Sei sulla pagina 1di 16

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

net/publication/328672891

Corrosion Characteristics of ASTM A106 Grade B Carbon Steel Pipelines


Exposed to Sodium Sulfate Solutions

Article · November 2018


DOI: 10.1520/MPC20180026

CITATIONS READS

2 1,260

4 authors, including:

Randa Abdel-Karim Michael Nabil


Cairo University Modern Sciences and Arts University
45 PUBLICATIONS   368 CITATIONS    18 PUBLICATIONS   73 CITATIONS   

SEE PROFILE SEE PROFILE

Y. Reda
Lecturer at higher institute of enginnering and technology tanta. And research at …
20 PUBLICATIONS   114 CITATIONS   

SEE PROFILE

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

Improvement of mechnical properties of7075 View project

Corrosion and mechanical properties of Al-Si/Al2O3 composites View project

All content following this page was uploaded by Y. Reda on 03 November 2018.

The user has requested enhancement of the downloaded file.


Materials Performance and Characterization

doi:10.1520/MPC20180026 / Vol. 7 / No. 1 / 2018 / available online at www.astm.org

R. Abdel-Karim,1 M. Nabil,2 Y. Reda,3 and S. El-Raghy2

Corrosion Characteristics of ASTM A106


Grade B Carbon Steel Pipelines Exposed
to Sodium Sulfate Solutions

Reference
Abdel-Karim, R., Nabil, M., Reda, Y., and El-Raghy, S., “Corrosion Characteristics of ASTM A106
Grade B Carbon Steel Pipelines Exposed to Sodium Sulfate Solutions,” Materials Performance and
Characterization, Vol. 7, No. 1, 2018, pp. 480–494, https://doi.org/10.1520/MPC20180026.
ISSN 2379-1365

ABSTRACT
Manuscript received February 6, Carbon steel is used in large-tonnage petroleum production and refining. In this
2018; accepted for publication
work, the corrosion behavior of ASTM A106 Grade B carbon steel subjected to
September 17, 2018; published
online November 2, 2018. sodium sulfate (Na2SO4) environments was studied in a range of sulfate
1
concentrations and different working temperatures. Increasing the sulfate ion
Department of Metallurgy,
Corrosion and Surface Treatment content had an inhibiting effect on the corrosion behavior of carbon steel. The
Lab, Cairo University, Gamaa St., highest corrosion rate was detected for electrolytes containing 0.5 M Na2SO4. The
Giza 12613, Egypt (Corresponding
author), e-mail: randaabdelkarim@
influence of temperature was significant on the oxide layer density and stability,
gmail.com, https://orcid.org/ while there was the appearance of thick cracked layer as the temperature and the
0000-0001-9608-3426
sulfate concentration increased.
2
Department of Metallurgy,
Corrosion and Surface Treatment
Lab, Cairo University, Gamaa St., Keywords
Giza 12613, Egypt carbon steel, polarization, scanning electron microscopy, weight loss, anodic dissolution, sodium
3
Chemical Engineering
sulfates
Department, Higher Institute of
Engineering and Technology, New
Damiattee St., 4th Zone, District Introduction
Area, New Damiatte 34517, Egypt
Pipelines form an integral part of the infrastructure of oil and gas production. The trans-
portation of crude oil, natural gas, and refined petroleum products is made possible
through pipelines, which are the preferred choice of transportion where large volumes
and long distances are involved [1]. Buried microalloyed pipeline steels’ integrity for
oil and gas transport depends greatly on soil environment interactions [2,3]. The corrosion
of steel in soil is a complex phenomenon with a multitude of variables involved. Some

Copyright © 2018 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959 480
ABDEL-KARIM ET AL. ON CORROSION 481

