Você está na página 1de 11

Corrosion Science 58 (2012) 181–191

Contents lists available at SciVerse ScienceDirect

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

Electrochemical evaluation of the corrosion behaviour of API-X100 pipeline steel


in aerated bicarbonate solutions
Faysal Fayez Eliyan a,⇑, El-Sadig Mahdi b, Akram Alfantazi a
a
Corrosion Group, Department of Materials Engineering, The University of British Columbia, Vancouver, BC, Canada V6T 1Z4
b
Mechanical Engineering Department, College of Engineering, Qatar University, P.O. Box 2713, Qatar, Doha, Qatar

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

Article history: This work is devoted to evaluating electrochemical corrosion aspects of API-X100 in bicarbonate solu-
Received 7 January 2011 tions. They were prepared to be aerated; chloride free and 3 wt.% chloride containing solutions of 0.1,
Accepted 24 January 2012 0.5, and 0.8 M bicarbonate at 20, 40, and 60 °C. Open circuit potentials, polarization potentiodynamics,
Available online 2 February 2012
and impedance spectroscopy were utilized. The behaviour exhibited a cathodic dependence on bicarbon-
ate and temperature where the corrosion rates consequently increased. Chloride prevented stable passiv-
Keywords: ation and increased the anodic sensitivity. EIS and polarization results were in agreement and impedance
A. Low alloy steel
in the presence of chloride was temperature dependent.
B. EIS
B. Polarization
Ó 2012 Elsevier Ltd. All rights reserved.
C. Kinetic parameters

1. Introduction traces on the corrosion of newly-developed high-strength steels


did not receive considerable interest. Our research work is devoted
Oil pipelines constructed from high strength low alloy (HSLA) to investigating electrochemical corrosion aspects of API-X100 ex-
steels are the transportation facilities through which sustainable, posed to natural aeration and bicarbonate in the presence and ab-
high pressure; high flow rate supplies of hydrocarbons are achieved. sence of chloride where the controversial role of bicarbonate in
During the ongoing operation time, they become more susceptible mild alkalinity is studied. Mildly alkaline pipeline flows can be
to many chemical degradation failures. Corrosion caused by car- established after excessive inhibitor injections and/or in the pres-
bon-carrying species is one of the most common problems that ence of miscellaneous ‘‘foreign’’ chemicals [10]. Most pipeline cor-
pipelines are prone to [1]. It was reported that almost 60% of oil rosion investigations did not consider oxygen or at least natural
pipeline corrosion failures in the USA are attributed to an inade- aeration. However, in the first stages of oil transportation, consider-
quate understanding of the corrosive capabilities of these able amounts getting leaked into the multiphase flows [11] can be a
dissolved species [2]. Many pipeline failures and disastrous fire great attribute towards failure tendency.
explosions were attributed to the continuous thinning of pipeline Bicarbonate was reported to be a key corrosive agent involved
sections caused, especially at low flow rates, by the electrochemical in anodic and cathodic reactions [12]. More specifically, the deter-
dissolution [3]. mining steps were found to be fundamentally associated with
Bicarbonate (HCO 3 ) plays a critical role in the dissolution bicarbonate as they are, for example, reduced directly to produce
reactions at internal and external sides of pipeline structures. Many adsorbed hydrogen atoms and/or hydrogen gas [13] represented
works were reported on the electrochemically-enhanced stress cor- by Eqs. (1)–(3) as:
rosion cracking when bicarbonate-saturated ground water is in
contact with pipeline surfaces [4,5]. Linter et al. reported that bicar- HCO3 þ e ! Hads þ CO2
3 ð1Þ
bonate is a major agent in the dissolution of pipeline steels
although of the reported controversy on its role in acidic CO2-satu- HCO3 þ Hads þ e ! H2 þ CO2 ð2Þ
3
rated flows [6]. In that detailed study, bicarbonate was also pro-
posed to contribute in formation and complexation of iron
carbonate. In most pipeline corrosion studies, the mechanisms 2HCO3 þ 2e ! H2 þ 2CO2
3 ð3Þ
conventionally involved H2CO3 in driving the cathodic reactions The corrosion rates and polarization characteristics were found
in deoxygenated media [7–9]. However, the significance of dependent on bicarbonate content [14]. In that study, a unique
chloride-free, bicarbonate-containing media with few oxygen anodic peak resulted at 610 mVSCE in 0.5 and 1 M bicarbonate
solutions, but the passivation ranges decreased in 0.05 and 0.1 M
⇑ Corresponding author. Tel.: +1 778 997 4878. solutions. In the dilute bicarbonate solutions in [15], the potential
E-mail address: faysal09@interchange.ubc.ca (F.F. Eliyan). at which the passivation onsets (Epass) was more noble, but the

0010-938X/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2012.01.015
182 F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191

