Está en la página 1de 8

Materials Chemistry and Physics 140 (2013) 508e515

Contents lists available at SciVerse ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Influence of temperature on the corrosion behavior of API-X100


pipeline steel in 1-bar CO2-HCO
3 solutions: An electrochemical
study
Faysal Fayez Eliyan*, Akram Alfantazi
Corrosion Group, Department of Materials Engineering, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada

h i g h l i g h t s

 An increase in the corrosion rates with bicarbonate concentration and temperature.


 High temperature anodic peaks are compared with standard potentials of formation of iron oxides.
 Adsorption-controlled and diffusion-controlled EIS interactions at the OCPs and 0.6 V vs. SCE, respectively.

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

Article history: This paper addresses on the influence of temperature, elucidated with a number of electrochemical
Received 12 October 2012 methods and immersion tests, on the corrosion behavior of API-X100 steel in CO2-saturated bicarbonate
Received in revised form solutions. Investigated by cyclic potentiodynamic polarization, the corrosion rates, which showed a
23 February 2013
sensible increase with 10 g L1 (0.16 mol L1), 30 g L1 (0.5 mol L1) and 50 g L1 (0.82 mol L1) bi-
Accepted 23 March 2013
carbonate concentrations, increased from about 500, to 1500 and 1800 mA cm2 at 20, 50 and 90  C,
respectively. Passivation at 50 and 90  C showed resistance to deteriorate against 100 ppm chloride ions,
Keywords:
of which anodic 0.5 V vs. SCE peaks exclusively appeared. Moreover, transpassivation occurred at 0.9 and
Metals
Electrochemical techniques
0.7 V vs. SCE, respectively, unlike with the 20  C cases whose chloride-induced-pitting-vulnerable,
Energy dispersive analysis of X-rays gradually-forming passive films transpassivated at 1 V vs. SCE. At different potentials, the potentiostatic
Corrosion currents increased with temperature, but their profiles suggested more effective passivation, accordingly.
The charge transfer resistance, calculated by electrochemical impedance spectroscopy, decreased with
temperature at the open circuit potentials and 0.6 V vs. SCE, where the interfacial interactions were
governed by adsorption, and diffusion-limited processes, respectively.
Crown Copyright Ó 2013 Published by Elsevier B.V. All rights reserved.

1. Introduction contributes to the reliable design of the pipelines to transport oil


and gas at variable demands and to be operated more efficiently
Corrosion in oil and gas pipelines is one of the most challenging and safely [6].
operation problems to predict and control. Its dependence on the In this research, the influence of temperature on the corrosion
environmental factors and flow rates, which are numerous and behavior of API-X100 pipeline steel is evaluated in CO2-saturated
intertwined for pipeline steel, and its nature as sustained chains of bicarbonate solutions. In many previous studies, it was not mostly
multiple reactions, in a flow condition on the long run, are the addressed by more than one method, especially with the passiv-
broad topics that still spur the ongoing research interests [1e4]. ation growth and the consequent interference with the interfacial
They have become over the years more interdisciplinary to achieve interactions. In addition, the investigations were carried out in
for better understanding and reliable prediction of the corrosion aerated or deoxygenated bicarbonate, and CO2-saturated solutions,
behavior, susceptibility and rates [5]. And albeit of the complexities with no consideration for a possible synergism imposed by bicar-
of the problem, it can be tackled from certain aspects in association bonate and the dissolved CO2 e a situation that can frequently be
to specificly simulated laboratory conditions, and appropriately encountered during internal pipeline corrosion. API-X100, a new-
selected and implemented experimental methods. Ultimately, that generation pipeline steel, received our interest in a number of
previous studies. In saline CO2-saturated solutions, the corrosion
* Corresponding author. Tel.: þ1 778 997 4878.
rates, and apparently the cathodic current densities, increased with
E-mail addresses: faysal09@interchange.ubc.ca, faysalfayez@yahoo.ie (F.F. Eliyan). temperature [7,8]. The passive currents and transpassivation, in

