Sie sind auf Seite 1von 16

A model for the corrosion of steel subjected to synthetic produced water

containing sulfate, chloride and hydrogen sulde


Peter Smith
a,n
, Sudipta Roy
a
, David Swailes
b
, Stephen Maxwell
c
, David Page
d
, John Lawson
d
a
Chemical Engineering and Advanced Materials, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom
b
Mechanical and Systems Engineering, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom
c
Commercial Microbiology Limited, Kettock Lodge, Campus 2, Aberdeen Science Park, Bridge of Don, Aberdeen, AB22 8GU, United Kingdom
d
Chevron North Sea Limited, Chevron House, Hill of Rubislaw, Aberdeen, AB15 6XL, United Kingdom
a r t i c l e i n f o
Article history:
Received 17 January 2011
Received in revised form
11 July 2011
Accepted 24 July 2011
Available online 6 August 2011
Keywords:
Corrosion
Electrochemistry
Mathematical modeling
Kinetics
Mass transfer
Produced water
a b s t r a c t
A model was developed for the prediction of corrosion rates associated with steel subjected to synthetic
produced water. The corrosive species included in the model, identied through water analysis
conducted in the eld, are sulfate, chloride and hydrogen sulde. The effect on corrosion of these
species was examined through polarization experimentation using a three electrode glass corrosion cell
and potentiostat. Samples of carbon steel, used in sub-sea pipeline systems, were used at the working
electrode and the experiments were carried out at similar physicochemical conditions observed in
pipeline systems in the eld. The model was based on heterogeneous reactions at the metal surface,
with electrochemical parameters determined through experimentation employed in the model to
describe the anodic and cathodic processes involved in the corrosion of steel. The model consists of a
system of equations with ButlerVolmer kinetics describing the charge transfer and the Nernst
diffusion model the mass transfer processes occurring in the corrosion system. The solution is based
on a charge balance between the reduction and the oxidation processes which occur at the steel
surface. Current density convergence criteria were used in the model to solve the system of equations
for corrosion potential, surface species concentration and component current densities. The corrosion
rate is determined as the rate of oxidation of iron at the surface and model results have been validated
using experimental data. The model demonstrates a reasonable qualitative match with corrosion data
collected in the potential region close to the corrosion potential in general, with good qualitative match
in the anodic region near the corrosion potential. Some deviation occurs between model and
experimental values where overpotentials become large but the model is shown to respond well to
changes in input parameter values and predicts the corrosion potential and corrosion rate for each
system within experimental variability and the accepted standards of accuracy.
& 2011 Elsevier Ltd. All rights reserved.
1. Introduction
Corrosion mitigation in the oileld industry has traditionally
been performed by combining methods for measuring corrosion
rates such as corrosion coupons and regular pipeline inspections
with prevention strategies such as cathodic protection and the
deployment of corrosion inhibiting chemicals. In more recent
times, these methods have been enhanced by the employment of
intelligent pigs which are able to detect the severity of corrosion
at particular sites and more effective in-situ corrosion probes for
corrosion measurement. Corrosion models and corrosion predic-
tion models are now proving to be increasingly important in the
oileld industry to further enhance the performance of corrosion
mitigation strategies and provide non-invasive methods for cor-
rosion assessment of oileld assets.
Several corrosion models have been developed for specic
corrosion mechanisms such as carbon dioxide (Nesic et al., 1996;
NORSOK, 2005) for which predictive modeling guidelines in the
eld have also been published (Nyborg, 2009). Work towards
corrosion modeling for other corrosion mechanisms such as that
associated with microbiologically inuenced corrosion (MIC)
(Maxwell, 2006; Pots et al., 2002) have also been conducted and
applied to oileld systems in the eld. More complex investiga-
tive models for MIC (Picioreanu and Loosdrecht, 2002) and
other more generalized corrosion mechanisms such as galvanic
corrosion (Morris and Smyrl, 1988, 1989) and differential aeration
of surfaces subjected to moisture lms (Alkire and Nicolaides,
1974) have proven successful at exploring the respective
corrosion mechanisms in more detail.
The motivation for this current work is based on recent
activity in the oileld industries to develop corrosion models that
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/ces
Chemical Engineering Science
0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.07.033
n
Corresponding author.
E-mail address: p.j.smith@hotmail.co.uk (P. Smith).
Chemical Engineering Science 66 (2011) 57755790
more effectively predict corrosion rates in oileld systems com-
pared to the models that are currently used. Corrosion mechan-
isms in oileld systems are complex and have been studied
widely showing high degrees of interaction between corrosion
species, products and oileld metallurgies. The interactions of
sulfate and chloride are of interest in this work since a presence of
sulfate in oileld produced water systems also facilitates further
corrosion mechanisms such as MIC by sulfate reducing bacteria
and chloride due to its catalytic action on corrosion. Abiotic
corrosion experimentation in this work as was conducted to
investigate the interaction between a sample of pipeline steel
and two main species, sodium sulfate and sodium chloride,
identied in the eld water analysis data presented in Table 1.
Firstly, it was determined if abiotic interactions could be mea-
sured in the system for these components. Then the effect of
synthetic hydrogen sulde in addition to these two components
was investigated. Presented within are the experimental proce-
dures and results of abiotic experimentation along with the
model development steps. The model is validated using corrosion
data collected in the experimental work.
By examination of the physicochemical data of produced water
encountered in a typical pipeline uid, presented in Table 1, a
number of abiotic corrosion inuencing components can be
initially identied based on knowledge of their corrosive proper-
ties and their concentration. The main abiotic corrosion para-
meter in Table 1 can be identied as sodium chloride. This is
present in higher concentrations as compared to the other
components present and is well known to have a signicant
inuence on the corrosion of steel. Another abiotic parameter of
interest in Table 1 is sulfate due to its inuence on the corrosion
mechanism of chloride and its potential to facilitate further
corrosion mechanisms. A sodium salt was chosen as a source of
sulfate to ensure any inuence observed on the corrosion system
Table 2
Summary of model species, parameters and reactions used in the model.
Model parameters Reactions
Kinetic Mass transport Heterogeneous Homogeneous
E1 V vs. SCE n
i
a
ai
a
ci
C
bi
(mol m
3
) D
i
(m
2
s
1
) d
Ni
(m)
Deaerated water system
Fe
2
0.681 2 0.4 0.6 7.9810
10
7.2310
6
Fe-Fe
2
2e

0.241 1 0.6 0.4 7.4 10


3
9.4710
9
1.6710
5
H

-1/2H
2
0.01 M Sulfate system
Fe
2
0.681 2 0.4 0.6 7.9810
10
7.2310
6
Fe-Fe
2
2e

0.241 1 0.6 0.4 2.5 10


3
9.4710
9
1.6710
5
H

-1/2H
2
SO
4
2
1.177 2 1.0 10
5
1.010
9
7.8710
6
SO
4
2
2e

H
2
O-SO
3
2
OH

SO
3
2
0.817 4 5.610
10
6.4910
6
2SO
3
2
4e

3H
2
O-S
2
O
3
2
6OH

S 0.900 4 SO
3
2
4e

3H
2
O-S6OH

S
2
0.688 2 S2e

-S
2
0.01 M Sulfate, 0.2 M chloride sSystem
Fe
2
0.681 2 0.25 0.75 7.9810
10
7.2310
6
Fe-Fe
2
2e

0.241 1 0.6 0.4 0.013 9.4710


9
1.6710
5
H

-1/2H
2
SO
4
2
1.177 2 1 10
5
1.010
9
7.8710
6
SO
4
2
2e

H
2
O-SO
3
2
OH

SO
3
2
0.817 4 5.610
10
6.4910
6
2SO
3
2
4e

3H
2
O-S
2
O
3
2
6OH

S 0.900 SO
3
2
4e

3H
2
O-S6OH

S
2
0.688 2 S2e

-S
2
0.01 M Sulfate, 0.2 M chloride system, 18% saturated hydrogen sulde
Fe
2
0.681 2 0.25 0.75 7.9810
10
7.2310
6
Fe-Fe
2
2e

0.241 1 0.6 0.4 6.2 10


3
9.4710
9
1.6710
5
H

-1/2H
2
SO
4
2
1.177 2 1 10
5
1.010
9
7.8710
6
SO
4
2
2e

H
2
O-SO
3
2
OH

SO
3
2
0.817 4 5.610
10
6.4910
6
2SO
3
2
4e

3H
2
O-S
2
O
3
2
6OH

S 0.900 4 SO
3
2
4e

3H
2
O-S6OH

H
2
S H
2
S
(aq)
-H

HS

HS

HS

-H

S
2
S
2
0.688 2 2.5 10
6
5.610
10
6.4910
6
S2e

-S
2
HS

-H

S
2
1. The dissociation of water to protons and hydroxyl ions is included implicitly in the pH.
2. Cl

is assumed to change reactions parameters as mentioned in the text, and therefore homogeneous or heterogeneous specic to Cl

species are not mentioned in


Table 2.
3. Values for j
0
are not mentioned in Table 2 as they are calculated in the model using Eq. (7) and model predicted values of C
s
for each component.
4. Parameters written in italics refer to experimentally calculated values, others refer to literature values.
Table 1
Physicochemical data for produced water collected at a North Sea sub-sea asset.
Element M Concentration
mg/l mol/l
Barium Ba
2
137.327 31 0.0002
Boron B 10.8 6 0.0006
Calcium Ca
2
40.08 284 0.0071
Iron Fe
3
55.85 8 0.0001
Magnesium Mg
2
24.31 161 0.0066
Phosphorous P
3
30.97 1 0.0000
Potassium K

39.1 50 0.0013
Sodium Na

22.99 4770 0.2075


Strontium Sr
2
87.62 83 0.0009
Chloride Cl

35.45 7480 0.2110


Bromide Br

79.9 20 0.0003
Sulphate SO
4
2
96 21 0.0002
Nitrate NO
3
62 0.5 0.0000
Hydroxyl OH

17 0 0.0000
Carbonate CO
3
2
60 0 0.0000
Bicarbonate HCO
3

61.02 500 0.0082


Dissolved CO
2
CO
2
44 92.4 0.0021
Specic gravity 1.014
pH 6.58
Resistivity 0.4405 Om
Total dissolved solids 13,453 mg/l
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5776
was due to the presence of sulfate and not the presence of a
different cationic components compared to sodium chloride. A
corrosion rate model based on these components can be
expressed conceptually as a function of the corrosion rates from
each of the contributing parameters as presented in
CR f CR
Cl

