You are on page 1of 29

2.

25 Corrosion in Acid Gas Solutions


S. Nesic
Institute for Corrosion and Multiphase Technology, Ohio University, Athens, OH, USA
W. Sun
ExxonMobil Upstream Research Company, Houston, TX, USA

ß 2010 Elsevier B.V. All rights reserved.

2.25.1 Introduction 1273


2.25.2 Aqueous CO2 Corrosion of Mild Steel 1273
2.25.2.1 Chemistry of CO2 Saturated Aqueous Solutions – Equilibrium Considerations 1273
2.25.2.2 Electrochemistry of Mild Steel Corrosion in CO2 Saturated Aqueous Solutions 1277
2.25.2.2.1 Oxidation of iron 1278
2.25.2.2.2 Reduction of hydronium ion 1278
2.25.2.2.3 Reduction of carbonic acid 1278
2.25.2.2.4 Reduction of water 1279
2.25.2.3 Transport Processes in CO2 Corrosion of Mild Steel 1279
2.25.2.4 Calculation of Mild Steel CO2 Corrosion Rate 1280
2.25.2.5 Successes and Limitations of Modeling of Aqueous CO2 Corrosion of Mild Steel 1280
2.25.2.6 Key Factors Affecting Aqueous CO2 Corrosion of Mild Steel 1281
2.25.2.6.1 Effect of pH 1281
2.25.2.6.2 Effect of CO2 partial pressure 1281
2.25.2.6.3 Effect of temperature 1282
2.25.2.6.4 Effect of flow 1283
2.25.2.6.5 Effect of corrosion inhibition 1284
2.25.2.6.6 Effect of organic acids 1285
2.25.2.6.7 Effect of glycol/methanol 1285
2.25.2.6.8 Effect of condensation in wet gas flow 1285
2.25.2.6.9 Nonideal solutions and gases 1286
2.25.2.7 Localized CO2 Corrosion of Mild Steel in Aqueous Solutions 1286
2.25.3 Aqueous H2S Corrosion of Mild Steel 1286
2.25.3.1 Chemistry of H2S Saturated Aqueous Solutions – Equilibrium Considerations 1287
2.25.3.2 Mild Steel Corrosion in H2S and Mixed H2S/CO2 Saturated Aqueous Solutions 1289
2.25.3.3 Calculation of Mild Steel H2S Corrosion Rate 1291
2.25.3.3.1 Pure H2S aqueous environment 1291
2.25.3.3.2 Mixed CO2/H2S environments 1292
2.25.3.4 Limitations of Modeling of Aqueous H2S Corrosion of Mild Steel 1292
2.25.3.5 Key Factors Affecting Aqueous H2S Corrosion of Mild Steel 1293
2.25.3.5.1 Effect of H2S partial pressure 1293
2.25.3.5.2 Effect of flow 1295
2.25.3.5.3 Effect of time 1295
2.25.3.6 Localized H2S Corrosion of Mild Steel in Aqueous Solutions 1297
References 1297

A=V Surface to volume ratio (m1)


Abbreviations
AðFeCO3 Þ Constant in the Arrhenius-type equation for
CR Corrosion rate (mm year1)
krðFeCO3 Þ
AH2 S Solid state diffusion kinetic constant for H2S
through mackinawite film,
Symbols AH2 S ¼ 2:0  105 mol m2 s1
A Surface area of the steel (m2)

1270
Corrosion in Acid Gas Solutions 1271

AHþ Solid state diffusion kinetic constant for Hþ DH2 CO3 Aqueous diffusion coefficient of H2 CO3
through mackinawite film, (m2 s1)
AHþ ¼ 4:0  104 mol m2 s1 DrefðH2 CO3 Þ Reference aqueous diffusion coefficient
ACO2 Solid state diffusion kinetic constant for CO2 of H2 CO3 , Dref;H2 CO3 ¼ 1.3  109 m2 s1 at
through mackinawite film, 25 C
ACO2 ¼ 2:0  106 mol m2 s1 DHþ Aqueous diffusion coefficient for Hþ
baðFeÞ Anodic Tafel slope for Fe oxidation (V) DrefðHþ Þ Reference aqueous diffusion coefficient for
bcðHþ Þ Cathodic Tafel slope for H+ ion reduction (V) Hþ , DrefðHþ Þ ¼ 2.80  108 m2 s1 at 25  C
bcðH2 CO3 Þ Cathodic Tafel slope for H2CO3 DH2 S Aqueous diffusion coefficient for dissolved
reduction (V) H2 S
bcðH2 OÞ Cathodic Tafel slope for H2O reduction (V) DCO2 Aqueous diffusion coefficient for dissolved
BðFeCO3 Þ Constant in the Arrhenius-type equation CO2 , DCO2 ¼ 1.96  109, m2 s1
for krðFeCO3 Þ (kJ mol1) E Potential (V)
cCO2 Bulk aqueous concentration of CO2 (kmol m3) Ecorr Corrosion (open circuit) potential (V)
cCO2 Bulk aqueous concentration of CO2 3 ions ErevðFeÞ Reversible potential of Fe oxidation,
3
(kmol m3) ErevðFeÞ ¼ 0.488 V
cFe2þ Bulk aqueous concentration of Fe2þ ions ErevðHþ Þ Reversible potential for Hþ ion reduction (V)
(kmol m3) ErevðH2 CO3 Þ Reversible potential for H2 CO3
cHþ Bulk aqueous concentration of Hþ ions reduction (V)
(kmol m3) ErevðH2 OÞ Reversible potential for H2 O reduction
csðHþ Þ ‘Near-zero’ concentration of Hþ underneath (A m2)
the mackinawite film at the steel surface, set fH2 CO3 Flow factor for the chemical reaction
to 1:0  107 (kmol m3) boundary layer
cHCO3 Bulk aqueous concentration of HCO 3 ions F Faraday’s constant, F ¼ 96 485 C mol1 e
(kmol m3) FluxH2 S Flux of H2 S (kmol m2 s1)
cH2 CO3 Bulk aqueous concentration of H2 CO3 FluxHþ Flux of Hþ ions (kmol m2 s1)
(kmol m3) FluxCO2 Flux of CO2 (mol m2 s1)
cH2 S Bulk aqueous concentration of H2 S (kmol m3) HsolðCO2 Þ Henry’s constant for dissolution of CO2
cHS Bulk aqueous concentration of HS ions (bar kmol m3)
(kmol m3) DHFe Activation enthalpy for Fe oxidation,
ci Bulk aqueous concentration of a given aqueous DHFe ¼ 50 kJ mol1
species (kmol m3) DHðHþ Þ Activation enthalpy for Hþ ion reduction,
ciðH2 SÞ Aqueous concentration of H2 S at the inner DHðHþ Þ ¼ 30 kJ mol1
sulfide film/outer sulfide layer interface DHðH2 CO3 Þ Activation enthalpy for H2 CO3 reduction,
(kmol m3) DHðH2 CO3 Þ ¼ 57.5 kJ mol1
cS2 Bulk aqueous concentration of S2 ions DHðH2 OÞ Activation enthalpy for H2 O reduction,
(kmol m3) DHðH2 OÞ ¼ 30 kJ mol1
csðH2 SÞ ‘Near-zero’ aqueous concentration of i Current density (A m2)
H2 S underneath the mackinawite film icorr Corrosion current density (A m2)
at the steel surface, set to 1:0  107 iaðFeÞ Anodic current density of iron oxidation
(kmol m3) (A m2)
coðH2 SÞ Aqueous concentration of H2 S at the icðHþ Þ Cathodic current density for Hþ ion reduction
outer sulfide layer/solution interface (A m2)
(kmol m3) icðH2 CO3 Þ Cathodic current density for H2 CO3
csðCO2 Þ Aqueous concentration of CO2 underneath reduction (A m2)
the mackinawite film at the steel surface icðH2 OÞ Cathodic current density for H2 O reduction
d Characteristic dimension for a given flow (A m2)
d
geometry (m) ilimðHþ Þ Mass transfer (diffusion) limiting current
dp Diameter of a pipe (m) density for Hþ ion reduction (A m2)
r
dc Diameter of a rotating cylinder (m) ilimðH2 CO3 Þ Chemical reaction limiting current density
D Diffusion coefficient of a given species (m2 s1) for H2 CO3 reduction (A m2)
1272 Liquid Corrosion Environments