variables that affect the corrosion rate of steel in soil are ground water contamination,
degree of aeration, pH, redox potential, resistivity, soluble ionic species, and microbiologi-
cal activity [4]. Salts dissolved in water have a marked influence on the corrosivity of water.
The anions most commonly found in water are chloride, sulphate, and bicarbonate [5].
The presence of sulfate ions in oilfield-produced water strongly influence corrosion
mechanisms [6]. These anions promote the following: corrosion and scaling in pipes,
structures and equipment, fouling and deposition in boilers, and acidification of soils
and blockage of soil pores, thereby retarding irrigation or water drainage [7].
In addition, sulfate ions increase the conductivity of the moisture film and reduce the
critical level of relative humidity where the film will begin to form [8]. In general, sulfate
ions decrease the susceptibility to pitting corrosion, and they may inhibit localized pitting
corrosion by developing a sulfate layer [9].
There are many investigations on the corrosion of carbon steel in neutral aerated salt
solutions, especially in sodium chloride (NaCl) solution. Few investigations have been
conducted on the corrosion of carbon steel in Na2SO4 salt solution [10–17].
Generally, there are different cases where the corrosion that is caused by sodium sul-
fate (Na2SO4) can occur. The corrosion of the turbine caused by a thin film deposit of
fused salt (Na2SO4) on the alloy surface (carbon steel) is an example of corrosion in
Na2SO4 [13–15]. During combustion in the gas turbine, sulfur from the fuel reacts with
NaCl from ingested air at elevated temperatures, leading to the formation of Na2SO4,
which then deposits on the hot-sections, such as nozzle guide vanes and rotary blades,
resulting in accelerated oxidation attack [16]. Na2SO4 also causes corrosion in the boiler,
in which sodium sulfite (Na2SO3) is utilized as an oxygen scavenger. When Na2SO3 is
reacted with oxygen at a low temperature and pressure, Na2SO4 is created, which causes
a severe attack on the tubes of the boiler [17].
This can be found in cooling water systems and in a lot of important industries, such
as oil refineries and petrochemical and energy plants; many factors, such as high temper-
ature, fluid turbulence, and quality of water, will have an impact on the rate of the cor-
rosion process. However, most cooling water systems also have high concentrations of
SO42− ions, which restrain the adsorption of chlorine ions on the passive film and increase
the pitting potential, which impedes the pitting corrosion [18,19].
In addition, the presence of Na2SO4 in a considerable concentration in crude oil
causes corrosion for the handing pipelines [4]. The aim of this work is to study the in-
fluence of temperature, time, and corrosion products layer on the corrosion rate and cor-
rosion potential of commercial steel pipe immersed in solutions with different
concentrations of Na2SO4. Additionally, the morphology and composition of the corrosion
product deposits are determined using scanning electron microscopy (SEM) and energy
dispersive spectroscopy.

Material and Methods


Experiments were conducted using ASTM A106, Standard Specification for Seamless
Carbon Steel Pipe for High-Temperature Service, Grade B carbon steel pipe that had an
outer diameter of 297 mm and a thickness of 9.6 mm; the pipe was provided by
Cairo Oil Refining Company in Cairo, Egypt. Test specimens in the form of rectangular
coupons with dimensions of 20 by 20 mm and a thickness of 9 mm were cut from this pipe.
The chemical composition and mechanical characteristics of this steel are illustrated in

Materials Performance and Characterization


482 ABDEL-KARIM ET AL. ON CORROSION

TABLE 1
Chemical composition of ASTM A106 Grade B carbon steel.

Element C Si P S Mn Ni Cr Cu Fe

Wt. % 0.3 0.5 0.003 0.003 0.5 0.2 0.2 0.2 Rem.

Tables 1 and 2. The measured hardness of the material was 276 HV, and the microstruc-
ture was composed of pearlite and ferrite. For corrosion measurements, the test coupons
were subjected to grinding using silicon carbide paper up to grid 1,200 and abrading using
alumina paste to obtain a mirror-like image. To avoid oxidation, each sample was rinsed
first with tap water and secondly with distilled water, and afterward the sample was de-
greased in acetone and kept in a dessicator until use.
The hardness of specimens was measured using a hardness tester machine (Zwick/
Roell ZHU250, Zwick Roell Group, Ulm, Germany). An average of three readings were
measured in Vickers scale. Meanwhile, the mechanical properties were determined by
using a universal tensile testing machine (SHIMADZU/Type H-1000kNX, Shimadzu,
Kyoto, Japan).
The chemical analyses of the testing electolyte [20–22] and the working parameters
are illustrated in Table 3. The corrosion behavior of carbon steel was studied in a solution
containing 0-, 0.2-, 0.5-, and 1-M Na2SO4 at different temperatures, namely 25°C, 40°C,
and 60°C. The pH was in the range of 8–9.
The corrosion performance of carbon steel coupons was evaluated by measuring the
weight loss as well as the open circuit potential (OCP) at time intervals of seven days for
three months and at room temperature. Then, specimens were rinsed with distilled water
and carefully dried before taking weight change measurements using a 0.0001-g accuracy
balance. The corresponding corrosion rates in mm/y, based on weight loss measurements,
were calculated using Eq 1 [23]:

ðK  WÞ
CR = (1)
ðA T  DÞ

where:
CR = corrosion rate;
W = the weight loss, gm;
D = the density of steel, gm/cm3;
A = the total surface area, cm2;
t = the total exposure time, hours; and
K = 8.76 * 104 is constant.
The potentiodynmic polarization tests were conducted using a standard three-
electrodes cell. Each specimen was used as a working electrode with an exposure area
of 1 cm2 in contact with the test solution. A saturated calomel electrode was used as a
reference electrode, and the counter electrode was a platinum electrode. The scanning
range was set from −500 mV to +500 mV with respect to OCP using a scanning rate
of 0.333 mV/sec. Computer software (Volta Master 4, Radiometer Analytical, Lyon,
France) was used to draw polarization curves and Tafel extrapolation.
Surface morphology and corrosion product film formation were examined by an op-
tical microscope (Olympus BX41M-LED, Olympus Corporation, Tokyo, Japan). For

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 483

TABLE 2
Tensile properties of ASTM A106 Grade B carbon steel sample.

Width, mm Thickness, mm Gauge Length, mm Yield Strength, N/mm2 UTS, N/mm2 Elongation %

13.3 8.8 71 445.03 500.40 23.94

TABLE 3
Chemical composition of testing electrolyte as well as working conditions.

Test Parameter Value

Chemical composition of testing electrolyte, g/L [21–23]


MgSO4 0.131
CaCl2 0.093
KCl 0.122
NaHCO3 0.483
Concentration of Na2SO4, M 0, 0.2, 0.5, 1
Temperature, °C 25, 40, 60
pH range 8–9

higher magnification, an SEM (Model Quanta 250 FEG, Thermo Fisher Scientific,
Waltham, MA) was used to demonstrate the microstructure and surface morphology
of the corrosion products. Energy dispersive X-ray analyses (EDAXs) were used to deter-
mine the chemical composition of corrosion products by a unit attached to the SEM.

Results and Discussion


OPEN CIRCUIT POTENTIAL
Fig. 1 summarize the variation of the OCP of ASTM A106 Grade B carbon steel as a
function of time while being immersed in solutions with different concentrations of
Na2SO4 at 25°C, namely 0-, 0.2-, 0.5-, and 1-M. After immersion for one week, the
OCP values were in the range of approximately −700 to −1,000 mV. The most negative

FIG. 1
OCP measurements after 1,100
immersion for three months 0.0 M 0.2 M 0.5 M 1M
at 25°C. 1,000

900
OCP, –mV

800

700

600

500

400
0 2 4 6 8 10 12 14
Time, weeks

Materials Performance and Characterization


484 ABDEL-KARIM ET AL. ON CORROSION

FIG. 2
Corrosion rates of test coupons 140
during immersion for three 0.0 M 0.2 M

Corrosion rate *10–3 mm/y


months as a function of Na2SO4 120
concentration based on weight
loss measurements. 100

80

60

40

20
0 2 4 6 8 10 12 14
Time, weeks

OCP was in the case of electrolyte containing 1-M Na2SO4. After immersion for three
months, the OCP values were in the range of −760 to −1,000 mV. The most negative
OCP was in the case of electrolytes containing 0.2 and 0.5-M Na2SO4.