0254-0584/$ e see front matter Crown Copyright Ó 2013 Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.matchemphys.2013.03.061
F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515 509

correlation to bicarbonate concentration and temperature in


aerated and deoxygenated solutions, showed discrepancy and the
nature of the interfacial interactions showed dependence on tem-
perature [9,10]. Jelinek et al. reported a decrease in the pitting
potentials with temperature for a mild steel, measured by
30 mV min1 potentiodynamic polarization sweeps, in deoxygen-
ated 0.1 M bicarbonate solutions of few chloride traces [11]. In
0.5 g L1 NaHCO3, CO2-saturated solutions, Louafi et al. reported an
increase of the corrosion rates of N80 steel from about 100 to
270 mA cm2, at 20 and 50  C, respectively [12]. Parkins et al.
correlated a number of parameters of kinetics and passivation of a
mild steel to bicarbonate concentration at 23  C in CO2-saturated,
but carbonate-containing solutions [13].
In this study, the influence of temperature, with a range of bi-
carbonate concentrations, is evaluated by a number of electro-
chemical methods and immersion. More comprehensively,
therefore, the corrosion rates, and the properties of the passive
films evaluated can serve for better predicting the corrosion
Fig. 1. Microstructure of as-received API-X100.
behavior of the new High Strength Low Alloy (HSLA) steels, and for
effective control of the corrosive pipeline flows of different
types [14]. 2.3. Test solutions

2. Experimental details The solutions were synthesized to simulate one of the most
typical types of water cuts that are saturated with CO2 which dis-
2.1. Test setup solves to contribute, in mildly alkaline conditions, to increased
concentrations of the coexisting bicarbonate and H2CO3 [1]. In our
The electrochemical tests were carried out in a 1-L, three- study, the solutions were first purged for 60 min with CO2, for
electrode, multi-port jacketed cell. The working electrode was the deoxygenation, and the purging was maintained continuous during
test material API-X100 machined into flat, thin, rectangular speci- the tests. To evaluate certain extents of aggressiveness of bicar-
mens which were sealed and immersed by a special holder. The bonate in 1 bar CO2-saturated solutions, three concentrations of
potentials were measured against the saturated calomel electrode NaHCO3, which is of analytical grade Fisher procured type, of 10, 30
(SCE) of þ0.241 V vs. SHE, which was kept isolated at room tem- and 50 g L1 were introduced to the solutions. The effect of
perature, and the counter electrode was made of graphite. To heat 100 ppm chloride was also studied in solutions of the same bicar-
the solutions to 50 and 90  C, the jacketed chamber of the cell was bonate concentrations. The pH, which was unbuffered, was be-
part of a hydraulic loop in which a constant-velocity water flow was tween 7.8 and 8.8. The tests were carried out at 20, 50 and 90  C,
maintained circulated by a Cole-Parmer heater. A regulated flow of maintained within 1  C.
high-purity carbon dioxide was purged into the solutions
throughout a time period of an electrochemical test. A computer-
synchronized Versastat 4 potentiostat was used to perform and 2.4. Tests
control the experiments and analyze the results.
After immersing the specimens, a potentiostatic 2 V vs. SCE
cathodic conditioning for 800 s was applied to dissolve any air-
2.2. Test material
formed oxides before running the experiments, which were
repeated at least for three times to ensure the reproducibility of the
The specimens were sequentially wet-ground by silicon carbide
results. The cyclic potentiodynamic experiments were performed
papers of 120, 320, and 600 grit, and then degreased ultrasonically
with a scan rate of 0.5 mV s1 from 1 to a vertex 1.2 V vs. SCE and
with ethyl alcohol for 10 min, rinsed with distilled water, and dried
then reversed to potentials below 1 V vs. SCE. The vertex 1.2 V vs.
with hot air. The chemical composition, shown in Table 1 on wt%
SCE potential was selected as a criterion of which the phase of
basis, was made by inductive coupled plasma (ICP) and LECO car-
passive film formation was assumed complete when the passive
bon analysis. A Nikon EPIPHOT 300 optical microscope was used to
currents show an abrupt increase at transpassivation. The poten-
reveal the microstructure of as-received sample. It was wet-ground
tiostatic polarization tests were performed at 0.6, 0, 0.6 and 1.2 V
with silicon carbide papers of up to 1200 grit, polished with 6 and
vs. SCE for 2000 s. The EIS tests were carried out at the open circuit
1 mm diamond suspensions, then immersed in 2% nital (2 ml
potentials and 0.6 V vs. SCE with a range of frequency of 0.01e
acid þ 98 ml ethyl alcohol) etchant, swapped with alcohol, and
10,000 Hz with a sampling rate of 10 points per decade.
dried in a hot air stream. The microstructure is shown in Fig. 1. It
The immersion tests were carried out for 6-day periods. The
mainly consists of a mixture of small amounts of ferrite and
weight loss was measured across after 2-day intervals. The bicar-
dispersed pearlite and bainite. There are slight microvariations in
bonate concentrations were 10 and 50 g L1 at 20 and 50  C.
colour with the microstructure attributed possibly to some micro-
heterogeneity in the alloying composition.
3. Results and discussion
Table 1
Chemical composition of the test material.
3.1. Cyclic potentiodynamic polarization and immersion
Composition (wt. %) C.E.