ioðFeÞ Exchange current density of iron oxidation KsolðCO2 Þ Solubility constant for dissolution of CO2
(A m2) (kmol m3 bar1)
ioðHþ Þ Exchange current density for Hþ ion reduction KspðFeCO3 Þ Solubility product constant for ferrous
(A m2) carbonate (kmol m3 bar1)
mackin
ioðH2 CO3 Þ Exchange current density for H2 CO3 KspðFeSÞ Solubility product constant for
reduction (A m2) mackinawite (kmol m3 bar1)
ioðH2 OÞ Exchange current density for water mos Mass of the outer sulfide layer (kg)
reduction (A m2) MFe Molecular mass of iron (kg kmol1 Fe )
ref
ioðFeÞ Reference exchange current density of Fe MFeS Molecular mass of ferrous sulfide
ref
oxidation, ioðFeÞ ¼ 1 A m2 (kg mol1 FeS )
ioðHþ Þ Reference exchange current density of Hþ
ref
n Number of electrons used in reducing or oxidizing
2
ref
oxidation, ioðH þ Þ ¼ 0.03 A m at Tc;ref ¼ 25  C a given species (kmole kmol1)
and pH 4 pCO2 Partial pressure of CO2 (bar)
ref
ioðH 2 CO3 Þ
Reference exchange current density for pH2 S Partial pressure of H2 S (bar)
H2 CO3 reduction, ioðH ref
2 CO3 Þ
¼ 0.06 A m2 at R Electrochemical reaction rate
Tc;ref ¼ 25 C, pH 5, and cH2 CO3 ;ref ¼ 104

(kmol m2 s1)
kmol m3 RFeCO3 Precipitation rate for iron carbonate
ref
ioðH2 OÞ Reference exchange current density for H2 O (kmol m3 s1)
ref
reduction, ioðH 2 OÞ
¼ 3  105 A m2 at R Universal gas constant, R ¼ 8.314 J mol1 K1

Tc;ref ¼ 20 C Re Reynolds number, Re ¼ vrH2 O d=mH2 O
iaðHþ Þ Charge transfer current density for Hþ ion Sc Schmidt number of a given species,
reduction (A m2) Sc ¼ mH2 O =ðrH2 O DÞ
iaðH2 CO3 Þ Charge transfer current density for H2 CO3 Shp Sherwood number of a given species
reduction (A m2) for a straight pipe flow geometry,
I Ionic strength kmol m3 Shp ¼ km dp =D
b
khyd Backward reaction rate of H2 CO3 dehydration Shr Sherwood number of a given species
reaction (1 s1), khyd b f
=khyd =Khyd for a rotating cylinder flow geometry,
f
khyd Forward reaction rate for the CO2 hydration Shr ¼ km dc =D
reaction (1 s1) SSðFeCO3 Þ Supersaturation of iron carbonate
kmðHþ Þ Aqueous mass transfer coefficient for Hþ ST Scaling tendency
(A m2) Tc Temperature ( C)
kmðH2 CO3 Þ Aqueous mass transfer coefficient for Tc;ref Reference temperature, Tc;ref = 25  C
H2 CO3 (A m2) Tf Temperature ( F)
kmðH2 SÞ Aqueous mass transfer coefficient for H2 S Tk Temperature (K)
(A m2) v Water characteristic velocity (m s1)
kmðCO2 Þ Aqueous mass transfer coefficient for CO2 zi Species charge of various aqueous species
(A m2) dmðH2 CO3 Þ Thickness of the mass transfer layer for
krðFeCO3 Þ Kinetic constant in the ferrous carbonate H2 CO3 (m)
precipitation rate equation (1 mol1 s1) drðH2 CO3 Þ Thickness of the chemical reaction layer
Khyd Equilibrium hydration constant for CO2 , for H2 CO3 (m)
Khyd ¼ khyd
f
=khyd
b
¼ 2:58  103 dos Thickness of the outer sulfide layer (m),
Kbi Equilibrium constant for dissociation of HCO 3 dos ¼ mos =ðrFeS AÞ
(kmol m3) Dt Time interval (s)
Kbs Equilibrium constant for dissociation HS mH2 O Water dynamic viscosity (Pa sÞ
(kmol m3) mH2 O;ref Reference water dynamic viscosity (Pa s) at
Kca Equilibrium constant for dissociation of H2 CO3 a reference temperature,
(kmol m3) mH2 O;ref ¼ 1:002  104 Pa s at 20  C
Khs Equilibrium constant for dissociation H2 S z H2 CO3 Ratio of the mass transfer layer and
(kmol m3) chemical reaction thicknesses for
KsolðH2 SÞ Solubility constant for dissolution of H2 S H2 CO3
(kmol m3 bar1) e Outer sulfide layer porosity
Corrosion in Acid Gas Solutions 1273

