitting parameters of P92 steels

Time/hr

0 0.5 1 24 48 126 296

1.0% NaCl Rs/X 20.230 19.170 19.140 17.200 15.610 16.490 10.940
R1/X 108 106 118 123 129 156 143
n1 0.80 0.80 0.80 0.86 0.81 0.75 0.81
R2/X 2859 2953 3768 7670 5315 4711 4397
n2 0.84 0.93 0.87 0.81 0.86 0.79 0.80
3.5% NaCl Rs/X 6.961 6.687 6.736. 5.567 3.590 3.287 3.206
R1/X 15 16 15 25 35 13 13
n1 0.80 0.81 0.82 0.80 0.85 0.84 0.86
R2/X 2613 2785 2830 1804 1637 1582 1446
n2 0.80 0.86 0.82 0.83 0.81 0.81 0.76

031009-4 / Vol. 137, JULY 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 8 The breakdown potentials Ebd as a function of NaCl
content

Fig. 9 The macroscopic profile of surface of specimen after 22


days immersion in water and chloride solutions: (a) distilled
water, (b) 1.0% NaCl solution, and (c) 3.5% NaCl solution

Fig. 7 (a) PD for P92 steel before exposure in water and chlo- the pits in Fig. 10(d) reveals that the corrosion products inside cor-
ride solution and (b) PD after 22 days exposure rosion pits are mostly iron oxide and chromium oxide. It has been
proved that the oxide scales formed on P92 steel after immersion
Table 5 Potentiodynamitic polarization parameters for differ- consist an outer layer Fe-rich magnetite and an inner layer Cr-rich
ent conditions Fe–Cr spinel [7], so it can be inferred that the corrosion products
in the pits contain residues originated from the breakdown of ox-
Conditions Solution Ecorr/V Ebd/V Erp/V ide film. In some pits, tungsten peaks are also present, which are
associated with the tungsten carbide precipitates or inclusions in
Original Distilled water 0.363 0.184 0.001 P92 steel. These inclusions may serve as preferential sites for pit
1.0% NaCl 0.345 0.228 0.597
initiation in these steels.
3.5% NaCl 0.344 0.292 0.620
After 22 days exposure Distilled water 0.305 0.251 0.402 The results of OM examination of specimens exposed to the
1.0% NaCl 0.724 0.682 0.614 chloride solutions after long-term exposure are shown in Fig. 11.
3.5% NaCl 0.763 0.710 0.618 The grains and grain boundaries can be distinguished clearly. It
can be seen in Fig. 11(a) that there are pits, like A, in the grains
and pits, like B, at the grain boundaries. Merged pits can also be
seen in the grain (Fig. 11(b)) in higher chloride concentration,
SEM was used to confirm the electrochemical results. Each which may due to the fast uniform corrosion and some pit growth
specimen was immersed for 22 days in the distilled water and re- coalescence. The inclusions in the grains and the carbide particles
spective chloride solution at RT, then rinsed gently with distilled at the grain boundaries are the sites most susceptible to pitting ini-
water, dried and stored in a desiccator before examination. Pit tiation. These sites exhibited a lower surface potential than the
analysis after tests, shown in Fig. 10, indicates the average pit size matrix, which adversely affected the pitting resistance [21]. Pas-
increases in higher chloride concentrations but the density of the sive film of boundaries between steel matrix and inclusions is
pits declines, which may probably because some corrosion pits weak and the early pitting corrosion may occur at these sites [22].
had coalesced to form bigger pits. No pitting was observed on the The size of inclusions in the grains, like MnS [23] silicate, alumi-
sample immersed in the distilled water (Fig. 10(a)). However, at num oxide, and sulphide type of inclusions [24], is about
Cl concentration higher than a threshold value, the morphology 25–35 lm, and it is easy to be peeled from the matrix during expo-
of the surface was quite different from the samples exposed to the sure. It can also be seen that merged pits (Fig. 11(c)) and scatter
lower concentration of chlorides and the rough surface (the pits (Fig. 11(d)) distribute along grain boundaries, which shows
amount of corrosion products and the number of pits) is virtually the grain boundary is also susceptible to pitting and becomes the
reduced (Fig. 10(c)). This conspicuous variation in the microstruc- weak site during exposure. The content of Cr at the grain bounda-
tures of the surface was mostly due to the fast uniform corrosion ries is not expected to decrease and the chromium depleted region
and some pit growth and coalescence [12,20]. Corrosion products generally does not appear for this type of steel [25]. Thus, the
are deposited inside the pits due to their low solubility in solutions scattered pits along the grain boundaries may be related to the car-
(Fig. 10(d)). Energy dispersive spectroscopy (EDS) analysis of bide particles along grain boundaries, which are produced during