WEIGHT LOSS MEASUREMENTS


Based on weight loss measurements, Fig. 2 shows corrosion rates obtained after
immersion in sulfates’ electrolytes as a function of time. After immersion for one week,
the corrosion rate decreased with an increase in the sulfate content. The carbon steel
samples immersed in 0.0-M Na2SO4 solution showed the highest corrosion rate
(108.93 × 10−3 mm/Y). Meanwhile, samples immersed in 1-M Na2SO4 solution showed
the lowest corrosion rate (36.54 × 10−3 mm/Y). The measured corrosion rates reached a
steady state after seven weeks. After twelve weeks of immersion, the corrosion process was
accelerated by increasing the sulphate ion concentration up to 0.5-M Na2SO4. The cor-
rosion rate was reduced with further increase in sulphate ion concentration. The registered
corrosion rate values after immersion for three month were 45.46, 77.97 × 10−3 and
41.31 × 10−3 mm/y for 0.0-, 0.5-, and 1-M Na2SO4 solutions, respectively.
According to the research conducted by Ikechukwu, Obioma, and Ugochukwu [24],
the corrosiveness of carbon steel, using the weight loss technique, in 1.0-M Na2CO3 was
the highest, which was mainly a function of its higher concentration, followed by the 0.5-M
solution of NaCl; the least was recorded in the 1.0-M solution of Na2SO4. This is due to the
buildup of more protective film.

POLARIZATION MEASUREMENTS
Fig. 3 shows the polarization curves obtained for ASTM A106 Grade B carbon steel ex-
amined in solutions with different concentrations of Na2SO4 and at different working tem-
peratures. Corrosion data obtained from these curves are illustrated in Table 4. Fig. 4
summarizes the corrosion rate values obtained from Tafel’s extrapolation as a function
of sulfate concentration as well as service temperature.
It is clear that the presence of sulphate ions causes a marked increase in corrosion
rates. It is noteworthy that the concentration of 0.5-M Na2SO4 is a critical concentration
that shows the highest corrosion rate; additionally, sulfate ions show an inhibition effect

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 485

FIG. 3
Polarization curves of A106
Grade B carbon steel as a
function of Na2SO4
concentration at different
temperatures: (a) 25°C,
(b) 40°C, and (c) 60°C.

Potential, V
(a)

Potential, V
(b)

Potential, V
(c)

Materials Performance and Characterization


486 ABDEL-KARIM ET AL. ON CORROSION

TABLE 4
Electrochemical parameters obtained from potentiodynamic polarization measurements for ASTM A106 Grade B carbon steel.

Concentration of Corrosion Rate


Na2SO4, M Temperature, °C pH βa, mV βc, mV E(i = 0), mV i Corrosion, μA/cm2 *10−3 mm/Y

0 ≈ 25 8.628 40.2 86.4 −694.2 5.67 44.22


0.2 9.097 39.4 63.1 −718.7 4.42 34.43
0.5 9.237 57.4 219.5 −731.5 6.70 77.87
1 9.244 46.3 69.4 −753.9 6.29 49.06
0 40 8.648 36.7 31 −697.4 3.91 30.59
0.2 9.021 49.9 74.4 −693.4 13.44 105.1
0.5 9.112 57.8 84.9 −744.2 31.45 245.2
1 9.15 63.8 90.1 −762.7 13.76 160.9
0 60 8.04 68.0 94.4 −717.9 18.87 220.7
0.2 8.99 57.2 110.0 −715.4 73.50 859.7
0.5 9.795 81.1 252.7 −774.6 78.00 912.0
1 8.91 57.3 89.9 −719.5 61.39 718.1

FIG. 4
Corrosion rate measurements 1,000
25°C 40°C 60°C
obtained from a
900
potentiodynamic polarizaion
Corrosion rate *10–3, mm/y

test of A106 Grade B carbon 800


steel as a function of Na2SO4 700
concentration as well as service
600
temperature.
500
400
300
200
100
0
0 0.2 0.4 0.6 0.8 1 1.2
Concentration of Na2SO4, (M/ L)

and the corrosion rate decreases with a further increasing of the concentration of sulphate
ions. The registered corrosion rates were 44.22, 77.87, and 49.06 mm/y for 0.0-, 0.5-, and
1-M Na2SO4 solutions, respectively. This behavior can be attributed to the increased
electro-conductivity because of the rise in salt content until the salt concentration is great
enough to cause an appreciable decrease in the oxygen solubility, resulting in a decrease in
the rate of corrosion with a further increase in the Na2SO4 concentration [10,12,25].
As indicated in Table 4, a negative shift in corrosion potential from −694.2 mV down
to −753.9 mV occurs by increasing the sodium sulfate concentration up to 1 M. This
tendency (shift) toward more negative potential values and the fact that the corrosion rate
increases when the Na2SO4 concentration is higher can be interpreted for a corrosion
process controlled by activation, which means a charge-transfer mechanism is controlling
the rate of anodic reaction [7].