c Outer sulfide layer tortuosity factor 2.25.2.1 Chemistry of CO2 Saturated


rH2 O Density of water (kg m3) Aqueous Solutions – Equilibrium
rFe Density of iron (kg m3) Considerations
rFeS Density of ferrous sulfide (kg m3)
CO2 gas is soluble in water:
Ksol
CO2ðgÞ , CO2 ½2
2.25.1 Introduction
For ideal gases and ideal solutions in equilibrium,
Henry’s law can be used to calculate the aqueous
As oil and gas emerge from the geological formation,
concentration of dissolved CO2, cCO2 , given that the
they are always accompanied by some water and
respective concentration in the gas phase (often
varying amounts of ‘acid gases’: carbon dioxide,
expressed in terms of partial pressure, pCO2 ) is known:
CO2, and hydrogen sulfide, H2S. This is a corrosive
combination, which affects the integrity of mild steel. 1 pCO2
HsolðCO2 Þ ¼ ¼ ½3
This has been known for over 100 years; aqueous CO2 KsolðCO2 Þ cCO2
and H2S corrosion of mild steel still represents a
The CO2 solubility constant, KsolðCO2 Þ , is a function of
significant problem for the oil and gas industry.1
temperature, Tf , and ionic strength, I 2:
Although corrosion resistant alloys that are able to
withstand this type of corrosion exist, mild steel is KsolðCO2 Þ
often the most cost effective construction material 14:5 3 6 2
used in this industry for these applications. All the ¼  10ð2:27þ5:6510 Tf 8:0610 Tf þ0:075I Þ
1:00258
pipelines, many wells, and much of the processing
equipment in the oil and gas industry are built out of ½4
mild steel. The cost of equipment failure due to internal Ionic strength, I , can be calculated as
CO2/H2S corrosion is enormous, both in terms of
1X 2 1
direct costs such as repair costs and lost production, I¼ ci zi ¼ ðc1 z21 þ c2 z22 þ   Þ ½5
as well as in indirect costs such as environmental cost, 2 i 2
impact on the downstream industries, etc. The concentration of CO2 in the aqueous phase is of
The following section summarizes the degree of the same order of magnitude as the one in the gas
understanding of the so-called ‘sweet’ CO2 corrosion phase. For example, at pCO2 ¼ 1 bar, at 25  C, the gas-
and the so-called ‘sour’ or H2S corrosion of mild steel eous CO2 concentration is 4 mol l1 (kmol1 m3)
exposed to aqueous environments. It also casts the while in the water it is about 3 mol l1. Since the
knowledge in the form of mathematical equations solubility of CO2 decreases with temperature, at
whenever possible. This should enable corrosion 100  C, the respective concentrations are 3.3 mol l1
engineers and scientists to build entry level corrosion in the gas and 1.1 mol l1 in water.
simulation and prediction models. A rather small fraction (about 1 in 500) of the
dissolved CO2 molecules hydrates to make a ‘weak’
carbonic acid, H2CO3:
2.25.2 Aqueous CO2 Corrosion of Khyd
Mild Steel CO2 þ H2 O , H2 CO3 ½6

Aqueous CO2 corrosion of carbon steel is an electro- due to a relatively slow forward (hydration) rate.
chemical process involving the anodic dissolution of Assuming that the concentration of water remains
iron and the cathodic evolution of hydrogen. The unchanged, the equilibrium concentration cH2 CO3 is
overall reaction is determined by:
cH CO
Khyd ¼ 2 3 ½7
Fe þ CO2 þ H2 O ! FeCO3 þ H2 ½1 cCO2
The equilibrium hydration/dehydration constant,
CO2 corrosion of mild steel is reasonably well under- Khyd ¼ 2:58  103 , does not change much across
stood. A number of chemical, electrochemical, and the typical temperature range of interest (20–100  C).3
transport processes occur simultaneously. They are Carbonic acid is considered to be ‘weak’ because
briefly described below. it only partially dissociates in water to produce
1274 Liquid Corrosion Environments

hydronium, Hþ ions and bicarbonate ions, HCO


3: two eqns [12] and [13], there are three unknowns:
Kca
cHþ , cHCO3 , and cCO2
3
, and therefore one more equa-
H2 CO3 , Hþ þ HCO
3 ½8 tion is needed to close the system: a constraint that
describes charge conservation, that is, electroneutral-
The HCO 3 dissociates further to give some more H
þ
ity of the solution. Clearly, chemical reactions [8] and
2
and carbonate ion, CO3 : [9], which involve ions, always remain balanced with
Kbi respect to charge and therefore one can write
HCO þ
3 , H þ CO3
2
½9
cHþ ¼ cHCO3 þ 2cCO2 ½14
The respective equilibrium relations can be written as 3

cHþ cHCO3 Now, the system of equations is closed and concentra-


Kca ¼ ½10 tions of all the aqueous species can be determined,
cH2 CO3
cHþ cCO2 including the cHþ and the corresponding pH. The pH
Kbi ¼ 3
½11 of pure water as a function of pCO2 at room tempera-
cHCO3 ture is shown in Figure 1.
The equilibrium constants can be calculated as func- If there are other ions in the aqueous solution,
tions of temperature Tf , and ionic strength, I as2 such as for example Fe2þ produced by corrosion of
steel, then eqn [14] is extended to read
ð6:411:594103 Tf þ8:52106
Kca ¼ 387:6  10 Tf2 3:07105 p14:70:4772I 0:5 þ 0:118I Þ 2cFe2þ þ cHþ ¼ cHCO3 þ 2cCO2
3
½15
½12 By inspecting the equations above, one can see that,
3 5
ð10:614:9710 Tf þ1:33110 Tf2 2:624105 as iron dissolution causes an increase in cFe2þ , it is
Kbi ¼ 10 p14:71:166I 0:5 þ 0:3466I Þ accompanied by a decrease of cHþ due to the cathodic
½13
reaction and a corresponding increase in pH. Other
One can use the equations above to calculate the pH cations and anions as well as other chemical reactions
for a pure aqueous CO2 saturated system. Assuming can be introduced into the mix in a similar way.
that the concentration of CO2 (or partial pressure, An example of a CO2 aqueous species distribution as
pCO2 ) in the gas phase is known, one can calculate a function of pH for an open system is given in Figure 2.
the concentration of aqueous/dissolved CO2 cCO2 , It is worth noting that this simple water chemistry
via eqn [3]. Then the concentration cH2 CO3 can be calculation procedure is valid only for the case when
determined via eqn [7]. However, in the remaining the concentration of gaseous CO2, i.e., the partial

5
pH

Pure H2O
3 wt% NaCl
4

3
0.001 0.01 0.1 1 10 100
pCO2 (bar)

Figure 1 Calculated pH of a pure aqueous solution saturated with CO2 as a function of partial pressure of CO2; T ¼ 25  C,
1 wt% NaCl.
Corrosion in Acid Gas Solutions 1275

1.E+00

1.E–01 CO2(g)

Species concentration (mol l–1) CO2


1.E–02
HCO3–
1.E–03

H2CO3
1.E–04

1.E–05
Ferrous carbonate

1.E–06 CO23

1.E–07
2 3 4 5 6 7
Mild steel pH
Figure 2 Calculated carbonic species concentrations as a function of pH for a CO2 saturated aqueous solution; pCO2 ¼1 bar,
25  C, 1 wt% NaCl.