Table 1 of 14 mm diameter and 4 mm thickness. They were welded to cop-


The chemical composition and carbon equivalent of the working electrode. per wires by conductive silver pastes and then mounted in hard,
Composition (wt. %) C.E. cold-curing epoxy resins. Prior to each corrosion test, they were
C Mn Mo Ni Al Cu Ti Nb Cr V sequentially wet-ground with 120, 320, and 600 grit silicon carbide
(SiC) emery papers and then degreased ultrasonically with ethyl
0.1 1.66 0.19 0.13 0.02 0.25 0.02 0.043 0.016 0.003 0.45
alcohol for 10 min. Afterwards, they were rinsed with double
distilled water and then finally dried in a stream of cool air. The
chemical cleaning procedures followed were in accordance with
G1-03 ASTM standard [26].
The chemical composition analysis was performed by inductive
coupled plasma (ICP) and LECO carbon analysis. The summarized
analysis results are shown in Table 1. Prior to microstructural
examination, the mounted specimens were wet ground up to
1200 grit silicon carbide finish and then polished with 6 and
1 lm diamond suspensions. The steel sample was etched by 2% ni-
tal (2 ml nitric acid +98 ml ethyl alcohol) and then treated with
alcohol swapping and finally dried by air stream.
The microstructural features, examined with optical micros-
copy, are shown in Fig. 1, revealing a complex microstructure of
acicular ferrite and bainite colonies with slight changes in colour,
probably related to local variations in the chemical composition.

2.2. Test solutions


Fig. 1. Optical micrograph of API-X100 steel showing martensite (dark), bainite,
and ferrite (light) colonies.
The corrosion behaviour is studied in naturally aerated solu-
tions synthesized from analytical grade Fisher procured reagent
transpassivation potential (Etranspass) was independent from of sodium bicarbonate (NaHCO3) added with 0.1, 0.5, and 0.8 M
bicarbonate. Additionally, the anodic current densities in the active concentration in double distilled deionized water. 3 wt.% chloride
regions increased with bicarbonate content and the associated was introduced to the same bicarbonate conditions considering
peaks were also larger [16]. In that work, passivation was the effect of chloride in a separate scheme. Although such a high
comprised of a mixture of oxides in films gradually forming as the concentration of bicarbonate of 0.8 M is seldom reached in typical
passive current density (ipass) showed a continuous decrease in pipeline environments, however, to be compliant with the consid-
the passive regime. Ikeda et al. proposed that pipeline corrosion is erations explained in the introduction, the study was performed
governed by the cathodic reactions in a manner dependent on the with an extended range of bicarbonate. The test temperatures were
amounts of the reducible species and the associated partial pres- 20, 40, and 60 °C and maintained within ±1 °C.
sures [17]. Zhang et al. confirmed the acceleration of the cathodic
reactions with bicarbonate and reported multiple anodic peaks in 2.3. Experimental setup
the passive regimes [18]. Temperature effect on mechanisms and
corrosion rates in the bicarbonate solutions was also investigated Corrosion tests were performed in a glass jacket test cell of a total
[19] and [20]. The role of intermediate complex species in the ano- volume of 0.6 L, open to the atmosphere. The working electrode was
dic dissolution changed at higher temperatures and the tendency the studied material, the counter electrode was made from graphite,
and properties of passivation were also temperature dependent and the measured potentials were in reference to the saturated cal-
[21,22]. Effective corrosion products developing with time led to omel electrode (SCE) of +0.241 VSHE. The reference electrode was iso-
a decrease in the corrosion rates [23] and the role of diffusion-lim- lated in a salt bridge and kept in the laboratory’s ambient
ited transport of oxygen in bicarbonate solutions was discussed. temperature. The electrochemical contact to the polarized electrode
Alves et al. showed that introducing 10 mM chloride eliminated was achieved through the Luggin capillary tube with a Vycor frit
the onset of passivation in some bicarbonate solutions attributing close to the steel surface. The cell was connected to a heater manu-
that to an accelerated generation of iron (III) hydroxychloro com- factured by Cole-Parmer and equipped with a digital controller to set
plexes [24]. Transpassivation potentials in chloride-containing test temperatures accurately. The three electrodes were connected
solutions exhibited a positive shift with carbonate at different to a Princeton applied research (PAR) Versastat 4 potentiostat/galva-
temperatures [25]. nostat controlled by VersaStudio v1.50.3712 software. The produced
In this work, the necessity to study the electrochemical corro- free potentials, potentiodynamics, and EIS data were processed and
sion of new generation pipeline steels is introduced. The corrosion analyzed to obtain the parameters of variations reflecting the
of API-X100 is evaluated in aerated; chloride free, and chloride electrochemical impact of the environmental conditions.
containing; 0.1, 0.5 and 0.8 M bicarbonate solutions at 20, 40, and
60 °C. Open circuit potentials, potentiodynamic polarization, and 2.4. Experimental methods
electrochemical impedance spectroscopy were utilized to study
the kinetics, passivation, and interfacial processes in reference to When the required temperature was reached in naturally aer-
a matrix of environmental factors. ated conditions, pH was measured and the electrochemical tests
were performed. The stable OCP profiles with the decaying fluctua-
2. Experimental details tions of about ±0.01 mV/s, were achieved within test time periods
of about 7100 s. Consequently, electrochemical impedance spec-
2.1. Test material troscopy tests were performed to study the electrochemical inter-
actions at the stable OCP conditions. The frequency range was
Test samples were cut from an oil pipeline constructed from from 0.01 to 10 kHz with a sampling rate of 10 points per decade.
API-X100 steel and machined into disks with nominal dimensions Analysis of EIS data and the fitting to equivalent circuits were
F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191 183

