You are on page 1of 11

Corrosion Science 74 (2013) 297–307

Contents lists available at SciVerse ScienceDirect

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

Corrosion of the Heat-Affected Zones (HAZs) of API-X100 pipeline steel


in dilute bicarbonate solutions at 90 °C – An electrochemical evaluation
Faysal Fayez Eliyan ⇑, Akram Alfantazi
Corrosion Group, Department of Materials Engineering, The University of British Columbia, Vancouver, BC, Canada V6T 1Z4

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

Article history: This research explores for key correlations between the microstructures of API-X100 steel HAZs, simu-
Received 24 October 2012 lated by GleebleÓ thermal cycles, and their electrochemical corrosion behavior in dilute bicarbonate
Accepted 2 May 2013 solutions at 90 °C. The potentiodynamic polarization revealed the role of ferrite of a HAZ cooled at
Available online 13 May 2013
10 °C/s with the lowest passive currents, and those of acicular ferrite and martensite of the 30 and
60 °C/s HAZs with unstable, thin passivation. The 0.5 V vs. SCE potentiostatic currents suggested also a
Keywords: slow passivation growth, of repetitive breakdowns and repassivations, for the 30 and 60 °C/s HAZs. EIS
A. carbon steel
equivalent circuits and time-dependent interfaces were proposed for each of the HAZ microstructures.
B. polarization
B. EIS
Crown Copyright Ó 2013 Published by Elsevier Ltd. All rights reserved.
B. potentiostatic
C. passive films

1. Introduction dic reactions, and to analyze the passivation growth of three


HAZs, collectively, to better understand the deterioration of a
Corrosion of oil and gas pipelines, from many undiscovered as- pipeline HAZ in its early stages. In literature, there appears a lack
pects, depends on the pipeline construction methods. Operation- of recent studies that take into account the significance of a HAZ
wise, however, it has been receiving a wide research interest since microstructure and the need for comprehensive investigations
the 1960s to predict for the corrosion rates, and to understand and that address the problem on an electrochemical basis. The elec-
control the physics and chemistry of the corrosive multiphase trochemical findings were a small part in almost all few such
flows [1–6]. Corrosion rates, if not susceptibility, remain difficult studies, in which evaluating the cracking susceptibility and stress
to predict when correlated to the metallurgy of a candidate steel. endurance were the main motivation to carry those studies out.
This is the case, specifically for instance, when localized critical In some cases, the electrochemical findings did not indicate spe-
changes with the microstructure and microcomposition occur to cific (or significant) differences in the corrosion behavior of the
a pipeline steel weldable with a variety of procedures in thermal HAZs in comparison to their base steels. The potentiodynamic
inputs and rates [7]. Corrosion then becomes more complex to ana- polarization tests reported by Mitsui et al. [9] and Zhang et al.
lyze, albeit not too difficult to detect – at the welding zones. It pro- [10], for instance, indicated virtually no difference in the corro-
ceeds onto at least three adjacent microstructures of the sion behavior of a number of HAZs from their welded base steels
weldment, the heat-affected zone (HAZ) and the base steel, with in bicarbonate–carbonate solutions. The slow potentiodynamic
uneven rates in a narrow region over the circumference of the polarization scans in this paper, however, revealed the signifi-
welding zones. Moreover, the galvanic conductance among these cance of each of the HAZs, in comparison to the API-X100 base
microstructures (which are of different electrochemical tendencies steel, with the stability of the passive films and the rate of the
to dissolve) contributes to enlarging micro-segregation of the ano- cathodic reactions. In an environmentally-assisted cracking study,
dic and cathodic reaction zones. Consequently, the corrosion over Mustapha et al. evaluated, with a range of potentiostatic poten-
time becomes preferential and increasingly localized, accelerating tials, the susceptibilities of the intergranular and transgranular
to cause continuous weakening and thinning of a pipeline becom- cracking and the hydrogen embrittlement for HAZs made from
ing highly more susceptible to dangerous ruptures [8]. API-X100 [11]. The tests were carried out in bicarbonate–carbon-
The emphasis, in this paper, is on the corrosion behavior of the ate solutions at 75 °C, and the authors pointed out the scarcity of
HAZs whose microstructures are primarily sensitive to the cooling the relevant studies in the field. Their HAZs were made by con-
rates during a regular pipeline welding. The electrochemical ventional heat treatments, while the HAZs in this paper are sim-
methods are implemented to study the dissolution and the catho- ulated by programmed GleebleÓ thermal cycles.
The test electrolytes synthesized for this study simulate
⇑ Corresponding author. Tel.: +1 778 997 4878. mildly-alkaline pipeline flows which can be encountered,
E-mail address: faysal09@interchange.ubc.ca (F.F. Eliyan). occasionally, after excessive inhibitor injection. Bicarbonate in

0010-938X/$ - see front matter Crown Copyright Ó 2013 Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.05.003
298 F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307

such conditions, especially at high temperature, exists with high 2.2. Test material
concentrations and can act as the main electro-active corrosive
species. In a pursuit on previous studies, investigating the electro- The specimens were wet ground in sequence of 120, 320, and
chemical corrosion behavior of API-X100 and API-X80 in bicar- 600 grit emery papers, ultrasonically degreased in ethyl alcohol,
bonate solutions is drawing our ongoing interest [12–14]. The
two new-generation HSLA steels, mainly due to their high-
pressure reliability, are currently considered in many strategic
projects, such as Keystone XL [15] between Canada and the US, 1000
and others in Qatar, Russia, Turkey, and the UK North Sea
[16,17]. Studying their corrosion behavior, especially when per-
taining to welding and pipeline construction, contributes to better 800
prediction and understanding of their corrosion behavior in the
field, and to the achievement – on the long run – the optimized

