Вы находитесь на странице: 1из 9

IBP1390_11

AERATED CORROSIO BEHAVIOR OF API-X100


PIPELIES STEEL AT HIGH TEMPERATURE –
ELECTROCHEMICAL IVESTIGATIO
Faysal Fayez M. Eliyan1, Akram M. Alfantazi 1

Copyright 2011, Brazilian Petroleum, Gas and Biofuels Institute - IBP


This Technical Paper was prepared for presentation at the Rio Pipeline Conference & Exposition 2011, held between September,
20-22, 2011, in Rio de Janeiro. This Technical Paper was selected for presentation by the Technical Committee of the event. The
material as it is presented, does not necessarily represent Brazilian Petroleum, Gas and Biofuels Institute’ opinion or that of its
Members or Representatives. Authors consent to the publication of this Technical Paper in the Rio Pipeline Conference &
Exposition 2011.

Abstract
The aim of this experimental work is to provide an insight to the main electrochemical perspectives related to carbon
dioxide corrosion behavior of API-X100 pipeline steel at high temperatures. The corrosion performance is investigated
in laboratory conditions simulating the extensive dissolution of carbon dioxide in water cuts at the early stages of
transmission where considerable amounts of oxygen is entrapped. Open Circuit Potentials (OCP), potentiodynamic
polarization, as well as Electrochemical Impedance Spectroscopy are all utilized in the electrochemical analysis. The
tests are performed in 0.05, 0.1, 0.5, and 1 mol/L bicarbonate solutions containing 3 wt% NaCl at 40, 60, and 80 oC in
conditions facilitating good oxygen diffusivity. OCP values were mainly influenced by the cathodic reduction of oxygen
and water and got nobler with increased bicarbonate content. Additionally, corrosion rates increased generally with
temperature showing sole anodic peaks before the extensive dissolution. The equivalent circuits proposed for the
electrochemical systems encountered achieved very good agreement with the experimental results as well as with the
polarization characteristics.

______________________________
1
Dept. Mat. Eng., Univ. of British Columbia, 6350 Stores Road, Vancouver, BC V6T 1Z4, Canada
Rio Pipeline Conference & Exposition 2011

1. Introduction

Carbon dioxide corrosion mechanism in oil carrying facilities has been reported to be considerably complex and the
different models proposed in the literature were valid for specific cases as discussed in [Schmitt, 1983] and [De Waard,
1975]. However, it is generally agreed that the following chemical reactions are the initial steps prior to the
electrochemical chain of carbon dioxide corrosion represented by equations (1), (2), and (3) with the thermodynamic
constants (pKeq) at 25 oC as reported in [Nesic, 2001], [George, 2007] and [Ogundle, 1986] as:
CO2 (g) ⇔ CO2 (aq) ; pKeq = 1.468 (1)
CO 2 (g) + H 2 O ⇔ H 2 CO 3 ; pKeq = 2.588 (2)
+ −
H 2 CO 3 ⇔ H + HCO 3 ; pKeq = 3.5 (3)
Carbon dioxide dissolves in the formation-water producing the weak carbonic acid, which in turn dissociates partially to
generate bicarbonate anions. In fact, bicarbonate species have been reported to be the key species involved in the anodic
dissolution and/or cathodic reduction reactions. In addition, carbon dioxide corrosion rate determining steps in mildly
alkaline media are governed by the bicarbonate species where they are reduced directly to produce adsorbed hydrogen
atoms and/or hydrogen gas as pointed out in [6] represented by equations (4), (5) and (6):
HCO 3- + e − ⇒ H ads + CO 32 − (4)
− 2−
HCO + H ads + e ⇒ H 2 + CO
-
3 3 (5)
− 2−
2HCO + 2e ⇒ H 2 + 2CO
-
3 3 (6)
Bicarbonate species at some concentrations can promote corrosion rates; however, they may also have an inhibiting
effect depending on the pH value. When pH increases, bicarbonate anions facilitate the formation of a protective layer
of iron carbonate (FeCO3) and other iron oxides, however, the corrosion rates increase when bicarbonate species are
involved in the cathodic reactions [Nazari, 2010]. It has been proposed that the cathodic reactions involving carbon
containing species have a noteworthy effect on the corrosion rates which are accelerated by increased carbon dioxide
partial pressures [Ikeda, 1984]. Zhang et al. [Zhang, 2006] found that cathodic reaction rates increase continuously with
increased concentrations of bicarbonate showing also relatively equal anodic dissolution rates before a single anodic
peak appeared when the bicarbonate species concentration was about 0.125 mol/L indicating the formation of a passive
layer. Similarly, Videm et al. [Videm, 1993] found that current densities in the active region increase proportionally with
bicarbonate concentrations; however, the maximum current density prior to passivation occurs in 0.01 mol/L
bicarbonate solution which was of the most dilute among all other test solution tests.
Temperature influences the ability of bicarbonate anions to increase anodic reaction rates. It has been reported that
anodic dissolutions become diffusion controlled due to an increase in bicarbonate concentration leading to complex ions
2- +
like Fe(CO 3 ) 2 and Fe(HCO 3 ) being formed at high temperatures [Davis, 1980] and [Castro, 1991]. In fact, these
ions are involved in the multistep reactions where iron bicarbonate and/or iron carbonate are finally formed. This
process however occurs away from the steel surface, but when the temperature is higher, bicarbonate ions facilitate the
formation of solid films on steel surfaces where scale thickness and grain size increase as found in [Lin, 2006], [Li,
1998], and [Savoye, 2001].
In the present experimental work, the corrosion behaviour of API-X100 pipeline steel has been investigated in the
regime where the influence of bicarbonate anions from thermodynamic and kinetic perspectives varies. Variation with
time of the open circuit potentials as well as anodic and cathodic characteristics may provide insight into the
predominant roles that bicarbonate anions play. Additionally, the effect of chloride anions along with temperature
changes in the established naturally aerated conditions may provide more information about the onset of passivation and
stability of passive films.