Fig. 2. Open circuit potentials (a) in chloride-free bicarbonate solutions and (b) as a function of bicarbonate content at 20, 40, and 60 °C in chloride-free and chloride-
containing solutions.

performed by a software package, ZSimpWinÒ. The potentiodynam- with temperature and they decreased upon introduction of chlo-
ic polarization sweeps were carried out within potential ranges at ride as shown in Fig. 2b. This indicates an enhanced anodic kinetics
which the key kinetic characteristics were fairly revealed. Prior to [29] and confirms with the polarization results when the acceler-
the polarization tests, the experimental setup and procedures were ated current densities were accompanied by lower corrosion
validated by practices in the ASTM Standard G 5-94 [27]. potentials. Similar behaviors with respect to bicarbonate and
temperature were exhibited and the chloride effect was more pro-
3. Results and discussion nounced in the dilute media and at high temperatures.
The free anodic dissolution can involve OH to form iron (II)
3.1. Open circuit potential (OCP) measurements hydroxide [30]:

The effects of mixed free potentials with minimal fluctuations Fe þ 2OH ! FeðOHÞ2 þ 2e ð4Þ
were investigated within a time period of about 7100 s. OCPs in chlo-
The cathodic reactions involve the simultaneous reduction of
ride-free conditions, shown in Fig. 2a, increased with bicarbonate
bicarbonate [31,32] and dissolved oxygen:
content at all temperatures. As to be discussed in the polarization
test results, the accelerated cathodic reactions with bicarbonate
HCO3 þ e ! CO2
3 þ 1=2H2 ð5Þ
and temperature could be responsible for those trends suggesting
the cathodic sensitivity towards temperature [28]. The variation
band between the extreme OCP values of 0.1 and 0.8 M broadened HCO3 þ e ! CO2
3 þ Hads ð6Þ
184 F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191

Fig. 3. Potentiodynamic polarization in 0.1, 0.5, and 0.8 M bicarbonate solutions at 20 °C in (a) chloride-free and (b) chloride-containing conditions.

Table 2
Potentiodynamic polarization test results in chloride free conditions.

Temperature Bicarbonate concentration Ecorr ipass Epass Etranspass


(oC) (M) (mVSCE) (lA/cm2) (mVSCE) (mVSCE)
20 0.1 766 4.2 603 894
0.5 754 9.3 538 970
0.8 743 13.8 470 1000
40 0.1 753 11.5 667 877
0.5 748 41 652 912
0.8 739 43.4 624 934
60 0.1 750 14.2 662 875
0.5 743 43.1 658 920
0.8 735 53.7 655 931

1=2O2 þ H2 O þ 2e ! 2OH ð7Þ 3.2.1. Chloride free test solutions


The potentiodynamic profiles at 20 °C are shown in Fig. 3a for
0.1, 0.5, and 0.8 M conditions with similar characteristics. As
3.2. Potentiodynamic polarization tests
shown in Table 2, the corrosion potentials were more noble with
the bicarbonate content at all temperatures as both anodic and
Anodic and cathodic corrosion kinetics as well as passivation
cathodic current densities showed an increase. This suggests the
were studied in chloride free and in 3 wt.% chloride containing;
great influence of cathodic consumption of bicarbonate on the
0.1, 0.5, and 0.8 M bicarbonate solutions at 20, 40, and 60 °C.
F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191 185

Fig. 4. Nyquist impedance representation in chloride free bicarbonate solutions.

corrosion reactions when it exists with greater amounts and/or at 3FeCO3 þ 4OH ! Fe2 O3 þ 2HCO3 þ H2 O þ 2e ð17Þ
higher temperatures. Independently from temperature, an
In our case, the passivation seemed more effective at lower
evidence for a multi-step dissolution was exhibited at a transition
temperatures with lower passive current densities and more noble
potential 750 mVSCE possibly due to physical interferences on the
passivation potentials (Epass) as shown in Table 2.
polarized, pre-passivated surfaces [33].
In the same proposed context, two distinct anodic peaks ap-
OH involvement in the dissolution processes results in a defec-
peared exclusively at 20 °C for 0.1 and 0.5 M solutions referring
tive hydrous film [34] capable of decelerating the current densities.
possibly to a pure chemical oxidation of the already formed prod-
The film could be hydroxide-based [35] developed within a short
ucts [42] as:
potential range, which was illustrated in a Fe–H–C–O Pourbiax dia-
gram, [36] as: 4FeCO3 þ O2 þ 4H2 O ! 2Fe2 O3 þ 4HCO3 þ 4Hþ ð18Þ