Temperature ( oC)
600
alloying compositions required for higher strength and corrosion
resistance, alike. 10oC/s
400 30oC/s
100oC/s 60oC/s
2. Experimental details
200

2.1. Corrosion test setup


0
The electrochemical tests were carried out in a 1-L, three-
electrode, multi-port jacketed cell. The working electrodes were
machined out of API-X100 pipeline shell into flat thin disks fitted 0 20 40 60 80 100 120
into a sample holder made of Tefzel, and sealed by a Karlez Time (s)
washer that maintain full, standard exposure of the surface to
the test solutions. The potentials were measured against the Fig. 2. The GleebleÓ thermal simulation cycles used to produce the HAZ micro-
Saturated Calomel Electrode (SCE) of +0.240 V vs. SHE, which structures by cooling at 10, 30, and 60 °C/s from a peak 950 °C.
was isolated in a salt bridge at the room temperature. The coun-
ter electrode was made from a slender graphite rod. The jacketed
chamber of the test cell was connected by plastic tubes to a cir-
culator that pumps a constant-velocity water flow of 95 °C to heat
the test solutions. The experiments were carried out by a Versas-
tat 4 potentiostat which was synchronized to a VersaStudio soft-
ware program, commercialized by Princeton Applied Research, to
control the experiments, and measure and analyze the electro-
chemical data.

Table 1
Chemical composition of the test API-X100 steel.

Composition (wt.%) C.E.


C Mn Mo Ni Al Cu Ti Nb Cr V
0.1 1.66 0.19 0.13 0.02 0.25 0.02 0.043 0.016 0.003 0.45

Fig. 3. The optical micrographs of the HAZ microstructures cooled at (a) 10, (b) 30,
Fig. 1. The optical micrograph of the as-received API-X100 microstructure. and (c) 60 °C/s.
F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307 299

rinsed in distilled water and dried with a stream of hot air before 2.3. Test solutions
immersion in test solutions. The chemical composition of API-
X100, presented in Table 1, was analyzed by inductive coupled The test solutions were synthesized from an analytical-grade,
plasma (ICP). A Nikon EPIPHOT 300 optical microscope produced Fisher-procured sodium bicarbonate (NaHCO3) reagent, with a
the optical micrographs of the as-received and HAZ samples. Prior concentration of 5 g/L, in double-distilled, deionised water. The
to the microstructural analysis, selected samples were wet-ground solutions were open to the atmosphere, of unbuffered pH between
up to 1200 grit finish and polished with 6 and 1 lm diamond sus- 8.1 and 8.4, at 90 °C of ±1 °C.
pensions. They, then, were immersed in a Nital etchant (2 mL of
70% nitric acid and 98 mL of anhydrous, denatured ethyl alcohol) 2.4. Electrochemical tests
and treated with alcohol swapping and dried in an air stream.
Fig. 1 shows the as-received microstructure. It consisted of a mix- Before commencing the experiments, a potentiostatic cathodic
ture of ferrite, bainite, and some intergranular ferrite. It showed conditioning of 2 V vs. SCE for 800 s was applied right after
some microvariations in color, which can be probably attributed immersing the samples in the heated solutions to eliminate the
to a micro-variation in the composition. A GleebleÓ 3500 thermal interference of any air-formed oxides. The experiments were re-
simulation machine was used to simulate three HAZs. The samples peated three times to ensure reproducibility. The cyclic potentio-
were heated at 100 °C/s up to 950 °C peak temperature, maintained dynamic polarization was scanned upward at 0.05 and 0.5 mV/s
for 0.5 s and then cooled with three controlled rates, which are pre- from 1.3 V vs. SCE to a vertex 1.2 V vs. SCE and then was scanned
sented in Fig. 2, of 10, 30, and 60 °C/s – cooling rates that can sim- downward to the open circuit potentials. The cathodic polarization
ulate roughly typical heat losses from a pipeline segment during a tests were also scanned upward with 0.5 mV/s from 2 V vs. SCE to
welding process. The HAZ microstructures are shown in Fig. 3. The the open circuit potentials. The potentiostatic polarization was car-
10 °C/s HAZ consisted of relatively large equiaxed ferrite, some dis- ried out at 0.5 V vs. SCE for 10,000 s. The electrochemical imped-
persed pearlite, and bainitic ferrite. The 30 °C/s HAZ consisted of a ance spectroscopy (EIS) tests were carried out after over 3000,
mixture of acicular ferrite and martensite/austenite (M/A), and the 6000, and 9000 s, at the open circuit potentials, with a range of fre-
microstructure of 60 °C/s HAZ was mainly martentistic with some quency from 10,000 to 0.001 Hz, with a sampling rate of 10 points
grain boundary ferrite. per decade.