C Mn Mo Ni Al Cu Ti Nb Cr V The potetiodynamic polarization experiments were carried out


0.1 1.67 0.21 0.13 0.02 0.25 0.01 0.043 0.016 0.003 0.47
to measure the corrosion rates, and investigate the influence of the
environmental factors on the anodic and cathodic reactions in the
510 F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515

a b

c d

e f

Fig. 2. Cyclic potentiodynamic polarization in: a) 20  C, chloride-free, b) 20  C, chloride-containing, c) 50  C, chloride-free, d) 50  C, chloride-containing, e) 90  C, chloride-free, f)
90  C, chloride-containing solutions.
F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515 511

Fig. 3. Corrosion current densities, corrosion potentials, passivation potentials, and transpassivation potentials as parameters depending on bicarbonate, temperature, and chloride.

kinetic region, and on the passivation. As shown in Fig. 2, temper- chloride-free and chloride-containing solutions, was unique from
ature appeared to be the controlling factor, and where the behavior, 50 and 90  C solutions’ during passivation. After the active disso-
relatively regardless of bicarbonate concentration and/or chloride lution of high current densities, broad anodic peaks appeared
at a given temperature, was similar. The behavior at 20  C, both in which were the largest in 10 g L1 solutions. The smaller size in
more concentrated solutions suggests an earlier passivation in
situations where the pH was higher [15]. The potentials of these
peaks were lower. The shape of the peaks was different from those
of the aerated and N2-saturated bicarbonate solutions [10,16],
which at the passivation onset, the currents decelerated abruptly to
almost two orders of magnitude. For about 350 mV above the

Fig. 4. The potentiodynamic profiles at 20, 50, and 90  C in chloride-free, 30 g L1 Fig. 5. The peak potentials and the equilibrium potentials of some possible passivation
solutions. reactions at 90  C.
512 F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515