Fe þ OH $ FeOHads ð8Þ
6FeCO3 þ O2 þ 6H2 O ! 2Fe3 O4 þ 6HCO3 þ 6Hþ ð19Þ
FeOHads ! FeOHads þ e ð9Þ
4FeðOHÞ2 þ O2 ! 2Fe2 O3 þ 4H2 O ð20Þ
FeOHads ! FeOHþads þ e ð10Þ
Transpassivation occurred at about 1 VSCE and it seemed inde-
FeOHþads þ OH ! FeðOHÞ2 ! Hydrous½FeðOHÞ2  ð11Þ pendent from bicarbonate and temperature requiring separate
investigations on the electrochemistry of O2 at high potentials in
Catalyzed by sufficient bicarbonate amounts, iron carbonate (FeCO3)
the aerated bicarbonate media. The accelerated bicarbonate con-
can form to result in a double-layered film between 100 to 150 mV
tent-dependent cathodic reactions in 0.8 M conditions seemed to
above the first retardation as already found in voltammetric studies
obliterate that of water reduction at least in the mixed control
in [37]. In our case, the acceleration in the current densities can be
mass/charge transfer regime identified at about 0.89 VSCE.
explained by the partial removal of iron hydroxide [38]:

FeðOHÞ2 þ HCO3 ! CO2  2þ 3.2.2. Chloride containing test solutions


3 þ OH þ Fe þ H2 O ð12Þ
The polarization profiles at 20 °C are shown in Fig. 3b for 0.1, 0.5,
Distinct anodic peaks appeared with intensities proportional to the and 0.8 M conditions where the addition of 3 wt% chloride induced
bicarbonate content at more noble potentials where the accelerated significant kinetic changes and corrosion rates consequently
dissolution delayed the onset of passivation.Depending on bicar- increased. Although bicarbonate preserved its cathodic influence,
bonate content and pH, FeCO3 forms, before do other iron oxides, however, chloride made the corroding surfaces anodically sensitive
by adsorbed complexes [FeHCO þ 
3 ] ads, [FeHCO3 ] ads, and [FeOH ]
[43] where the corrosion potentials were lower in relative to the
[39] as: chloride-free conditions. The anodic regimes were apparently active
without retardations where chloride seemed to prevent pre-passiv-
Fe þ HCO3 $ FeHCO3 ð13Þ
ation steps as previously reported in [44]. That was also the case at
40 and 60 °C where the chances for effective passivation or for ano-
FeHCO3 ! FeHCO3 þ e ð14Þ
dic peaks to appear were inconsiderable. Ecorr was less sensitive to
higher temperatures especially in the dilute conditions where both
FeHCO3 þ OH ! FeCO3 þ H2 O þ e ð15Þ
anodic and cathodic reactions showed comparable kinetics [45].
The passive films finally comprised Fe3O4 and Fe2O3 with the Chloride ions seemed to get involved in an autocatalytic generation
contribution of reaction complexes and OH [40] and [41] as: of cation/oxygen pairs across the unstable pre-passive films [46] of
localized pH gradients. As a result, the anodic processes were accel-
3FeCO3 þ 5OH ! Fe3 O4 þ 3HCO3 þ H2 O þ 2e ð16Þ erated and the pitting susceptibility became greater [47].
186 F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191

Fig. 5. Impedance module representation plots in 0.1, 0.5, and 0.8 M chloride free bicarbonate solutions at (a) 20 °C and (b) 40 °C.

3.3. Electrochemical impedance spectroscopy (EIS)

EIS was utilized to study the possible interfacial processes [48]


in open circuit potentials with respect to bicarbonate, temperature,
and chloride. The equivalent circuits [49] comprised Rct, Rf, YW, and
RS representing charge transfer resistance, resistance across the
passive film, diffusion parameter, and solution resistance, respec-
tively. Chloride made the impedance dependent on bicarbonate
and temperature revealing the effects of adsorption and induction.
The new elements were then Ra and L representing adsorption Fig. 6. Equivalent circuits proposed for the electrochemical impedance response in
resistance and inductance, respectively and Qd and Rd were intro- chloride-free bicarbonate solutions.
duced in a third time constant to account for a supplementary
layer developed at higher temperatures.
kinetics and diffusion occurs suggest a presence of localized active
sites [50]. The complex plots achieved a good agreement with the
3.3.1. Chloride free test solutions polarization results as the size decreased with temperature.
Nyquist impedance plots are shown in Fig. 4 accounting for Passivation seemed to dominate the impedance in a manner where
bicarbonate and temperature with a relative similarity and reflect- the interactions became a function of the intrinsic, temperature
ing the significance of kinetics and diffusion-controlled processes. dependent properties of the passive films rather than of adsoprtion
The capacitive arcs expanded with bicarbonate and the variations and/or relaxation of intermediate species. As shown in Fig. 5a,
in the medium frequency regions, where a parallel control of impedance moduli |Z| in 0.5 and 0.8 M solutions at 20 °C exhibited
F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191 187

Table 3
Electrochemical Impedance Spectroscopy (EIS) component values in chloride free bicarbonate solutions at 20, 40, and 60 oC.