2. Experimental Details
Test samples were cut from an oil pipeline made from API-X100 steel and machined into disks having the nominal
dimensions of 14 mm diameter and 4 mm thickness. They were then connected to the electrochemical setup via copper
wires attached to the samples by conductive silver paste. Then, they were mounted in hard, cold-curing epoxy resins.
Prior to each corrosion test, they were sequentially wet-ground with 120 grit, 320 grit, and 600 grit silicon carbide (SiC)
emery papers and then degreased ultrasonically with ethyl 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 to G1-03 ASTM standard [ASTM, 2004]. Corrosion behaviour is investigated in naturally aerated

2
Rio Pipeline Conference & Exposition 2011

electrolytes synthesized from analytical grade Fisher procured reagent of sodium bicarbonate of concentrations of 0.05,
0.1, 0.5, and 1 mol/L and double distilled deionized water with the addition of 3 wt % NaCl at 40, 60, and 80 oC.
Corrosion tests were performed in a jacket glass test cell of a total volume of 0.6 L, open to the atmosphere. The
working electrode was the studied material, the counter electrode was made from graphite, and the measured potentials
were in reference to the Saturated Calomel Electrode (SCE) of +0.241 VSHE. Open circuit potentials were measured in
bicarbonate solutions at different temperatures attaining the final OCP values after about 7100 s when the free potential
variations were within ± 0.01 mVSCE/s. The potentiodynamic polarization measurements were carried out at a scan rate
of 0.50 mVSCE/s. Electrochemical Impedance Spectroscopy (EIS) technique was then applied to study the
electrochemical systems established at the stabilized OCP states. The frequency range considered was from 0.01 to
10,000 Hz with a sampling rate of 10 points per decade. The microstructural features, examined by optical microscopy,
are shown in Figure 1 comprising of acicular ferrite and bainite.

Figure 1. Optical microscopy image of the microstructure of the examined test material

3. Results and Discussion

3.1. Open Circuit Potential (OCP) Variations

OCP variations were monitored to account first for the changes that mixed potentials could undergo before other
electrochemical aspects are investigated. OCP profiles were taken within test time periods of about 7100 seconds at
which relatively stable OCP’s are achieved with minimal fluctuations. OCP in bicarbonate containing solutions at the
corresponding temperatures are shown selectively for 0.1, 0.5, and 1 mol/L in Figure 2.

Figure 2. OCP variations with temperature in the aerated bicarbonate solutions

3
Rio Pipeline Conference & Exposition 2011

Apparently, OCP exhibited an increase with the bicarbonate content at all temperatures. As discussed in the polarization
test results section, the accelerated bicarbonate content – dependent cathodic reactions could be greatly responsible for
that behavior where the higher temperatures exerted also the same influence in these mildly alkaline media. In addition,
this temperature-dependent behavior suggests that the cathodic reactions, rather than the anodic ones, are more sensitive
to the higher temperatures [Brossia, 2000].
Interestingly, the band of variation between the highest and lowest OCP values at 0.1 and 1 mol/L respectively
broadened with the higher temperatures. The anodic dissolutions are proposed to involve hydroxyl (OH-) forming iron
(II) hydroxide (Fe(OH)2) in these mildly alkaline media [Fang, 2006] as shown in equation (7) as:
Fe + 2OH - → Fe(OH)2 + 2e − (7)
The cathodic branch is considered to involve the simultaneous reduction of bicarbonate and dissolved oxygen,
represented by equation (10). Bicarbonate is reduced substantially in these conditions, [Ahmad, 2006] and [Paoinelli,
2008] represented by equations (8) or (9) as:

HCO 3- + e - → CO 32- + 1/2H 2 (8)


HCO + e → CO + H ads
-
3
- 2-
3 (9)
1/2 O 2 + H 2O + 2e - → 2OH - (10)

The standard half cell potentials are calculated from the standard Gibbs free energy [Dean, 2000] of the species involved
in the corrosion reactions and, by utilizing Nernst equation [Vaidva, 2007], the theoretical anodic and cathodic potential
limits are calculated taking into account temperature, pH, and bicarbonate concentration. OCP variations and the
associated theoretical potential limits are selectively presented for chloride free at 60 oC in Table 1. Apart from the
theoretical assumptions, the accelerated bicarbonate content-dependent cathodic reactions seemed to make OCP more
sensitive to the cathodic reduction of dissolved oxygen, existing appreciably with the same amount in these bicarbonate
solutions, with the absence and presence of chloride [Stansbury, 2000].

Table 1. Open Circuit Potentials (OCP) and the calculated half cell equilibrium potentials at 60.

Reaction
Reaction (7) equilibrium Reaction (9) equilibrium (10)
Temperature HCO 3- OCP
pH potential potential equilibrium
concentration potential
(oC) (mol/L) (VSCE) (VSCE) (VSCE) (VSCE)
0.1 7.7 -0.751 -0.968 -0.762 0.506
60 0.5 7.9 -0.743 -0.974 -0.763 0.519
1 8.4 -0.67 -0.988 -0.768 0.526

3.2. Potentiodynamic polarization results

The potentiodynamic polarization profiles at 60 oC, are shown in Figure 3 for 0.05, 0.5, and 0.1 mol/L bicarbonate
concentrations. Similar polarization behaviours were exhibited at 40 and 60 oC at the characteristic regions. It is
apparent that the anodic and cathodic current densities accelerated with the increased bicarbonate contents. Kinetically,
corrosion current density (icorr), and anodic and cathodic Tafel slopes (βa), (βc) are iteratively, with the best fit with the
experimental data, calculated from Butler-Erdey-Volmer (BEV) equation for the purely charge-transfer controlled
polarization [23] as:
  2.3ηs   - 2.3ηs  
i = i corr  exp   − exp    (11)
 β β
  a   c 

The corrosion rate, or (icorr), was proportional with the bicarbonate content and confirming with OCP findings, the
corrosion potential (Ecorr) increased accordingly at all temperatures as illustrated in Table 2. These two preliminary
trends suggest the considerable sensitivity of the corrosion reactions to the cathodic evolution of hydrogen generated
from electrochemical bicarbonate consumption. Interestingly, the higher temperatures seemed to induce a similar
electrochemical influence where the corrosion rates progressively increased as did the corrosion potentials accompanied

4
Rio Pipeline Conference & Exposition 2011

by a noticeable decrease of ( β c ) with slightly appreciable changes in ( β a ). Additionally, the influence of bicarbonate
content on the accelerated corrosion reactions was less significant as the temperature increased from 40 to 80 oC where
the acceleration factor was almost 4:1 respectively.

Figure 3. Potentiodynamic polarization profiles from 0.05, 0.5, and 1 M conditions at 60 oC

Starting with the active dissolution regime, an evidence for a multi-step corrosion mechanism appeared where two
anodic slopes resulted, more noticeably with the concentrated bicarbonate solutions, in the pre-passivation range.
Dissolution reactions showed a retardation at an anodic transition overpotential of about -550 mVSCE appreciably
irrespective from temperature where the consecutive pre-passive slopes increased. This behaviour, in these mildly
alkaline media, is very attributable to an introduced physical influence at the active corrosion interface [Talbot, 1998].