,CR
SO
2
4
CR
H
2
S
1
The abiotic corrosion model presented within this study
reects a modeling strategy that would allow the inclusion of
further processes in the future. The model is based on causal
electrochemical processes that describe the charge and mass
transfer processes occurring at the steel surface. Similar methods
have been implemented in several of the models cited within, but
this model is based on parameters which can be easily measured
in the eld allowing the model to be tailored to individual oileld
systems and implemented effectively within corrosion mitigation
strategies. This holistic modeling strategy allows the inclusion of
further corrosion parameters specic to individual corrosion
systems. This concept may also provide a basis for the future
enhancement of current MIC models used in pipeline systems by
modeling together the synergistic relationship (Crolet, 2005;
Hamilton, 2002; Hardy and Bown, 1984; Lee and Characklis,
1993) between abiotic corrosion parameters measured in the
eld and those contributing to MIC.
2. Experimental
Polarization experiments have been employed in this study to
examine processes occurring in the corrosion system with the use
of a three electrode corrosion cell and a potentiostat. A Radio-
meter EDI101 rotating disk electrode (RDE) was employed as the
working electrode (WE) in this system with the tip containing a
sample of pipeline steel. The tip was manufactured by setting a
cylindrical sample of pipeline steel within a cylindrical PTFE
collar. The tip surface was machined ush and attached to the
RDE body by a screw thread. A spring loaded metallic contact
within the RDE body provided electrical connection between the
RDE body and the steel sample in the tip. The steel disk exposed
at the RDE tip had diameter of 0.6 cm with the total disk diameter
of 1.1 cm. A counter electrode (CE) manufactured from platinised
titanium with surface area of 3 cm
2
was employed as a source/
sink of electrons, and a saturated calomel electrode (SCE) refer-
ence electrode (RE) was utilized as a datum for potential mea-
surement. A schematic showing the experimental setup for this
system is presented in Fig. 1. The electrode conguration adopted
with the WE positioned centrally in the cell ensures effective
transport of electrolyte to the RDE surface. This schematic also
shows the presence of a nitrogen sparger used to purge the
corrosion cell of oxygen prior to polarization measurements.
Hydrogen sulde was introduced to the glass corrosion cell
using the experimental setup shown in Fig. 2. A quantity of 1.0 M
sulfuric acid is injected into vessel (a) via a syringe which reacts
with iron sulde in the vessel to form hydrogen sulde as
described by
FeSH
2
SO
4
-H
2
SFeSO
4
2
A feed port allows the delivery of hydrogen sulde to corrosion
cell, vessel (b). An airlock was utilized to monitor the production
of hydrogen sulde in vessel (a). Positioned between vessels
(a) and (b), the airlock contained a white solution of lead acetate
which is in the presence of hydrogen sulde forms of a gray
precipitate of lead sulde. This detector was used in this location
to determine a presence of hydrogen sulde before the feed to
vessel (b) was opened. Vessel (c) contains a 1.0 M sodium
hydroxide and is used as a scrubber to neutralize any hydrogen
sulde passing through vessel (b). There is also a hydrogen sulde
detector located on the exit port of vessel (c) to detect any
unreacted hydrogen sulde exiting the system.
Potentiostat
RE
WE
CE
RE
WE
CE
Electrolyte
WE
N
2
Electrolyte
Air lock
Nitrogen
sparger
Fig. 1. (a) Experimental setup of three electrode corrosion cells and potentiostat from front elevation and (b) side elevation showing airlock and nitrogen gas sparger
conguration.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5777
2.1. Electrolyte preparation
Electrolytes have been prepared using water analysis data of
the produced water presented in Table 1. The major corrosion
parameters present in this eld sample includes 0.2 M sodium
chloride and 0.0002 M sodium sulfate. It was necessary to prepare
electrolytes containing individual and then mixed components to
assess the inuence of each component on corrosion.
The corrosion cell was setup with the RE, CE, gas sparger and
air lock in position. The WE orice was utilized as a ll hole with
the aid of funnel. Using low conductivity water produced in a
Purite select 300 unit, the corrosion cell was then lled to a level
2 cm above the RE electrode tip, approximately 50 cm
3
. A glass
stopper was used to seal the ll hole. An air lock containing low
conductivity water was then placed at the top of the cell as in
Fig. 1(b). Each of the ground glass joints in the cell were coated
with silicone grease prior to sparing to ensure that an airtight cell
is established. The sparger connected to a nitrogen supply was
submerged to the base of the corrosion cell and used to gently
sparge nitrogen through the electrolyte. Sparging was maintained
for 20 min to allow full evacuation of oxygen from the cell. The
sparger was then raised to just below the electrolyte surface. The
WE, set to a rotation speed of 800 rpm, was then lowered into
position. This rotation minimizes the chance of nitrogen bubbles
adhering to the surface of the WE during positioning. Sparging
was maintained for a further 5 min to ensure that any oxygen
entered during the WE positioning is removed. The gas sparger
was raised to just above the electrolyte level during data collec-
tion and a nitrogen atmosphere was maintained throughout the
experiment to prevent any oxygen inux.
2.2. Corrosion data collection
Ej data was collected for this corrosion system by employing
potentiodynamic polarization in order to deduce parameters for
the model. The corrosion potential, E
corr
, of the corrosion system
was initially measured during a 30 s stabilization period prior to
imposing a potential on the system. A potential scan was then
initiated at a sweep rate of 0.01 V s
1
from a cathodic over-
potential potential of 1.5 V w.r.t. E
corr
to an anodic overpotential
of 5.0 V w.r.t. E
corr
. The potential was then reversed at the same
rate back to the start potential. This procedure was performed
three times for each corrosion system using a freshly prepared
RDE tip and fresh electrolyte for each replicate.
Ohmic drop compensation was applied to the data collected
for this system, retrospectively (Oelner et al., 2006), using
measured values of electrolyte conductivity and the interelec-
trode gap between the WE and the RE. Ohmic drop compensated
data have been presented for each system.
2.3. Materials characterization
Prior to experimental work the sample of steel was standar-
dised and characterized by assessing the microstructure and its
elemental composition. A sample of carbon steel used as the WE
in this corrosion system was encapsulated in Struers Epox resin
and the surface was mechanically ground using grades P#500,
P#1200, P#2400 and P#4000 silicone carbide grit paper consecu-
tively lubricated with 96% ethanol. The sample was washed with
96% ethanol after each paper grade to reduce contamination of
ne grit paper with larger grit particles. After the grinding stages,
the sample was ultrasonicated in ethanol for 3 min to remove any
further surface particulates. The sample was then chemically
etched using 2% nital for 60 s, washed in 96% ethanol and dried
in a ow of nitrogen.
Energy dispersive X-ray analysis (EDX), optical microscopy and
scanning electron microscopy (SEM) were employed to assess the
elemental and microstructure properties of the steel samples
used in experimental work. The steel was determined to contain
low concentrations of carbon, a small fraction of manganese with
few other alloying components. The microstructure showed
characteristics of ferrite and pearlite grain structure with grains
in the 212 mm diameter range.
3. Experimentation results
The effect of sodium chloride and sodium sulfate on the abiotic
corrosion system have been investigated to establish the con-
tribution that each of these components have on the overall
Electrolyte
FeS Powder
Lead Acetate
for detection of
H
2
S
H
2
SO
4
Lead Acetate
for detection o
H
2
S
NaOH
H
2
S
H
2
S
H
2
S
H
2
S
WE
Fig. 2. A schematic showing the experimental setup used to quantify the inuence of synthetic hydrogen sulde. (a) H
2
S production vessel, (b) corrosion cell and (c) H
2
S
neutralisation.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5778
corrosion rate. Since the mechanism of corrosion due to chloride
and the inhibition effects observed with sulfate are widely
studied, and respective concentrations in eld water are well
known, only a small range of concentrations were investigated in
these experiments. Since the electrolytes have been prepared
using low conductivity water and have then been deaerated, the
corrosion effects due to deaerated water have also been examined
in this work.
3.1. Deaerated low conductivity water system
The initial corrosion system under investigation was a system
of steel subjected to an electrolyte of pure deaerated low
conductivity water. This system represents the most simple
corrosion system where the oxidation of iron from steel is
balanced by the reduction of protons in the electrolyte to form
hydrogen at the surface. The concentration of dissolved carbon
dioxide in the electrolyte was assumed to be common across all
systems and its affect on corrosion rate was indirectly taken into
account in the model through measurement of the pH and
conductivity for each system. The electrolyte conductivity in this
system was measured for all three replicates and ranges from
65.4 to 104 mS cm
1
. Ohmic drop was calculated using the RDE
radius, knowledge of the conductivity and the interelectrode gap
between the RE and WE. For this system the Ohmic drop was
calculated to be 9669 kO. Due to the Ohmic drop, an applied
potential of 4.5 V w.r.t. the open circuit potential (OCP) is
required in polarization experiments to impose an anodic polar-
ization of 1.5 V at the working electrode in the experimental
system.
Ej data for three potential sweeps on the low conductivity
water system were carried out in triplicate. One such set of Ej
data is presented in Fig. 3. The forward and backward sweep
directions have been indicated showing that the start point of the
potential sweep was located in the cathodic region of Ej data
minimizing the oxidation reaction of iron at the surface. The
corrosion potential, E
corr
, for the system is quoted as 0.742 V
using the corrosion potential measured for the forward potential
sweep, E
corr,f
, as identied in Fig. 3 rather than E
corr,b
, because in
the backward sweep the steel is partially oxidized due to anodic
polarization. This was observed in all polarization experiments
and reects the change in iron (II) concentration in the electrolyte
due to the oxidation of iron during anodic polarization. Since the
iron (II) concentration in the electrolyte for all polarization
experiments was initially zero and that there was little change
in electrolyte pH before and after each potential sweep, the
potential shift observed in E
corr
can be attributed to an increase
in the concentration of iron (II) in the electrolyte.
The magnitude of potential shift observed in pure deaerated
low conductivity water is approximately 0.160 V which can be
translated into a change in iron (II) concentration of approxi-
mately 210
6
M with use of the Nernst equation. The corrosion
current density, j
corr
, for the low conductivity water system was
determined by Tafel analysis for each replicate and ranges from
0.107 Am
2
in replicate 1 to 0.180 Am
2
in replicate 3 which is
comparable to literature values (Ateya et al., 2002).
3.2. Sodium sulfate system
The effect of adding sulfate to pure deaerated low conductivity
water was investigated by using a 0.01 M sodium sulfate electro-
lyte. The Ej data shown in Fig. 4 for this corrosion system
showed signicant variation across the three replicates although
a similar general trend in cathodic and anodic Ej data was seen
across each replicate. The reason for variation in Ej data in this
sulfate system is not clear but there was also signicant noise in
the Ej data recorded, particularly at high anodic overpotential
regions. Similar noise has been reported as current uctuations in
literature (Cheng et al., 1999; Simard et al., 2001).
It was observed in the 0.01 M sodium sulfate electrolyte
system that a black precipitate was formed in the electrolyte
when the anode traversed to high overpotential. The identica-
tion of this precipitate was not possible without chemical testing
but on exposure to air, there was a color change from black to
orange suggesting that the precipitate was an iron based com-
pound possibly iron (II) oxide which could oxidize to iron (III)
oxide in air. Corrosion products of Fe(OH)
2
and Fe(OH)
3
have been
detected in sulfate systems under similar deaerated conditions