Fig. 4. The profiles of the cyclic potentiodynamic polarization scanned at 0.05 mV/s for the as-received, and 10, 30, and 60 °C/s HAZs.
300 F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307

3. Results and discussion The appearance of relatively broad peaks extending from about
0.77 to 0.5 V vs. SCE is attributed to the intersection of the
3.1. Cyclic potentiodynamic polarization tests cathodic currents of oxygen reduction (Eq. (6)) with the passive
currents at least at two potentials.
The notably microstructure-dependent passive currents and the O2 þ 2H2 O þ 4e ! 4OH ð6Þ
interference of oxygen reduction with the active–passive transition
are the two main features revealed by the slow 0.05 mV/s poten- They can be viewed as a coalescence between the first anodic
tiodynamic scan, as shown in Fig. 4. The cathodic currents, domi- and cathodic peaks (which occurred in a highly reproducible man-
nantly representing water reduction at 1.3 V vs. SCE, were ner), except for the 30 °C/s HAZ, which initially showed an anodic
nearly 5 mA/cm2 as high, except for the acicular-ferritic 30 °C/s peak below multiple, smaller peaks below 0.2 V vs. SCE. The
HAZ, of nearly 130 lA/cm2. The anodic dissolution, which was of cathodic behavior, which showed the minimum currents in that
high currents below 0.7 V vs. SCE, incorporates bicarbonate and case, was not reported as distinct for similar acicular-ferritic
hydroxyl species – in mechanisms reported as seemingly indepen- microstructures when polarized at higher scan rates [10,27]. The
dent from the minimal oxygen traces [18,19]. The nature of disso- minimum cathodic activity of such a microstructure was reported
lution (given it was controlled and slow), as implied from the from harsher corrosive conditions [28].
current densities, was independent from a HAZ microstructure Above the mixed-control regions at almost 0.2 V vs. SCE, the
type. In literature, the extent of involvement of bicarbonate and significance of microstructure appeared the most during passiv-
hydroxyl, in such relatively dilute bicarbonate solutions, is unclear. ation, which is of electronic and morphological properties on
At the active–passive transition, the nature of the dissolution prod- whom largely could depend [10,29]. The as-received and 10 °C/s
ucts and the proportions of their constituents of Fe(OH)2 and FeCO3 HAZ, of the lowest 3 lA/cm2 passive currents, are of considerably
in a morphology in which they can be multilayered or intermixed large ferrite grains that dissolve much more readily. That facilitates
had received an earlier interest, but the findings carried some con- the corrosion products to precipitate into the conductive,
troversy. Electrochemically, it is mainly on whether it is the bicar-
bonate or hydroxyl initially (or more dominantly) drives the
charge-transfer steps after which the growth of the corrosion prod- a
ucts depend at higher potentials. Reported from voltammetric
scans for iron and mild steels in bicarbonate-based solutions of
pH of no less than 8.4, Fe(OH)2 forms as a defective, hydrous layer
during the ‘‘first stage’’ of oxidation [20,21,18]. Taking first into ac-
count the thermodynamic tendency of formation; the tempera-
ture-independent, charge-transfer steps of Fe(OH)2 formation
were reported as:
Fe þ H2 O $ FeðOHÞ þ Hþ þ e ð1aÞ

FeðOHÞ $ FeðOHÞþ þ e ð1bÞ

FeðOHÞþ þ OH $ FeðOHÞ2 ! Hydrous FeðOHÞ2 ð1cÞ


Lu et al. reported step (1b) as rate-determining, in a mechanism
hydroxyl acts as an active adsorbent, and in which Fe(OH) is a pre-
cursor of dissolution, reporting the overall reaction of Fe(OH)2 for-
mation as [22]:
Fe þ 2OH ! FeðOHÞ2 þ 2e ð2Þ
On the other hand, Simard et al. reported, as concluded from
b
voltammetric scans of rotating mild steel electrodes in bicarbonate
solutions, that bicarbonate directly oxidizes iron (Eq. (3)) in pro-
portion with its concentration [23].
Fe þ HCO3 ! FeHCOþ3 þ 2e ð3Þ
That was associated to the higher critical currents of the first
anodic peaks, at 0.65 V vs. SCE, with the bicarbonate concentra-
tion – higher critical currents whom Castro et al., on the other
hand, attributed to the partial dissolution of Fe(OH)2 by bicarbon-
ate as [21]:

FeðOHÞ2 þ HCO3 ! CO2 


3 þ OH þ Fe

þ H2 O ð4Þ
In general, and regardless of the controversy on the dissolution
steps, the interfaces are considered to be passivated, more rapidly
in our case at 90 °C, with a mixture of Fe(OH)2 and FeCO3. FeCO3,
whom Niu et al. reported to form in aerated bicarbonate solutions
[24], and as being stable in Pourbaix systems [25,26], can grow
mainly by carbon-carrying intermediates [22], or, as an outer layer
[23], forming as a result of Fe(OH)2 reacting with bicarbonate as [18]: Fig. 5. The profiles of (a) the cyclic potentiodynamic polarization, and (b) the
cathodic polarization scanned at 0.5 mV/s for the as-received, and 10, 30, and 60 °C/
FeðOHÞ2 þ HCO3 ! FeCO3 þ H2 O þ OH ð5Þ s HAZs.
F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307 301