Journal of Engineering Materials and Technology JULY 2015, Vol. 137 / 031009-5

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 10 Scanning electron micrographs of P92 steel after 22 days immersion in water and
chloride solutions: (a) distilled water; (b) 1.0% NaCl solution; and (c) and (d) 3.5% NaCl
solution

the martensite transformation. The size of carbide particles at the can be explained by the bigger size of the inclusions in the grains.
grain/lath boundaries, like M23C6 [26], is about 0.1–0.2 lm, and it The scattered pits join together and become the merged pits, dis-
can also be the pitting initiation site during the exposure. More tributed along the grain boundaries (Fig. 11(d)), with the time of
pits are observed in the grains than that at grain boundaries, which the immersion in chloride solutions. The corroded grain

Fig. 11 Metallographic microstructure of P92 steel after 22 days immersion in chloride solu-
tions: (a) and (c) 1.0% NaCl solution and (b) and (d) 3.5% NaCl solution

031009-6 / Vol. 137, JULY 2015 Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use


boundaries will cause easy transgranular crack initiation and prop- [6] Jin, S. X., Guo, L. P., Li, T. C., Chen, J. H., Yang, Z., Luo, F. F., Tang, R.,
agation when it undergoes stress in the service process. Qiao, Y. X., and Liu, F. H., 2012, “Microstructural Evolution of P92 Ferritic/
Martensitic Steel Under Arþ Ion Irradiation at Elevated Temperature,” Mater.
Charact., 68, pp. 63–70.
[7] Xu, H., Yuan, J., Zhu, Z. L., Zhang, Q., and Zhang, N. Q., 2014, “Oxidation
Summary Behavior of Ferritic-Martensitic Steel P92 Exposed to Supercritical Water at
Based on the results of test above, the main conclusions are 600  C/25 MPa,” J. Chin. Soc. Corros. Prot., 34(2), pp. 119–124.
[8] Hayakawa, M., Hara, T., Matsuoka, S., and Tsuzaki, K., 2000, “Microstructural
drawn as follows: Observation of Tempered Martensite in Medium-Carbon Low-Alloy Steel by
(1) The change in the OCP and LPR results display the corro- Atomic Force Microscopy,” J. Jpn. Inst. Met., 64(6), pp. 460–466.
[9] Hayakawa, M., Matsuoka, S., and Tsuzaki, K., 2001, “Observations of Prior
sion rate of P92 steel increases as the chloride concentra- Austenite Grain Boundaries and Carbides in the Same Area of Tempered Mar-
tion increases in the solution. This kind of steel has weaker tensite in Medium-Carbon Steel by Atomic Force Microscopy,” J. Jpn. Inst.
corrosion resistance and higher pitting corrosion tendency Met., 65(8), pp. 734–741.
in higher chloride concentrations. EIS tests exhibit that the [10] Banas, J., Stypula, B., Banas, K., Swiatowska-Mrowiecka, J., Starowicz, M., and
Lelek-Borkowska, U., 2009, “Corrosion and Passivity of Metals in Methanol Sol-
polarization resistance has a peak value during immersion, utions of Electrolytes,” J. Solid State Electrochem., 13(11), pp. 1669–1679.
which is related to the formation and breakdown of a pas- [11] Cavalcanti, E., Wanderley, V. G., Miranda, T. R. V., and Uller, L., 1987, “The
sive film at the metal–solution interface. PD tests show that Effect of Water, Sulphate and pH on the Corrosion Behavior of Carbon Steel in
there exists a critical value of NaCl concentration, the pas- Ethanolic Solutions,” Electrochim. Acta, 32(6), pp. 935–937.
[12] Lou, X. Y., and Singh, P. M., 2010, “Role of Water, Acetic Acid and Chloride
sive film on the surface of specimen is easy to breakdown on Corrosion and Pitting Behaviour of Carbon Steel in Fuel-Grade Ethanol,”
when the NaCl concentration exceeds the limit shown. Corros. Sci., 52(7), pp. 2303–2315.
(2) The pit morphology shows isolated pits with corrosion [13] Liu, F. G., Du, M., Zhang, J., and Qing, M., 2008, “Inhibition Mechanism of
products inside are produced on the surface of specimen in Imidazoline Derivative Inhibitor for Q235 Steel in Saltwater Saturated With
CO2,” Acta Phys.-Chim. Sin., 24(1), pp. 138–142.
chloride solutions. The inclusions inside grains and carbide [14] Wang, B., Du, M., Zhang, J., and Gao, C. J., 2011, “Electrochemical and Sur-
particles at grain boundaries are the preferred sites suscepti- face Analysis Studies on Corrosion Inhibition of Q235 Steel by Imidazoline De-
ble to pitting initiation. The scattered pits can join together rivative Against CO2 Corrosion,” Corros. Sci., 53(1), pp. 353–361.
and become the merged pits, which will cause easy trans- [15] Luo, H., Dong, C. F., Li, X. G., and Xiao, K., 2012, “The Electrochemical