but in more alkaline solutions (Simard et al., 2001). The Pourbaix
diagram for iron (Pourbaix, 1974) also suggests that Fe(OH)
2
is
the likely corrosion product of iron within this anodic potential
region.
3.3. Sodium sulfate and sodium chloride system
The corrosion effect due to chloride was measured by prepar-
ing a 0.01 M sodium sulfate electrolyte with the addition of 0.2 M
sodium chloride. Ej data in triplicate can be seen for the
corrosion system in Fig. 5. A similar trend in the data, i.e. the
2 1 0 1 2
110
4
110
3
0.01
0.1
1
10
100
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
corr
= -0.742V
j
corr
= 0.107Am
-2
E
corr, f
E
corr, b
Fig. 3. Ej data for pure deaerated low conductivity water pH5.15 in triplicate.
Fresh electrolyte and RDE surfaces were prepared prior to each potential sweep.
( ) replicate 1, ( ) replicate 2 and ( ) replicate 3. E
corr,f
and E
corr,b
refer to the corrosion potentials measured for the potential sweep in the
anodic (forwards) direction and that in the cathodic (backwards) direction,
respectively.
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
corr
= -0.863V
Fig. 4. Ej data for three separate replicates of a corrosion system of deaerated
0.01 M sodium sulfate. ( ) replicate 1, ( ) replicate 2 and
( ) replicate 3.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5779
shift in potential between the forward and backward potential
sweep directions, to that of the low conductivity water and
sulfate system, can be observed. The magnitude of the potential
shift, however, in this system is considerably greater and equal to
0.3 V compared to 0.151 V in the 0.01 M sulfate system.
The corrosion potential recorded for the forward potential
sweep is also observed to shift in the cathodic direction by
0.104 V with the addition of 0.2 M sodium chloride, from
0.863 V measured in the 0.01 M sulfate system, to 0.967 V
measured in the 0.2 M chloride and 0.01 M sulfate system. An
increase in the corrosion current density, j
corr
, was also measured
with a j
corr
of 3.0 Am
2
for the 0.2 M chloride and 0.01 M sulfate
system which is comparable to literature (Ateya et al., 2002) and
an order of magnitude greater than that of the 0.01 M sulfate
system with j
corr
of 0.5 Am
2
. The shift in corrosion potential and
increase in corrosion current density can be attributed to the
addition of 0.2 M chloride in the system, since chloride is well
known to facilitate lm breakdown at the surface of steel
(Al-Khara et al., 2002; Ateya et al., 2002; El-Egamy and Badaway,
2004; Laycock and Newman, 1997; Meguid and Latif, 2004).
A plateau in the current response can be seen in the anodic
region of the forward potential sweep between the potentials
0.967 and 0.650 V. This plateau suggests that the corrosion
potential falls within the passivation region of the anodic character-
istics of iron. The current plateau can be attributed to the develop-
ment of a metal oxide at the surface, or due to the deaerated nature
of the electrolyte and low alloying content of the steel, a protective
salt lm (Landolt, 2007). The potential, E
F
, which is called the Flade
potential, at 0.650 V, denotes where the lmbegins to break down
and suggests the onset of localized corrosion which increases in rate
as the anodic overpotential increases.
The effect of sodium sulfate concentration was examined using
deaerated low conductivity water electrolytes containing 0.2 M
sodium chloride and two different concentrations of sodium
sulfate; 0.0002 and 0.01 M against the 0.2 M sodium sulfate only
solution. The effect of sulfate concentration can be seen in the Ej
data for the two systems of different sulfate concentrations and
the control presented in Fig. 6. It can be seen that the increase in
sodium sulfate concentration causes the equilibrium potential to
shift in the cathodic direction from 0.882 V in the control,
to 0.900 V in the system containing 0.0002 M sulfate, and then to
0.946 V with sulfate concentration of 0.01 M. A cathodic shift in
corrosion potential correspondingly results in an increase in j
corr
.
Ej data showing the effect of increasing sodium chloride
concentration can be seen in Fig. 13. Three 0.0002 M sodium
sulfate electrolytes with increasing sodium chloride concentra-
tions of 0.025, 0.2 and 0.8 M have been examined. An increase in
j
corr
with increasing chloride concentration can be initially
observed, as can be slight but consistent anodic shift in the
corrosion potential with increasing chloride concentration. The
pH measured for these corrosion systems of increasing chloride
concentration also increases with values of 5.03, 5.36 and 5.70 for
0.025, 0.2 and 0.8 M chloride, respectively. As proton reduction is
a primary cathodic process in this system it is expected that the
cathodic reaction also increases in rate to balance the increase in
the anodic corrosion process. It appears that the plateau in anodic
Ej data depicting a salt lm is more prominent in where the
chloride concentration is greater. This may also be due to the
increased dissolution rate of iron (II) from the surface which in
turn facilitates the formation of a salt lm (Landolt, 2007). The
Flade potential, E
F
, for this system is observed at 0.560 V,
beyond which pitting is expected to occur, along with a sharp
increase in current density, as shown in Fig. 7.
3.4. Hydrogen sulde system
The effect of synthetically produced hydrogen sulde on a
corrosion system was examined by the addition of hydrogen
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
Increasing sulfate
concentration
E
corr (3)
= -0.946V
(3)(2)(1)
E
corr (2)
= -0.900V
E
corr (1)
= -0.882V
Fig. 6. Ej data for two corrosion systems of increasing sulfate concentration.
( ) 0.2 M sodium chloride, pH5.61 ( ) 0.0002 M and sodium
sulfate 0.2 M sodium chloride, pH5.36 and ( ) 0.01 M sodium sulfate
0.2 M and sodium chloride, pH4.99.
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
F
= -0.650V
E
corr
= -0.967V
Fig. 5. Ej data for three separate replicates of a corrosion system of steel
subjected to an electrolyte of deaerated 0.01 M sodium sulfate and 0.2 M sodium
chloride in triplicate. pH4.99. ( ) replicate 1, ( ) replicate 2 and
( ) replicate 3.
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
Increasing chloride
concentration
E
F
= -0.560V E
corr (3)
= -0.891V
(2) (3)
E
corr (2)
= -0.904V
E
corr (1)
= -0.916V
(1)
Fig. 7. Ej data for three corrosion systems of increasing sodium chloride concentra-
tion. ( ) 0.0002 M sodium sulfate and 0.025 M sodium chloride, pH5.03
( ) 0.0002 M sodium sulfate 0.2 M and sodium chloride, pH5.36 and
( ) 0.0002 M sodium sulfate and 0.8 M sodium chloride, pH5.70.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5780
sulde to a deaerated low conductivity water electrolyte contain-
ing 0.01 M sodium sulfate and 0.2 M sodium chloride. Ej data for
this system can be seen in triplicate in Fig. 8 and shows the
presence of noise in current values and greater deviation between
measured corrosion potentials, E
corr
, of each replicate. The range
in E
corr
measured for the three replicates was 0.123 V with mean
E
corr
of 0.912 V. Although noisy data were observed in the
sodium sulfate system, the magnitude of noise and variation in
E
corr
is greater in the presence of hydrogen sulde, which shows
the formation of unstable surface lms. In fact a dark gray/black
precipitate was observed to develop in each corrosion system
both with and without added hydrogen sulde. In the system
with hydrogen sulde, however, the precipitate adhered to the
surface and was resistant to removal by ultrasonication suggest-
ing the precipitate had changed into an adherent protective lm
at the surface.
A study conducted to determine the corrosion products of iron
with hydrogen sulde (Shoesmith et al., 1980) reported large varia-
tion in corrosion potential measurement of over 0.030 V conducting
galvanostatic potentiometry experiments. The corrosion potential
was observed to shift in the initial 10 min of their experiments by
up to 0.150 V which is comparable to ours. In that study X-ray
diffractometry was used (Shoesmith et al., 1980) to characterise the
corrosion products at the surface yielding three ferrous monosulde
phases, mackinawite, troilite (FeS) and cubic ferrous sulde. It was
concluded (Shoesmith et al., 1980) that mackinawite and cubic
ferrous sulde form unpassivating layers at the surface which allow
local corrosion attack until more stable slowly forming troilite forms a
stable passive layer at the surface protecting the surface from further
corrosion. A similar study (Huang et al., 1996) identied that sulde
lm morphology is dependent on time.
Fig. 9 shows a comparison in Ej data between deaerated low
conductivity water sodium sulfatesodium chloride electrolytes
with added hydrogen sulde. It can be seen that the j
corr
for the
corrosion system containing hydrogen sulde, measured at
approximately 4.0 Am
2
, is approximately twice that of the
corrosion system with no hydrogen sulde, measured at
2.1 Am
2
. The current density response at high overpotentials
however, in the anodic potential region of Ej data, indicates
hydrogen sulde decreases the rate of corrosion. It is also evident
that an oxidation peak, E
ox
, is present in the anodic region for the
corrosion system containing hydrogen sulde which is not pre-
sent in systems without hydrogen sulde. This suggests that a
protective lm forms on the surface when overpotential is high in
the presence of hydrogen sulde, which can be oxidized as was
observed by Shoesmith et al. (1980).
The effect of lowering the sulfate concentration in the pre-
sence of hydrogen sulde was examined by using a deaerated low
conductivity water electrolyte containing 0.2 M sodium chloride
with synthetic hydrogen sulde but in the absence of sodium
sulfate. There was signicant noise in the Ej data for this system.
The standard deviation in the corrosion potential for this system
is calculated to be 0.021 V which is low compared to 0.050 V for
the corrosion system containing 0.01 M sodium sulfate but shows
that signicant variation remains present and conrms that this
variation is due to the presence of aqueous hydrogen sulde.
The notable item was that the oxidation peak observed at
0.700 V, in Ej data for the corrosion system containing 0.01 M
sodium sulfate and 0.2 M sodium chloride with the addition of
hydrogen sulde, was also present in Ej data for this system. The
magnitude of the peak, however, was lower. Ej data presented in
literature (Shoesmith et al., 1978) show a similar oxidation peak
to the one observed in our study but at a more anodic potential of
0.405 V, in alkaline conditions of pH 912. Black precipitates of
black iron sulde similar to those in our study, shown in Fig. 11,
were also reported to form at the surface in their experiments
where X-ray diffractometry was used to deduce that the pre-
cipitates were mackinawite.
The corrosion potential measured for the experimental system
containing hydrogen sulde appeared to occur in a more anodic
position compared to that of the system without sulde which is
consistent with experimental studies presented in literature
(Videm and Kvarekval, 1995). A dark gray/black precipitate was
observed to form in this system and a dark gray/black lm
developed at the RDE surface. The current density response for
the system containing hydrogen sulde showed little deviation
from the system without hydrogen sulde in the region close to
E
corr
. At high overpotentials however, the current density was
lower suggesting that the lm formed offered protection against
corrosion which has also been observed in literature (Shoesmith
et al., 1978; Shoesmith et al., 1980).
4. Model development
4.1. Corrosion model: deaerated water
In this corrosion system, the steel surface is in contact with
low conductivity water electrolyte forming a solidliquid inter-
face. It is assumed that the electrons formed in iron oxidation
process in this system described by Eq. (3) are consumed by
the proton reduction process at the surface described by Eq. (4).
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
F
= -0.657V
Fig. 8. Ej data for the corrosion system of deaerated 0.01 M sodium sulfate and
0.2 M sodium chloride and H
2
S in triplicate. ( ) replicate 1, ( )
replicate 2 and ( ) replicate 3.
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
ox
= -0.700V
Current density is
greater in the
system with no
hydrogen sulfide
at high anodic
overpotential
Oxidation peak
Fig. 9. A comparison between a corrosion system containing ( ) 0.01 M
sodium sulfate and 0.2 M sodium chloride, and a system containing ( )
0.01 M sodium sulfate and 0.2 M sodium chloride and hydrogen sulde.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5781
This allows a balance to be made between the oxidation and
reduction processes in the system where the processes are related
by the transfer of electrons or charge.
Fe-Fe
2
2e