anchoring cementite (Fe3C) platelets as continually thickening, 2FeCO3 þ 4OH ! Fe2 O3 þ 2HCO3 þ H2 O þ 2e ð9Þ
compact passive films. For the 30 and 60 °C/s HAZs, however, the
After reversing the currents, at the vertex 1.2 V vs. SCE, all sam-
greater lattice defects of their more rapidly cooled, ferritic-bainitic,
ples exhibited negative hysteresis loops, suggesting considerably
and martensitic microstructures seemed to make the passive films
stable, pitting-immune passive films. In a short communication,
more active and of greater donor density. As a result, they become
Von Fraunhofer attributed the negative loops to the reduction of
thinner, whose currents were on average 35 and 10 lA/cm2,
O2 to OH [32], although that the reduction of Fe2O3 and/or
respectively, showing also a gradual decrease with the upscan.
Fe3O4 can still not be discounted in our similar case.
Transpassivation occurred at between 0.45 and 0.6 V vs. SCE. There
The 0.5 mV/s polarization profiles are shown in Fig. 5a. The oxy-
were peaks that appeared more reproducibly for the 30 and 60 °C/s
gen-reduction cathodic peaks did not appear and the passivation
HAZs at 0.9 V vs. SCE, which can be ascribed, most likely, to further
behavior of all samples was relatively comparable and the cyclic
oxidations of the FeCO3-based films to Fe2O3 and/or Fe3O4.
loops were negative. It is worth mentioning that the corrosion
Although such transformations were found evident to take place
potentials, of almost between 0.8 and 0.77 V vs. SCE, were in
in a number of electrochemical and characterization studies, the
few millivolts difference with those of the slower 0.05 mV/s poten-
reactions and, in turn, the responsible species for driving the trans-
tiodynamic scan. In independent tests, the significance of a micro-
formations were many and different. That actually adds – given
structure type with the cathodic currents is shown by 0.5 mV/s
that their correlation to a microstructure type is very difficult to
potentiodynamic scans from 2 V vs. SCE to the corrosion poten-
establish for such conditions – to the original difficulty in dealing
tials, of profiles shown in Fig. 5b. The reduction was initially under
with the chemical synergism imposed by HCO 3 ;O2 ;H2 O; and OH .

mass-transfer control of currents of nearly 45 and 70 mV/s, for the
These species can be involved in electro-active roles at different
30 °C/s HAZ, and 10 and 60 °C/s HAZs, respectively. Further studies
potentials at which different paths for a certain transformation
are encouraged to study the interrelation between a HAZ micro-
can be proposed. Considering, for example, Fe2O3, below is shown
structure and the rate of the cathodic reactions, with slower scans
reactions proposed on its transformation from FeCO3 in electro-
especially when the reduction becomes under a mixed control of
chemical studies carried out in bicarbonate-based solutions, illus-
charge and mass transfer.
trated by equations (7) [30], (8) [31], and (9) [32] as:
4FeCO3 þ O2 þ 4H2 O ! 2Fe2 O3 þ 4HCO3 þ 4Hþ ð7Þ 3.2. Potentiostatic polarization tests

2FeCO3 þ 3H2 O ! Fe2 O3 þ 2CO2 þ


3 þ 6H þ 2e

ð8Þ The potentiostatic polarization experiments were carried out to
measure the currents, and observe for the manner they change,

3.0 25
As-received 10oC/sec HAZ
2.5
20
Current density (mA/cm 2)

Current density (µA/cm2)

2.0
15

1.5

10
1.0

5
0.5

0.0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Time (s) Time (s)

50 50
o
30oC/sec HAZ 60 C/sec HAZ

40 40
Current density (µA/cm2)

Current density (µA/cm2)

30 30

20 20

10 10

0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Time (s) Time (s)

Fig. 6. The profiles of the potentiostatic polarization at 0.5 V vs. SCE for the as-received, and 10, 30, and 60 °C/s HAZs.
302 F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307

showed a decay, and the intensity of the fluctuations was dramat-


ically dampening over the time to be almost 20 and 10 lA/cm2,
respectively. The shape of these fluctuations, in which the currents
abruptly increased, and slowly decreased, can be ascribed to a
repetitive mode of breakdowns and repassivations [33] of the pas-
sive films, under which the substrate dissolution was simulta-
neously inhibited. For all samples, the final potentiostatic
currents were generally in the same order of magnitude of the
potentidynamic currents.

3.3. Electrochemical impedance spectroscopy tests

The EIS tests were carried out to evaluate the time-dependent


electrochemical influence of the growing passive films on the
interfacial interactions, in reference to a HAZ microstructure, at
Fig. 7. The SEM surface observation of the corrosion product formed on the 10 °C/s the open circuit potentials. The charge-transfer resistance, and
HAZ after 9000 s of immersion in a 5 g/L NaHCO3 solution at 90 °C.
the resistance and capacitance of the passive films, fitted with
equivalent circuits, were calculated by ZSimpWin 3.30 dÓ for EIS
over 10,000 s, at 0.5 V vs. SCE – a potential at which all samples data collected over after 3000, 6000, and 9000 s. In the first place,
underwent a passive state. The lowest 6 lA/cm2 currents, as shown the rate of dissolution of ferrite, the surface area covered by the ac-
in Fig. 6, of fluctuations limited almost at between 3 and 10 lA/ tive adsorbent complexes [34], and, in turn, the steps (and rate-
cm2 over the time, were exhibited by the 10 °C/s HAZ. The currents determining steps) of the passive film formation [22] are the main
of the as-received sample were initially as high as 1.6 mA/cm2, on factors that can correlate its thickness, essentially, over a period of
average, during the first 2500 s, before a continuous decay to reach time, to a microstructure type [35,36]. In our case, the passive films
100 lA/cm2. The more stable currents of the 10 °C/s HAZ suggest, appeared as an intact, dark precipitation only on the 10 °C/s HAZ,
as explained in Section 3.3, that the passive films grew to be thick- as shown in Fig. 7, while they appeared as slightly transparent,
er/compact in a direct dependence on the initial dissolution of its brownish films on the other samples after 9000 s. It is shown in
more abundant ferrite content. The 30 and 60 °C/s HAZs exhibited Figs. 8 and 9 the Nyquist spectra at the end of the immersion,
a similar behavior during which the potentiostatic currents and the Bode phase spectra after specific time intervals.

