Sunteți pe pagina 1din 12

Corrosion Science 59 (2012) 186–197

Contents lists available at SciVerse ScienceDirect

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

Discussion of the CO2 corrosion mechanism between low partial pressure and
supercritical condition
Yucheng Zhang a, Xiaolu Pang a, Shaopeng Qu a, Xin Li b,c, Kewei Gao a,⇑
a
Department of Materials Physics and Chemistry, University of Science and Technology Beijing, Beijing 100083, China
b
China Petroleum Pipeline Coating Engineering Co., Ltd., Langfang 065000, Hebei, China
c
Department of Materials Science and Engineering, China University of Petroleum, Beijing 102249, China

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

Article history: The corrosion behaviour of X65 pipeline steel at various temperatures for different immersion time under
Received 16 September 2011 low CO2 partial pressure and supercritical CO2 condition were investigated by weight loss measurements
Accepted 1 March 2012 and surface analysis techniques. By comparing the characteristics of CO2 corrosion product scale formed
Available online 9 March 2012
under experimental conditions and the variation rule of corrosion rate with temperature, the CO2 corro-
sion mechanism under low partial pressure and supercritical condition was studied. To explain the big
Keywords: difference of corrosion rate between low CO2 partial pressure and supercritical CO2 condition, thermody-
A. Steel
namic calculation of the solubility of CO2 in H2O were discussed.
B. SEM
B. XRD
Ó 2012 Elsevier Ltd. All rights reserved.
B. Weight loss
C. Acid corrosion

1. Introduction 1710
log V corr ¼ 5:8  þ 0:67 logðpCO2 Þ ð1Þ
T
Low cost carbon steels are often used as construction material
where, Vcorr is the corrosion rate of carbon steels, mm/y; T is tem-
in oil producing and transportation pipelines, which are, however,
perature, K; pCO2 is the CO2 partial pressure, Pa.
very susceptible to corrosion in CO2 environments [1]. CO2 in an
This simple model which only includes CO2 partial pressure and
aqueous environment is a corrosive species that has caused several
temperature has been revised by de Waard and Milliams in 1991
failures related to the oil and gas pipelines as well as equipments
[15], 1993 [16] and 1995 [17] by introducing parameters such as
[2,3]. The presence of CO2 causes ‘‘sweet’’ corrosion due to the
pH, water chemistry, scaling tendency, total pressure, hydraulic
formation of corrosive carbonic acid, H2CO3, which significantly
diameter, fluid flow velocity, oil wetting, glycol concentration
increases the corrosion rates of carbon steels. Under suitable con-
and steel composition. According to de Waard–Milliams model,
ditions, corrosion products can form scales (FeCO3 or other iron
the corrosion rate of carbon steel increases with an increase of
compounds) on the corroded surface [4,5]. The corrosion rate of
CO2 partial pressure and the predicted corrosion rate is well accord
carbon steels in CO2 environments depends, to a great extent, on
with the laboratory and field data ranging from very low pressures
the environmental conditions involved, such as temperature, pres-
up to 1 MPa.
sure, pH, solution chemistry, water/oil ratio, flow, steel type, corro-
After de Waard–Milliams model, several oil companies and re-
sion inhibitor and the characteristics of the surface scales [6–12].
search institutions contributed with a great number of prediction
In order to describe the effect of these environmental conditions
models based either on empirical correlations with field and labo-
on the CO2 corrosion rate of carbon steels, a number of mathemat-
ratory data or on mathematical modeling considering chemical,
ical prediction models of CO2 corrosion related to experimental
electrochemical and transport processes. These models include
parameters were established. The first and most widely used mod-
NORSOK model, CORMED model, LIPUCOR model, HYDROCOR
el for predicting CO2 corrosion of carbon steels in wells and pipe-
model, KSC model, TULSA model, PREDICT model, SweetCor model,
lines was proposed by de Waard and Milliams in 1975 [13,14],
CORPOS model, OHIO model, ULL model, DREAM model, OLI model
shown as following [14]:
and FREECORP model, which have been described and discussed in
much detail in literature [18].
However, it should be noted that these models were developed
preferentially for service conditions in oil and gas production and
⇑ Corresponding author. Tel./fax: +86 10 62334909. pipeline transport, and therefore mainly focused on the corrosion
E-mail addresses: kwgao@yahoo.com, kwgao@mater.ustb.edu.cn (K. Gao). systems with CO2 partial pressures of 0–2 MPa. For high pressures,

0010-938X/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2012.03.006
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 187