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 487

The mechanism of corrosion in sulfate solutions can be explained based on some


analysis from Arzola, Palomar-Pardavé, and Genesca [7].Thermodynamically, mild steel
corrosion in sulfate media proceeds by means of two different reactions depending on the
sulfate concentration:
if log [SO4 2−] < −2.25, the thermodynamically favored corrosion process is men-
tioned in Eq 2:

FeðSÞ ↔ Fe2+ ðaqÞ + 2e (2)

whereas if log [SO42−] > −2.25, mild steel corrosion in Na2SO4 aqueous media will
proceed according to Eq 3:

SO2 2− ðaqÞ + FeðSÞ ↔ FeSO4ðaqÞ + 2e (3)

Sulfate can act as a depolarization agent of the reaction (Eq 3), thus increasing the
corrosion rate when [SO4 2−] > 0.1 % [7].
As stated by Gui and Devine [26], the effect of SO42− on the rate of oxidation of iron
in mildly alkaline solutions results from its ability to form soluble complexes with either
Fe2+ or Fe3+.

FIG. 5 Optical microscope images of test specimens after polarization tests in different test solutions at 25°C: (a) 0.0 M, (b) 0.2 M,
(c) 0.5 M, and (d) 1 M.

Materials Performance and Characterization


488 ABDEL-KARIM ET AL. ON CORROSION

Florianovich, Sokolova, and Kolotyrkin [27] illustrated that sulfate ions accelerate the
active dissolution of iron. This means that iron will dissolve in sulfate solutions at a high
rate, leading to the formation of supersaturated solution and allowing precipitation of
films. These films were found to be nonprotective [9]. Brasher [28] stated that all anions
are corrosive when present in dilute solutions; however, they become inhibitive at suffi-
ciently high concentrations.
The relation between temperature and corrosion rate is also illustrated in Fig. 3. As
the temperature increased, the corrosion rate increased. For example, for 0.5-M Na2SO4
solution, the corrosion rates were 77.87, 245.2, and 912.0 mm/y for temperatures of 25°C,

FIG. 6
SEM images of test specimens
after potentiodynamic
polarization tests conducted in
0.5-M Na2SO4 electrolyte at
different service temperatures:
(a) 25°C, (b) 40°C, and (c) 60°C.

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 489

FIG. 6 Continued

40°C, and 60°C, respectively. This could be due to the increase in oxygen solubility with
increasing temperature, which causes the highest increase in the corrosion rate [12].

SURFACE MORPHOLOGY
Fig. 5 shows optical microscope images of the test specimens’ surface after potentiody-
namic testing at 25°C. Carbon steel samples subjected to 0.0-M Na2SO4 were covered by a
dense protective layer Fig. 5a.
Fig. 5b shows an optical microscopy of a coupon that was tested in a 0.2-M Na2SO4
solution, illustrating a spongy partially protective layer. Samples subjected to 0.5-M Na2SO4

Materials Performance and Characterization


490 ABDEL-KARIM ET AL. ON CORROSION

FIG. 6 Continued

were characterized by the formation of a spongy nonprotective layer, see Fig. 5c. Fig. 5d
shows an optical microscopy of a coupon that was immersed in 1-M Na2SO4 solution, and
the image shows a dense protective layer with few porous nonprotective islands.
An electrochemical system such as Fe/Na2SO4 can be considered as a simple corro-
sion process in which corrosion takes place uniformly. Vatankhah et al. [18] studied the
effect of sulfate ions on the electrochemical behavior of iron electrodes. They pointed out
that iron experiences uniform corrosion in solutions containing sulfate ions only and that
sulfate ions are aggressive.

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 491

FIG. 7
EDAX analysis of the corrosion
product layers after a
potentiodynamic test
conducted in 0.5-M Na2SO4
solution at (a) 25°C, (b) 40°C,
and (c) 60°C.

Materials Performance and Characterization


492 ABDEL-KARIM ET AL. ON CORROSION

TABLE 5
EDAX analysis showing the effect of temperature on the chemical analysis of the corrosion product
layer formed on carbon steel.