Fig. 8. The Nyquist EIS spectra for the as-received, and 10, 30, and 60 °C/s HAZs after 9000 s of immersion.
F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307 303

Fig. 9. The Bode EIS spectra for the as-received, and 10, 30, and 60 °C/s HAZs after 3000, 6000 and 9000 s of immersion.

Dissolution of ferrite and subsequent precipitation and growth of


the passive films of the As-received and 10 °C/s HAZ could have
proceeded initially, due to their relative microstructural resem-
blance, with similar manners. That is, as explained by Dugstad
[37], the vulnerability for ferrite of the two samples to selectively
dissolve, allow Fe3C to be an active unattacked area becoming, with
time, from being a region onto which the adsorbent intermediates
drive precipitation in an accelerating process, to a region onto
which the cathodic reduction is slower and reaction-controlling.
While it is difficult, in our case of mildly alkaline bicarbonate solu-
tions, to determine on the role(s) of the cathodic reduction of O2,
HCO 3 ; and H2 O (unlike with the case of the predominant reduction
of H+ in lower-pH, CO2-saturated solutions [6]), the homogeneity Fig. 10. The equivalent circuit fitted with the experimental EIS data for the as-
received sample.
and thickness of the passivation growing into Fe3C seemed to dom-
inantly affect the interactions, and in turn the charge-transfer
resistance. For the As-received sample, although that the depth np np
of Fe3C (between 35 and 55 lm) could not have been as large as Y ¼ YQ xn cos þ j YQ xn sin ð10Þ
2 2
that of 10 °C/s HAZ, the thickness of film precipitation seemed
not to change substantially over time. The passive film, as con- where x is the angular frequency and n is the CPE exponent. The
firmed by fitting the EIS data with the {R(Q((QR)R))} equivalent cir- solution resistance (Rs) was almost 11 O.cm2 and as shown in Ta-
cuit, formed a mixture with Fe3C at a thickness beyond which part ble 2, the double layer, as indicated from nch values, was pseudo-
of Fe3C remained a porous layer, as represented in Fig. 10. To ac- capacitive. As shown in Fig. 9, the phase values beyond the low-fre-
count for the interfacial heterogeneities [38], a constant phase ele- quency peaks at nearly 0.02 Hz increased with time, associated, as
ment (CPE) was used for the double layer and the passive film, demonstrated mathematically by Mansfeld [40], and experimen-
whose admittance (Y) is expressed as [39]: tally by Kinsella et al. [41], with the increase of charge-transfer
304 F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307

Table 2 resistance (Rch). The resistance of the porous Fe3C (Rp) did not
Equivalent circuit fitting for the EIS data of the as-received sample. change from an average of about 130 O cm2.
Elements Time (s) For the 10 °C/s HAZ, the interface transformed over time to
3000 6000 9000
make the interactions multi-time-constant based, as indicated
from the changes with the size and number of the phase peaks
Rs (O cm2) 11.64 10.56 10.71
Yfilm (S sn/cm2) 4.16E04 2.70E04 2.44E04
in Fig. 9. During immersion, the passive film grew to occupy Fe3C
nfilm 0.75 0.78 0.79 to constitute a second time constant in a {R(QR)(QR)} circuit, which
Ydl (S sn/cm2) 3.10E03 2.60E03 2.59E03 is shown in Fig. 11b. This circuit was used by Alves et al. in more
nch 0.62 0.76 0.77 concentrated bicarbonate solutions to study the passivation of a
Rch (O cm2) 9255 60,800 122500
number of steels at different potentials [42]. The passive film
Rp (O cm2) 140.6 113.5 122.8
Chi-square 5.53E05 5.47E05 3.84E05 showed evidence (see Fig. 7) to grow further, forming a deposit
layer, to account for a third-time constant as shown in Fig. 11c,

Fig. 11. The equivalent circuits fitted with the experimental EIS data, collected after (a) 3000, (b) 6000, and (c) 9000 s, for the 10 °C/s HAZ sample.

Table 3
Equivalent circuit fitting for the EIS data of the 10 °C/s HAZ sample.

Elements Time (s) Elements Time (s) Elements Time (s)


3000 6000 9000
Rs (O cm2) 11.50 Rs (O cm2) 11.58 Rs (O cm2) 11.70
Yfilm (S sn/cm2) 8.34E06 Ydl (S sn/cm2) 6.25E05 Ydl (S sn/cm2) 4.85E06
nfilm 0.842 nch 0.66 nch 0.93
Ydl (S sn/cm2) 3.61E05 Rch (O cm2) 266900 Rch (O cm2) 285300
nch 0.76 Yfilm (S sn/cm2) 1.65E05 Yfilm (S sn/cm2) 7.65E05
Rch (O cm2) 211300 nfilm 0.91 nfilm 0.60
Rp (O cm2) 1711 Rfilm (O cm2) 576.6 Rfilm (O cm2) 813.5
Chi-square 2.15E05 Chi-square 6.20E05 Ydeposit (S sn/cm2) 4.74E05
ndeposit 0.74
Rdeposit (O cm2) 267.3
Chi-square 8.74E05
F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307 305