these models are not applicable. For example, at 40 °C and 7.5– (precision 0.1 mg) and then stored in a desiccator until use. Before
9 MPa, the corrosion rates of X65 and 0.5Cr carbon steels in aque- the specimens were put into autoclave, they were fixed in a spec-
ous CO2 conditions ranged in the order 1–6 mm/y [19], however imen holder (made by polytetrafluoroethene (PTFE)), only one sur-
the corrosion rates of both steels under these conditions predicted face (10  10 mm2) was exposed to corrosion medium and the
by KSC model [20] and NORSOK model [21] were 10 and 17 mm/y, other surfaces were sealed by silicone rubber.
respectively. Thus, the corrosion rates measured in the experi- All corrosion experiments were carried out in a high tempera-
ments at high pressures were generally much lower than those ture and high pressure autoclave and the schematic experimental
predicted by the CO2 corrosion models. setup is shown in Fig. 1. The corrosion medium was distilled water
Possible explanations to the big difference between measured together with high purity (99.99%) CO2. Prior to the experiments,
and predicted corrosion rates of carbon steels could ascribe to a the water was deaerated by N2 bubbling for 4 h and then intro-
change in CO2 solubility with pressure and that liquid or supercrit- duced into the autoclave. The autoclave was deoxygenated again
ical CO2 interacts differently from gaseous CO2 with water [22]. In by purging with CO2 gas with a purity of 99.99% for 2 h and then
general, in the case of scale-free CO2 corrosion, an increase of CO2 heated to the required temperatures. After that, the autoclave
partial pressure typically leads to an increase in the corrosion rate, was pressed to the required pressures and statically kept for cer-
which has been verified by many predicted models. The commonly tain immersion time. In this study, the experiments under low
accepted explanation is that with an increase of pressure the CO2 partial pressure and supercritical CO2 condition are constant
concentration of H2CO3 increases and accelerates the cathodic at 1 and 9.5 MPa, respectively. For constant immersion time exper-
reactions, and ultimately enhances the corrosion rate [2]. Surpris- iments, the immersion time was 168 h and the temperatures were
ingly, when the CO2 partial pressure continues to increase up to 50, 80, 110 and 130 °C, respectively. For the tests with different
supercritical condition while the temperature and the pressure immersion time, the temperature was kept at 80 °C and the
are over 31.1 °C and 7.38 MPa, respectively, severe corrosion was immersion time was 0.5, 2, 7, 23, 48, 96 and 168 h, respectively.
encountered at carbon steels as well as corrosion resistant alloys All the experimental conditions in this study are listed in Table 1.
(more than 10 mm/y for carbon steels [23–26] and even After the experiments, all specimens were taken out and imme-
0.8 mm/y for 13Cr stainless steel [23]). This is due to the high diately cleaned by deionized water and acetone. To remove the CO2
corrosiveness of aqueous phases coexisting with supercritical corrosion product scale, the specimen was pickled in 10% hydro-
CO2, which has already been observed by many researchers chloric acid (HCl) inhibited with 10 g/L hexamethylenetetramine
[23–28]. (urotropine), rinsed subsequently in deionized water and acetone
It is common knowledge that supercritical CO2 corrosion exists and dried. Then, the weight loss of the specimen was measured
in tertiary oil and gas recovery like EOR (Enhanced Oil Recovery) to calculate the corrosion rate by the following equation [39]:
[29] or EGR (Enhanced Gas Recovery) [30] where injection of CO2
is applied to extend the life of oil and gas reserves. Furthermore, 8:76  104 Dm
V corr ¼ ð2Þ
the upcoming CCS (Carbon Capture and Storage) technologies S qt
[31] raises even more frequently the question of materials integrity
where Vcorr is the corrosion rate, mm/y; Dm is the weight loss, g; S is
under high temperature and high pressure CO2 conditions up and
the surface area of specimen, cm2; q is the density of test steels, g/
into the supercritical region [32]. So, in the implement of EOR,
cm3; t is the immersion time, h. The final value of the corrosion rate
EGR and CCS, supercritical CO2 corrosion problem has to be solved
in this study was averaged of three specimens.
urgently and more researches need to be carried out on supercrit-
After the experiments, the chemical composition and morphol-
ical CO2 corrosion.
ogy of CO2 corrosion product scale were analyzed by means of
The question arose is that why the present CO2 prediction mod-
Scanning Electron Microscopy (SEM), X-ray diffraction (XRD) and
els were widely accord with laboratory and field data at low pres-
Energy Dispersive X-ray Spectroscopy (EDS). As the CO2 corrosion
sures (usually below 1 MPa) but were not applicable at high
product scale is not electricity-conductive and also contains the
pressures? Also, why the supercritical CO2 corrosion caused more
severe corrosion to carbon steels than that at low pressures? This
may imply different CO2 corrosion mechanism between low CO2
partial pressure and supercritical CO2 conditions. Although the
CO2 corrosion mechanisms at CO2 partial pressures relevant for
the oil and gas transport (usually below 1 MPa) have been widely
studied [2,4,5,13,33–38], however, very few investigations so far
have been performed on the CO2 corrosion mechanism under high
pressures, especially supercritical CO2 conditions.
Thus, the objective in this work is to find the difference of CO2
corrosion mechanism between low partial pressure and supercrit-
ical conditions by comparing the CO2 corrosion behaviour of X65
pipeline steel at various temperatures and with different immer-
sion time under low CO2 partial pressure and supercritical CO2
conditions.

2. Experimental procedure

The material used in this study was X65 pipeline steel and its
chemical composition (wt.%) is 0.04 C, 1.5 Mn, 0.2 Si, 0.02 Mo
and Fe balance. The test specimens were machined to a size of
Fig. 1. Experimental setup used in this study 1: specimen; 2: specimen holder; 3:
10  10  3 mm3. Prior to the experiments, the surface of the spec- insulation jacket; 4: autoclave body; 5: heating jacket; 6: gas valve; 7: pressure
imen was ground with silicon carbide (SiC) paper progressively up gauge; 8: addition of brine and additives; 9: stirring motor; 10: thermocouple; 11:
to 800 grit. After drying in hot air, the specimen was weighed gas valve; 12: medium.
188 Y. Zhang et al. / Corrosion Science 59 (2012) 186–197

Table 1
Experimental conditions.

No. Pressure (MPa) Temperature (°C) Immersion time (h)


1 9.5 50 80 110 130 168 (0.5, 2, 7, 23, 48, 96)
a a a
1
a a a
2
a a a
3
a a a
4
a a a
5
a a a
6
a a a
7
a a a
8
a a
9–14 (a)
a a
15–20 (a)
a
For each immersion time, one experiment was carried out.

carbon element, a spray gold processing was made on the speci- iron was detected, while the diffraction peaks of FeCO3 appeared
men before SEM observations. After that, the specimen was sealed and substituted for that of pure iron after 48 h. It is noticed that
by epoxy, and then cut along the cross-section. The cross-sectional after 7 h immersion, only iron peaks from the substrate were de-
analysis was carried out to examine the cross-section morphology tected at low CO2 partial pressure, while the specimen had been
and the thickness of the CO2 corrosion product scale. covered by FeCO3 scale under supercritical CO2 condition.
The surface and cross-section morphologies of the corroded
specimens at 80 °C under low CO2 partial pressure and supercriti-
3. Results cal CO2 condition for different immersion time are shown in Figs. 3
and 4, respectively. By comparing the appearance of CO2 corrosion
3.1. CO2 corrosion behaviour of X65 steel at 80 °C with different product scale (shown in Figs. 3 and 4) with the XRD analysis
immersion time (shown in Fig. 2), it can be seen that the surface coverage by iron
carbonate scale increased with the increase of immersion time.
Fig. 2(a) and (b) shows the XRD spectra on the corroded surface In both conditions, there was no scale formed before 2 h, as shown
of X65 pipeline steel immersed in water saturated with CO2 under in Figs. 3(a) and (b) and 4(a) and (b). However, during 2–7 h, the
low partial pressure and supercritical condition at 80 °C for various scale formed under supercritical CO2 condition, while no scale
immersion time, respectively. There is no obvious difference ob- was found at low CO2 partial pressure, as shown in Figs. 3(c) and
served between these two conditions. In the first 2 h, only pure 4(c), which indicates that increasing pressure is more benefit to