3
H

-1=2H
2
4
The transfer of charge in this system occurs only at the steel
surface due to the nature of electron transport but the rate of
transfer is dependent on the reaction kinetics and mass transfer of
protons to the steel surface from the electrolyte. The overall
corrosion reaction of this initial system is described by
Fe2H

-Fe
2
H
2
5
The schematic in Fig. 12 for the overall corrosion reaction
shows that protons at the steel surface consume electrons and
cause iron (II) ions to be released from the steel into the
electrolyte. The protons are then released from the surface as
hydrogen gas.
The electrolyte in this corrosion system is assumed to be well
mixed in the bulk with a diffusion layer established at the steel
surface. The thickness of the diffusion layer can be determined by
employing the Nernst Diffusion model. At the point d
N
from the
surface, the concentration of reacting species is the same as that
in the bulk. The diffusion layer thickness however, is specic to
each species since they have differing diffusivities.
4.1.1. Charge transfer to the surface
The charge transfer at the steel surface for each species is
described by the ButlerVolmer equation. In general, this is
written as in Eq. (6), with subscript i denoting reacting species
such as Fe or H

.
j
i
j
0,i
exp
a
a,i
n
i
F
RT
Z
i

exp
a
c,i
n
i
F
RT
Z
i

6
The kinetic parameters for each species, exchange current
density, j
0
, and transfer coefcients for the cathodic and anodic
processes, a
c
and a
a
, respectively, were determined from a Tafel
analysis applied to Ej data collected in the laboratory. If atten-
tion is focused on region (I) in Fig. 13, where the cathodic
overpotential is high, the predominating process at the steel
surface is the cathodic reduction of protons. The cathodic Tafel
slope of region (I) therefore, was used to determine kinetic
parameters for the reduction of protons at the steel surface. The
cathodic transfer coefcient for proton reduction, a
c,H
, was
determined from a Tafel analysis to be 0.4. The anodic transfer
coefcient for iron, a
a,Fe
, was determined in the same manner by
focusing attention on region (II) in Fig. 13, where the predominat-
ing process is the anodic oxidation of iron. The Tafel slope of
region (II) yielded a value for a
a,Fe
of 0.4. The experimental data
cannot be used to predict a
c,H
or a
a,Fe
, because there is consider-
able interference from each reaction near the corrosion potential.
The exchange current density, j
0
, was determined by using
experimental data and the denition of j
0
in Eq. (7) (Newman and
Thomas-Alyea, 2004). The bulk concentration, C
b
, of iron (II) was
initially set at 110
9
mol dm
3
to simulate iron (II) concentra-
tion of approximately zero, with iron (II) surface concentration, C
s
,
set at 110
7
mol dm
3
.
j
0,i

e
0,i
A
C
s,i
C
b,i

g
i
7
The concentration sensitivity parameter, g
i
, was determined
for iron by examining the change in concentration of iron (II) at
the surface. In the Ej data plot for deaerated water, a potential
shift can be observed in the reverse potential scan direction as a
result of the change in iron (II) surface concentration. The
magnitude of the potential shift, approximately 0.12 V, was used
with the Nernst equation, and a surface concentration value of
iron (II) for the forward potential scan of 110
7
mol dm
3
,
to
estimate the change in iron (II) concentration at the surface to be
110
3
mol dm
3
. The concentration sensitivity parameter for
iron (II), g
Fe
, was determined to be 0.3.
The bulk proton concentration was determined from pH
measurement in the laboratory. The surface proton concentration
was assumed to be close to zero in this system due to the proton
reduction current approaching its diffusion limiting value. The
concentration sensitivity parameter for proton reduction was set
to 0.75 in order for the concentration terms in the exchange
current density and reversible potential calculation to be in
agreement.
The reversible potential E
rev
for each component was deter-
mined using the Nernst equation, where the activity is assumed
to be the same as the surface concentration, C
s
, as in Eq. (8), since
the electrolyte is a dilute solution. Literature values for the
standard potentials, U
0
, were used (Landolt, 2007) in the model.
The charge transferred is expressed as the number of electrons
transferred, n, in each reaction, Faradays constant, F, tempera-
ture, T and the universal gas constant, R, are also required.
E
rev,i
E
i
1
2:3RT
n
i
F
logC
s,i
8
The overpotential, Z, for each species was calculated using
Eq. (9) and the application of mixed potential theory. Steady-state
oxidation and reduction reactions are established in the corrosion
system between the oxidation of iron and the reduction of
protons. This yielded a balance in the transfer of charge and at a
measured current density of zero. The potential, E, at which this
occurs is the corrosion potential, E
corr
. The resulting overpotential
for each species was calculated using Eq. (10).
Z
i
EE
rev,i
9
Z
i
E
corr
E
rev,i
10
4.1.2. Mass transfer to the surface
The transfer of reduction species to the surface and oxidation
products from the surface in a corrosion system affects the rate of
corrosion signicantly. A material balance shown in Eq. (11) was
used to describe the ux, N, of species at the surface.
@C
i
@t
rN
i
R
i
11
Since steady-state charge transfer was established at the steel
surface and this simple corrosion system contains no homoge-
neous reactions in the bulk electrolyte, since the protons released
from water dissociation are already included as per the pH
measurement. The net ux at the surface is equal zero, describing
the system at steady-state. In the absence of an external applied
potential eld the, ux of each species is described using
N
i
D
i
rC
i
12
Since the ux for each species, N, in the corrosion system is
related to the current density, j, using the value of charge transfer,
n and Faradays constant, F, as shown in
N
i

j
i
n
i
F
13
A ux balance arising from the dependency on diffusion and
charge consumption is expressed in Eq. (14) for each species. The
negative term on the right-hand side denotes that the ux is a
vector and that the ux occurs in the opposite direction to
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5782
concentration gradient.
j
i
n
i
F
D
i
C
b,i
C
s,i
d
Ni

14
4.1.3. Solution technique and convergence criteria
The model was set up to seek the corrosion potential based on
the calculation of surface concentrations for each species and by
employing convergence criteria. Since the number of electrons
created in the oxidation process is equal to the number of
electrons consumed by the reduction process, and each reaction
can be expressed as current, the total current in the system i.e.
sum of the current due to oxidation and reduction, must equate to
zero. This is expressed in Eq. (15) in terms of the oxidation and
reduction component current densities.
j
T
j
Fe
j
H
15
where j
T
0, or smaller than a tolerance or residual. This was used
as a convergence criterion in the model to locate the corrosion
potential of the system. The model calculates the partial currents
based on the surface concentrations required to meet the con-
vergence criteria. The corrosion rate was then equal to the current
density for iron oxidation.
This solution technique was written using a commercial soft-
ware package, Mathcad 14. The Solve block facility was employed
in the workspace for the input of modeling equations and
specication of convergence criteria. The function Find was then
congured to solve the set of non-linear equations to determine
the corrosion potential and component currents in the corrosion
system. Initial values in the model solution were based on
estimates of corrosion potential and surface concentration of
components in the model. The convergence tolerance for the
solution algorithm in Mathcad 14 (which controls the precision to
which differentials and integrals are evaluated) was xed at a
value of 0.001.
4.2. Reaction system: sulfate
There are many reactions involving sulfate or sulfur com-
pounds that are possible in this corrosion system due to the
ability of sulfur to adopt oxidation states ranging from 6 when
present in sulfate compounds to 2 when present as sulde.
Since the polarization experiments conducted on this system
involve scanning the potential in the anodic direction from a start
potential of 1.5 V cathodic to the measured open circuit potential
(OCP), through the OCP to a potential 1.0 V anodic to the OCP, it is
possible that sulfate could be reduced at the surface. Since the
standard potentials of sulfatesulte and other redox couples are
known, a network of possible reactions containing sulfur species
from an initial source of sulfate occurring at the surface, at
different applied overpotential, has been devised.
The standard potential of the sulfatesulte couple, described
in Eq. (16) is reported as 0.936 V vs. SHE (Bard et al., 1985),
which equates to 1.177 V vs. SCE.
SO
2
4
2e