Fig. 13. The equivalent circuit fitted with the experimental EIS data for the 60 °C/s
HAZ sample.

slightly extended interface under the deposit layer. Rch increased


with time, as shown in Table 3, and the resistance of the passive
films (Rf) were almost between 600 and 800 O cm2.
For the 30 °C/s HAZ, the interface, as depicted from the final Ny-
c quist and Bode spectra, was of no more than two-time-constant
based. Fitting the EIS data provided an evidence for unstable, or
defective, passive films which could have underwent an intermit-
tent mode in covering the substrate. The acicular-ferritic micro-
structure in our case, as explained earlier, and as validated by
Zhang and Cheng [10], therefore did not facilitate with the passive
film growth to be thicker or more compact. The role of Fe3C is dif-
ficult to articulate on, but the thin passive films seemed, after a suf-
ficiently long period in the middle of immersion time, to have been
porous through which the solution was, more significantly, in con-
tact with the substrate. Such an interface was proposed by Qin
et al. [46] and Lee et al. [34], and the equivalent circuit
{R(Q((QR)R))}, as shown in Fig. 12b, was applicable to simulate it,
and the circuit achieved a fairly precise fitting with the EIS data.
It is important to indicate to the relatively high charge-transfer
resistance of the 30 °C/s HAZ, as shown in Table 4 – intrinsically
in relation to its microstructure; highly confirming with the poten-
Fig. 12. The equivalent circuits fitted with the experimental EIS data, collected after tiodynamic results that showed the minimum anodic and cathodic
(a) 3000, (b) 6000, and (c) 9000 s, for the 30 °C/s HAZ sample.
currents. The passive film of the 60 °C/s HAZ seemed to be more
stable, and its interface, as suggested from the Bode spectra, was
which was already used by Li et al., Hamadou et al., and Zhou et al., two-time-constant based, fitted with the equivalent circuit
to account for the absorption and insertion phenomena which oc- {R(QR)(QR)} shown in Fig. 13. The phase values shifted to lower
cur specifically within [43–45]. Since the morphology of, and espe- frequencies with time, associated to a slight continual increase in
cially cohesion between the inner and deposit layers were not to be the charge-transfer resistance, as shown in Table 5, by a passive
investigated for this case, the inner layer is denoted in Fig. 11c as a film of resistance of nearly 160 O cm2.

Table 4
Equivalent circuit fitting for the EIS data of the 30 °C/s HAZ sample.

Elements Time (s) Elements Time (s) Elements Time (s)


3000 6000 9000
Rs (O cm2) 10.2 Rs (O cm2) 11.8 Rs (O cm2) 11.5
Ydl (S sn/cm2) 3.85E05 Yfilm (S sn/cm2) 7.66E05 Ydl (S sn/cm2) 3.44E05
nch 0.98 nfilm 0.69 nch 0.77
Rch (O cm2) 117900 Ydl (S sn/cm2) 0.0045 Rch (O cm2) 928000
Yfilm (S sn/cm2) 1.03E05 nch 0.96 Yfilm (S sn/cm2) 8.72E05
nfilm 0.68 Rch (O cm2) 175400 nfilm 0.61
Rfilm (O cm2) 12,280 Rp (O cm2) 112500 Rfilm (O cm2) 8512
Chi-square 7.71E05 Chi-square 1.72E05 Chi-square 1.37E5
306 F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307