pressure, pCO2 is known, constant, and independent phases as conditions change. When one accounts for
from what is happening in the aqueous phase. This is this, an extra equation is obtained:
often referred to as an open system. It is relevant to
nCO2 ðgÞ þ nCO2 ðaqÞ þ nH2 CO3 ðaqÞ þ nHCO3 ðaqÞ
field situations where there is an overwhelming
amount of CO2 in the gas phase (such as seen in þ nCO2 ¼ Const: ½16
3 ðaqÞ
wet gas lines, multiphase pipelines, gas/liquid sep-
arators, etc.). In the lab setting, this condition is where n denotes the number of moles of a particular
easily achieved by continuous purge of a vessel with species in a gaseous or aqueous phase of a closed
gaseous CO2. system.
In contrast, there are many systems where there The dissociation steps [8] and [9] are very fast
is a limited amount of CO2 in the gas phase com- compared to all other processes occurring simulta-
pared to the amount in the liquid phase, such as in neously in corrosion of mild steel, thus preserving
oil well tubing, oil transportation lines, liquid/liquid chemical equilibrium. However, the CO2 dissolution
separators, etc. In the lab, aqueous systems with a reaction [2] and the hydration reaction [6] are much
limited gas phase are frequently found in high- slower. When such chemical reactions proceed
pressure autoclaves and flow loops. Consequently slowly, other faster processes (such as electrochemi-
they are often referred to as closed systems, and in cal reactions or diffusion) can lead to local nonequi-
principle can have varying gas/liquid volume ratios. librium in the solution.
An open system can be seen as a closed system with an Either way, the occurrence of chemical reactions
infinitely large gas/liquid volume ratio. In closed can significantly alter the rate of electrochemical pro-
systems, the concentration of gaseous CO2, that is, cesses at the surface and the rate of corrosion. This is
the partial pressure, pCO2 , is not known explicitly particularly true when, due to high local concentra-
and typically depends on the aqueous chemistry. In tions of species, the solubility limit of salts is exceeded
mathematical terms, this means that there is one and precipitation of a surface layer occurs. In a precip-
more unknown: pCO2 , and therefore one needs one itation process, heterogeneous nucleation occurs first
more equation to be able to solve for species con- on the surface of the metal or within the pores of an
centrations. The extra equation comes from the existing layer since homogenous nucleation in the
additional constraint: in a closed system, the total bulk requires a much higher concentration of species.
amount of carbonic species remains constant; they Nucleation is followed by crystalline layer growth.
are just redistributed between the gas and aqueous
1276 Liquid Corrosion Environments

Under certain conditions, surface layers can become


very protective and reduce the rate of corrosion.
In CO2 corrosion, when the concentrations of
Fe2þ and CO2 3 ions exceed the solubility limit,
they form solid ferrous carbonate according to
KspðFeCO3 Þ
Fe2þ þ CO2
3 , FeCO3ðsÞ ½17
Ferrous carbonate
where the solubility product constant for ferrous
carbonate KspðFeCO3 Þ is4
ð59:34980:041377Tk ð2:1963=Tk Þ Mild steel
KspðFeCO3 Þ ¼ 10 þ24:5724 log Tk þ2:518I 0:5 0:657I Þ ½18

Actually ferrous and carbonate ions are frequently


found in the aqueous solution at concentrations much
higher than predicted by the equilibrium KspðFeCO3 Þ .
This is termed supersaturation and is a necessary con-
dition before any substantial precipitation can occur.
The ferrous carbonate supersaturation, SSðFeCO3 Þ , is
defined as:
cFe2þ cCO2
SSðFeCO3 Þ ¼ 3
½19
KspðFeCO3 Þ
The precipitation process can be seen as the process
of the solution returning to equilibrium and is driven
by the magnitude of supersaturation. The rate of the
precipitation (<FeCO3 ) is therefore often expressed as
A  
<FeCO3 ¼ krðFeCO3 Þ KspðFeCO3 Þ SSðFeCO3 Þ  1 ½20 Figure 3 SEM images showing a cross-section and a top
V view of a ferrous carbonate layer formed on mild steel;
80  C, pH 6.6, pCO2 ¼ 0.5 bar, stagnant conditions.
where krðFeCO3 Þ is a kinetic constant, which can be
derived from the experimental results as a function of
temperature, using an Arrhenius’ type equation5
  tendency’ (ST) can be used to quantify the relative
BðFeCO3 Þ rates of precipitation (<FeCO3 ) and corrosion (CR)
krðFeCO3 Þ ¼ exp AðFeCO3 Þ  ½21
RTk expressed in the same volumetric units:
where AðFeCO3 Þ ¼ 28.2 and BðFeCO3 Þ ¼ 64 851 J mol1. <FeCO3
ST ¼ ½22
SEM images of a crystalline ferrous carbonate CR
layer formed on a mild steel substrate are shown in For ST  1, porous and unprotective films are likely
Figure 3. The ferrous carbonate layer can slow down to form. Conversely, for ST 1, conditions become
the corrosion process by presenting a diffusion barrier favorable for the formation of dense, protective fer-
for the species involved in the corrosion process, rous carbonate films. However, the use of scaling
thereby changing the conditions at the steel surface. tendency is not as straightforward as it appears, as it
The effective protectiveness of a solid ferrous carbon- requires simultaneous calculation of <FeCO3 and cor-
ate layer depends on its porosity, which hangs in the rosion CR.
balance of the precipitation rate and the underlying In some cases, other salts can be detected in the
corrosion rate. For high precipitation rates, and low surface layers that form on mild steel. In high tem-
corrosion rates, a dense and protective ferrous car- perature CO2 corrosion magnetite (Fe3O4) has been
bonate layer is obtained, and conversely, low precipi- detected. In the presence of oxygen, a ferric oxide
tation rates and high corrosion rates lead to the hematite (Fe2O3), which is very insoluble but offers
formation of porous unprotective ferrous carbonate little protection from corrosion forms. Oilfield scales
layers. A non-dimensional parameter termed ‘scaling also include calcium carbonate, calcium sulfate,
Corrosion in Acid Gas Solutions 1277