H
2
O-SO
2
3
2OH

E
B
1 1:177 V vs: SCE 16
With use of the Nernst equation, the reversible potential of the
sulfatesulte couple in this system was calculated to be
1.345 V vs. SCE when the bulk concentration of sulfate is xed
at 0.01 M. The applied potential during the Ej data collection for
this sulfate system was initiated at 1.447 V vs. SCE. which is
conducive for sulfate reduction to sulte. Since sulte was then
present at the surface it was possible for this to be reduced via
two pathways. Either to form thiosulfate, Eq. (17), as observed in
crevice corrosion studies of stainless steel containing sulfur
deposits (Lott and Alkire, 1989) or via a multistep pathway via
reactions described in Eqs. (18) and (19) forming sulde at the
surface.
2SO
2
3
4e

3H
2
O-S
2
O
2
3
6OH

E
B
1
0:817 V vs: SCE
17
SO
2
3
4e

3H
2
O-S6OH

E
B
1
0:900 V vs: SCE 18
S2e

-S
2
E
1
0:688 V vs: SCE 19
The standard potential suggests pathway via Eqs. (18) and (19)
is most likely and allowing the formation of sulde at the surface,
which was observed as a dark gray precipitate at the electrode
surface and in the electrolyte. The abiotic corrosion model allows
either mechanism to be investigated based on knowledge of the
bulk concentration of sulfate in the system.
Proton and sulfate reduction reactions are therefore assumed
to form the cathodic processes in this system. Since the pH in the
sulfate system was 5.60, proton reduction is another reduction
reaction that needs to be included.
Iron oxidation was considered to be the dominant oxidation
reaction in this system and the current contribution of iron
oxidation was determined as in the low conductivity water
system.
4.3. Reaction system: sulfate and chloride
The addition of chloride to the corrosion system containing
low conductivity water and sulfate promotes localized corrosion.
This has been well documented in literature and observed in this
study during abiotic corrosion experimentation. The effect that
chloride has on Ej data is described in the experimental results
for the 0.01 M sodium sulfate and 0.2 M sodium chloride corro-
sion system in previous sections. It was observed that the
presence of chloride increases the corrosion current density, j
corr
,
of the system but there was little shift, at this chloride concen-
tration, in the anodic Tafel slope which is comparable to observa-
tions from other experimental studies (AlKhara et al., 2002;
Ateya et al., 2002). This indicates that the iron oxidation reaction
remains largely unchanged when chloride was present at this
concentration other than an increase in the current density, j
corr
.
The effect of chloride at this concentration can be therefore
modeled by adjusting the exchange current density value for
the iron oxidation process in the model to match that of the
experimental system.
4.4. Reaction system: sulfate, chloride and synthetic Hydrogen
sulde
Hydrogen sulde in the experiments was introduced to a
corrosion system of 0.01 M sodium sulfate and 0.2 M sodium
chloride synthetically by the reaction of iron sulde with sulfuric
acid. The reaction took place in a separate reaction vessel and
hydrogen sulde was delivered to the RDE surface by sparging
through the electrolyte. Hydrogen sulde was included in the
model by assessing the reactions that were likely to occur at the
surface. If the reaction system was through a redox process then it
was likely, from knowledge of the redox systems, that sulfur
species in this system will be present at the surface as either
sulfur or sulde. This is because the E
corr
of the experimental
system, Fig. 8, lies close to the equilibrium potential of the sulfur
sulde redox system. Two redox couples identied in literature
(Bard et al., 1985) linking sulfur with hydrogen sulde which are
presented in Eqs. (20) and (21). Each equation describes a redox
system which could be used to incorporate hydrogen sulde at
the surface and therefore a means of including hydrogen sulde
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5783
in the model.
Ss2e

2H

-H
2
Saq E
3
0:097 V vs: SCE 20
Ss2e

2H

-H
2
Sg E
3
0:067 V vs: SCE 21
It can be seen from these two reactions, however, that high
concentrations of protons i.e. a low pH, is required to proceed.
Since the experimental system was conducted at pH approaching
neutral, protons were only present in low concentrations. Reac-
tions described in Eqs. (20) and (21) would therefore occur very
slowly which suggest that hydrogen sulde would be controlled
by the supply of protons. The equilibrium potentials for the
processes described by Eqs. (20) and (21) are also signicantly
more positive than the corrosion potential of the experimental
system. This tells us that the reduced species, hydrogen sulde, in
Eqs. (20) and (21) is more stable.
In this case then, hydrogen sulde will then dissociate to
produce protons and bisulde ions as described by Eqs. (22) and
(23) (Bard et al., 1985).
H
2
Saq-H

HS

K
1
9:1 10
8
22
HS

-H

S
2
K
2
1:2 10
12
23
The presence of sulde ions as described by Eq. (23) allows the
inclusion in the model of a redox system between sulfur and
sulde at the surface as described in Eq. (19).
This reaction scheme shows that presence of hydrogen sulde
can contribute to the concentration of sulde at the surface by
dissociation rather than charge exchange with the surface. The
sulde concentration at the surface is, therefore, dependent on
the concentration of aqueous hydrogen sulde in the electrolyte
as well as the reduction processes of sulfate.
The corrosion cell in the experimental system for hydrogen
sulde inuence was initially sparged with nitrogen prior to
polarization. Hydrogen sulde was then fed into the system via
a sparger in the electrolyte which bubbled into the space above
the electrolyte. It is assumed that this space becomes enriched
with hydrogen sulde and equilibrium between hydrogen sulde
in the gas phase and hydrogen sulde in the liquid phase forms. If
this condition holds true then the concentration of hydrogen
sulde in the electrolyte can be calculated using the Henrys law
(Henry, 1803). Vapor equilibrium calculations are required in the
model to determine the mole fraction of hydrogen sulde in
the gas phase. As exit ports for the gas are present in the cell and
the cell temperature was xed at 30

C, constant temperature and


pressure conditions were used to determine the vapor fractions, y,
of hydrogen sulde and nitrogen in the gas phase. The total
pressure, P
T
, in the experimental system was xed at atmospheric
pressure. This allows the partial pressures, P, of nitrogen and
hydrogen sulde to be determined by using.
P
T
y
H
2
S
P
H
2
S
y
N
2
P
N
2
24
The gas phase total volume, V
T
, was determined using Eq. (25)
with knowledge of vapor fractions, y, and partial volumes, V, of
nitrogen and hydrogen sulde.
V
T
y
H
2
S
V
H
2
S
y
N
2
V
N
2
25
It was assumed in this problem that ideal gas behavior holds
true for each component. This allows the universal gas law for
each component to be used to determine the number of moles of
each component, n, in the system, Eqs. (26) and (27) for hydrogen
sulde and nitrogen, respectively.
P
H
2
S
V
H
2
S
n
H
2
S
RT 26
P
N
2
V
N
2
n
N
2
RT 27
The vapor fractions of hydrogen sulde and nitrogen can then
be determined from Eqs. (28) and (29), respectively, with knowl-
edge of the total number of moles in the system, n
T
.
y
H2S

n
H
2
S
n
T
28
y
N
2

n
N
2
n
T
29
Since the vapor fractions summate to unity and the total
number of moles is calculated by summation of the number of
moles of each component. The total pressure and volume in the
system was determined using the universal gas law for the
system expressed in
P
T
V
T
n
T
RT 30
A system of equations was then solved to determine the mole
fraction of hydrogen sulde in the gas phase. In the model the
system of equations was solved using the nd function in
Mathcad.
The Henrys law (Henry, 1803) outlined in Eq. (31) was used in
the model to determine the mole fraction, x, of hydrogen sulde
in the electrolyte.
x
H
2
S
H
H
2
S
y
H
2
S
P
T
31
A Henrys constant, H, of 60.889 MPa mole fraction
1
(Carroll
and Mather, 1988) for hydrogen sulde at a temperature of 30 1C
and 101.325 kPa was used. The number of moles of hydrogen
sulde entering the cell was determined using the stoichiometry
outlined in Eq. (2) used to produce hydrogen sulde from sulfuric
acid and iron sulde. With the assumption that an equilibrium
state holds true for the reaction outlined in Eq. (2), a reaction
balance was set up in Mathcad using Eq. (32) based on the
conversion, X, of sulfuric acid where the initial concentration of
sulfuric acid was evaluated at t0.
X
H
2
SO
4
1
C
H
2
SO
4
C
H
2
SO
4
9
t 0
32
The model was implemented with conversions of 50%, 75% and
100% of sulfuric acid to hydrogen sulde to deliver a range of
concentrations of hydrogen sulde to the corrosion cell. The
equilibrium balance for the reaction was expressed in Eq. (33)
where the initial concentrations of sulfuric acid and hydrogen
sulde are both evaluated at t0.
C
H
2
SO
4
9
t 0
C
H
2
SO
4
C
H
2
S
C
H
2
S9
t 0
33
The change in concentration of sulfuric acid from the initial
concentration of sulfuric acid gives rise to the change in concen-
tration of hydrogen sulde. As the initial concentration of hydro-
gen sulde was assumed to be zero, the concentration of
hydrogen sulde produced at 50%, 75% and 100% conversion
was determined. This value was then used as a known parameter
in the system of equations used to determine the mole fraction of
hydrogen sulde in the gas phase and subsequently the concen-
tration of hydrogen sulde in the electrolyte.
Assuming that the electrolyte consists of hydrogen sulde and
water only, the mole fraction of hydrogen sulde in the electro-
lyte was used in the model to determine the concentration of
hydrogen sulde in the electrolyte. Using the molar volume of
water and the total volume of the cell, the number of moles in the
electrolyte was determined and the concentration of hydrogen
sulde calculated. The solubility of hydrogen sulde of
0.158 mol% (Carroll and Mather, 1988) was used to ensure that
electrolyte saturation was not exceeded in the model.
Hydrogen sulde dissociation in the model was calculated
using Eqs. (22) and (23) and a molar balance of sulfur containing
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5784
species expressed in.
C
H
2
S9
t 0
C
HS