Element, wt.% 25°C 40°C 60°C

Fe 66.36 67.46 60.55


Na 1.79 4.81 5.08
S 0.84 0.86 0.32
Cl 0.46 0.69 0.38
Mn 0.78 0.78 0.65
O 25.59 17.88 27.35
Others 4.18 7.52 5.67

The mechanism of corrosion attack in sulphate solutions occurring on the surface of


the mild steel after a certain exposure can be regarded as partially uniform. This attack
started in isolated locations and spread laterally without developing any pits. The effect of
SO4−2 on the rate of oxidation of iron in mildly alkaline solutions results from its ability to
form soluble complexes with Fe+2 or Fe+3 ions. Furthermore, sulphate ions accelerate the
active dissolution of iron and the films formed in sulfate solutions resulting from super-
saturated solutions of iron salts, most probably iron hydroxysulfate. Accordingly, iron
dissolves in sulphate solutions at a high rate, creating a supersaturated solution and
allowing precipitation of a nonprotective film [12,29].
Fig. 6 illustrates the SEM images for a carbon steel sample subjected to 0.5-M Na2SO4
solution as a function of service tempaerature. The degree of attack was more severe with a
lot of crack formation on the substrate as the temperature increased.
It can be concluded that the morphology of the corrosion attack after exposure to
Na2SO4 solution can be regarded as partially uniform where there was visible film for-
mation. According to Yang et al. [30], the corrosion pattern of carbon steel in the solution
containing only sulfate ions is close to uniform corrosion with just a few tiny pits. The
corrosion rate in the solution containing 3 % NaCl and 0.5 % NaSO4 is much smaller than
that in the solution containing only 3 % NaCl.
The temperature influence was significant on the oxide layer density and stability with
the appearance of spongy porous layer as the sulfate concentration increased.
EDAX analysis of the corrosion product layers after a potentiodynamic test con-
ducted in a 0.5-M Na2SO4 solution at different temperatures of 25°C, 40°C, and 60°C
are presented in Fig. 7 and Table 5. As the temperature increases, the weight percentage
of iron, sulfer, and carbon decreases and the percent of oxygen increases, which indicates
an increase of oxide film and an increase of sulfate attack matching the increase of the
corrosion rate at 0.5-M Na2SO4 at all temperatures [15].

Conclusions
(1) According to immersion and potentiodynamic test results, the corrosion process
was accelerated by increasing the sulphate ion concentration up to 0.50-M Na2SO4.
Further increments in the sulfate ion concentration had an inhibiting effect on the
corrosion process.
(2) Increasing the content of sulfates leads to a shift in the open OCP toward more
negative values. The most negative OCP was in the case of electolytes containing
0.2- and 0.5-M Na2SO4.

Materials Performance and Characterization


ABDEL-KARIM ET AL. ON CORROSION 493

(3) From the potentiodynamic test, the corrosion rates were drasically increased with
increases in the working temperature.
(4) The morphology of the corrosion attack after three months of exposure can be
regarded as partially uniform.
(5) Temperature influence was significant on oxide layer density and stability, with the
appearance of spongy cracked layer as the sulfate concentration increased to 0.5 M
and the temperature increased to 60°C.