barium sulfate, strontium sulfate, etc. The presence often referred to as ‘direct’ reduction of carbonic
of calcium carbonate, in particular, can have a bene- acid9–11 and is written as
ficial effect upon corrosion and upon the stability of
the FeCO3 scale. Finally, in the presence of H2S, 2H2 CO3 þ 2e ! H2 þ 2HCO
3 ½25
various types of sulfides form as discussed in a sepa-
rate section below. Clearly, the addition of the reactions [8] and [24] gives
the reaction [25] proving that the overall reaction is
the same and the distinction is only in the pathway,
that is, in the sequence of reactions. The rate of
2.25.2.2 Electrochemistry of Mild Steel reaction [25] is limited primarily by the slow hydra-
Corrosion in CO2 Saturated Aqueous tion step [6]11,12 and in some cases by the slow CO2
Solutions dissolution reaction [2].
The electrochemical dissolution of iron in a water It can be conceived that in CO2 solutions at
solution: pH > 5 the direct reduction of the bicarbonate ion
becomes important13:
Fe ! Fe2þ þ 2e ½23
2HCO 
3 þ 2e ! H2 þ 2CO3
2
½26
is the dominant anodic reaction in CO2 corrosion.
The reaction is pH dependent in acidic solutions which seems plausible, as the concentration of HCO 3

with a reaction order with respect to OH between increases with pH and can exceed that of H2CO3
1 and 2, decreasing toward 1 and 0 at pH > 4, which is as seen in Figure 2. However, it is difficult to dis-
the typical range for CO2 corrosion. Measured Tafel tinguish experimentally the effect of this particular
slopes are typically 30–80 mV. This subject, which is reaction pathway for hydrogen evolution from the
still somewhat controversial with respect to the two previously discussed (eqns [8] and [25]). In
mechanism, has been reviewed for acidic corrosion6,7 addition, evidence exists that suggests that the rate
and CO2 solutions.8 of this reaction is comparatively low and can be
The presence of CO2 increases the rate of corro- neglected. For example, as the pH increases, the
sion of mild steel in aqueous solutions primarily by amount of HCO 3 increases as well (see Figure 2),

increasing the rate of the hydrogen evolution reac- suggesting that the corrosion rate should follow the
tion. It is well known that in strong acids, which are same trend, if one is to believe that the direct reduc-
fully dissociated, the rate of hydrogen evolution tion of the bicarbonate ion [26] is a significant
occurs according to cathodic reaction. Experimental evidence does not
support this scenario and shows the opposite trend:
2Hþ þ 2e ! H2 ½24 the corrosion rate actually decreases with an increas-
ing pH, even if no protective ferrous carbonate layer
and is, for the case of mild steel corrosion, limited by forms.
the rate at which Hþ ions are transported from the Hydrogen evolution by direct reduction of water:
bulk solution to the steel surface (mass transfer limi-
tation). In CO2 solutions, where typically pH > 4, this 2H2 O þ 2e ! H2 þ 2OH ½27
limiting flux would be small, and therefore it is the
presence of H2CO3 which enables hydrogen evolu- is always possible, but is comparatively very slow and
tion at a much higher rate. Thus, for pH > 4, the is important only at pCO2  0:1 bar and pH > 6.14,15
presence of CO2 leads to a much higher corrosion Therefore, this reaction is rarely a factor in practical
rate than would be found in a solution of a strong acid CO2 corrosion situations.
at the same pH. The various electrochemical processes described
This can be readily explained by considering that above can be quantified using the well established
the homogenous dissociation of H2CO3, as given by electrochemical theory. The rate of the electrochem-
reaction [8], serves as an additional source of Hþ ions, ical reactions, < in kmol m2 s1, can be readily exp-
which are subsequently adsorbed at the steel surface ressed in terms of current density, i in A m2, since
and reduced according to reaction [24].1 A different the two are directly related: for example, during
pathway is also possible, where the H2CO3 first hydrogen evolution [24] for every kmol of Hþ
adsorbs at the steel surface followed by heterogeneous 1 kmol of electrons is used (n ¼ 1 kmole kmol1),
dissociation and reduction of the Hþ ion. This is while for every kmol of iron dissolved [23] two
1278 Liquid Corrosion Environments

kmoles of electrons are used (n ¼ 2 kmole kmol1). The reversible potential for H+ reduction ErevðHþ Þ is a
Therefore, one can write function of temperature and pH:
i ¼ nF < ½28 2:303RðTc þ 273:15Þ
ErevðHþ Þ ¼  pH ½36
F
2.25.2.2.1 Oxidation of iron The cathodic Tafel slope bc ðHþ Þ is calculated as
In the corrosion of mild steel, the oxidation (dissolu-
2:303RðTc þ 273:15Þ
tion) of iron [23] is the dominant anodic reaction. bcðHþ Þ ¼ ½37
0:5F
The anodic dissolution of iron at the corrosion
d
potential (and up to 200 mV above) is under charge The limiting mass transfer current density ilimðHþ Þ
is
+
transfer control. Thus, pure Tafel behavior can be related to the rate of transport of H ions from the
assumed close to the corrosion potential: bulk of the solution through the boundary layer to
the steel surface:
iaðFeÞ ¼ ioðFeÞ 10ðEcorr ErevðFeÞ Þ=baðFeÞ ½29
ðHþ Þ ¼ kmðH Þ FcH ½38
d
ilim þ þ

The exchange current density of iron oxidation is a


function of temperature: where the mass transfer coefficient, kmðHþ Þ can be
calculated from a correlation of the Sherwood, Rey-
ioðFeÞ ¼ iorefðFeÞ nolds, and Schmidt numbers as explained in the
   following section.
DHFe 1 1
exp  ½30
R Tc þ 273:15 Tc;ref þ 273:15
2.25.2.2.3 Reduction of carbonic acid
The Tafel slope of this reaction is given by The carbonic acid reduction reaction [25] can be
2:303RðTc þ 273:15Þ under charge transfer control or limited by the
baðFeÞ ¼ ½31 slow chemical reaction–hydration step [6], preceding
1:5F
it.11,12 The rate of this reaction in terms of current
density is
2.25.2.2.2 Reduction of hydronium ion 1 1 1
In general, the Hþ ion reduction reaction [24] can be ¼ þ r ½39
icðH2 CO3 Þ ia ðH2 CO3 Þ ilim ðH2 CO3 Þ
either under charge transfer or mass transfer (diffu-
sion) control, therefore, one can write: The charge transfer current density iaðH2 CO3 Þ is cal-
culated as
1 1 1
¼ þ ½32 ðEcorr ErevðH Þ=bcðH
icðHþ Þ iaðHþ Þ d
ilim ðHþ Þ
iaðH2 CO3 Þ ¼ ioðH2 CO3 Þ  10 2 CO3 Þ 2 CO3 Þ ½40
The exchange current density ioðH2 CO3 Þ depends on
The charge transfer current density can be calculated
pH, H2CO3 concentration, and temperature:
by
ðEcorr ErevðHþ Þ Þ=bcðHþ Þ @ log ioðH2 CO3 Þ
iaðHþ Þ ¼ ioðHþ Þ  10 ½33 ¼ 0:5 ½41
@pH
The exchange current density ioðHþ Þ is a function of @ log ioðH2 CO3 Þ
¼1 ½42
@cH2 CO3
pH and temperature. The pH dependence is
@ log ioðHþ Þ ioðH2 CO3 Þ DHðH2 CO3 Þ
¼ 0:5 ½34 ref
¼ exp 
@pH ioðH2 CO3 Þ R
 