anodic peaks at 20  C, the passivation currents showed a gradual


decrease, until about 0.5 V vs. SCE the currents up to trans-
passivation were relatively stable. Independent from the bicar-
bonate concentration, transpassivation occurred at around 1 V vs.
SCE and negative hysteresis loops resulted upon scan reversal.
Chloride seemed to cause destabilization onto passivation and
initiated many repassivated pits, as evident from the positive hys-
teresis loops. Nevertheless, although it is difficult to articulate on
how thicker a passive film would be with higher bicarbonate con-
tent e so that it would be more immune to deterioration e bicar-
bonate, which is also an active anion, seemed to reduce the critical
concentration of chloride ions adsorbed onto a passive film to
deteriorate. Small positive loops appeared, but the passivation
range was notably broad in 50 g L1 solutions, suggesting that the
bicarbonate contributed to thicken the passivation or it adsorbed
on the majority of the surface likely sustaining for further oxidation
reactions. The cathodic branch at low temperature showed small
inner cathodic peaks, which were more apparent in chloride-
containing solutions, possibly associated to bicarbonate reduction,
obliterating that of charge transfer limited reduction of water
at 0.9 V vs. SCE.
At 50 and 90  C, the current densities were higher, the anodic
peaks were smaller and blunt, and the passivation commenced at
lower potentials. The physical effect of temperature on the
passivation structure was not to be resolved in this study, but
although of the given controversy in many previous studies [17e
20], the higher temperature generally hinders the chloride ions
from deteriorating the passive films, with no clear mechanism to be
suggested. The small second and third anodic peaks, which
appeared at 50  C, were notably larger at 90  C. Some parameters,
shown in Fig. 3, were extracted from the polarization profiles to
Fig. 6. The morphologies of the corrosion products formed during the potentiody-
better correlate the environmental factors of bicarbonate, temper-
namic polarization to 0.6 V vs. SCE in 30 g L1 solutions, at a) 50, and b) 90  C.
ature and chloride with the corrosion behavior. At 20  C, the
corrosion rates increased with bicarbonate concentration from
about 400 to 600 mA cm2, which were significantly higher than the

Fig. 7. Elemental EDS mapping for the corrosion surface in a 30 g L1 solution at 50  C.
F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515 513

aerated or N2-saturated bicarbonate systems’ [21,22], suggesting anodic peaks were a result of transformation from Fe3O4 and/or
that H2CO3 might have a strong effect in accelerating the reactions FeCO3 to g-Fe2O3. Fig. 6 shows the morphologies of the corrosion
in CO2-saturated bicarbonate solutions, relatively regardless of the products formed during the potentiodynamic polarization up to
pH. This is well backed by a model published by Han et al. to predict 0.6 V vs. SCE at 50 and 90  C in 30 g L1 solutions. Regardless of
the corrosion rates in similar solutions reporting that H2CO3 exists the presence or absence of chloride, the corrosion products at
with comparable concentrations of about 1.8 x 105 mol kg1 in 50  C seemed more compact and of larger grains, while those of
solutions containing 4.9 x 106 to 0.95 mol kg1 bicarbonate [23]. 90  C appeared dispersed. This could be well relevant to the high
The corrosion potentials decreased from about 730 to 760 mV passive current densities and the low transpassivation potentials