Behaviour of 2205 Duplex Stainless Steel in Alkaline Solutions With Different
granular crack initiation and propagation when it is undergo pH in the Presence of Chloride,” Electrochim. Acta, 64, pp. 211–220.
stress. [16] Raistrick, I. D., Franceschetti, D. R., and MacDonald, J. R., 2005, “Theory,”
Impedance Spectroscopy: Theory, Experiment, and Application, 2nd ed., E.
From above investigations, it is clear that the 9Cr Barsoukov, and J. R. MacDonald eds., Wiley, Hoboken, NJ, Chap. 2.
ferritic–martensitic steels are sensitive to corrosion in solutions [17] Kosec, T., Merl, D. K., and Milosev, I., 2008, “Impedance and XPS Study of
with chloride at RT. During the storage, transportation, and serv- Benzotriazole Films Formed on Copper, Copper–Zinc Alloys and Zinc in Chlo-
ride Solution,” Corros. Sci., 50(7), pp. 1978–1997.
ice, these kind steels and their products or fabricated components [18] Macdonald, D. D., 1992, “Point Defect Model for the Passive State,” J. Electro-
should be well protected from humidity or salted environment like chem. Soc., 139(12), pp. 3434–3449.
ocean or sea. [19] Kocijan, A., Merl, D. K., and Jenko, M., 2011, “The Corrosion Behavior of
Austenitic and Duplex Stainless Steels in Artificial Saliva With the Addition of
Fluoride,” Corros. Sci., 53(2), pp. 776–783.
Acknowledgment [20] Wang, B., Zhang, L. W., Su, Y., Xiao, Y., and Liu, J., 2013, “Corrosion
Behavior of 5A05 Aluminum Alloy in NaCl Solution,” Acta Metall. Sin., 26(5),
This study was jointly supported by the Shanghai Key Labora- pp. 581–587.
tory for R&D and Application of Metallic Functional Materials in [21] Zheng, S., Li, C., Qi, Y., Chen, L., and Chen, C., 2013, “Mechanism of
Tongji University and The Corrosion and Materials Chemistry (Mg,Al,Ca)-Oxide Inclusion-Induced Pitting Corrosion in 316L Stainless Steel
Exposed to Sulphur Environments Containing Chloride Ion,” Corros. Sci., 67,
Research Laboratory in Georgia Institute of Technology. pp. 20–31.
[22] Zhang, C. Y., Chen, X. Q., Chen, D. B., Li, G. M., and Pan, R. Y., 2001,
“Research of Pitting Susceptibility in Low Carbon Steels and Mechanism of
References Pitting Initiation,” J. Chin. Soc. Corros. Prot., 10(5), pp. 265–272.
[1] Yi, Y., Lee, B., Kim, J., and Jang, J., 2006, “Corrosion and Corrosion Fatigue [23] Chiba, A., Muto, I., Sugawara, Y., and Hara, N., 2013, “Pit Initiation Mecha-
Behaviors of 9Cr Steel in a Supercritical Water Condition,” Mater. Sci. Eng., nism at MnS Inclusions in Stainless Steel: Synergistic Effect of Elemental Sul-
A, 429(1–2), pp. 161–168. fur and Chloride Ions,” J. Electrochem. Soc., 160(10), pp. 511–520.
[2] Toloczko, M. B., Hamilton, M. L., and Maloy, S. A., 2003, “High Temperature [24] Shashank Dutt, B., Nani Babu, M., Shanthi, G., Venugopal, S., Sasikala, G.,
Tensile Testing of Modified 9Cr-1Mo After Irradiation With High Energy Pro- and Bhaduri, A. K., 2012, “Influence of Microstructural Inhomogeneities on the
tons,” J. Nucl. Mater., 318, pp. 200–206. Fracture Toughness of Modified 9Cr-1Mo Steel at 298-823 K,” J. Nucl. Mater.,
[3] Ennis, P. J., and Czyrska-Filemonowicz, A., 2003, “Recent Advances in Creep- 421(1–3), pp. 15–21.
Resistant Steels for Power Plant Applications,” Sadhana, 28(3–4), pp. 709–730. [25] Zhang, B. H., 2005, Electrochemical Corrosion and Protection of Metal,
[4] Vyrostkova, A., Kroupa, A., Janovec, J., and Svoboda, M., 1998, “Carbide Higher Education Press, Beijing, p. 84.
Reactions and Phase Equilibria in Low Alloy Cr–Mo–V Steels Tempered at [26] Sikka, V. K., Ward, C. T., and Thomas, K. C., 1983, “Modified 9Cr-1Mo
773–993 K—Part I: Experimental Measurements,” Acta Mater., 46(1), pp. Steel—An Improved Steel for Steam Generator Application,” Ferritic Steels for
31–38. High Temperature Applications: Proceedings of ASM International Conference
[5] Kim, B.-J., Kim, H.-J., and Lim, B.-S., 2008, “Creep-Fatigue Damage and Life on Production, Fabrication, Properties and Application of Ferritic Steels for
Prediction in P92 Alloy by Focused Ultrasound Measurements,” Met. Mater. High Temperature Applications, Warren, PA, 6–8 October 1981, A. K. Khare,
Int., 14(4), pp. 391–395. ed., ASM, Metals Park, OH, pp. 65–84.

Journal of Engineering Materials and Technology JULY 2015, Vol. 137 / 031009-7

DownloadedViewFrom:
publicationhttp://materialstechnology.asmedigitalcollection.asme.org/
stats on 12/06/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like