Fig. 2. XRD analysis of CO2 corrosion product scales formed at 80 °C for various immersion time under low CO2 partial pressure (a) and supercritical CO2 condition (b).
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 189

a1 b1 c1 d1

20 µm 20 µm 20 µm 20 µm

a2 b2 c2 d2
A
A A
A B
C C
20 µm 20 µm C 20 µm C 20 µm

e1 f1 g1

20 µm 20 µm 20 µm

A e2 A f2 A g2
B
B B
C 20 µm C 40 µm 40 µm
C
Fig. 3. The surface and cross-section morphologies of CO2 corrosion product scales formed at low CO2 partial pressure at 80 °C for various immersion time (a: 0.5 h; b: 2 h; c:
7 h; d: 23 h; e: 48 h; f: 96 h; g: 168 h) (1: surface morphology; 2: cross-section morphology; A: epoxy; B: scale; C: steel substrate).

a1 b1 c1 d1

20 µm 20 µm 20 µm 20 µm

a2 b2 c2 d2
A A
A A
B
B
C C
20 µm 20 µm C 20 µm C 20 µm

e1 f1 g1

20 µm 20 µm 20 µm

A e2 A f2 A g2
B B
B
C
C 40 µm C 80 µm 80 µm

Fig. 4. The surface and cross-section morphologies of CO2 corrosion product scales formed under supercritical CO2 condition at 80 °C for various immersion time (a: 0.5 h; b:
2 h; c: 7 h; d: 23 h; e: 48 h; f: 96 h; g: 168 h) (1: surface morphology; 2: cross-section morphology; A: epoxy; B: scale; C: steel substrate).
190 Y. Zhang et al. / Corrosion Science 59 (2012) 186–197

forming scale [40,41]. From the cross-section morphologies shown 35


in Figs. 3(d) and (e) and 4(d) and (e), it can be seen that although in
30 28.3 9.5 MPa
both cases steel surface was covered with corrosion product scale

Corrosion rate, mm/y


27.5 1 MPa
for immersion time from 23 to 48 h, the corrosion product scale 25
was not dense and had some local defects. When the immersion 20.56
20 17.2
time was 96 h or longer, the iron carbonate scales formed in both
cases obviously became denser, as can be seen from Figs. 3(f) 15 16.5 12.87
and (g) and 4(f) and (g). 16.2
10 8.5
The fact that the formation of iron carbonate scale with immer- 7.35 7.26
sion time was earlier appeared under supercritical CO2 condition 5
8.36
(see Fig. 4(c)) than that at low CO2 partial pressure (see Fig. 3(c)) 4.03
could ascribe to the iron carbonate precipitation and growth kinet- 0 1.75 1.64
ics. It is no doubt that the prerequisite for the scale formation is a 0 20 40 60 80 100 120 140 160 180
super saturated solution. In the case of CO2 corrosion, when the Immersion time, h
concentrations of Fe2+ and CO32 ions exceed the solubility product
of iron carbonate, solid iron carbonate precipitates on the steel sur- Fig. 5. The corrosion rates of X65 steel exposed in water saturated with CO2 under
low partial pressure and supercritical condition at 80 °C for various immersion
face [33,42]:
time.
Fe2þ þ CO2
3 ! FeCO3 ð3Þ
According to the models proposed by Johnson and Tomson in the case of scale-free CO2 corrosion, an increase of CO2 partial pres-
1991 [43] and van Hunnik and Hendriksen in 1996 [44], two fol- sure typically leads to an increase in the corrosion rate, as the con-
lowing equations (Eq. (5) [44] is a revised version of Eq. (4) [43] centration of H2CO3 increases with an increase of pressure, which
for the use in a wide range of supersaturation) are used to describe consequently accelerates the cathodic reaction, and ultimately en-
the kinetics of iron carbonate precipitation in CO2 corrosion: hances the corrosion rate [2,16,45]. That is why the corrosion rate
at low CO2 partial pressure (17 mm/y) was much lower than that
RP ¼ f ðTÞfðSS Þ0:5  1g2 ð4Þ under supercritical CO2 condition (28 mm/y).
Furthermore, as shown in Fig. 5, in the immersion time range of
RP ¼ f ðTÞðSS  1Þð1  S1
S Þ ð5Þ 2–7 h, the corrosion rate under supercritical CO2 condition de-
creased sharply while it only slightly decreased at low CO2 partial
where, RP is the precipitation rate of iron carbonate, kmol m s1;3
pressure, which ascribes to the formation of corrosion product
SS is the supersaturation of iron carbonate; and f(T) (kmol m3 s1)
scale, as can be seen from the morphologies in Figs. 3(c) and 4(c)
is the temperature impact factor that can be obtained by the kinetic
and XRD analysis in Fig. 2. In general, the decrease of CO2 corrosion
constant kr (kg2 mol1 m2 s1) which obeys Arrhenius law.
rates was mainly due to the passivation of the surface by CO2 cor-
Iron carbonate supersaturation SS is defined as follows [34,40]:
rosion product scales formed on the surfaces of steels.
½Fe2þ ½CO2
3 
In addition, in both conditions the corrosion rates decreased
SS ¼ ð6Þ sharply in the immersion time range of 7–48 h while the corrosion
K sp
rates decreased more smoothly with the immersion time from 48
where [Fe2+] represents the equilibrium ferrous ion concentration, to 96 h. When the immersion time increased up to 96 h or longer,
mol/L; [CO32] represents the equilibrium concentration of carbon- the corrosion rates decreased very slightly and nearly kept con-
ate ion, mol/L; Ksp is the solubility product of iron carbonate. stant. It could be explained that with an increase of the immersion
In the experiments with different immersion time, the temper- time, the surface coverage percentage by iron carbonate increased,
ature was constant, thus Ksp could be approximately considered as at the same time, the iron carbonate scales became denser and
a constant. However, with increasing pressure, the concentration more protective, as shown in Figs. 3(c)–(g) and 4(c)–(g). This is
of H2CO3 increases, which accelerates the cathodic reactions to in good agreement with the experimental results of Cui et al. [46].
generate more CO32 in the solution [24,25]. To keep the electric The error bars shown in the plots of Fig. 5 were used for the
charge balance, the anodic reaction accordingly accelerates to experimental uncertainty analysis of corrosion rate measurement.
release more Fe2+ into the solution. With an increase of the concen- It could be seen that after certain immersion time (96 h in this
trations of Fe2+ and CO32 ions, the supersaturation of iron carbon- study, as shown in Fig. 5), the corrosion rate changed very slightly
ate correspondingly increases according to Eq. (6). In this case, and nearly kept constant. Therefore, the following investigations
according to Eqs. (4) and (5), the precipitation rate of iron carbon- were carried out at longer and constant immersion time (168 h).
ate increases with an increase of supersaturation. In general, rapid
precipitation of iron carbonate can promote the scale to form. This
is the reason why the CO2 corrosion product scale under supercrit- 3.2. CO2 corrosion behaviour of X65 steel at various temperatures with
ical CO2 condition was formed during 2–7 h while it can be only constant immersion time
formed after 23 h at low CO2 partial pressure, as can be seen from
Figs. 4(c) and 3(c) and (d). The chemical composition of CO2 corrosion product scale
Fig. 5 shows the corrosion rates of X65 pipeline steel exposed in formed at various temperatures for 168 h immersion was analyzed
water saturated with CO2 under low partial pressure and supercrit- by means of EDS, as shown in Fig. 6. These results revealed that
ical condition at 80 °C for 0.5, 2, 7, 23, 48, 96 and 168 h, respec- both under low CO2 partial pressure and supercritical CO2 condi-
tively. In general, the corrosion rates of X65 pipeline steel tion, the element compositions of the scales formed at various
decreased with an increase of the immersion time, no matter under temperatures were the same and consisted of Fe, O and C.
low CO2 partial pressure or supercritical CO2 condition. In the first In Fig. 7, for both conditions, the XRD spectra of the scales
2 h immersion when there was no scale formed on the steel formed at various temperatures were similar. All the diffraction
surface, the corrosion rates only decreased slightly and kept a high peaks of the scales were corresponding to FeCO3 crystal and only
level, e.g. about 17 mm/y at low CO2 partial pressure and about the relative intensities of FeCO3 peaks at various temperatures
28 mm/y under supercritical CO2 condition. It is known that in were different. No iron crystal peaks were detected. This indicated
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 191