References
[1] Ikpi, M. E. and Okonkwo, B. O., “Electrochemical Investigation on the Corrosion of
API 5L X52 Carbon Steel in Simulated Soil Solutions,” J. Mater. Environ. Sci., Vol. 8,
No. 11, 2017, pp. 3809–3816.
[2] Bansode, V. M., Vagge, S. T., and Kolekar, A. B., “Relationship between Soil
Properties and Corrosion of Steel Pipe in Alkaline Soils,” Int. J. Res. Sci. Innov.,
Vol. 2, No. 11, 2015, pp. 57–61.
[3] Tanthapanichakoon, W., Veawab, A., and McGarvey, B., “Electrochemical
Investigation on the Effect of Heat-Stable Salts on Corrosion in CO2 Capture
Plants Using Aqueous Solution of MEA,” Ind. Eng. Chem. Res., Vol. 45, No. 8,
2006, pp. 2586–2593, https://doi.org/10.1021/ie050575a
[4] Duan, D., Choi, Y. S., Jiang, S., and Nesic, S., “Corrosion Mechanism of Carbon Steel
in MDEA-Based CO2 Capture Plants,” Corrosion, Vol. 37, 2013, pp. 890–901.
[5] Strazisar, B. R., Anderson, R. R., and White, C. M., “Degradation Pathways for
Monoethanolamine in a CO2 Capture Facility,” Energy Fuels, Vol. 17, No. 4,
2003, pp. 1034–1039, https://doi.org/10.1021/ef020272i
[6] Freitas, V., Lins, C., Leonard, M., Ferreira, M., and Saliba, P. A., “Corrosion
Resistance of API X52 Carbon Steel in Soil Environment,” J. Mater. Res. Technol.,
Vol. 1, 2012, pp. 161–166, https://doi.org/10.1016/S2238-7854(12)70028-5
[7] Arzola, S., Palomar-Pardavé, M. E., and Genesca, J., “Effect of Resistivity on the
Corrosion Mechanism of Mild Steel in Sodium Sulfate Solutions,” J. Appl.
Electrochem., Vol. 33, No. 12, 2003, pp. 1233–1237, https://doi.org/10.1023/B:
JACH.0000003855.95788.12
[8] Florianovitch, G. M., Sokolova, L. A., and Kolotyrkin, Y. M., “On the Mechanism of
the Anodic Dissolution of Iron in Acid Solutions,” Electrochimica Acta, Vol. 12,
No. 7, 1967, pp. 879–887, https://doi.org/10.1016/0013-4686(67)80124-5
[9] Szklarska-Smialowska, S. and Mrowczynski, G., “Ellipsometric Study of Film
Formation on Iron in Sodium Sulphate Solutions in the pH Range 6–12,” Br. Corr.
J., Vol. 10, No. 4, 2013, pp. 187–191, https://doi.org/10.1179/000705975798320486
[10] Emori, W., Jiang, S. L., Duan, D. L., Ekerenam, O. O., Zheng, Y. G., Okafor, P. C., and
Qiao, Y. X., “Corrosion Behavior of Carbon Steel in Amine-Based CO2 Capture
System: Effect of Sodium Sulfate and Sodium Sulfite Contaminants,” Mater.
Corr., Vol. 68, No. 6, 2017, pp. 674–682, https://doi.org/10.1002/maco.201609245
[11] Abd El Kader, J. M., El Warraky, A. A., and Abd El Aziz, A. M., “Corrosion Inhibition
of Mild Steel by Sodium Tungstate in Neutral Solution Part 2: Behaviour in NaCl and
Na2SO4,” Br. Corr. J., Vol. 33, No. 2, 1998, pp. 145–151, https://doi.org/10.1179/bcj.
1998.33.2.145
[12] Hasan, B. O. and Sadek, S. A., “Corrosion Behavior of Carbon Steel in Oxygenated
Sodium Sulphate Solution under Different Operating Conditions,” Adv. Chem. Eng.
Res., Vol. 2, No. 3, 2013, pp. 61–71.
[13] Bornstein, N. S. and DeCrescente, M. A., “The Role of Sodium in the Accelerated
Oxidation Phenomenon Termed Sulfidation,” Metall. Trans., Vol. 2, No. 10, 1971,
pp. 2875–2883, https://doi.org/10.1007/BF02813266
[14] Goebel, J. A., Pettit, F. S., and Goward, G. W., “Mechanisms for the Hot Corrosion of
Nickel-Base Alloys,” Metall. Trans., Vol. 4, No. 1, 1973, pp. 261–269, https://doi.org/
10.1007/BF02649626