vs. SCE indicating the anodic sensitivity towards the increased bi- at 90  C. An elemental EDS mapping for iron, oxygen, carbon,
carbonate concentration. The corrosion rates increased by as much and manganese of the corrosion surface formed at 50  C is
as two to three times at 50 and 90  C, with which the corrosion shown in Fig. 7. Fe was semi-qualitatively analyzed to constitute
potentials respectively decreased, indicating the acceleration, pre- almost 50 wt% of the surface rich with O and C of about 40 and
dominantly, of the anodic reactions. The effect of chloride was more 10 wt%, respectively. Some other elements like manganese and
apparent in accelerating the corrosion rates at higher temperatures silicon were found to exist with about 0.81 and 0.6 wt%,
than 20  C, while the corrosion potentials in both cases were mostly respectively.
less than the chloride-free cases. Passivation onset potentials (Epass) The corrosion rates from the immersion tests are shown in Fig. 8.
were the highest in 10 g L1 solutions but their dependence on In 10 g L1 solutions, as shown in Fig. 8a, the corrosion rates were
higher bicarbonate concentrations was not clear, similar to the notably higher than that of the 50 g L1 solutions’. They decreased
discrepancy of the locations and shapes of the anodic peaks re- over time from about 12 g m2 day1 to 5 g m2 day1 and
ported by Gonzalez-Rodrigues et al. [24]. The higher temperature, 0.5 g m2 day1 at 20 and 50  C, respectively. The sample surfaces
however, made the passivation to onset at lower potentials, were shiny clean in 50 g L1 solutions and were, however,
regardless of the bicarbonate concentrations and chloride, completely covered with corrosion scales in 10 g L1 solutions
decreasing from about 300 mV to 600 mV vs. SCE. The chloride growing to decelerate the corrosion rates. It seems that the bicar-
effect was not as clear but it seemed to make the anodic peaks bonate can act as an effective inhibitor to protect the pipeline
bigger, or to appear overlapped of multiple peaks. The currents in surfaces if it exists with sufficiently high concentrations.
the passivation regimes at 50 and 90  C were almost one order of
magnitude higher than the 20  C cases’, being almost within 110e
280 mA cm2. The critical current densities of the first anodic peaks
a
were almost independent from the environmental factors, and they
were between 4.5 and 7 mA cm2. The transpassivation potentials
(Etranspass), which reflect on the quality of the passive films, were
the highest at 20  C of about 1000, decreasing to 900 and 700 mV
vs. SCE, at 50 and 90  C, respectively. The reasons behind that highly
reproducible trend encourage special physico-electrochemical in-
vestigations to be carried out in the future.
The peaks which appeared above the first anodic peaks at
around 0.2, and more notably at 0.5 V vs. SCE at 90  C drew our
attention. Although we cannot articulate on the quality of the
passive films at that high temperature, these peaks could less
represent repassivated ruptures [11] than the formation of other
oxides having greater tendency at high temperatures to constitute
multi-layered passivation. As shown in Fig. 4, the three peaks were
annotated EP1, EP2, and EP3 indicating them with comparisons to 50
and 20  C cases in 30 g L1 solutions. From the passivation reactions
utilized to analyze the polarization behavior and the chemistry of
the corrosion products in CO2-saturated bicarbonate solutions, by b
Parkins et al., Castro et al., and Lu et al. [13,25e27], we calculated
the equilibrium potentials of the reactions (1) to (5), plotted along
with the peak potentials in Fig. 5.