pCO2=1MPa o
pCO2=9.5MPa
130 C O Fe
Fe Au o
130 C
O Au C Fe
C Fe
Au Fe
O
C OFe
o
110 C Au Fe o
110 C
C Fe

CPS
CPS

o
Au Fe 80 C O
COFe
o
C Fe Au Fe 80 C

O
O Fe o Au Fe
50 C C Fe o
Au 50 C
C Fe

0 2 4 6 8 0 2 4 6 8
E, keV E, keV
(a) (b)
Fig. 6. EDS analysis of CO2 corrosion product scale formed at various temperatures for 168 h immersion under low CO2 partial pressure (a) and supercritical CO2 condition (b).

basically identical with that under supercritical CO2 condition (see


Fig. 9), both of which were a crystalline layer, which has already
been identified as FeCO3 in the previous EDS and XRD analysis
shown in Figs. 7 and 8.
At 50 and 80 °C, the scale composed of crystal grains with het-
erogeneous size in both conditions, as shown in Figs. 8(a) and (b)
and 9(a) and (b). However, with increasing temperatures up to
110 and 130 °C, the grain size decreased significantly and the scale
became more compact, as can be seen from Figs. 8(c) and (d) and
9(c) and (d). This phenomenon was especially obvious under
supercritical CO2 condition. Smaller grains could closely accumu-
late together and form a compact FeCO3 scale, which can also be
identified by the cross-section morphologies shown in Figs. 10
and 11.
The phenomenon that CO2 corrosion product scale formed at
high temperatures looks more compact than that at low tempera-
tures could be still ascribed to the iron carbonate precipitation and
growth kinetics at various temperatures. According to Eqs. (4)–(6),
the precipitation rate depends on the iron carbonate solubility
product. A number of researches [47–52] revealed that the solubil-
ity product of iron carbonate Ksp is a function of temperature (Tk),
expressed in the following Eqs. (7) [50], (8) [51] and (9) [52]:

1365:17
log K sp ¼ 14:66 þ ð7Þ
Tk

2:1963
log K sp ¼ 59:2385  0:041377T k 
Tk
þ 24:5724 log ðT k Þ ð8Þ

 
DH 0 1 1 1
log K sp ¼ log K 298:15   
2:303R T k 298:15 2:303RT k
Z Tk Z Tk
Fig. 7. XRD spectra of CO2 corrosion product scales formed at various temperatures
1
 DC dT k þ DC d ln T k ð9Þ
for 168 h immersion under low CO2 partial pressure (a) and supercritical CO2 298:15 2:303RT k 298:15
condition (b).
where Ksp is the solubility product of iron carbonate; Tk is the tem-
perature, K; R is the gas constant, 8.3145 J mol1 K1; DH0 is the
that the scales formed under these conditions were very thick and standard enthalpy of reaction, J/mol; DC is the standard heat capac-
compact. Combining the EDS analysis and XRD results shown in ity of reaction, J mol1 K1. According to Eqs. (7)–(9), the solubility
Figs. 6 and 7, it could be concluded that the composition of CO2 product of iron carbonate decreases with an increase of tempera-
corrosion product scale formed under these conditions was only ture, which has been verified by experiments [47,48,51].
FeCO3, without other corrosion phases such as Fe3O4, a-FeOOH As the solubility product of iron carbonate Ksp decreases with an
and so on. increase of temperature, the iron carbonate supersaturation SS
The surface morphologies of CO2 corrosion product scales accordingly increases according to Eq. (6) and consequently iron
formed at various temperatures are shown in Figs. 8 and 9. It carbonate precipitation rate RP increases from the precipitation
was obvious that the surface morphologies of the CO2 corrosion kinetics shown in Eqs. (4) and (5). The higher the temperature,
product scale formed at low CO2 partial pressure (see Fig. 8) were the bigger the supersaturation SS and the higher the iron carbonate
192 Y. Zhang et al. / Corrosion Science 59 (2012) 186–197

Fig. 8. Surface morphologies of CO2 corrosion scales formed at low CO2 partial pressure for 168 h immersion at 50 °C (a), 80 °C (b), 110 °C (c) and 130 °C (d).

Fig. 9. Surface morphologies of CO2 corrosion scales formed under supercritical CO2 condition for 168 h immersion at 50 °C (a), 80 °C (b), 110 °C (c) and 130 °C (d).

precipitation rate RP are. As the precipitation rate increases with pressure leads to an increase in bicarbonate and carbonate ion con-
the increase of temperature, the scale forms more rapid and centration and a higher supersaturation (Eq. (6)), which accelerates
becomes more compact and protective at higher temperatures. the precipitation of FeCO3 and the formation of scale [40,41].
Similar trend was also found in the experimental results provided The corrosion rates of X65 pipeline steel immersed in water sat-
by Sun and Nesic [34] and Cui et al. [46]. urated with CO2 under low partial pressure and supercritical con-
The variation of the thickness of CO2 corrosion product scale dition for 168 h at different temperatures are shown in Fig. 13. In
with temperature is shown in Fig. 12. No matter under low CO2 both conditions, the corrosion rate increased with an increase of
partial pressure or supercritical CO2 condition, the thickness of temperature from 50 to 80 °C and then decreased from 80 to
the scale decreased with an increase in the scale forming temper- 130 °C. The maximum corrosion rate was observed at 80 °C. The
ature. However, at a given temperature, the thickness of the scale variation rule of corrosion rate with temperature is in agreement
formed under supercritical condition was bigger than that under with our previous studies [23] and the results from Ikeda et al.
low CO2 partial pressure, which suggests that increasing pressure [53]. However, it should be noted that although the variation of
is more benefit to forming scale, because higher CO2 partial corrosion rate with temperature was the same, the corrosion rate
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 193