Components 20 °C 40 °C 60 °C
0.1 M 0.5 M 0.8 M 0.1 M 0.5 M 0.8 M 0.1 M 0.5 M 0.8 M
Rs (X cm2) 8.09 2.955 2.471 6.263 1.421 1.1275 2.805 1.6 3.532
Qdl (lF/cm2) 51.92 43.2 16.78 93.69 38.56 14.82 70.1 31.72 16.16
ndl 0.8339 0.8186 0.7514 0.783 0.878 0.602 0.9365 0.722 0.763
Rct (X cm2) 580.44 1513.2 1973 22.5 83.1 132.157 21.1 77 96.21
Qf (lF/cm2) 26.74 13.84 12.12 28.4 16.1 11.27 23.79 99.86 10.47
nf 0.9841 0.914 0.84 0.995 0.987 0.9383 0.981 0.987 0.9724
Rf (X cm2) 11.3 160.1 246.3 3.7 32.11 82 2.12 47.7 64.9
YWf (X1 S0.5) 0.003404 0.000483 0.0003182 0.004346 0.0009547 0.0005373 0.004606 0.0009633 0.0005331
Chi-square 3.39  105 2.97  104 2.14  105 6.37  105 7.87  105 1.36  105 4.97  105 4.86  105 3.11  105

Fig. 7. Experimental and calculated impedance module representations of impedance for selected chloride-free bicarbonate solutions.

similar variations with respect to frequency and the bode profiles Table 4
showed slight inflections at 1 Hz suggesting a possible change in Electrochemical Impedance Spectroscopy (EIS) component values in chloride con-
the governing mechanism(s) [51]. The high-frequency bode phase taining bicarbonate solutions at 20oC.

peaks were broad and more discernable with the bicarbonate Components 0.1 M 0.5 M 0.8 M
content and the phase angles were high at lower frequencies Rs (X cm2) 1.809 1.436 1.328
suggesting the significance of processes occurring outside the Qdl (lF/cm2) 59.46 45.14 32.92
double layer. ndl 0.8819 0.8848 0.8834
The proposed equivalent circuit for the chloride free conditions Rct (X cm2) 133.6 220 1781
Qa (lF/cm2) 57.96 44.11 19.24
is shown in Fig. 6 where the double layer and passivation contribu- na 0.6883 0.6221 0.6814
tions are considered in two separate time constants in a circuit of Ra (X cm2) 771.8 1679 2021
the configuration {R(QR)(Q(RW))}. The applicability of nested Chi-square 1.78  105 5.56  105 4.91  105
equivalent circuits of the configurations {R(Q(R(Q(RW))))} or
{R(C(R(Q(RW))))} was limited where they could be more compliant
adsorption processes [52]. Although these two circuits would be are shown in Table 3. Charge transfer resistance decreased with
also applicable for similar passivation conditions [53,54], but a good temperature but increased with bicarbonate in conditions where
fitting in our case would be achieved if two depressed semi-circles the passive films seemed more effective. Qdl exhibited an opposite
or a characteristic diffusion extension were exhibited at low trend with bicarbonate reflecting the capacitive character of the
frequencies. In addition, a third-time circuit {R(QR)(QR)(QR)} was double layer. Resistance across the passive films decreased about
attempted to account for a possible adsorption and/or insertion an order of magnitude at higher temperatures and it followed a
processes [55], but the applicability was limited and so was the case similar trend of that of Rct with bicarbonate, apparently indicating
for {R(QR)(QR)(Q(RW))}. the enhanced passivation [56] in the free conditions. The low Rf
A Constant Phase Element (CPE) was considered at the double values in the dilute solutions and/or at high temperatures, were
layer (Qdl) to account for the surface heterogeneities expressed as: accompanied by greater diffusion parameters (Yw). That suggests
 1 the greater mobility of the electroactive species through the
Z CPE ¼ Q ðjxÞn ð21Þ
pffiffiffiffiffiffiffi increased transfer channels in the passive films. The good applica-
(x) represents frequency, (j) equals to 1; and (n) is (CPE) expo- bility of the proposed circuit is shown in Fig. 7 for selected
nent and the electric components for the chloride-free conditions impedance data.
188 F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191

Fig. 8. Nyquist impedance representation in chloride-containing bicarbonate solutions at (a) 20 °C, (b) 40 °C, and (c) 60 °C.
F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191 189

Fig. 9. Iimpedance module representation plots in 0.1, 0.5, and 0.8 M chloride-containing bicarbonate solutions at (a) 20 °C and (b) 40 °C.

Table 5
Electrochemical Impedance Spectroscopy (EIS) component values in chloride containing bicarbonate solutions at 40oC.

Components 0.1 M Components 0.5 M Components 0.8 M


Rs (X cm2) 0.9526 Rs (X cm2) 0.9652 Rs (X cm2) 0.9323
Qdl (lF/cm2) 63.56 Qdl (lF/cm2) 55.94 Qdl (lF/cm2) 59.0
ndl 0.7824 ndl 0.801 ndl 0.9182
Rct (X cm2) 1.56 Rct (X cm2) 67.5 Rct (X cm2) 132.7
RL (X cm2) 53.3 Qa (lF/cm2) 34.64 Qf (lF/cm2) 18.29
L (H/cm2) 120.4 na 0.9021 nf 0.731
Chi-square 3.20  105 Ra (X cm2) 1303 Rf (X cm2) 24.41
Chi-square 3.83  105 Qd (lF/cm2) 13.5
nd 0.77
Rd (X cm2) 68.2
YWd (X1 S0.5 0.002754
Chi-square 1.62  105

3.3.2. Chloride containing test solutions consisted of two fairly overlapped loops at medium-to-high and
Nyquist impedance plots are shown in Fig. 8a–c for 20, 40, and low frequency regimes and the plot sizes increased with bicarbon-
60 °C conditions, respectively showing the induced effects of ate where the chances for diffusible passive films to form seemed
3 wt% chloride on the free interactions. At 20 °C, the profiles limited. Direct bicarbonate adsorption or more possibly relaxation
190 F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191

a Table 6
Electrochemical Impedance Spectroscopy (EIS) component values in chloride con-
taining bicarbonate solutions at 60oC.