Fe þ HCO þ
3 /FeCO3 þ H þ 2e

(1)

3FeCO3 þ 4H2 O/Fe3 O4 þ 3HCO þ


3 þ 5H þ 2e

(2)

2Fe3 O4 þ H2 O/3Fe2 O3 þ 2Hþ þ 2e (3)

2FeCO3 þ 3H2 O/g  Fe2 O3 þ 2HCO þ


3 þ 4H þ 2e

(4)

2Fe3 O4 þ H2 O/3g  Fe2 O3 þ 2Hþ þ 2e (5)

The figure implies that the first anodic peaks are likely Fig. 8. Corrosion rates calculated from immersion tests in a) 10 and b) 50 g L1
associated with FeCO3 oxidation, while the second and third solutions.
514 F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515

Fig. 9. Potentiostatic polarization in chloride-free, 10 g L1 solutions at 20  C.

3.2. Potentiostatic polarization but the profile of 0.6 V vs. SCE, although of the highly intense
fluctuations within the other potentials in the passivation range,
The potentiostatic experiments were performed at four poten- was different notably from that of the 20  C case.
tials of 0.6, 0, 0.6, and 1.2 V vs. SCE to observe for the variations of
the current densities in relation to the dissolution and passivation 3.3. Electrochemical impedance spectroscopy
across almost 2000 s. The potentiostatic profiles of 10 g L1 solu-
tions at 20  C are shown in Fig. 9. In 0.6 V vs. SCE, the currents The electrochemical and other diffusion-limited interactions
increased quickly in the first 120 s to reach 2.45 mA cm2 (it was [29] during the open circuit potentials and 0.6 V vs. SCE were
almost 1.2 mA cm2 during the potentiodynamic polarization) and investigated by EIS. As shown in Fig. 11, it was evident that the
the substrate underwent extensive dissolution. During 0 V vs. SCE, interfacial reactions were mainly governed by adsorption during
the currents were less by almost three orders of magnitude, of the OCP conditions. Raising the temperature to 90  C did not
about 1.5 mA cm2, corresponding to a passive state across which change the nature of the interactions but the overlapped,
the potentiodynamic currents were gradually decreasing. The depressed, capacitive arcs were smaller, reflecting less charge
profile showed extensive current fluctuations (or spikes) repre- transfer resistance and confirming with the polarization results.
senting, as reported by Tang et al. [28], ruptures and consequent Zhang et al. and Farelas et al. discussed the role(s) of certain in-
repairs of metastable passive films. The spikes dampened with time termediate species in similar corrosion environments [30,31].
until the currents became stable in the last 400 s. Interestingly, the Complex species such as FeOHþ get adsorbed to drive the disso-
currents at 0.6 V vs. SCE were significantly more stable and less lution with rates changing over time, and as the reductions of
than that of the 0 V vs. SCE case, indicating a better passivation. The H2CO3, HCO þ
3 and H simultaneously take place with slower rates.
currents at 1.2 V vs. SCE increased steadily with time, but remained The impedance modules were noticeably less at higher tempera-
less than 300 mA cm2, corresponding possibly to parallel processes tures across the full frequency range, and the sole phase peaks
of transpassivation and O2 evolution. The currents in other bicar- appeared to shift to higher frequencies at higher temperature. The
bonate solutions were generally within comparable ranges, but in EIS response in the other solutions, regardless of the bicarbonate
the presence of chloride, the currents showed discrepancy, espe- concentration, and to a high extent of the chloride, was not quite
cially at the high potentials. At 90  C, whose profiles are shown in different. The experimental data were simulated with the equiva-
Fig. 10 representing for 10 g L1 solutions, the currents were higher lent circuit {R(Q(R(QR)))} which accounts for charge transfer at the

Fig. 10. Potentiostatic polarization in chloride-free, 10 g L1 solutions at 90  C.


F.F. Eliyan, A. Alfantazi / Materials Chemistry and Physics 140 (2013) 508e515 515

typical pipeline flow. The influence of temperature on the


a
corrosion rates and passivation was investigated by electro-
chemical methods and immersion in CO2-saturated bicarbonate
solutions. The corrosion rates showed sensible dependence on bi-
carbonate concentration, but they increased from about 500 to
1500 and 1800 mA cm2 at 20, 50 and 90  C, respectively. At 20  C,
the passivation onset was at high potentials, showed evidence of
gradual formation and deteriorated by 100 ppm chloride ions in
dilute bicarbonate solutions. At 50 and 90  C, the passivation was
resistant against the chloride pitting, and the transpassivation
occurred at lower potentials of 0.9 and 0.7 V vs. SCE, respectively.
The potentiostatic currents, although of being higher with tem-
perature at different potentials, suggested effective passivation
accordingly. The charge transfer resistance decreased with the
higher temperature, as revealed by the EIS measurements, and the
interactions were governed by adsorption and diffusion-limited
processes at the OCPs and 0.6 V vs. SCE, respectively.

Acknowledgment

This publication was made possible by NPRP grant # 09-211-2-


b
089 from the Qatar National Research Fund (a member of Qatar
Foundation). The statements made herein are solely the re-
sponsibility of the authors.

References

[1] S. Nesic, Corr. Sci. 49 (2007) 4308e4338.