Fig. 10. Cross-section morphologies of CO2 corrosion product scales formed under low CO2 partial pressure for 168 h immersion at 50 °C (a), 80 °C (b), 110 °C (c) and 130 °C
(d) (A: epoxy, B: scale, C: steel substrate).

under supercritical CO2 condition was much higher than that at solution can still approach to the steel surface through the defects
low CO2 partial pressure at all temperatures. such as pore in the scale and therefore the scale showed poor pro-
It is known that the temperature accelerates all the processes tectiveness. Due to the poor protectiveness of the corrosion prod-
(electrochemical, chemical, transport, etc.) involved in corrosion. uct scale at lower temperatures, the corrosion rates were still
One would expect then that the corrosion rate steadily increases much higher than that at higher temperatures, which indicates
with temperature, and this is the case when precipitation of iron that the compactness of CO2 corrosion product scale plays domi-
carbonate or other protective scales does not occur. However, the nant role in decreasing the corrosion rate while the thickness of
situation changes markedly once the concentrations of Fe2+ and CO2 corrosion product scale had little effect on the corrosion rate.
CO32 ions exceed the solubility product of iron carbonate. In that
case, increasing temperature accelerates the kinetics of precipita- 4. Discussion
tion and the formation of the protective scale. Since iron carbonate
scales formed faster and were more protective at higher tempera- 4.1. CO2 corrosion mechanism
tures, thereby the corrosion rate decreased more with the increase
of temperature. That is why the corrosion rates at 110 and 130 °C In the absence of water, no water molecules are available to hy-
were lower compared with 50 and 80 °C. drate Men+ metal ions. Hence, no electrochemical corrosion can
By comparing the thickness of CO2 corrosion product scale (see take place in dry CO2 both under low pressures and supercritical
Fig. 12) and the corrosion rates at various temperatures (see conditions [23,27].
Fig. 13), it could be found that although the thicknesses of CO2 cor- In the presence of water, CO2 dissolves into water to form car-
rosion product scale formed at low temperatures (50 and 80 °C) bonic acid (H2CO3). The initial step is presented by the following
were bigger than that at high temperatures (110 and 130 °C), the reaction:
corrosion product scale had some local defects and failures, as
CO2 þ H2 O ! H2 CO3 ð10Þ
shown in Figs. 10(a) and (b) and 11(a) and (b). In this case, the
194 Y. Zhang et al. / Corrosion Science 59 (2012) 186–197

Fig. 11. Cross-section morphologies of CO2 corrosion product scales formed under supercritical CO2 condition for 168 h immersion at 50 °C (a), 80 °C (b), 110 °C (c) and 130 °C
(d) (A: epoxy, B: scale, C: steel substrate).
Thickness of CO2 product scale,µm

100 8
9.5 MPa
9.5 MPa 1 MPa
80 1 MPa
Corrosion Rate, mm/y

60
4
40

2
20

0 0
40 60 80 100 120 140 40 60 80 100 120 140
Temperature,°C Temperature,°C

Fig. 12. Variation of the thickness of CO2 corrosion product scale formed under low Fig. 13. The corrosion rates of X65 pipeline steel immersed in water saturated with
CO2 partial pressure and supercritical CO2 condition for 168 h immersion as a CO2 under low partial pressure and supercritical condition for 168 h at different
function of scale forming temperature. temperatures.
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 195

A typical CO2 corrosion system composes of three cathodic reac- supercritical condition with increasing pressure, it will lead to dif-
tions and one anodic reaction. The cathodic reactions include the ferent interaction with water, i.e. CO2 solubility in water will not
reduction of carbonic acid into bicarbonated ions (Eq. (11)), the follow Henry’s law in liquid or supercritical CO2 conditions, which
reduction of bicarbonate into carbonated ions (Eq. (12)) and the leads CO2 solubility in water to increase significantly with increas-
reduction of hydrogen ions (Eq. (13)): ing pressure to supercritical region [25], as shown in the following
discussion.
2H2 CO3 þ 2e ! H2 þ 2HCO3 ð11Þ
4.2. The reason why the corrosion rates of steels under low CO2 partial
HCO3 ! Hþ þ CO2
3 ð12Þ pressure and supercritical CO2 condition show so big difference?