Table 2. Potentiodynamic polarization test results at 40, 60, and 80 oC


Bicarbonate
Temp concentration Ecorr icorr βa βc ipass Epass
o 2 2
( C) (mol/L) (mV) (µA/cm ) (mV /d) (mV /d) (µA/cm ) (mV)
0.1 -766 9.6 58 56 4.2 -603
40 0.5 -754 20 53 46 9.3 -538
1 -743 45 52 55 13.8 -470
0.05 -753 33 51 48 11.5 -667
60 0.5 -748 47 53 44 41 -652
1 -739 63 55 42 43.4 -624
0.05 -750 56 51 45 14.2 -662
80 0.5 -743 62 52 43 43.1 -658
1 -735 73 49 40 53.7 -655

5
Rio Pipeline Conference & Exposition 2011

Hydroxyl (OH-) species, getting involved in the dissolution mechanism, could contribute to this active behaviour in our
case where a defective iron hydroxide passive film resulted. Despite of the possibly defective nature [Zhang, 2009] of
this hydrous film, it was capable to interfere with a decreased rate of anodic dissolution.
The formation of such a hydroxide-based film [26] in a short range of potential in Pourbiax diagram of Fe-H-C-O
[Himyi, 2001] is illustrated as:
Fe + OH - ↔ FeOH -ads (12)
FeOH -ads → FeOH ads + e - (13)
+
FeOH ads → FeOH ads +e -
(14)
+
FeOH ads + OH - → Fe(OH) 2 → Hydrous [Fe(OH) 2 ] (15)
Iron carbonate (FeCO3) can be consequently incorporated with the preliminary formed film with sufficient bicarbonate
contents but at a different transition point where the electrochemical retardation became temporary at potentials, in our
case, between 100 to 150 mV above the previous transition. It is shown below in Figure 4 the corrosion product
morphology after the polarization showing dispersed oxide colonies with localized pitting spots in 1 M bicarbonate
condition at 60 oC.

Figure 4. Corrosion products morphology after the polarization test in 1 M bicarbonate condition at 60 oC

3.3. Electrochemical Impedance Spectroscopy (EIS) results

Nyquist impedance representation plots for chloride containing test solutions are shown in Figure 5 for all bicarbonate
and temperature conditions established. The relatively similar profiles seemed to indicate the significance of kinetics as
well as parallel diffusion-limited processes across possibly effective passive films. More specifically, the high-frequency
capacitive arc expands at different extents, in greater association with the bicarbonate content, to produce a second
capacitive semicircle. The other part at low and medium-frequency regions flattens to indicate a diffusion-influenced
process represented by the diffusion parameter (YW). In comparison with the polarization results, Nyquist semicircles
achieved a clear agreement in terms of temperature where the charge transfer resistance and the other associated
resistances were higher at lower temperatures; i.e. with larger Nyquist semicircles. The medium-frequency capacitive
arcs, produced where the charge transformation of electrochemical process is controlled in parallel with other diffusion-
limited process, are related in most cases to localized active sites where the corrosion rates are high [28]. Although that
in conditions where carbon carrying species dominating the electrochemical processes could induce a separate time
constant, where the significance of adsorption is specified for [Lopez, 2003],[Hong, 2000], adsorption and/or relaxation

6
Rio Pipeline Conference & Exposition 2011

of reaction intermediates of these species were excluded. The passive film formation, which was facilitated with greater
amounts of bicarbonate, induced the prevalent electrochemical influence over any other associated processes which
might occur when no effective passivation is established. Therefore, the similar electrochemical interactions seemed to
be fundamentally a function of the passive film compactness whose properties are, for example, intrinsically related to
temperature.

Figure 5. EIS profiles in 0.1, 0.5, and 1 mol/L conditions at 40, 60, and 80

4. Conclusion

This electrochemical investigation performed on the corrosion behavior of API-X100 pipeline steel in the aerated
bicarbonate test solutions at 40, 60, and 80 oC provided a new understanding on the general electrochemical
considerations that corroded pipelines could be exposed to. In the mildly unbuffered alkaline media established, the
corrosion rates showed an increasing corrosion rate with the increased bicarbonate contents and with higher
temperatures. In addition, anodic peaks appeared and passive current densities in the passive regions in a rate depending
on the bicarbonate accordingly. EIS findings as well as that of OCP were confirmative with the polarization findings
from either potential or current perspectives. The free potential was nobler with the increased bicarbonate content and
temperature and the charge transfer resistance as well as that associated with the passive film increased accordingly.

7. Acknowledgements
The authors would like to thank Qatar National Research Fund (QNRF) and National Science and Engineering
Research Council (NSERC) for providing the financial support for this published work.

8. References
Ahmed Z., Principles of corrosion engineering and corrosion control, first ed., Butterworth-Heinemann, 2006.
7
Rio Pipeline Conference & Exposition 2011

ASTM Standard G 1-03, Standard practice for preparing, cleaning, and evaluating corrosion test specimens, Annual
Book of Standards, ASTM, Pennsylvania, 2004.

Brossia C., and Cragnolino G., Effect of Environmental Variables on Localized Corrosion of Carbon Steel, Corrosion.
56 (2000) 505-514.