The temperature dependence of the exchange cur- 1 1
 ½43
rent density can be calculated via an Arrhenius-type Tc þ 273:15 Tc;ref þ 273:15
relation:
 The cathodic Tafel slope bc ðH2 CO3 Þ is
ioðHþ Þ DHðHþ Þ
¼ exp  2:303RðTc þ 273:15Þ
ref
ioðHþ Þ R bcðH2 CO3 Þ ¼ ½44
  0:5F
1 1
 ½35 Since the reductions of H2CO3 and H+ are equivalent
Tc þ 273:15 Tc;ref þ 273:15
Corrosion in Acid Gas Solutions 1279

thermodynamically, the reversible potential for


H2CO3 reduction ErevðH2 CO3 Þ is calculated as ðEcorr ErevðH OÞ Þ=bcðH OÞ
icðH2 OÞ ¼ ioðH2 OÞ 10 2 2 ½53
2:303RðTc þ 273:15Þ
ErevðH2 CO3 Þ ¼  pH ½45 Since the reduction of H2O and Hþ are equivalent
F thermodynamically, they have the same reversible
The chemical reaction limiting current density potential at a given pH:
r
ilim ðH2 CO3 Þ can be calculated from :
16
2:303RðTc þ 273:15Þ
ErevðH2 OÞ ¼  pH ½54
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi F
r
ilim ðH2 CO3 Þ ¼ Fc CO 2
f H2 CO 3
DH2 CO3 Khyd khyd f ½46
The exchange current density for water reduction
ioðH2 OÞ depends on temperature:
The diffusion coefficient for carbonic acid DH2 CO3 as

a function of temperature can be calculated using ioðH2 OÞ DHðH2 OÞ
Einstein’s relation: ref
¼ exp 
ioðH2 OÞ R
    
Tc þ 273:15 mH2 O;ref 1 1
D ¼ Dref ½47  ½55
Tc;ref þ 273:15 mH2 O Tc þ 273:15 Tc;ref þ 273:15

where T is temperature and m is dynamic viscosity. The Tafel slope for H2O reduction was found to be
The forward reaction rate for the CO2 hydration the same as that for Hþ reduction:
f
reaction khyd is calculated as 2:303RðTc þ 273:15Þ
bcðH2 OÞ ¼ ½56
0:5F
f
khyd ¼ 10169:253:0 log ðTc þ273:15Þð11715=ðTc þ273:15ÞÞ ½48

The flow factor fH2 CO3 is


2.25.2.3 Transport Processes in CO2
fH2 CO3 ¼ coth zH2 CO3 ½49 Corrosion of Mild Steel

where From the description of the electrochemical processes


above, it is clear that certain species in the solution are
dmðH2 CO3 Þ ‘produced’ at the metal surface (e.g., Fe2+) while others
zH2 CO3 ¼ ½50
drðH2 CO3 Þ are depleted (e.g., Hþ). The established concentration
gradients lead to molecular diffusion of the species
and toward and away from the surface. In cases when the
diffusion processes are much faster than the electro-
DH2 CO3 chemical processes, the concentration change at the
dmðH2 CO3 Þ ¼ ½51
kmðH2 CO3 Þ metal surface is small. In contrast, when the diffusion is
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi unable to ‘keep up’ with the rate of the electrochemi-
DH2 CO3
drðH2 CO3 Þ ¼ b
½52 cal reactions, the concentration of species at the metal
khyd surface can become very different from that in the
bulk solution. The rate of the electrochemical pro-
The carbonic acid mass transfer coefficient kmðH2 CO3 Þ
cesses depends on the concentration of the reactants
is discussed in Section 2.25.2.3.
at the surface. Therefore, there exists a two-way cou-
pling between the electrochemical processes at the
2.25.2.2.4 Reduction of water metal surface (corrosion) and processes in the adjacent
Unless water is mixed with methanol or glycol to solution layer (i.e., diffusion in the boundary layer).
prevent hydrate formation or somehow diluted oth- The same is true for chemical reactions, which interact
erwise, it can be assumed that water molecules are with both the transport and electrochemical processes
present in virtually unlimited quantities at the steel in a complex way.
surface, and the reduction rate of H2O is controlled In most practical systems, the water solution
by the charge-transfer process and, hence, pure Tafel moves with respect to the metal surface. Therefore,
behavior: the effect of convection on transport processes cannot
1280 Liquid Corrosion Environments

be ignored. Turbulent eddies can penetrate deep into which expresses the simple fact that at steady state all
the hydrodynamic boundary layer and significantly the electrons generated by the oxidation processes are
alter the rate of species transport to and from the consumed by the sum of the reduction processes. By
surface. Very close to the surface no turbulence can substituting the expressions for the various currents
exist and the species are transported solely by diffu- given by eqns [33], [40], [53], and [29] into eqn [59] a
sion. The effect of turbulent flow is captured most single nonlinear equation is now obtained with Ecorr as
easily by using the concept of mass transfer coeffi- the only unknown, which can be easily solved. When
cient, described below. the calculated value of Ecorr is now returned to eqns
In turbulent flow of dilute ideal solutions, a mass [33], [40], [53], and [29], the rate of each individual
transfer coefficient km for a given species (Hþ ions, reaction can be explicitly computed. This also
H2CO3 etc.) can be calculated from a correlation, such includes the corrosion current density obtained from
as the straight pipe correlation of Berger and Hau25: eqn [29]:
Shp ¼ 0:0165Re 0:86 Sc 0:33 ½57 icorr ¼ iaðFeÞ ½60
26
or the rotating cylinder correlation of Eisenberg et al. : Finally, the CO2 corrosion rate is recovered by using
Shr ¼ 0:0791Re Sc 0:7 0:356
½58 Faraday’s law:

or any other similar correlation for the flow geometry at icorr MFe
CR ¼ ½61
hand. It should be noted that most of the mass transfer rFe nF
correlations found in the literature (including the two
listed above) are suited only for single-phase flow. where M is the molecular mass and r is the density. If
Therefore, extension of this approach to multiphase the unit amperes per square meters is used for the
flow situations needs to be done with careful corrosion current density icorr, then conveniently
consideration. the corrosion rate for iron and steel expressed in
Overall, CO2 corrosion of mild steel is not very millimeter per year takes almost the same numerical
sensitive to flow, at least not so when compared to value, precisely, CR ¼ 1:155icorr .
mild steel corrosion in strong acids. This is due to the
fact that the main corrosive species in CO2 corrosion
is H2CO3 which can easily be depleted due to a slow
2.25.2.5 Successes and Limitations of
chemical step which precedes it: the hydration reac-
Modeling of Aqueous CO2 Corrosion of
tion [6]. Therefore, the limiting rate of CO2 corrosion
Mild Steel
is primarily affected by the rate of this chemical
reaction [46], which is a function of temperature and Evidence that our basic understanding of the pro-
CO2 partial pressure and not very sensitive to flow. cesses underlying CO2 corrosion of mild steel is
reasonably sound can be found by comparing the
predictions made by the mechanistic model outlined
2.25.2.4 Calculation of Mild Steel CO2 above with experimental values. In Figure 4, below,
Corrosion Rate one can see the comparison of a potentiodynamic
Leading to this point, the main processes underpin- sweep obtained in the experiments and the one pre-
ning CO2 corrosion were defined: the speciation of dicted by the model. Many other comparisons of the
the aqueous CO2 solution using the thermodynamic predicted and measured corrosion rates are given in
approach outlined in Section 2.25.2.1, the electro- the following section, where the effect of key factors
chemical theory described in Section 2.25.2.2, and in CO2 corrosion of mild steel is discussed.
the transport processes as covered in Section 2.25.2.3. Despite the relative progress we have made in
Using this information, the corrosion rate of mild steel understanding and modeling of aqueous CO2 corro-
can now be calculated. The unknown corrosion sion of mild steel, many questions persist. One is the
potential Ecorr in [33], [40], [53], and [29] can be issue of localized CO2 corrosion, which is still a topic
found from the current (charge) balance equation at of intense ongoing research. Effect of other factors
the steel surface: such as steel metallurgy, organic acids, oxygen, mul-
tiphase flow, and inhibitors are challenges that need
icðHþ Þ þ icðH2 CO3 Þ þ icðH2 OÞ ¼ iaðFeÞ ½59 further effort. Some of those are discussed in the
following sections.
Corrosion in Acid Gas Solutions 1281

–0.1

–0.2

–0.3 Total cathodic

–0.4
E vs. SHE (V)

Model sweep
Ecorr
–0.5 Experimental
Total anodic sweep
–0.6 (Fe dissolution)

H2O reduction
–0.7

–0.8

–0.9 H2CO3 reduction


H+ reduction
–1
0.1 1 icorr 10
i (A m–2)
Figure 4 Potentiodynamic sweep, experimental (points) vs. model (lines); 20  C, pCO2 ¼ 1 bar, pH 4, 2 m s1.

2.25.2.6 Key Factors Affecting Aqueous supersaturations are shown in Figure 5. At lower
CO2 Corrosion of Mild Steel supersaturations obtained at the lower pH of 6, shown
in Figure 5, the corrosion rate does not change much
Armed with the understanding and the ability to
with time, even if some ferrous carbonate precipitation
calculate CO2 corrosion rates, as described in the
occurs, reflecting the fact that a relatively porous,
sections above, in this section, the effect of key factors
detached and unprotective layer is formed (low scaling
which affect the rate of CO2 corrosion are discussed,
tendency ST). The higher pH of 6.6 results in higher
and the predictions made by the model are compared
supersaturation, faster precipitation, and formation of
to empirical results.
more protective ferrous carbonate, reflected by a rapid
decrease of the corrosion rate with time. There are
2.25.2.6.1 Effect of pH other indirect effects of pH, and by almost all accounts,
The pH has a significant influence on the CO2 corro- higher pH leads to a reduction of the corrosion rate,
sion rate. Lower pH leads to higher corrosion rates and making the ‘pH stabilization’ (meaning: pH increase)
vice versa, just like in many other acidic solutions. technique an attractive way of managing CO2 corro-
Typical pH in CO2 saturated condensed water is sion. The drawback of this technique is that it can lead
about pH 4 while in buffered brines, one frequently to excessive scaling and can rarely be used with forma-
encounters 5 < pH < 7. At pH 4 or below, direct reduc- tion water systems.
tion of Hþ ions, reaction [24], is important, particularly
at lower partial pressures of CO2, when direct reduc- 2.25.2.6.2 Effect of CO2 partial pressure
tion of carbonic acid, reaction [25], can be ignored. In In the case of scale-free CO2 corrosion, an increase of
that case, the pH has a direct effect on the corrosion pCO2 typically leads to an increase in the corrosion
rate. Another important effect of pH is indirect and rate. The commonly accepted explanation is that
relates to how pH changes conditions for the formation with increasing pCO2 the concentration of H2CO3
of ferrous carbonate layers. Higher pH (5 < pH < 7) increases and accelerates the cathodic reaction, eqn
results in a decreased solubility of ferrous carbonate [25], and ultimately the corrosion rate. The detri-
and leads to an increased precipitation rate and a mental effect of pCO2 at a constant pH is illustrated in
higher scaling tendency. The effect of various pH and Figure 6. The model described above reasonably
1282 Liquid Corrosion Environments

3.00

Corrosion rate (mm year –1)


2.50

2.00 SS = 9

1.50

1.00 SS = 7

0.50 SS = 150
SS = 37
SS = 30
0.00
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70
Time (h)
Figure 5 Effect of ferrous carbonate supersaturation SSðFeCO3 Þ on corrosion rate obtained at a range of pH 6.0–6.6, for
5 ppm < cFe2þ < 50 ppm at T ¼ 80  C, under stagnant conditions. Error bars represent minimum and maximum values
obtained in repeated experiments. Data taken from Chokshi et al.17

50

Model
40
CR (mm year–1)

30

Experimental

20

10

0
1 10 100
pCO2 (bar)

Figure 6 The effect of CO2 partial pressure, pCO2 on bare steel corrosion rate, comparison of experimental results
(points) and model (line); 60  C, pH 5, 1 m s1, 100 mm ID single-phase pipe flow.

captures well this trend up to approximately 2.25.2.6.3 Effect of temperature


pCO2 ¼ 10 bar. However, when other conditions are Temperature accelerates all the processes involved in
favorable for the formation of ferrous carbonate corrosion: electrochemical, chemical, transport, etc.
layers, increased pCO2 can have a beneficial effect. One would expect then that the corrosion rate
At a high pH, higher pCO2 leads to an increase in steadily increases with temperature, and this is the
bicarbonate and carbonate ion concentration and a case at low pH when precipitation of ferrous carbon-
higher supersaturation which accelerates precipita- ate or other protective layers does not occur. An
tion and protective layer formation. The effect of example is shown Figure 8. The situation changes
pCO2 on the corrosion rate in the presence of ferrous markedly when solubility of ferrous carbonate is
carbonate precipitation is illustrated in Figure 7 exceeded, typically at a higher pH. In that case,
where in stratified wet gas flow, corrosion rate is increased temperature rapidly accelerates the kinet-
reduced both at top and bottom of the pipe with the ics of precipitation and protective layer formation,
increase partial pressure of CO2. decreasing the corrosion rate. The peak in the
Corrosion in Acid Gas Solutions 1283