2Hþ þ 2e ! H2 ð13Þ It is known that Henry’s law can be applied in order to calculate
the vapor–liquid equilibrium of CO2 at low pressures [55]. How-
The anodic reaction is the dissolution of iron shown in Eq. (14):
ever, at high pressures, Henry’s law cannot be used to calculate
Fe ! Fe2þ þ 2e ð14Þ the concentration of CO2 in the solution [24]. Thus, under super-
critical CO2 conditions, the mutual solubility of CO2 and H2O could
So, the whole CO2 corrosion reaction can be written as follows: be obtained from the following equations provided by Spycher
et al. [56] and Choi et al. [24,25]:
CO2 þ H2 O þ Fe ! FeCO3 þ H2 ð15Þ
1B
These electrochemical reactions of CO2 corrosion shown in Eqs. yH2 O ¼ ð16Þ
1=A  B
(10)–(15) are often accompanied by the formation of FeCO3 scales
(and/or other iron compounds such as Fe3O4), which can be protec-
xCO2 ¼ Bð1  yH2 O Þ ð17Þ
tive or non-protective depending on the experimental conditions
under which they are formed. The effect of environmental condi- where xCO2 is the solubility of CO2 in water (molar fraction) and yH2 O
tions (such as temperature, pressure, pH, solution chemistry, flow is the solubility of H2O in CO2 (molar fraction).
and metallurgy of the steel) on the CO2 corrosion rate of steels Parameters A and B in Eqs. (16) and (17) are defined as follows:
mainly depends on their effect on the formation and morphological !
characteristics of corrosion product scale [6–12]. This type of CO2 K 0H2 O ðP  P 0 ÞV
H O
A¼ exp 2
ð18Þ
corrosion mechanism at CO2 partial pressures relevant for the oil /H2 O Ptot RT
and gas transport (usually below 1 MPa) have been widely studied
!
and accepted by the researchers, as shown in many literatures /CO2 Ptot ðP  P0 ÞV CO2
[2,4,5,13,33–38]. B¼ exp  ð19Þ
In this study, under both low CO2 partial pressure and supercrit- 55:508K 0CO2 ðgÞ RT
ical CO2 condition, the characteristics of CO2 corrosion product where Vi is the average partial molar volume of the pure condensed
scale formed under experimental conditions (at various tempera- component i over the pressure interval P0 to P, cm3/mol; K is the
tures and different immersion time) were basically identical. true equilibrium constant; /i is the fugacity coefficient; Ptot is the
Moreover, although the corrosion rates of X65 pipeline steel under total pressure, bar; P is the partial pressure, bar; P0 is the reference
experimental conditions had quantitative difference between low pressure, 1 bar; R is the gas constant, 8.3145 J mol1 K1; T is the
CO2 partial pressure and supercritical CO2 condition, it showed temperature, K.
the same variation trend with temperature – increased first and As an example, the solubility of CO2 in water at 50 °C and pres-
then decreased, with a maximum corrosion rate at 80 °C. These re- sures to 600 bar calculated using Eq. (17) in terms of mole fractions
sults imply that supercritical CO2 corrosion is basically identical are shown in Fig. 14 [56]. The CO2 solubility in water increases
with the commonly known typical CO2 corrosion patterns at low sharply with pressure up to the critical point and smoothly after
partial pressure shown in Eqs. (10)–(15). this point. The CO2 solubility trend with pressure reflected two sol-
In addition, Wang et al. [54] studied the effect of CO2 partial ubility curves for two distinct phases: supercritical CO2 above crit-
pressures on the Nyquist impedance plots and found that the ical pressure and gaseous CO2 below this pressure. This resulted in
CO2 partial pressure obviously changed the ‘‘size’’ of the semicircle a sharp change in slope on the overall solubility trends, as shown in
but it imparted little change in the shape of the impedance plot.
This implies that with a change of CO2 partial pressure, the corro-
sion rates were altered while the corrosion mechanism was not
0.03
changed, because the ‘‘size’’ of the semicircle corresponds to the
scale resistance and the corrosion rate while the shape of the
CO2 Mole Fraction in H2O

impedance plot determines the corrosion mechanism. Therefore,


it could be concluded that the corrosion mechanism under super-
0.02
critical CO2 conditions is identical with a known typical CO2 corro- Ref. 58
sion system at low pressures. Ref. 59
However, it should be noted that in this study the corrosion Ref. 60
rates of X65 pipeline steel under supercritical CO2 conditions were Ref. 61
0.01
much higher (about 3–4 times) compared with low CO2 partial Ref. 62
pressure, as shown in Fig. 13. A similar high corrosion rate
(>10 mm/y) was also reported for carbon steels by Choi and Nesic
[24,25] and Cui et al. [26]. The question arose is that now that the 0.00
corrosion mechanism under supercritical CO2 conditions is similar 0 100 200 300 400 500 600
with the typical CO2 corrosion, which factor causes so big differ- Pressure, bar
ence in corrosion rate between low CO2 partial pressure and super- Fig. 14. The solubility of CO2 in H2O at 50 °C and pressures to 600 bar in terms of
critical CO2 condition? A commonly accepted view for this big mole fractions [56]. Experimental data are shown as symbols, with sources given in
difference is that since CO2 changes from gaseous to liquid or Figures. Solubilities calculated using Eq. (17) are shown as solid lines.
196 Y. Zhang et al. / Corrosion Science 59 (2012) 186–197