Table 5 [4] K. Videm, Fundamental studies aimed at improving models for prediction of
Equivalent circuit fitting for the EIS data of the 60 °C/s HAZ sample. CO2 corrosion, Proceedings from 10th European Corrosion Congress, Progress
in the Understanding and Prevention of Corrosion, vol. 1, Institute of Metals,
Elements Time (s) London, 1993. p. 513.
[5] S. Nesic, J. Postlethwaite, S. Olsen, An electrochemical model for prediction of
3000 6000 9000
corrosion of mild steel in aqueous carbon dioxide solutions, Corrosion 52
Rs (O cm2) 11.3 10.6 11.4 (1996) 280–294.
Ydl (S sn/cm2) 1.60E04 1.12E04 2.42E04 [6] F. Eliyan, F. Mohammadi, A. Alfantazi, An electrochemical investigation on the
nch 0.66 0.65 0.64 effect of the chloride content on CO2 corrosion of API-X100 steel, Corros. Sci. 64
Rch (O cm2) 18,510 20,910 22,380 (2012) 37–43.
Yfilm (S sn/cm2) 1.41E04 1.21E04 9.25E05 [7] E. Shashi Menon, Pipeline Planning and Construction Field Manual, Gulf
Professional Publishing, USA, 2011.
nfilm 0.77 0.83 0.76
[8] G. Papadakis, Major hazard pipelines: a comparative study of onshore
Rfilm (O cm2) 155.5 149.50 182.9
transmission accidents, J Loss Prevent Proc 12 (1999) 91–107.
Chi-square 6.36E05 4.78E05 5.37E05
[9] H. Mitsui, R. Takahashi, H. Asano, N. Taniguchi, M. Yui, Susceptibility to stress
corrosion cracking for low-carbon steel welds in carbonate-bicarbonate
solution, Corrosion 64 (2008) 939–948.
[10] G. Zhang, Y. Cheng, Micro-electrochemical characterization and Mott-Schottky
4. Conclusion analysis of corrosion of welded X70 pipeline steel in carbonate/bicarbonate
solution, Electrochim. Acta 55 (2009) 316–324.
[11] A. Mustapha, E. Charles, D. Hardie, Evaluation of environment-assisted
This research presented an electrochemical study on the corro- cracking susceptibility of a grade X100 pipeline steel, Corros. Sci. 54 (2012) 5–9.
sion behavior of a number of HAZs, made of API-X100 pipeline [12] F. Eliyan, E. Mahdi, A. Alfantazi, Electrochemical evaluation of the corrosion
steel, in naturally aerated, 5 g/L NaHCO3 solutions at 90 °C. It ad- behaviour of API-X100 pipeline steel in aerated bicarbonate solutions, Corros.
Sci. 58 (2012) 181–191.
dressed on the corrosion susceptibility and passivation stability [13] F. Mohammadi, F. Eliyan, A. Alfantazi, Corrosion of simulated weld HAZ of API
of the HAZs, of a welded pipeline steel, whose microstructures X-80 pipeline steel, Corros. Sci. 63 (2012) 323–333.
are primarily affected by the cooling rates. The HAZs were simu- [14] F. Eliyan, E. Mahdi, Z. Farhat, A. Alfantazi, Interpreting the passivation of HSLA
steel from electrochemical corrosion investigations in bicarbonate-oil aqueous
lated by GleebleÓ thermal cycles by which the samples were
emulsions, Int. J. Electrochem. Sci. 8 (2013) 3026–3038.
cooled with controlled rates of 10, 30, and 60 °C/s from a peak [15] A. Otazo, PDVSA’S possible sale of CITGO and the subsequent ramifications on
950 °C. The slow potentiodynamic scans revealed the important US foreign and energy policy, Energy Policy 42 (2012) 89–94.
[16] M. Ozturk, Y. Yuksel, N. Ozek, A bridge between east and west: turkey’s natural
roles of ferrite, in the 10 °C/s HAZ, and acicular ferrite and martens-
gas policy, Renew. Sustain. Energy Rev. 15 (2011) 4286–4294.
ite in the 30 °C/s and 60 °C/s HAZs, respectively with the passiv- [17] A. Kemp, An assessment of UK North Sea oil and gas policies Twenty-five years
ation process and the cathodic currents. The 10 °C/s HAZ and on, Energy policy 18 (1990) 599–623.
as-received samples exhibited the lowest passive current densities, [18] C. Rangel, R. Leitão, Voltammetric studies of the transpassive dissolution of
mild steel in carbonate/bicarbonate solutions, Electrochim. Acta 34 (1989)
while those of the 30 and 60 °C/s HAZs showed an evidence of thin, 255–263.
unstable passive films. Oxygen reduction currents intersected with [19] D. Davies, G. Burstein, Effects of bicarbonate on the corrosion and passivation
the passivation regimes, and appeared as broad cathodic peaks. A of iron, Corrosion 36 (1980) 416–422.
[20] C. Valentini, C. Moina, The electrochemical behaviour of iron in stagnant and
short review on the controversy in the literature on the mecha- stirred potassium carbonate-bicarbonate solutions in the 0–75 °C temperature
nisms of anodic dissolution, and the active–passive transition, range, Corros. Sci. 25 (1985) 985–997.
which were independent from a HAZ microstructure in this study, [21] E. Castro, C. Valentini, C. Moina, J. Vilche, A. Arvia, The influence of ionic
composition on the electrodissolution and passivation of iron electrodes in
was carried out. The 30 °C/s HAZ, notably, exhibited the minimum potassium carbonate-bicarbonate solutions in the 8.4–10.5 pH range at 25 °C,
cathodic current densities. The potentiostatic currents of the 30 Corros. Sci. 26 (1986) 781–793.
and 60 °C/s HAZs, at 0.5 V vs. SCE, suggested a slow growth of [22] Z. Lu, C. Huang, D. Huang, W. Yang, Effects of a magnetic field on the anodic
dissolution, passivation and transpassivation behaviour of iron in weakly
the passive films which possibly underwent repetitive breakdowns
alkaline solutions with or without halides, Corros. Sci. 48 (2006) 3049–3077.
and repassivations. By measuring the impedance over after 3000, [23] S. Simard, M. Drogowska, H. Menard, Electrochemical behavior of 1024 mild
6000, and 9000 s of immersion at the open circuit potentials, an steel in slightly alkaline bicarbonate solutions, J. Appl. Electrochem. 27 (1997)
317–324.
electrochemical evidence of the growth of the passive films was
[24] L. Niu, Y. Cheng, Corrosion behaviour of X-70 pipe steel in near-neutral pH
found. Taking into account the roles of the porous Fe3C, and the solution, Appl. Surf. Sci. 253 (2007) 8626–8631.
changing thickness of the passive films, a set of equivalent circuits [25] S. Hirnyi, Anodic hydrogenation of iron in a carbonate-bicarbonate solution,
and interfaces were proposed to fit the multi-time-constant EIS Mater. Sci. 37 (2001) 491–498.
[26] M. Sazzadur Rahman, S. Divi, D. Chandra, J. Daemen, Effect of different salts on
data, in reference to a HAZ microstructure. The charge-transfer the corrosion properties of friction type A607 steel rock bolt in simulated
resistance decreased with time as a result of the growing passive concentrated water, Tunn. Undergr. Sp. Technol. 23 (2008) 665–673.
films, but the 30 °C/s HAZ showed an intrinsic high charge-transfer [27] J. Gonzalez-Rodriguez, M. Espinosa-Medina, C. Angeles-Chavez, T. Zeferino-
Rodriguez, SCC of X-52 and X-60 weldments in diluted NaHCO3 solutions with
resistance, confirming with the potentiodynamic polarization chloride and sulfate ions, Mater. Corros. 58 (2007) 599–603.
results. [28] E. Ramírez, J. González-Rodriguez, A. Torres-Islas, S. Serna, B. Campillo, G.
Dominguez-Patiño, J. Juárez-Islas, Effect of microstructure on the sulphide
stress cracking susceptibility of a high strength pipeline steel, Corros. Sci. 50
Acknowledgment (2008) 3534–3541.
[29] C. Kwok, F. Cheng, H. Man, Microstructure and corrosion behavior of laser
surface-melted high-speed steels, Surf. Coat. Technol. 202 (2007) 336–348.
This publication was made possible by NPRP grant # 09-211-2- [30] A. Fu, Y. Cheng, Electrochemical polarization behavior of X70 steel in thin
089 from the Qatar National Research Fund (a member of Qatar carbonate/bicarbonate solution layers trapped under a disbonded coating and
its implication on pipeline SCC, Corros. Sci. 52 (2010) 2511–2518.
Foundation). The statements made herein are solely the responsi-
[31] R. Parkins, S. Zhou, The stress corrosion cracking of C–Mn steel in
bility of the authors. CO2  HCO 2
3  CO3 solutions. II: Electrochemical and other data, Corros. Sci.
39 (1997) 175–191.
[32] J. von Fraunhofer, The polarization behaviour of mild steel in aerated and de-
References aerated 1M NaHCO3, Corros. Sci. 10 (1970) 245–251.
[33] Y. Tang, Y. Zuo, The metastable pitting of mild steel in bicarbonate solutions,
[1] J. Bockris, D. Drazic, A. Despic, The electrode kinetics of the deposition and Mater. Chem. Phys. 88 (2004) 221–226.
dissolution of iron, Electrochim. Acta 4 (1961) 325–361. [34] C. Lee, Z. Qin, M. Odziemkowski, D. Shoesmith, The influence of groundwater
[2] C. de Waard, D. Milliams, Carbonic acid corrosion of steel, Corrosion 31 (1975) anions on the impedance behaviour of carbon steel corroding under anoxic
177–181. conditions, Electrochim. Acta 51 (2006) 1558–1568.
[3] G. Schmitt, B. Rothman, Untersuchungen zum Korrosionsmechanismus von [35] G. Schmitt, M. Mueller, M. Papenfuss, E. Strobel-Effertz, Understanding
unlegiertem Stahl in sauerstofffreien Kohlensäurelösungen. Teil I. Kinetik der localized CO2 corrosion of carbon steel from physical properties of iron
Wasserstoffabscheidung, Werkst. Korros. 28 (1977) 816–822. carbonate scales, CORROSION/99, Paper No. 38, NACE, Houston, TX, 1999.
F.F. Eliyan, A. Alfantazi / Corrosion Science 74 (2013) 297–307 307