[2] S. Papavinasam, A. Doiron, A. Revie, Corrosion 65 (2009) 663e673.
[3] B. Mishra, S. Al-Hassan, D. Olson, M. Salama, Corrosion 53 (1997) 852e859.
[4] C. de Waard, U. Lotz, D. Milliams, Corrosion 47 (1991) 976e985.
[5] P. Jackman, L. Smith, Advances in Corrosion Control and Materials in Oil and
Gas Production, IOM Communications, London, 1997.
[6] A. Peabody, Peabody’s Control of Pipeline Corrosion, second ed., NACE,
Houston, 2001.
[7] F. Eliyan, F. Mohammadi, A. Alfantazi, Corr. Sci. 64 (2012) 37e43.
[8] F. Eliyan, A. Alfantazi, J. Appl. Electrochem. 42 (2012) 233e248.
[9] F. Eliyan, E. Mahdi, Z. Farhat, A. Alfantazi, Int. J. Electrochem. Sci. 8 (2013)
3026e3038.
[10] F. Eliyan, E. Mahdi, A. Alfantazi, Corr. Sci. 58 (2012) 181e191.
Fig. 11. EIS, represented by a) Nyquist and b) Bode plots, in chloride-free, 10 g L1 [11] J. Jelinek, P. Neufeld, Corr. Sci. 20 (1980) 489e496.
solutions at 20 and 90  C, at the open circuit potentials and 0.6 V vs. SCE. [12] Y. Louafi, M. Ladjouzi, K. Taibi, J. Solid State Electrochem. 14 (2010) 1499e1508.
[13] R. Parkins, S. Zhou, Corr. Sci. 39 (1997) 175e191.
[14] T. Miesner, W. Leffler, Oil and Gas Pipelines in Nontechnical Language, first
double layer, and the adsorption resistance. The charge transfer ed., PennWell, Oklahoma, 2006.
[15] K. Videm, A. Koren, Corrosion 49 (1993) 746e754.
resistance was between 20 and 95 U.cm2, and at higher tempera- [16] X. Mao, X. Liu, R. Revie, Corrosion 50 (1994) 651e657.
ture the overall behavior was more capacitive. [17] Z. Cui, S. Wu, S. Zhu, X. Yang, Appl. Surf. Sci. 252 (2006) 2368e2374.
When polarizing the interfaces at 0.6 V vs. SCE, the EIS response [18] Y. Zhang, X. Pang, S. Qu, X. Li, K. Gao, Int. J. Greenh. Gas Con. 5 (2011) 1643e1650.
[19] G. Schmitt, M. Mueller, Critical Wall Shear Stresses in CO2 Corrosion of Carbon
was multi-time-constant based. As shown in Fig. 11, the Nyquist Steel. CORROSION/1999. Paper no. 44, NACE, 1999.
profiles were very large, in comparison to those of the OCP condi- [20] E. Dayalan, F. de Moraes, J. Shadley, S. Shirazi, E. Ribicki, CO2 Corrosion Pre-
tions, showing extended diffusion tails. The interactions were diction in Pipe Flow Under FeCO3 Scale-forming Conditions. CORROSION/98,
Paper no. 51, NACE, 1998.
governed by diffusion-limited transport through thick passive [21] V. Alves, C. Brett, Corr. Sci. 44 (2002) 1949e1965.
films. The impedance was simulated with the equivalent circuit [22] M. Mohorich, J. Lamb, D. Chandra, J. Daemen, R. Rebak, Metall. Mater. Trans. A
{R(QR)(QR)} and it achieved a very good fitting. The charge transfer 41A (2010) 2563e2574.
[23] J. Han, J. Zhang, J. Carey, Int. J. Greenh. Gas Con. 5 (2011) 1680e1683.
resistance was between 800 and 2300 U.cm2, and two semi- [24] J. Gonzalez-Rodriguez, M. Casales, V. Salinas-Bravo, J. Albarrarn, L. Martinez,
overlapped phase peaks appeared. The response, at this potential, Corrosion 58 (2002) 584e590.
was similar in the other bicarbonate solutions at 20 and 90  C. [25] R. Parkins, S. Zhou, Corros. Sci. 39 (1997) 159e173.
[26] E. Castro, J. Vilche, A. Arvia, Corr. Sci. 32 (1991) 37e50.
[27] Z. Lu, C. Huang, D. Huang, W. Yang, Corr. Sci. 48 (2006) 3049e3077.
4. Conclusion [28] Y. Tang, Y. Zuo, Mater. Chem. Phys. 88 (2004) 221e226.
[29] R. Cottis, S. Turgoose, Electrochemical Impedance and Noise, NACE Interna-
The influence of temperature on the corrosion behavior of a tional, USA, 1999.
[30] G. Zhang, Y. Cheng, Corr. Sci. 51 (2009) 1589e1595.
new-generation API-X100 steel was investigated in solutions [31] F. Farelas, M. Galicia, B. Brown, S. Nesic, H. Castaneda, Corr. Sci. 52 (2010)
simulating one of the considerably corrosive conditions in a 509e517.

También podría gustarte