Materials Performance and Characterization


494 ABDEL-KARIM ET AL. ON CORROSION

[15] Tang, Y., Liu, L., Fan, L., Li, Y., and Wang, F., “The Corrosion Behavior of Pure Iron
under Solid Na2SO4 Deposit in Wet Oxygen Flow at 500°C,” Materials, Vol. 7, No. 9,
2014, pp. 6144–6157, https://doi.org/10.3390/ma7096144
[16] Stringer, J., “Hot Corrosion in Gas Turbines,” High Temperature Corrosion and
Materials Application, ASM International, Materials Park, OH, 2007, pp. 249–258.
[17] Huijbregts, W., “Acid Corrosion Resistance of Boiler Steels,” Mater. Perform., Vol. 16,
2007, pp. 23–27.
[18] Vatankhah, G., Drogowska, M., Menard, H., and Brossard, L., “Electrodissolution of
Iron in Sodium Sulfate and Sodium Bicarbonate Solutions at pH 8,” J. Appl.
Electrochem., Vol. 28, 1998, pp. 173–183, https://doi.org/10.1023/A:1003230725414
[19] Elsayed, A. A. and ElShazly, Y. M. S., “Corrosion Inhibition of Carbon Steel in
Chloride and Sulfate Solutions,” Int. J. Eng. Res. Appl., Vol. 6, No. 2, 2016, pp. 1–6.
[20] Torres-Islas, A., Serna, S., Campillo, B., Colin, J., and Molina, A., “Hydrogen
Embrittlement Behavior on Microalloyed Pipeline Steel in NS-4 Solution,” Int. J.
Electrochem. Sci., Vol. 8, 2013, pp. 7608–7624.
[21] Abdel-Karim, R., Farag, M. A., Ahmed, H. A.-A., and El-Raghy, S., “Corrosion
Resistance of API5L X52 Carbon Steel in Sulfide Polluted Environments,” Mater.
Sci. Appl., Vol. 7, No. 1, 2015, pp. 39–50, https://doi.org/10.4236/msa.2016.71005
[22] Pumtree, A. and Lambert, S. B., “Stress Corrosion Crack Growth of Pipeline Steels in
NS4 Solution,” ASME Proc. Corr. Integr., Vol. 1, 1996, pp. 565–571, https://doi.org/
10.1115/IPC1996-1861
[23] Pound, B. G., Wright, G. A., and Sharp, R. M., “The Anodic Behavior of Iron in
Hydrogen Sulfide Solutions,” Corrosion, Vol. 45, No. 5, 1989, pp. 386–392,
https://doi.org/10.5006/1.3582033
[24] Ikechukwu, A. S., Obioma, A. S. E., and Ugochukwu, N. H., “Studies on Corrosion
Characteristics of Carbon Steel Exposed to Na2CO3, Na2SO4, and NaCl Solutions of
Different Concentrations,” Int. J. Eng. Sci., Vol. 3, No. 10, 2014, pp. 48–60.
[25] Hasan, B. O. and Sadek, S. A., “The Effect of Temperature and Hydrodynamics on
Carbon Steel Corrosion and Its Inhibition in Oxygenated Acid–Salt Solution,”
J. Ind. Eng. Chem., Vol. 20, No. 1, 2014, pp. 297–307, https://doi.org/10.1016/j.jiec.
2013.03.034
[26] Gui, J. and Devine, T. M., “A SERS Investigation Of The Passive Films Formed On
Iron In Mildly Alkaline Solutions Of Carbonate/Bicarbonate and Nitrate,” Corr. Sci.,
Vol. 37, No. 8, 1995, pp. 1177–1189, https://doi.org/10.1016/0010-938X(94)00179-A
[27] Florianovitch, G. M., Sokolova, L. A., and Kolotyrkin, Y. M., “On the Mechanism of
the Anodic Dissolution of Iron in Acid Solutions,” Electrochimica Acta, Vol. 12,
No. 7, 1967, pp. 879–887, https://doi.org/10.1016/0013-4686(67)80124-5
[28] Brasher, D. M., “Role of the Anion in Relation to Metallic Corrosion and Inhibition,”
Nature, Vol. 193, 1962, pp. 868–869, https://doi.org/10.1038/193868a0
[29] Arzola-Peralta, S., Llongueras, J. G., Palomar-Pardavé, M., and Romero-Romo, M.,
“Study of the Electrochemical Behaviour of a Carbon Steel Electrode in Sodium
Sulfate Aqueous Solutions Using Electrochemical Impedance Spectroscopy,”
J. Solid State Electrochem., Vol. 7, No. 5, 2003, pp. 283–288, https://doi.org/10.
1007/s10008-002-0344-x
[30] Yang, L., Xu, Y., Zhu, Y., Liu, L., Wang, X., and Huang Y., “Evaluation of Interaction
Effect of Sulfate and Chloride Ions on Reinforcements in Simulated Marine
Environment Using Electrochemical Methods,” Int. J. Electrochem. Sci., Vol. 11,
2016, pp. 6943–6958, https://doi.org/10.20964/2016.08.51

Copyright by ASTM International (all rights reserved), pursuant to License Agreement. No further reproduction authorized.

View publication stats

Potrebbero piacerti anche