Castro E., and Vilche J., Electrooxidation/electroreduction processes at composite iron hydroxide layers in carbonate-
bicarbonate buffers, Journal of Applied Electrochemistry. 21 (1991) 543-551.

Castro E., Vilche J., Arvia A., Iron dissolution and passivation in K2CO3-KHCO3 solutions; Rotating ring disc electrode
and XPS studies, Corrosion Science. 32 (1991) 37-50.

Davis D., Burstein G., The Effect of bicarbonate on the corrosion and passivation of iron, Corrosion. 36 (1980) 416-
422.

De. Waard C., Milliams D., Prediction of carbonic acid corrosion in natural gas pipelines, First international
conference for internal and external protection of pipes. Paper no. F1, University of Durham, 1975.

Dean J., Lange’s Handbook of Chemistry, fifteenth ed., McGraw-Hill, New York, 2000.

Fang H., Nesic S., and Brown B., General CO2 corrosion in high salinity brines, CORROSION/2006. Paper no. 6372,
NACE, 2006.

George K., Nešić S., Investigation of carbon dioxide corrosion of mild steel in the presence of acetic acid-part 1: basic
mechanisms, Corrosion. 63 (2007) 178-186.

Hirnyi S., Anodic hydrogenation of iron in a carbonate-bicarbonate solution, Materials Science. 37 (2001) 491-498.

Hong T., Shi H, Wang H., Gopal M. and Jepson W. P., EIS study of corrosion product film in pipelines,
CORROSION/2000. Paper no. 44, NACE, 2000.

Ikeda A., Ueda M., Mukai S., CO2 behaviour of carbon and Cr steels, Advances in CO2 Corrosion. Paper no. 5, NACE,
1984.

Li J., Meier D., An AFM study of the properties of passive films on iron surfaces, Journal of Electro analytical
Chemistry. 454 (1998) 53-58.

Lin G., Zheng M., Bai Z., Zhao X., Effect of temperature and pressure on the morphology of carbon dioxide corrosion
scales, Corrosion. 62 (2006) 501-507.

Lo´pez D., Pe´rez T., Simison S., The influence of microstructure and chemical composition of carbon and low alloy
steels in CO2 corrosion. A state-of-the-art appraisal, Materials and Design. 24 (2003) 561–575.

Nazari M., Allahkaram S., Kermani M., The effects of temperature and pH on the characteristics of corrosion product in
CO2 corrosion of grade X70 steel, Materials and Design. 31 (2010) 3559-3563.

Nešic´S., Nordsveen M., Nyborg R., Stangeland A., A mechanistic model for CO2 corrosion with protective iron
carbonate films, CORROSION/01, Paper no. 40, NACE, 2001.

Ogundele G., and White W., Observations on the influence of dissolved hydrocarbon gases and variable water
chemistries on corrosion of an API-L80 steel, Corrosion. 43 (1987) 665-673.

Ogundle G., White W., Some observations on corrosion of carbon steel in aqueous environments containing carbon
dioxide, Corrosion. 42 (1986) 71-82.

Paolinelli L., Pérez T., and Simiso S., The effect of pre-corrosion and steel microstructure on inhibitor performance in
CO2 corrosion, Corrosion Science. 50 (2008) 2456–2464.
8
Rio Pipeline Conference & Exposition 2011

Savoye S., Legrand L., Sagon G., Lecomte S., Chausse A., Messina R., Toulhoat P., Experimental investigations on iron
corrosion products formed in bicarbonate/carbonate – containing solutions at 90 oC, Corrosion Science. 43 (2001) 2049-
2064.

Schmitt G., Fundamental aspects of CO2 corrosion, CORROSION/83. Paper no. 3, NACE, 1983.

Stansbury E.E., and R.A. Buchanan, Fundamentals of electrochemical corrosion, ASM International, USA, 2000.

Talbot D., and J. Talbot, Corrosion science and technology, second ed., CRC Press LLC, 1998.

Vaidya R., Using electrochemical monitoring to predict metal release in drinking water distribution systems, PhD.
Thesis, University of Central Florida, (2007).

- -
Videm K., Koren A., Corrosion, passivity, and pitting of carbon steel in aqueous solutions of HCO 3 , CO 2 , and Cl ,
Corrosion. 49 (1993) 746-754.

Zhang G., Lu M., Chai C., Wu Y., Effect of HCO 3- concentration on CO2 corrosion in oil and gas fields, Materials. 13
(2006) 44-49.

Zhang L., Li X., Du C., and Cheng Y., Corrosion and stress corrosion cracking behaviour of X70 pipeline steel in a
CO2-containing solution, Journal of Materials Engineering and Performance. 18 (2009) 319-323.

Вам также может понравиться