Fig. 14. The CO2 critical pressure is recognized as a turning point for [2] S. Nesic, Key issues related to modelling of internal corrosion of oil and gas
pipelines – a review, Corros. Sci. 49 (2007) 4308–4338.
the concentrations of H2CO3 and HCO3. This is because after the
[3] J. Villarreal, D. Laverde, C. Fuentes, Carbon steel corrosion in multiphase slug
critical point, the gas CO2 will be changed into supercritical CO2 flow and CO2, Corros. Sci. 48 (2006) 2363–2379.
which exhibits special physicochemical properties of both a liquid [4] M. Gao, X. Pang, K. Gao, The growth mechanism of CO2 corrosion product films,
and a gas CO2-filling its container like a gas but with a density like Corros. Sci. 53 (2011) 557–568.
[5] S. Nesic, K.-L.J. Lee, A mechanistic model for carbon dioxide corrosion of mild
that of a liquid. At the critical point (31.1 °C, 7.38 MPa), the density steel in the presence of protective iron carbonate films, Corros. Sci. 6 (2003)
of the CO2 can increase drastically up to 0.4–0.5 g cm3 [57]. That 616–628.
is why the solubility of CO2 in water under supercritical CO2 con- [6] G.A. Zhang, Y.F. Cheng, On the fundamentals of electrochemical corrosion of
X65 steel in CO2-containing formation water in the presence of acetic acid in
ditions is much higher than that at low pressures. petroleum production, Corros. Sci. 51 (2009) 87–94.
A comparison between the calculated solubility of CO2 in H2O at [7] X.P. Guo, Y. Tomoe, The effect of corrosion product layers on the anodic and
50 °C according to Eq. (17) and the experimental data available in cathodic reactions of carbon steels in CO2-saturated mdea solutions at 100 °C,
Corros. Sci. 41 (1999) 1391–1402.
the literatures [58–62] is also represented in Fig. 14. The results [8] G.A. Zhang, Y.F. Cheng, Electrochemical characterization and computational
presented in Fig. 14 show a good agreement between the calcu- fluid dynamics simulation of flow-accelerated corrosion of X65 steel in a CO2-
lated data and the experimental data at various pressures. There- saturated oilfield formation water, Corros. Sci. 52 (2010) 2716–2724.
[9] J.K. Heuer, J.F. Stubbins, An XPS characterization of FeCO3 films from CO2
fore, it is suitable to use Eq. (17) to calculate the solubility of CO2 corrosion, Corros. Sci. 41 (1999) 1231–1243.
in water under supercritical CO2 conditions. [10] B. Wang, M. Du, J. Zhang, C.J. Gao, Electrochemical and surface analysis studies
As shown in Fig. 14, when the CO2 partial pressure increases, on corrosion inhibition of Q235 steel by imidazoline derivative against CO2
corrosion, Corros. Sci. 53 (2011) 353–361.
the solubility of CO2 in water phase increases, which consequently
[11] K. Gao, F. Yu, X. Pang, G. Zhang, L. Qiao, W. Chu, M. Lu, Mechanical properties of
enhances the concentrations of carbonic species (H2CO3, HCO3 CO2 corrosion product scales and their relationship to corrosion rates, Corros.
and CO32) in the solution. Based on the studies by Choi and Nesic Sci. 50 (2008) 2796–2803.
[24,25], with an increase of pressure, the concentrations of H2CO3 [12] G.A. Zhang, Y.F. Cheng, Localized corrosion of carbon steel in a CO2-saturated
oilfield formation water, Electrochim. Acta 56 (2011) 1676–1685.
and HCO3 increases. For example, in the experimental pressures [13] C. de Waard, D.E. Williams, Carbonic acid corrosion of steel, Corrosion 31
of this study, the H2CO3 concentration under supercritical CO2 con- (1975) 177–181.
dition (9.5 MPa, 206 ppm) is 5.4 times than at low pressure (1 MPa, [14] C. de Waard, D.E. Milliams, Prediction of carbonic acid corrosion in natural gas
pipelines, in: Proceedings of the First International Conference on the Internal
38 ppm). and External Protection of Pipes, Paper F1, University of Durham, Durham, UK,
With an increase of the concentrations of H2CO3 and HCO3, the 1975.
processes of the electrochemical corrosion reactions shown by Eqs. [15] C. de Waard, U. Lotz, D.E. Milliams, Predictive model for CO2 corrosion
engineering in wet natural gas pipelines, Corrosion 47 (1991) 976–985.
(11)–(15) will be accelerated accordingly and ultimately the corro- [16] C. de Waard, U. Lotz, Prediction of CO2 corrosion of carbon steel, Corrosion/93,
sion rate was increased [2]. Consequently, the corrosion rate of X65 NACE International, Houston/TX, 1993, Paper No. 69.
pipeline steel under supercritical CO2 condition was much higher [17] C. de Waard, U. Lotz, A. Dugstad, Influence of liquid flow velocity on CO2
corrosion: a semi-empirical model, Corrosion/95, NACE International,
than that at low CO2 partial pressure. Houston/TX, 1995, Paper No. 128.
[18] R. Nyborg, Overview of CO2 corrosion models for wells and pipelines,
Corrosion/2002, NACE International, Houston/TX, 2002, Paper No. 233.
5. Conclusion [19] M. Seiersten, Material selection for separation, transportation and disposal of
CO2, Corrosion/2001, NACE International, Houston/TX, 2001, Paper No. 42.
[20] S. Nesic, M. Nordsveen, R. Nyborg, A. Stangeland, A mechanistic model for CO2
In this study, the CO2 corrosion behaviour of X65 pipeline steel corrosion with protective iron carbonate films, Corrosion/2001, NACE
at various temperatures for different immersion time under low International, Houston/TX, 2001, Paper No. 40.
CO2 partial pressure and supercritical CO2 condition were investi- [21] A.M.K. Halvorsen, T. Sntvedt, CO2 corrosion model for carbon steel including a
wall shear stress model for multiphase flow and limits for production rate to
gated by weight loss measurements and surface analysis tech- avoid mesa attack, Corrosion/99, NACE International, Houston/TX, 1999, Paper
niques. The following conclusions can be drawn: No. 42.
[22] M.B. King, A. Mubarak, J.D. Kim, T.R. Bott, The mutual solubilities of water with
supercritical and liquid carbon dioxide, J. Supercrit. Fluids 5 (1992) 296–302.
1. Under both low CO2 partial pressure and supercritical CO2 con-
[23] Y.C. Zhang, K.W. Gao, G. Schmitt, Effect of water on steel corrosion under
dition, the corrosion behaviour of X65 pipeline steel, including supercritical CO2 conditions, Mater. Performance 50 (2011) 62–68.
characteristics of CO2 corrosion product scale and the variation [24] Y.S. Choi, S. Nesic, Corrosion behaviour of carbon steel in supercritical CO2–
rule of corrosion rate with temperature, were similar. water environments, Corrosion/2009, NACE International, Houston/TX, 2009,
Paper No. 256.
2. The change in CO2 partial pressure does not change the corro- [25] Y.S. Choi, S. Nesic, Determining the corrosive potential of CO2 transport
sion mechanism. Supercritical CO2 corrosion was basically iden- pipeline in high pCO2–water environments, Int. J. Greenh. Gas Con. 5 (2011)
tical with the typical CO2 corrosion patterns at low pressures. 788–797.
[26] Z.D. Cui, S.L. Wu, C.F. Li, S.L. Zhu, X.J. Yang, Corrosion behavior of oil tube steels
3. However, due to the significantly higher concentration of corro- under conditions of multiphase flow saturated with supercritical carbon
sive carbonic acid under supercritical CO2 condition, the corro- dioxide, Mater. Lett. 58 (2004) 1035–1040.
sion rates of X65 pipeline steel under supercritical CO2 [27] E.M. Russick, G.A. Poulter, C.L.J. Adkins, N.R. Sorensen, Corrosive effects of
supercritical carbon dioxide and cosolvents on metals, J. Supercrit. Fluids 9
condition were much higher than that at low CO2 partial (1996) 43–50.
pressure. [28] F.W. Schremp, G.R. Roberson, Effect of supercritical carbon dioxide (CO2) on
construction materials, J. Soc. Pet. Eng. 15 (1975) 227–233.
[29] P.M. Jarrel, C.E. Fox, M.H. Stein, S.L. Webb, Practical aspects of CO2 flooding, SPE
Monograph, vol. 22, Society of Petroleum Engineers (SPE), Richardson, TX,
2002.
Acknowledgment
[30] C.M. Benson, S.H. Stevens, S.M. Benson, Economic feasibility of carbon
sequestration with enhanced gas recovery, Energy 29 (2004) 1413–1422.
The authors would like to express their thanks to Prof. Changf- [31] E. Rubin, L. Meyer, H. Coninck, Carbon dioxide capture and storage: technical
eng Chen (Department of Materials Science and Engineering, China summary, IPCC Special Report, 2005.
[32] M. Seiersten, K.O. Kongshaug, Materials selection for capture, compression,
University of Petroleum, Beijing, China) for his help in supercritical transport and injection of CO2, in: D.C. Thomas, S.M. Benson (Eds.), Carbon
CO2 corrosion experiments. Dioxide Capture for Storage in Deep Geologic Formations, vol. 2, Elsevier Ltd.,
2005, p. 937.
[33] A. Dugstad, Mechanism of protective film formation during CO2 corrosion of
References carbon steel, Corrosion/98, NACE International, Houston/TX, 1998, Paper No. 31.
[34] W. Sun, S. Nesic, Basics revisited: kinetics of iron carbonate scale precipitation
in CO2 corrosion, Corrosion/2006, NACE International, Houston/TX, 2006,
[1] M.B. Kermani, A. Morshed, Carbon dioxide corrosion in oil and gas production
Paper No. 365.
– a compendium, Corrosion 59 (2003) 659–683.
Y. Zhang et al. / Corrosion Science 59 (2012) 186–197 197