Components 0.1 M 0.5 M 0.8 M


Rs (X cm2) 1.027 1.138 0.6287
Qdl (F/cm2) 0.0000535 0.0000511 0.0000112
ndl 0.833 0.9793 0.8826
Rct (X cm2) 3.32 76.8 103.7
Qf (F/cm2) 0.0000212 0.0000765 0.000069
b nf
Rf (X cm2)
0.54
0.115
0.773
38.3
0.802
42.6
Qd (F/cm2) 0.0000206 0.00001312 0.00006335
nd 0.68 0.7606 0.8432
Rd (X cm2) 133.3 546.2 400.9
YWd (X1 S0.5) 0.00424 0.004924 0.00603
Chi-square 4.46  105 5.93  105 6.03  105

c with an equivalent circuit {R(QR)(QR)(Q(RW))}which is shown in


Fig. 10c. At 60 °C, the interactions were similar independently from
bicarbonate and they reflected the significance of diffusion-limited
processes. Complying with the circuit proposed in Fig. 10c, the pas-
sive films seemed to exhibit more apparently the diffusion charac-
ter with the increased bicarbonate content. As shown in Tables 4–6,
Rct values were noticeably lower than those in the chloride free con-
ditions, and they decreased with temperature. However, it showed
Fig. 10. Equivalent circuits proposed for the electrochemical impedance response an opposite trend with the bicarbonate content, suggesting a de-
at (a) 20 °C and at 40 °C in 0.5 M chloride containing bicarbonate test solutions and cayed effect of chloride. Qdl seemed to be more capacitive in the
(b) at 40 °C in 0.1 M chloride containing bicarbonate solution. chloride-containing conditions and it showed an opposite trend of
Rct. Interestingly, Rf and Ra along with the CPEs, followed similar
trends. The confirmative fittings between the proposed circuits
of carbon carrying intermediate species [57] could be responsible and some selected data are shown in Fig. 11.
for this behavior. Single phase peaks appeared with intensities pro-
portional to bicarbonate content at 10 Hz reflecting the significance 4. Conclusion
of adsorption but not necessarily in a multi-time fashion as shown
in Fig. 9a. The equivalent circuit {R(Q(R(QR)))} for 20 °C conditions The cathodic reactions controlled the corrosion phenomena in a
is shown in Fig. 10a and it achieved a good fit with the experimental manner dependent on bicarbonate and temperature where the cor-
data. At 40 °C, chloride made the interactions dependent on bicar- rosion rates showed an increase. The cathodic effect was preserved
bonate similarly to CO2 saturated conditions studied in [58]. In when chloride was introduced but the steel surfaces showed an
0.1 M condition, inductive fields resulted as already found in [59] anodic sensitivity and the passivation steps were eliminated. EIS
modeled with an equivalent circuit {R(Q(R(RL)))} which is shown and polarization results were in a good agreement and the interfa-
in Fig. 10b. However, the competitive passive film formation cial interactions were governed by passivation and adsorption in
resulted in a third-time impedance in 0.8 M condition modeled chloride-free and chloride-containing conditions, respectively.

Fig. 11. Experimental and calculated iimpedance module representations of impedance for selected chloride-free bicarbonate test solutions.
F.F. Eliyan et al. / Corrosion Science 58 (2012) 181–191 191