C
H
2
Saq
C
S
2 34
A species charge constraint expressed in Eq. (35) was also
included to maintain electroneutrality in the electrolyte.
C
H
C
HS

C
S
2 35
The concentration of protons and sulde were then included in
the bulk concentrations of corresponding redox processes occur-
ring at the surface.
5. Model results
5.1. Deaerated water
The corrosion model for a system of deaerated water was
based on a charge balance between the oxidation of iron within
steel, releasing iron (II) to the electrolyte and the reduction of
protons to hydrogen at the surface. Mass transport effects in the
model have been included by calculating the Nernst diffusion
layer thickness using literature values and measured values for
diffusion coefcients of reduction species in the electrolyte.
Corrosion parameters collected during the potential scan in the
anodic direction of the Ej data collected for this system have
been used in model development, it is important to note there-
fore, that model validation has been conducted using the para-
meters collected from potential the scan in the cathodic direction
and then to test the performance of the model under different
conditions. This means that the model t will be examined
against the corrosion potential (the value measured on the back-
ward potential scan i.e. the more anodic corrosion potential
presented in the experimental data).
The model prediction for the corrosion potential, E
corr
, for this
and each system was determined by setting the convergence
criteria, j
T
E0. In this system, a value of 110
4
Am
2
was used
as a tolerance level. This yielded a model value for E
corr
of
0.556 V vs. SCE. E
corr
measured for the potential sweep in the
cathodic direction in the experimental system holds values
ranging from 0.631 to 0.578 V vs. SCE with a mean of
0.609 V vs. SCE. The model value for E
corr
for this system is
0.053 V more positive than the mean experimental value for the
potential sweep in the cathodic direction. Since the Find sub-
routine in Mathcad allows for the search of local minima, no
further renement of the starting values for overpotential is
needed.
Fig. 14 shows that the model results appear to be offset in the
anodic direction compared to the experimental Ej data values for
a typical polarization experiment. This is likely due to discrepan-
cies between the estimated species surface concentrations in the
model and the surface concentration in the experimental system.
As mentioned earlier the bulk species concentrations in the
model are xed using the measured pH for proton concentration
and the Nernst equation with the magnitude of the observed
potential shift between the forward and backward potential
sweeps in the experimental data for iron (II) bulk concentration,
this may lead to over or under estimation of each component in
the model.
A range of convergence criteria, i.e. i
T
varying between 10
4
and 2.0 Am
2
, were examined to assess the limits of the model
and give insight into the accuracy of the model. The qualitative
trend yielded from this shows the model to t well with the
experimental data and it converged to nearly the same value of
E
corr
but there is a clear offset in the model results in the anodic
direction compared to the experimental data.
It can be seen for this system in Fig. 14 that the model
response cathodic to E
corr
follows closely the experimental Ej
data although as the cathodic overpotential increases, the model
values for potential begin to diverge from the experimental data.
Divergence from experimental data of a similar magnitude has
also been observed in a corrosion model (Ca ceres et al., 2007)
where oxygen reduction is the principal cathodic process and it is
reported that the model shows a good t with the experimental
data up to cathodic overpotentials of approximately 0.250 V,
thereafter divergence occurs. It was concluded in that model
(Ca ceres et al., 2007) that divergence between the model and the
experimental data at high overpotentials approaching 0.250 V
in the cathodic region is due to the onset of other cathodic
processes. Since the divergence in that case as here is far from
the corrosion potential, this does not signicantly affect the
model from locating the corrosion potential.
The model values at potentials cathodic to 0.700 V where
the current density convergence criterion in the model is set to
0.35 Am
2
tend to very large numbers, because this region is
under mass transfer control. The mass transfer to the surface in
the model is dependent on the diffusion (Landolt, 2007; Newman
and Thomas-Alyea, 2004) of protons to the surface which is xed
at 9.4710
9
m
2
s
1
in the model using literature (Newman and
Thomas-Alyea, 2004). Another value for the diffusion coefcient
of protons of 9.31210
9
m
2
s
1
, cited in the same literature
(Newman and Thomas-Alyea, 2004) was employed in the model
to test the sensitivity to changes in diffusion coefcient and
determine if this is the cause of divergence in the cathodic region
of the data. The corrosion potential for the model of 0.556 V,
employing a proton diffusion coefcient of 9.4710
9
m
2
s
1
,
remained unchanged from that employing a diffusion coefcient
of 9.3110
9
m
2
s
1
, showing that changes in diffusion coef-
cients have little effect on the value of the corrosion potential, as
has also been observed in literature (Ca ceres et al., 2007).
In the anodic direction, where iron oxidation occurs, shows a
reasonable t with experimental Ej data for potentials up to
0.500 V vs. SCE. The model then appears to diverge from
experimental values and suggesting that there is an increasing
of resistance within the system, which could be due to lm
formation at the surface which is not incorporated in the model.
At high overpotentials the overpotential values, again, tend to
very large numbers, due to passivation of the surface.
As shown in Fig. 14, the corrosion potential is predicted to be
0.556 V vs. SCE and the corrosion current is 0.129 Am
2
, which
equates to 0.15 mm per year. This is greater than the experimen-
tally measured corrosion current of 0.107 Am
2
by 0.022 Am
2
or 20%. The over estimation of corrosion current by the model for
this system is due to the anodic processes in the model not fully
taking into account the passivation properties of the steel when
subject to anodic polarization. Most design is carried out using a
range for materials corrosion data which is based on an order of
magnitude (Perry and Green); our results are well within the
current industry standards.
5.2. Sulfate system
A comparison made between the model results shows that the
trends in Ej data produced by the model yield a corrosion
potential of E
corr
for pathway (2) of 0.740 V vs. SCE and the
E
corr
estimated for pathway (1) of 0.734 V vs. SCE which are
virtually indistinguishable. The corrosion current density esti-
mated in pathway (2) is 0.12 Am
2
which was greater than that
of pathway (1), 0.1 Am
2
, the difference, however, was very small
and equates to a corrosion rate of 0.02 mm per year. It can be
seen, however, that the anodic Ej model values in pathway
(2) follow the trend in experimental data in the anodic region
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5785
more closely than that of pathway (1) which provides greater
credibility to pathway (2).
The model results appear to under predict the current density
in the cathodic region compared to the Ej data collected for the
experimental sulfate system. This suggests that the cathodic
reactions in the experimental system occur at a greater rate than
described by the model. The Ej data collected for the experi-
mental systems containing sulfate, however, showed signicant
noise and variation in both the current density and E
corr
values for
each replicate suggesting that the interaction processes at the
surface are complex and metastable Fig. 15.
5.3. Sulfate and chloride system
The model conguration for the corrosion system containing
0.01 M sodium sulfate and 0.2 M sodium chloride was similar to
that of the system containing only sulfate. Since chloride facil-
itates iron oxidation (Al-Khara et al., 2002; Ateya et al., 2002;
El-Egamy and Badaway, 2004; Laycock and Newman, 1997;
Meguid and Latif, 2004), the addition of chloride was incorporated
in the model by increasing the exchange current density of the
iron oxidation process at the surface. The model for this system
also included the effect of each reaction pathway for the reduc-
tion of sulfate, Eqs. (16)(19).
The model matches well with the experimental Ej for this
system up to anodic potentials of 0.500 V. The divergence at
higher potentials may be due to the assumption that only
oxidation of iron from the steel surface is included. It is possible
at the high overpotentials that other anodic processes are occur-
ring. It was also shown in a previous corrosion modeling study
(Ca ceres et al., 2007) that the iron (II) anodic transfer coefcient
can decrease with increasing concentration of chloride in the
system. Fig. 16 shows the model data with the anodic transfer
coefcient for iron (II) set to 0.25 as opposed to 0.4 which was
initially used in the low conductivity water system. The model t
in the anodic region is considerably better at high overpotentials
using an a
Fe
2 value of 0.25 as compared to 0.4.
The model prediction for E
corr
in this system with a
Fe
2 of 0.4 is
0.702 V vs. SCE which is 0.061 V cathodic to the experimentally
measured E
corr
. The model implemented with a
Fe
2 of 0.25 yields
a corrosion potential prediction of 0.721 V vs. SCE which is
0.080 V cathodic to the experimentally measured value, which
shows that the transfer coefcient is an important parameter.
It should be noted that value of the anodic transfer coefcient
for iron dissolution in earlier experimental work has also been
reported with values of 0.70 and even approaching 1.0 (Morris
et al., 1980). In the presence of chloride, since iron dissolution is
catalyzed, there may be a variation of j
0
with time and the anodic
transfer coefcient may also vary. Since this is difcult to
quantify, the results from our model can be considered to be in
reasonable agreement.
The model response in the cathodic region for this system
shows that calculated values of I
corr
are approximately one order
of magnitude lower than the experimental Ej data. This may be
due to the catalytic effects and lm forming nature of chloride
and sulfate system which are not described in the model. The
experimental Ej data also showed variation of a similar magni-
tude in the cathodic region for this system.
5.4. Sulfate, chloride and hydrogen sulde system
Three corrosion scenarios were developed to assess the model
predictions for a corrosion system containing hydrogen sulde.
An equilibrium calculation was set up in the model to determine
the concentration of protons and sulde produced by hydrogen
sulde for three conversions of 50%, 75% and 100% for the reaction