[36] S. Al-Hassan, B. Mishra, D. Olson, M. Salama, Effect of microstructure on [42] V. Alves, C. Brett, Characterisation of passive films formed on mild steels in
corrosion of steels in aqueous solutions containing carbon dioxide, Corrosion bicarbonate solution by EIS, Electrochim. Acta 47 (2002) 2081–2091.
54 (1998) 480–491. [43] D. Li, Y. Feng, Z. Bai, J. Zhu, M. Zheng, Influence of temperature, chloride ions
[37] A. Dugstad, Mechanism of protective film formation during CO2 corrosion of and chromium element on the electronic property of passive film formed on
carbon steel, CORROSION/98, Paper No. 31, NACE, Houston, TX, 1998. carbon steel in bicarbonate/carbonate buffer solution, Electrochim. Acta 52
[38] K. Juttner, Electrochemical impedance spectroscopy (EIS) of corrosion (2007) 7877–7884.
processes on inhomogeneous surfaces, Electrochim. Acta 35 (1990) 1501– [44] L. Hamadou, A. Kadri, N. Benbrahim, Characterisation of passive films formed
1508. on low carbon steel in borate buffer solution (pH 9.2) by electrochemical
[39] B. Boukamp, Equivalent Circuit User’s Manual, second ed., University of impedance spectroscopy, Appl. Surf. Sci. 252 (2005) 1510–1519.
Twente, Einshede, 1989. [45] J. Zhou, X. Li, C. Du, Y. Pan, T. Li, Q. Li, Passivation process of X80 pipeline steel
[40] F. Mansfeld, Recording and analysis of AC impedance data for corrosion in bicarbonate solutions, Int. J. Min. Met. Mater. 18 (2011) 178–185.
studies, Corrosion 36 (1981) 301–307. [46] Z. Qin, B. Demko, J. Noel, D. Shoesmith, F. King, R. Worthingham, K. Keith,
[41] B. Kinsella, Y. Tan, S. Bailey, Electrochemical impedance spectroscopy and Localized dissolution of millscale-covered pipeline steel surfaces, Corrosion 60
surface characterization techniques to study carbon dioxide corrosion product (2004) 906–914.
scales, Corrosion 54 (1998) 835–842.

You might also like