Acknowledgements [30] L. Paolinelli, T. Pérez, S. Simiso, The effect of pre-corrosion and steel
microstructure on inhibitor performance in CO2 corrosion, Corros. Sci. 50
(2008) 2456–2464.
We would like to express our appreciation to Qatar National Re- [31] A. Veawab, A. Aroonwilas, Identification of oxidizing agents in aqueous amine–
search Fund (QNRF) and to the Natural Sciences and Engineering CO2 systems using a mechanistic corrosion model, Corros. Sci. 44 (2002) 967–
987.
Research Council of Canada (NSERC) for providing the financial
[32] R. Vaidya, Using electrochemical monitoring to predict metal release in
support for this work. drinking water distribution systems, PhD. Thesis, University of Central Florida,
2007
[33] D.E.J. Talbot, J.D.R. Talbot, Corrosion Science and Technology, Second ed., CRC
References
Press, LLC, 1998.
[34] L. Zhang, X. Li, C. Du, Y. Cheng, Corrosion and stress corrosion cracking
[1] S. Nesic, Key issues related to modeling of internal corrosion of oil and gas behaviour of X70 pipeline steel in a CO2 containing solution, J. Mater. Eng.
pipelines – A review, Corros. Sci. 49 (2007) 4308–4338. Perform. 18 (2009) 319–323.
[2] M. Kermani, L. Smith, CO2 corrosion control in the oil and gas production; [35] E. Castro, J. Vilche, Electrooxidation/electroreduction processes at composite
design considerations, first ed., The Institute of Materials, London, 1997. iron hydroxide layers in carbonate-bicarbonate buffers, J. Appl. Electrochem.
[3] Z. Xia, K. Chou, Z. Smialowska, Pitting corrosion of mild steel in CO2 21 (1991) 543–551.
containing NaCl brine, Corrosion 45 (1989) 636–642. [36] S. Hirnyi, Anodic hydrogenation of iron in a carbonate-bicarbonate solution,
[4] J. Gonzalez-Rodriguez, M. Casales, V. Salinas-Bravo, J. Albarran, L. Martinez, Mater. Sci. 37 (2001) 491–498.
Effect of microstructure on the stress corrosion cracking of X-80 pipeline steel [37] E. Castro, C. Valentini, C. Moina, J. Vilche, A. Arvia, The influence of ionic
in diluted sodium bicarbonate solutions, Corrosion 58 (2002) 584–590. composition on the electrodissolution and passivation of iron electrodes in
[5] B. Harle, J. Beavers, Technical note: low-ph stress corrosion crack propagation potassium carbonate-bicarbonate solutions in the 8.4–10.5 pH range at 25 °C,
in API X-65 line pipe steel, Corrosion 49 (1993) 861–863. Corros. Sci. 26 (1986) 791–793.
[6] R. Linter, G.T. Burstein, Reactions of pipeline steels in carbon dioxide solutions, [38] L. Niu, Y. Cheng, Corrosion behaviour of X-70 pipe steel in near-neutral pH
Corros. Sci. 41 (1999) 117–139. solution, Appl. Surf. Sci. 253 (2007) 8626–8631.
[7] J. Villarreal, D. Laverde, C. Fuentes, Carbon-steel corrosion in multiphase slug [39] L. Moiseeva, Carbon dioxide corrosion of oil and gas field equipment, Prot. Met.
flow and CO2, Corros. Sci. 48 (2006) 2363–2379. 41 (2005) 76–83.
[8] K. Efird, R. Jasinski, Effect of the crude oil on corrosion of steel in crude oil/brine [40] V. Alves, C. Brett, Influence of alloying on the passive behaviour of steels in
production, Corrosion 45 (1989) 165–171. bicarbonate medium, Corros. Sci. 44 (2002) 1949–1965.
[9] K. Masamura, S. Hashizume, K. Nunomura, J. Sakai, I. Matsushima, Corrosion of [41] Z. Lu, C. Huang, D. Huang, W. Yang, Effects of a magnetic field on the anodic
carbon and alloy steels in aqueous CO2 environments, CORROSION/1984. dissolution, passivation and transpassivation behaviour of iron in weakly
Paper no. 14, NACE, 1984. alkaline solutions with or without halides, Corros. Sci. 48 (2006) 3049–3077.
[10] W. Lyons, G. Plisga, Standard handbook of petroleum & natural gas [42] G. Zhang, Y. Cheng, Micro-electrochemical characterization and Mott-Schottky
engineering, Elsevier Inc., UK, 2005. analysis of corrosion of welded X70 pipeline steel in carbonate/bicarbonate
[11] F. Manning, R. Thompson, Oilfield Processing of Petroleum: Crude oil, Penn solution, Electochim. Acta. 55 (2009) 316–324.
Well Publishing Company, USA, 1995. [43] J. Jelinek, P. Neufeld, Temperature effect on pitting corrosion of mild steel in
[12] M. Nazari, S. Allahkaram, M. Kermani, The effects of temperature and pH on de-aerated sodium bicarbonate-chloride solutions, Corros. Sci. 20 (1980) 489–
the characteristics of corrosion product in CO2 corrosion of grade X70 steel, 496.
Mater. Des. 31 (2010) 3559–3563. [44] F. Al-Kharafi, B. Ateya, R. Abdallah, Electrochemical behaviour of low carbon
[13] G. Ogundele, W. White, Observations on the influence of dissolved steel in concentrated carbonate chloride brines, J. Appl. Electrochem. 32 (2002)
hydrocarbon gases and variable water chemistries on corrosion of an API- 1363–1370.
L80 steel, Corrosion 43 (1987) 665–673. [45] M. El-Naggar, Effects of Cl-, NO 2
3 and SO4 anions on the anodic behaviour of
[14] X. Mao, X. Liu, R. Revie, Pitting corrosion of pipeline steel in dilute bicarbonate carbon steel in deaerated 0.5 M NaHCO3 solutions, Appl. Surf. Sci. 252 (2006)