presented in Eq. (2) which equates to hydrogen sulde saturation
levels of 18%, 30% and 35% in the electrolyte, respectively. The
model results for the case of 18% hydrogen sulde saturation are
show in Fig. 17. The trend in the data t for the other conversions
was very similar.
The model response in the anodic region shows a reasonable
t with the experimental data until overpotentials of 0.600 V
are reached, thereafter a divergence was observed. However,
there was signicant noise and variation in the experimental
data, Fig. 8, suggesting that the processes occurring at the surface
are unsystematic due to formation of surface lms. These Ej
measurements coincide with other studies (Videm and Kvarekval,
1995) showing the unsystematic effects of aqueous hydrogen
sulde on Ej data. The divergence is attributed to the limited
description of passivity and lm formation in the model which
has been identied as the cause of increased resistance at the
surface (Huang et al., 1996). The model values for the cathodic
region in this system show similar characteristics to the other
systems presented in this work, where the cathodic current
densities are somewhat lower than the experimental data
6. Discussion
The results for the abiotic model systems have been presented
here which show that the qualitative trend in the model was in
broad agreement with the experimental data. There is also good
quantitative agreement with potentials higher than E
corr
for up to
0.500 V. It has been shown that the model responds effectively to
changes in the input parameters such as the anodic transfer
coefcient for iron. The model response in the region cathodic
to E
corr
shows limited agreement which suggesting that water
reduction and reduction of dissolved carbon dioxide maybe
required for more accurate predictions. The model was also used
to assess reaction pathways for the abiotic reduction of sulfate.
Although there was little difference in the model response for the
two different pathways, pathway (2) showed somewhat better
agreement. Further assessment of these differences could be
addressed in future work.
The initial deaerated water corrosion system was utilized to
demonstrate the most simple corrosion system containing the
anodic dissolution of iron to form iron (II) and the cathodic
reduction of protons to form hydrogen at the steel surface. The
Ej data collected from this system allowed the corrosion current
of approximately 0.14 Am
2
to be determined along with a
corrosion potential of 0.742 V vs. SCE. Corrosion current density
values of this magnitude represent corrosion rates of 0.16 mmy
1
which is relatively low when the threshold corrosion rate of
0.1 mmy
1
(Crolet, 2005) for mitigated corrosion systems is
considered.
The Ej data for the deaerated water system presented for
three replicate potential sweep experiments, shows that the
corrosion potential measured at 0.742 V vs. SCE was fairly
constant but there was variation in the current response between
each of the three replicates. This variation in current response is
due to the effect of low conductivity in the electrolyte which
varies from 65.4 to 104 mS cm
1
and may reect a systematic
error in the data accumulated through the electrolyte prepara-
tion. A variation in conductivity of 38.6 mS cm
1
measured for
between replicates induces a change in Ohmic drop up to
4000 O(Oelner et al., 2006) for electrodes of 0.6 cm diameter
used here. This range in Ohmic drop between replicates may be
reected in the variation observed in the Ej data.
The addition of 0.01 M sodium sulfate to the corrosion system
gave rise to an increase in the corrosion current density to
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5786
0.50 Am
2
which equates to 0.58 mm y
1
and is comparable to
other corrosion systems of similar pH and sulfate concentration
(Hernandez-Espejel et al., 2010). A signicant level of noise was
also recorded in the anodic current response in the presence of
sodium sulfate. Noise of similar character was also observed in
several studies presented in literature (Cheng et al., 1999; Simard
et al., 2001) where it was attributed to the formation and break-
ing of passive lms on the surface due to the presence of salt
precipitates. It is reasonable to draw similar conclusions in the
data for the sulfate corrosion system presented in this study as
black precipitates were also observed to form at the surface and
in the electrolyte. The formation of precipitates at the surface in
this system was also reected by the large variation in corrosion
potential of each replicate which was also observed in similar
studies (Hernandez-Espejel et al., 2010). Since it is known that
sulfate is able to form many species in solution (Bard et al., 1985)
it is likely that these species and their varying morphologies
interact with the surface giving rise to the variance in potential
observed in the experimental measurement. It was shown in one
study (Hernandez-Espejel et al., 2010) that the variation in open
circuit potential was dependent on immersion time of steel in an
electrolyte prepared to induce sulfate based lm formation at the
surface, there was, however, no obvious trends reported.
Experimental work in this study was also conducted to assess
the effect of chloride addition to the corrosion system. Since the
overall objective was to assess the corrosion inuence due to
components in the produced water, presented in Table 1, an
electrolyte of 0.01 M sodium sulfate and 0.2 M sodium chloride
was used. The effects on corrosion of this electrolyte were
observed as an increase in corrosion current density from
0.5 Am
2
measured in the sulfate system to 3.0 Am
2
with the
inclusion of chloride which is comparable to corrosion currents
measured in other corrosion studies (Ca ceres et al., 2007).
There was also evidence in the Ej data of pitting corrosion
which is widely understood to be facilitated by the presence of
chloride ions at the steel surface (Al-Khara et al., 2002; Ateya
et al., 2002; El-Egamy and Badaway, 2004; Laycock and Newman,
1997; Meguid and Latif, 2004). The characteristic lm breakdown
effect which causes pitting can be seen in the Ej data plots
containing chloride. A plateau can be observed in the anodic
current response close to the corrosion potential in both plots
which, as the overpotential is increased, suddenly gives rise to an
increase in current density. The potential at which this occurs is
called the Flade potential and pitting is expected to occur at
values of overpotential greater than this. The Flade potential for
the corrosion system containing 0.01 M sodium sulfate and 0.2 M
sodium chloride was recorded at 0.650 V and for the system
containing 0.0002 M sodium sulfate and 0.8 M sodium chloride
was 0.560 V. These values for the Flade potential are more
negative than several published in literature (Ateya et al., 2002;
Videm and Kvarekval, 1995) but this is likely to reect differences
in the nature of the lms at the surface and their ability to resist
pitting induced by chloride at the surface.
It was clear that a presence of hydrogen sulde in the
electrolyte induced major changes in the interactions at the
surface compared to corrosion systems in the absence of hydro-
gen sulde. Figs. 810 show experimental Ej data for the
hydrogen sulde systems where there is evidence of lm forma-
tion at the surface through the development of an oxidation peak
at 0.700 V vs. SCE and noise in the current density response.
This Ej data also show that the lms formed at the surface are
protective in nature due to the reduced magnitude in current
response as reported in several other studies (Hernandez-Espejel
et al., 2010; Shoesmith et al., 1978; Shoesmith et al., 1980).
The Ej data for the hydrogen sulde system shows a two-fold
increase in corrosion current density from 2.1 to 4.0 Am
2
. This
reects an increase in corrosion activity at the surface due to
increased presence of aggressive species in the electrolyte but the
Ej data also describes an increase in the resistance at the surface
with increases in overpotential in both the cathodic and anodic
directions compared to the sulfate-chloride system. This shows
the development of a protective lm at the surface of the RDE
during polarization, which were observed optically (Smith, 2010).
The modeling method is, however, different in the other
electrochemical corrosion models where experimental Ej data
is required in the model calculation or calibration. These required
steps may limit the usage of these models where experimental Ej
is not available. One benet of the model presented in this paper
is that an estimation of corrosion potential and corrosion current
density can be made without the need of collection of new
experimental data. Providing that estimation of the Tafel coef-
cients is available, the model will provide a prediction of corro-
sion rate. The accuracy of this method, however, is dependent on
the components included in the model. It is necessary to ensure
that all components in the corrosion system under investigation
are included in the model before predictions are sought.
1.5 1 0.5 0
110
4
110
3
0.01
0.1
1
10
100
110
3
110
4
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
F
= -0.629V
Fig. 10. Ej data for three separate replicates of a corrosion system of deaerated
0.2 M sodium chloride and H
2
S. The standard deviation in the corrosion potential
measured for this system was 0.021 V. The Flade potential, E
F
, showing lm
breakdown at the surface during the potential sweep in the anodic direction can
be seen at a potential of 0.629 V vs. SCE.
Dark Area
Dark
Area
Light Area
Fig. 11. Optical micrograph of RDE tip after exposure to a deaerated low
conductivity water electrolyte containing 0.01 M sodium sulfate electrolyte and
polarization experimentation at rotation speed of 800 rpm. Dark areas outlined
show areas of lm formation on the surface.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5787
Application of the model in the eld for corrosion prediction
without addressing this issue could lead to inaccuracies.
When the model is compared to other corrosion models based
on electrochemical processes (Ca ceres et al., 2007; Nesic et al.,
1996; Picioreanu et al., 2007) which present current density Ej
data in the cathodic region, it is apparent that these models include
reduction components such as carbon dioxide explicitly rather
than indirectly through pH and conductivity measurements.
7. Conclusions
The usefulness of this model lies in the fact that it treats the
chemical reaction system in a holistic fashion. Although there are
elaborate models for pitting and other localized corrosion
mechanisms, they require extensive knowledge about the mate-
rial, its origin and the process. This model uses the reaction
I II
1 0.8 0.6 0.4
110
7
110
6
110
5
110
4
110
3
0.01
0.1
1
10
Potential vs. SCE / V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|