[35] D.H. Zheng, D.F. Che, Y.H. Liu, Experimental investigation on gas–liquid two- [49] W. Sun, S. Nesic, R.C. Woollam, The effect of temperature and ionic strength on
phase slug flow enhanced carbon dioxide corrosion in vertical upward iron carbonate (FeCO3) solubility limit, Corros. Sci. 51 (2009) 1273–1276.
pipeline, Corros. Sci. 50 (2008) 3005–3020. [50] G.M. Marion, D.C. Catling, J.S. Kargel, Modeling aqueous ferrous iron chemistry
[36] X.Y. Zhang, F.P. Wang, Y.F. He, Y.L. Du, Study of the inhibition mechanism of at low temperatures with application to Mars, Geochim. Cosmochim. Acta 67
imidazoline amide on CO2 corrosion of Armco iron, Corros. Sci. 43 (2001) (2003) 4251–4266.
1417–1431. [51] J. Greenberg, M. Tomson, Precipitation and dissolution kinetics and equilibria
[37] A. Munoz, J. Genesca, R. Duran, J. Mendoza, Mechanism of FeCO3 formation on of aqueous ferrous carbonate vs. temperature, Appl. Geochem. 7 (1992) 185–
API X70 pipeline steel in brine solutions containing CO2, Corrosion/2005, NACE 190.
International, Houston/TX, 2005, Paper No. 297. [52] H.C. Helgeson, Thermodynamics of hydrothermal systems at elevated
[38] A. Eslami, R. Kania, B. Worthingham, G.V. Boven, R. Eadie, W. Chen, Effect of temperatures and pressures, Am. J. Sci. 267 (1969) 729–804.
CO2 and R-ratio on near-neutral pH stress corrosion cracking initiation under a [53] A. Ikeda, M. Ueda, S. Mukai, CO2 behavior of carbon and Cr steels, in: R.H.
disbonded coating of pipeline steel, Corros. Sci. 53 (2011) 2318–2327. Hausler, H.P. Giddard (Eds.), Advances in CO2 Corrosion, vol. 1, NACE, Houston,
[39] ASTM Standard G 31, Standard practice for laboratory immersion corrosion 1984, pp. 39–51.
testing of metals, in: Annual Book of ASTM Standards, vol. 03. 02. ASTM [54] S.H. Wang, K. George, S. Nesic, High pressure CO2 corrosion electrochemistry
International, West Conshohocken, PA, 1994. and the effect of acetic acid, Corrosion/2004, NACE International, Houston/TX,
[40] Y. Sun, S. Nesic, A parametric study and modeling on localized CO2 corrosion in 2004, Paper No. 375.
horizontal wet gas flow, Corrosion/2004, NACE International, Houston/TX, [55] B. Brown, K.-L.J. Lee, S. Nesic, Corrosion in multiphase flow containing small
2004, Paper No. 380. amounts of H2S, Corrosion/2003, NACE International, Houston/TX, 2003, Paper
[41] S. Nesic, K.-L.J. Lee, V. Ruzic, A mechanistic model of iron carbonate film No. 341.
growth and the effect on CO2 corrosion of mild steel, Corrosion/2002, NACE [56] N. Spycher, K. Pruess, J.E. King, CO2–H2O mixtures in the geological
International, Houston/TX, Paper No. 237. sequestration of CO2, I. Assessment and calculation of mutual solubilities
[42] A. Dugstad, Formation of protective corrosion films during CO2 corrosion of from 12 to 100 °C and up to 600 bar, Geochim. Cosmochim. Acta 67 (2003)
carbon steel, in: Proceedings of the Conference of European Federation of 3015–3031.
Corrosion (EFC), Institute of Materials, London, UK, 1997. [57] R.C. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and Liquids, fourth
[43] M.L. Johnson, M.B. Tomson, Ferrous carbonate precipitation kinetics and its ed., McGraw-Hill Book Company, New York, 1987.
impact CO2 corrosion, Corrosion/91, NACE International, Houston/TX, 1991, [58] R. Wiebe, V.L. Gaddy, The solubility in water of carbon dioxide at 50, 75, and
Paper No. 268. 100 °C, at pressures to 700 atmospheres, J. Am. Chem. Soc. 61 (1939) 315–318.
[44] E.W.J. van Hunnik, E.L.J.A. Hendriksen, The formation of protective FeCO3 [59] K. Todheide, E.U. Frank, Das Zweiphasengebiet und die kritische Kurve im
corrosion product layers, Corrosion/96, NACE International, Houston/TX, 1996, System Kohlendioxid-Wasser bis zu Drucken von 3500 bar, Z. Phys. Chem. 37
Paper No. 6. (1963) 387–401.
[45] A. Dugstad, L. Lunde, K. Videm, Parametric study of CO2 corrosion of carbon [60] J.A. Briones, J.C. Mullins, M.C. Thies, Ternary phase equilibria for acetic acid
steel, Corrosion/94, NACE International, Houston/TX, 1994, Paper No. 14. water mixtures with supercritical carbon dioxide, Fluid Phase Equilib. 36
[46] Z.D. Cui, S.L. Wu, S.L. Zhu, X.J. Yang, Study on corrosion properties of pipelines (1987) 235–246.
in simulated produced water saturated with supercritical CO2, Appl. Surf. Sci. [61] R. D’Souza, J.R. Patrick, A.S. Teja, High pressure phase equilibria in the carbon
252 (2006) 2368–2374. dioxide–n-hexadecane and carbon dioxide–water systems, Can. J. Chem. Eng.
[47] C.A.R. Silva, X. Liu, F.J. Millero, Solubility of sidertie (FeCO3) in NaCl solutions, J. 66 (1988) 319–323.
Solution Chem. 31 (2002) 97–108. [62] R. Dohrn, A.P. Bünz, F. Devlieghere, D. Thelen, Experimental measurements of
[48] A. Dugstad, The importance of FeCO3 supersaturation on the CO2 corrosion of phase equilibria for ternary and quaternary systems of glucose, water, CO2 and
carbon steels, Corrosion/92, NACE International, Houston/TX, 1992, Paper No. 14. ethanol with a novel apparatus, Fluid Phase Equilib. 83 (1993) 149–158.

S-ar putea să vă placă și