solution with chloride ions, Corrosion 50 (1994) 651–657. 6179–6194.
[15] A. Torres-Islas, J. Gonzalez-Rodriguez, J. Uruchurtu, S. Serna, Stress corrosion [46] D. Li, Y. Feng, Z. Bai, J. Zhu, M. Zheng, Influence of temperature, chloride ions
cracking study of micro-alloyed pipeline steels in dilute NaHCO3 solutions, and chromium element on the electronic property of passive film formed on
Corros. Sci. 50 (2008) 2831–2839. carbon steel in bicarbonate/carbonate buffer solution, Electrochim. Acta. 52
[16] K. Videm, A. Koren, Corrosion, passivity, and pitting of carbon steel in aqueous (2007) 7877–7884.
solutions of HCO 3 , CO2, and Cl-, Corrosion 49 (1993) 746–754. [47] C. Ren, X. Wang, L. Liu, H. Yang, N. Xian, Lab and field investigations on
[17] A. Ikeda, M. Ueda, S. Mukai, CO2 behaviour of carbon and Cr steels, Advances in localized corrosion of casing, Mater. Corros. 61 (2010) 1–5.
CO2 Corrosion. Paper No. 5, NACE, 1984. [48] E. Castro, Analysis of the impedance response of passive iron, Electrochim.
[18] G. Zhang, M. Lu, C. Chai, Y. Wu, Effect of HCO 3 concentration on CO2 corrosion Acta 39 (1994) 2117–2123.
in oil and gas fields, Materials 13 (2006) 44–49. [49] R. Cottis, S. Turgoose, Electrochemical impedance and noise, NACE
[19] M. Kermani, A. Morshed, Carbon dioxide corrosion in oil and gas production-A International, USA, 1999.
Compendium, Corrosion 59 (2003) 659–683. [50] L. Mu, W. Zhao, Investigation on carbon dioxide corrosion behaviour of
[20] D. Davis, G. Burstein, The Effect of bicarbonate on the corrosion and HP13Cr110 stainless steel in simulated stratum water, Corros. Sci. 52 (2010)
passivation of iron, Corrosion 36 (1980) 416–422. 82–89.
[21] J. Li, D. Meier, An AFM study of the properties of passive films on iron surfaces, [51] H. Ma, X. Cheng, S. Chen, C. Wang, J. Zhang, H. Yang, An ac impedance study of
J. Electroanal. Chem. 454 (1998) 53–58. the anodic dissolution of iron in sulfuric acid solutions containing hydrogen
[22] S. Savoye, L. Legrand, G. Sagon, S. Lecomte, A. Chausse, R. Messina, P. Toulhoat, sulfide, J. Electroanal. Chem. 451 (1998) 11–17.
Experimental investigations on iron corrosion products formed in bicarbonate/ [52] Z. Guo-xian, L. Xiang-hong, X. Jian-min, H. Yong, Formation characteristic of
carbonate – containing solutions at 90 °C, Corros. Sci. 43 (2001) 2049–2064. CO2 corrosion product layer of P110 steel investigated by SEM and
[23] L. Cáceres, T. Vargas, L. Herrera, Influence of pitting and iron oxide formation electrochemical techniques, J. Iron. Steel Res. Int. 16 (2009) 89–94.
during corrosion of carbon steel in unbuffered NaCl solutions, Corros. Sci. 51 [53] Y. Chen, W. Jepson, EIS measurement for corrosion monitoring under
(2009) 971–978. multiphase flow conditions, Electrochim. Acta. 44 (1999) 4453–4464.
[24] V. Alves, C. Brett, The influence of alloying on the passive behaviour of steels in [54] P. Liang, X. Li, C. Du, X. Chen, Stress corrosion cracking of X80 pipeline steel in
bicarbonate medium studied by electrochemistry and XPS, Key Eng. Mater. simulated alkaline soil solution, Mater. Des. 30 (2009) 1712–1717.
232 (2002) 436–439. [55] V. Alves, C. Brett, Characterization of passive films formed on mild steels in
[25] M. Ergun, A. Turan, Pitting potential and protection potential of carbon steel bicarbonate solution by EIS, Electrochim. Acta. 47 (2002) 2081–/2091.
for chloride ion and the effectiveness of different inhibiting anions, Corros. Sci. [56] X. Liu, X. Mao, Electrochemical polarization and stress corrosion cracking
32 (1991) 1137–1142. behaviours of a pipeline steel in dilute bicarbonate solution with chloride ion,
[26] ASTM Standard G 1–03, Standard practice for preparing, cleaning, and Scr. Metall. Mater. 33 (1995) 145–150.
evaluating corrosion test specimens, Annual Book of Standards, ASTM, [57] S. Wu, Z. Cui, G. Zhao, M. Yan, S. Zhu, X. Yang, EIS study of the surface film on
Pennsylvania, 2004. the surface of carbon steel from supercritical carbon dioxide corrosion, Appl.
[27] ASTM Standard G 5–94, Standard reference test method for making Surf. Sci. 228 (2004) 17–25.
potentiostatic and potentiodynamic anodic polarization measurements, [58] G. Lin, M. Zheng, Z. Bai, X. Zhao, Effect of temperature and pressure on the
Annual Book of Standards, ASTM, Pennsylvania, 2004. morphology of carbon dioxide corrosion scales, Corrosion 62 (2006) 501–507.
[28] C. Brossia, G. Cragnolino, Effect of environmental variables on localized [59] F. Farelas, A. Ramirez, Carbon dioxide corrosion inhibition of carbon steels
corrosion of carbon steel, Corrosion 56 (2000) 505–514. through bis-imidazoline and imidazoline compounds studied by EIS, Int. J.
[29] H. Fang, S. Nesic, B. Brown, General CO2 corrosion in high salinity brines, Electrochem. Sci. 5 (2010) 797–814.
CORROSION/2006. Paper No. 6372, NACE, 2006.

Você também pode gostar