/

A

m
I II
E
corr
Fig. 13. Tafel slopes used in model development. In Region (I), the cathodic
reduction of protons predominates. In region (II), the anodic dissolution of iron
predominates.( ) Experimental data, (- - - - - -) proton reduction Tafel
slope and ( ) iron dissolution Tafel slope.
1.5 1 0.5 0
110
5
110
4
110
3
0.01
0.1
1
10
100
110
3
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
E
corr
model = -0.556V
E
corr
cathodic sweep
experimental data =
-0.609V
Anodic model
shows similar
trend to
experimental
data
Fig. 14. Model results for a corrosion system containing deaerated low conduc-
tivity water. ( ) Experimental Ej data ( ) model response in potential due to
imposed current density convergence criteria. Diffusion coefcient for protons,
D
H
9.4710
9
m
2
s
1
.
1.5 1 0.5
110
5
110
4
110
3
0.01
0.1
1
10
100
110
3
Potential vs. SCE / V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|

/

A

m
-
2
Fig. 15. Model results for a corrosion system containing deaerated 0.01 M sodium
sulfate. ( ) Experimental Ej data, ( ) model response in potential due to
imposed current density convergence criteria where sulfate is assumed to reduce
at the surface to sulte which in turn reduces to thiosulfate. ( ) Model response
when sulfate is assumed to reduce to sulte, sulfur and subsequently sulde at the
surface.
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
1.5 1 0.5
110
5
110
4
110
3
0.01
0.1
1
10
100
110
3
Fig. 16. Model results for a corrosion system containing deaerated 0.01 M sodium
sulfate and 0.2 M sodium chloride. ( ) Experimental Ej data ( ) model response
in potential due to imposed current density convergence criteria when a
Fe
0.4.
( ) Model response when a
Fe
0.25.
H
+
OH
-
H
+
OH
-
H
+
OH
-
H
+
OH
-
H
2
O H
+
+OH
-
H
+
OH
-
H
+
OH
-
H
2
H
+
OH
-
Fe
2+
2e
-
Fe
Diffusion
near
surface
Convection
in bulk
D
i
s
t
a
n
c
e

f
r
o
m

R
D
E

s
u
r
f
a
c
e

N
Steel RDE Tip
Fig. 12. A schematic showing the surface interactions of steel with deaerated low
conductivity water. The Nernst diffusion layer thickness is also shown d
N
from the
steel surface.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5788
network data with process parameters and provides reasonable
agreement with corrosion rates measured experimentally.
A model was developed based on experimental data collected
for corrosion systems containing sulfate, chloride and hydrogen
sulde. Using the mixed potential theory the corrosion model
aimed to balance the charge transferred in reduction and oxida-
tion reactions at the surface and calculate the corrosion current
density. Mass transport of reduction species to the surface in the
abiotic model, is described by using the Nernst diffusion model
with knowledge of diffusion coefcients and the Nernst diffusion
layer thickness. Charge transfer at the surface is described using
ButlerVolmer kinetics. The rate of the iron oxidation reaction is
dependent on the sum of rates for all the reduction reactions
occurring at the surface. The abiotic corrosion model outputs, Ej
data, corrosion potential and corrosion current density, for the
model system are determined in the abiotic corrosion model
subject to current density convergence criteria.
The Ej data predicted by the abiotic model developed in this
work was found in general to match well with qualitative trends
observed in the experimental Ej data for the corrosion systems.
The quantitative t of the model to experimental data, however,
was variable over the range of Ej data. There was a good match
of E
corr
and I
corr
(within experimental variability) and small over-
potentials in the anodic region. There was a poor quantitative t
in general between the model and experimental Ej data in the
cathodic region. Deviations between the model and experimental
Ej data were recorded at high anodic overpotentials but in
several systems these deviations were eliminated through adjust-
ment of the anodic Tafel slope values used by the model.
Nomenclature
List of symbols
A electrode surface area (m
2
)
CR corrosion rate (mm y
1
)
C concentration of species (mol m
3
)
D diffusion coefcient (m
2
s
1
)
e
0
exchange current density transfer parameter
E potential (V)
E
corr
free corrosion potential (V)
E
F
ade Potential
E
rev
reversible potential for species (V)
E
0
equilibrium potential (V)
E1 standard electrode potential (V)
F Faraday constant 96485 C mol
1
H Henrys coefcient
j current density (A m
2
)
j
0
exchange current density (A m
2
)
M molar mass (g mol
1
)
N ux of reaction species
n number of electrons transferred in the process
n number of moles of species (moles)
P system or partial species pressure (Pa)
R universal gas constant8.314 J K
1
mol
1
R rate of homogeneous production (mol m
3
s
1
)
T temperature /(K)
t time (s)
V system or partial volume (m
3
)
X conversion
x mole fraction in the liquid phase
y mole fraction in the vapor phase
a transfer function for redox process
g concentration sensitivity parameter
d
N
Nernst diffusion layer thickness (m)
Z overpotential (V)
List of subscripts
a anodic process parameter
c cathodic process parameter
i reaction species
T total system pressure, volume, moles parameter
References
Al-Khara, F.M., Ateya, B.G., Abdallah, R.M., 2002. Electrochemical behaviour of
low carbon steel in concentrated carbonate chloride brines. Journal of Applied
Electrochemistry 32, 13631370.
Alkire, R., Nicolaides, G., 1974. Differential aeration corrosion of a passivating
metal under a moist lm of locally variable thickness. Journal of the Electro-
chemical Society: Electrochemical Science and Technology 121, 183190.
Ateya, B.G., Khara, F.M.A., Abdalla, R.M., 2002. Electrochemical behavior of low carbon
steel in slightly acidic brines. Materials Chemistry and Physics 78, 534541.
Bard, A.J., Parsons, R., Jordan, J., 1985. Standard Potentials in Aqueous Solution,
1st Edition, CRC Press.
Ca ceres, L., Vargas, T., Herrera, L., 2007. Determination of electrochemical para-
meters and corrosion rate for carbon steel in un-buffered sodium chloride
solutions using a superposition model. Corrosion Science 49, 31683184.
Carroll, J.J., Mather, A.E., 1988. The solubility of hydrogen sulde in water from 0 to
90 1C and pressures to 1 MPa. Geochimica et Cosmochimica Acta 53, 11631170.
Cheng, Y.F., Wilmott, M., Luo, J.L., 1999. Transition criterion of metastable pitting
towards stability for carbon steel in chloride solutions. British Corrosion
Journal 34, 280284.
Crolet, J.-L., 2005. Microbial Corrosion in the Oil Industry: A Corrosionists View,
Petroleum Mircobiology. ASM Press, Washington, DC, pp. 143169.
El-Egamy, S.S., Badaway, W.A., 2004. Passivity and passivity breakdown of 304
stainless steel in alkaline sodium sulphate solutions. Journal of Applied
Electrochemistry 34, 11531158.
Hamilton, W.A., 2002. Microbially inuenced corrosion as a model system for the
study of metal microbe interactions: A unifying electron transfer hypothesis.
Biofouling 19, 6576.
Hardy, J.A., Bown, J.L., 1984. The corrosion of mild steel by biogenic sulphide lms
exposed to air. Corrosion 40, 650654.
Henry, W., 1803. Experiments on the quantity of gasses absorbed by water, at
different temperatures, and under different pressures. Philosophical Transac-
tions of the Royal Society London 93 (2943), 274275.
Hernandez-Espejel, A., Dominguez-Crespo, M.A., Cabrera-Sierra, R., Rodrigues-
Meneses, C., Arce-Estrada, E.M., 2010. Investigations of corrosion lms formed
on API-X52 pipeline steel in acid sour media. Corrosion Science 52,
22582267.
Huang, H.H., Tsai, W.T., Lee, J.T., 1996. Electrochemical behaviour of A516 carbon
steel in solutions containing hydrogen sulde. Corrosion 52, 708713.
Landolt, D., 2007. Corrosion and Surface Chemistry of Metals, Translated from the
revised second French edition ed EPFL Press.
Laycock, N.J., Newman, R.C., 1997. Localised dissolution kinetics, salt lms and
pitting potentials. Corrosion Science 39, 17711790.
1.5 1 0.5 0
110
5
110
4
110
3
0.01
0.1
1
10
100
110
3
Potential vs. SCE, V
|
C
u
r
r
e
n
t

D
e
n
s
i
t
y
|
,

A

m
-
2
Fig. 17. Model results for a corrosion system containing deaerated 0.01 M sodium
sulfate and 0.2 M sodium chloride and 18% hydrogen sulde saturation. ( )
Experimental Ej data ( ) Model response in potential due to imposed current
density convergence criteria.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5789
Lee, W., Characklis, W.G., 1993. Corrosion of mild steel under anaerobic biolm.
Corrosion 49, 186199.
Lott, S.E., Alkire, R.C., 1989. The role of inclusions on initiation of crevice corrosion
of stainless steel. Journal of the Electrochemical Society 136, 973979.
Maxwell, S., 2006. Monitoring the mitigation of MIC risk in pipelines. NACE, 2006.
Meguid, E.A.A.E., Latif, A.A.A.E., 2004. Electrochemical and SEM study on type 254
SMO stainless steel in chloride solutions. Corrosion Science 46, 24312444.
Morris, D.R., Sampaleanu, L.P., Veysey, D.N., 1980. The corrosion of steel by
aqueous solutions of hydrogen sulde. Journal of The Electrochemical Society
127, 12281235.
Morris, R., Smyrl, W., 1988. Galvanic interactions on periodically regular hetero-
geneous surfaces. AIChE Journal 34, 723732.
Morris, R., Smyrl, W., 1989. Galvanic interactions on random heterogeneous
surfaces. Journal of the Electrochemical Society 136, 32373248.
Nesic, S., Postlethwaite, J., Olsen, S., 1996. An electrochemical model for prediction
of corrosion of mild steel in aqueous carbon dioxide solutions. Corrosion 52,
280294.
Newman, J., Thomas-Alyea, K.E., 2004. Electrochemical Systems, Third ed. John
Wiley & Sons, Inc.
NORSOK, 2005. CO
2
corrosion rate calculation model.
Nyborg, R., 2009. Guidelines for Prediction of CO
2
Corrosion in Oil and Gas
Production Systems. Institute for Energy Technology.
Oelner, W., Berthold, F., Guth, U., 2006. The iR DropWell-known but often
underestimated in electrochemical polarisation measurements and corrosion
testing. Materials and Corrosion 57, 455466.
Perry, R.H., Green, D.W., Perrys Chemical Engineers Handbook. McGraw-Hill.
Picioreanu, C., Head, I.M., Katuri, K.P., Loosdrect, M.C.M.v., Scott, K., 2007.
A computational model for biolm-based microbial fuel cells. Water Research
41, 29212940.
Picioreanu, C., Loosdrecht, M.C.M.v., 2002. A mathematical model for initiation of
microbiologically inuenced corrosion by differential aeration. Journal of The
Electrochemical Society 149, B211B223.
Pots, B.F., John, R.C., Rippon, I.J., Thomas, M.J.J.S., Kapusta, S.D., Girgis, M.M.,
Whitman, T., 2002. Improvements on De Waard-Milliams corrosion prediction
and applications to corrosion management. NACE, 2002.
Pourbaix, 1974. Atlas of Electrochemical Equilibria in Aqueous Solutions.
Shoesmith, D.W., Bailey, M.G., Ikeda, B., 1978. Electrochemical formation of
Mackinawite in alkaline sulde solutions. Electrochimica Acta 23, 13291339.
Shoesmith, D.W., Taylor, P., Bailey, M.G., Owen, D.G., 1980. The formation of ferrous
monosulde polymorphs during the corrosion of iron by aqueous hydrogen
sulde at 21 1C. Journal of the Electrochemical Society 127, 10071015.
Simard, S., Odziemkowski, M., Irish, D.E., Brossard, L., Menard, H., 2001. In situ
micro-Ramen spectroscopy to investigate pitting corrosion product of 1024
mild steel in phosphate and bicarbonate solutions containing chloride and
sulfate ions. Journal of Applied Electrochemistry 31, 913920.
Smith, P.J., 2010. A Predictive Model for Microbiologically Inuenced Corrosion
(MIC) in Sub-Sea Production Pipelines, School of Chemical Engineering and
Advanced Materials. Newcastle University, Newcastle upon Tyne.
Videm, K., Kvarekval, J., 1995. Corrosion of carbon steel in carbon dioxide
saturated solutions containing small amounts of hydrogen sulde. Corrosion
Science 51, 260269.
P. Smith et al. / Chemical Engineering Science 66 (2011) 57755790 5790

Das könnte Ihnen auch gefallen