100
Bottom
Top
15

Corrosion rate (mm year–1)


10

0.2 0.2

0.1 0.06

0.01
P = 3.8 bar P = 10.6 bar
CO2 partial pressure

Figure 7 Experimental measurements of the corrosion rate at the top and bottom of the pipe in stratified gas–liquid
flow showing the effect of CO2 partial pressure, pCO2 on formation of ferrous carbonate layer. Test conditions: 90  C, pH 6,
100 mm ID, Vsg ¼ 10 m s1, Vsl ¼ 0.1 m s1. Data taken from Sun and Nešić.18

25 lead to an increase in the corrosion rate as illustrated


in Figure 9. At lower pH 4, the effect is much more
Corrosion rate (mm year –1)

20 pronounced as the dominant cathodic reaction is


15 direct Hþ ion reduction [24], which is under mass
transfer control (see eqn [38]).
10 When protective ferrous carbonate layers form
(typically at higher pH in produced water) or when
5
inhibitor films are present on the steel surface, the
0 above-mentioned effect of flow becomes insignificant
0 20 40 60 80 100 120
Temperature (⬚C)
as the main resistance to corrosion is now in the surface
layer or inhibitor film. In this case, the effect of flow is
Figure 8 The effect of temperature on CO2 corrosion rate
to interfere with the formation of protective surface
of mild steel; pH 4, pCO2 ¼ 1 bar, 100 mm ID single phase
pipe flow. Points are experimental values and the solid line layers or to remove them once they are in place, often
is the model. The dotted line is a model simulation of the leading to an increased risk of localized attack.
same conditions at pH 6.6 accounting for protective ferrous The two flow accelerated corrosion effects dis-
carbonate film formation. cussed above are frequently aggravated by flow dis-
turbances such as valves, constrictions, expansions,
corrosion rate is usually seen between 60 and 80  C bends, etc. where local increases of near-wall turbu-
depending on water chemistry and flow conditions as lence and wall-shear stress are seen. However, flow
shown in Figure 8 (dotted line). can lead to onset of localized attack only when given
the ‘right’ set of circumstances as discussed in a
2.25.2.6.4 Effect of flow separate heading below.
There are two main ways in which flow may affect The effect of multiphase flow on CO2 corrosion is
CO2 corrosion, which can be distinguished based on complicated by the different flow patterns that exist,
whether or not other conditions are conducive to the most common being stratified, slug, and annular-
protective layer formation or not. mist flow. In the liquid phase, water and oil can flow
In the case of corrosion where protective layers do separated or mixed with either phase being continu-
not form (typically at low pH as found in condensed ous with the other flowing as a dispersed phase.
water and in the absence of inhibitors), the main role Different flow patterns lead to a variety of steel
of turbulent flow is to enhance transport of species surface wetting mechanisms: stable water wetting,
toward and away from the metal surface. This may stable oil wetting, intermittent wetting, etc., which
1284 Liquid Corrosion Environments

4
pH = 4

CR (mm year–1)
2

4
pH = 5

3
CR (mm year–1)

4
pH = 6

3
CR (mm year–1)

0
0 2 4 6 8 10 12 14
Velocity (m s–1)
Figure 9 Predicted (line) and experimentally measured corrosion rates (points) showing the effect of velocity in the
absence of ferrous carbonate layers. Test conditions: 20  C, pCO2 ¼ 1 bar, 15 mm ID single-phase pipe flow. Experimental
data taken from Nešić et al.19

greatly affect corrosion. In annular mist flow, the Corrosion inhibitors


liquid droplets move at high velocity and can lead Describing the effect of corrosion inhibitors is not a
to protective layer damage at points of impact such as straightforward task due to the enormous complexity
bends, valves, tees, constrictions/expansions, and of the subject. Quantifying them and predicting their
other pipe fittings. Slug flow can lead to significant behavior are even harder. There is a plethora of
short-lived fluctuations in the wall-shear stress, approaches in the open literature, varying from the
which can help remove a protective surface layer of use of simple inhibitor factors and inhibition efficiencies
ferrous carbonate or possibly affect an inhibitor film. to the application of complicated molecular modeling
techniques to describe inhibitor interactions with the
steel surface and ferrous carbonate layer. A middle-
2.25.2.6.5 Effect of corrosion inhibition
of-the-road approach is based on the assumption that
The two most common sources of corrosion inhibi-
corrosion protection is achieved by surface coverage,
tion need to be considered:
that is, that the inhibitor adsorbs onto the steel sur-
(a) inhibition by addition of corrosion inhibitors and face and slows down one or more electrochemical
(b) inhibition by components present in the crude oil. reactions by ‘blocking.’ The degree of protection is
Corrosion in Acid Gas Solutions 1285

assumed to be directly proportional to the fraction of 60


the steel surface blocked by the inhibitor. In this type 50

CR (mm year–1)
of model, one needs to establish a relationship 40
between the surface coverage y and the inhibitor
30
concentration in the solution cinh. This is most com-
20
monly done by the use of adsorption isotherms.
10

0
Corrosion inhibition by crude oil 1 10 100 1000
It has been known for a while that CO2 corrosion Undissociated aqueous HAc concentration (ppm)
rates seen in the field in the presence of crude oil are Figure 10 Predicted (line) and experimentally measured
much lower than those obtained in laboratory condi- data (points) showing the effect of the concentration of
tions where crude oil was not used or synthetic crude undissociated acetic acid (HAc) on the CO2 corrosion rate,
oil was used. One can identify two main effects of 60  C, pCO2 ¼ 0.8 bar, pH 4, 12 mm OD rotating cylinder flow
at 1000 rpm. Experimental data taken from Sun et al.20
crude oil on the CO2 corrosion rate.
The first is a wettability effect and relates to a
hydrodynamic condition where crude oil entrains
the protectiveness of ferrous carbonate layers; how-
the water and prevents it from wetting the steel
ever, the mechanism is still not clear.
surface (continuously or intermittently).
The second effect is corrosion inhibition by compo-
2.25.2.6.7 Effect of glycol/methanol
nents of the crude oil that reach the steel surface
Glycol and methanol are often added to flowing
either by direct contact or by first partitioning into
systems in order to prevent hydrates from forming.
the water phase. Various surface active organic com-
The quantities are often significant (50% of total
pounds found in crude oil (typically oxygen, sulfur
liquid phase is not unusual). In the very few studies
and nitrogen containing molecules) have been iden-
available, it has been assumed that the main ‘inhibi-
tified to directly inhibit corrosion of mild steel in
tive’ effect of glycol/methanol on corrosion comes
CO2 solutions.
from dilution of the water phase, which leads to
a decreased activity of water. However, there are
2.25.2.6.6 Effect of organic acids many unanswered questions such as the changes in
The low molecular weight organic acids are primarily mechanisms of CO2 corrosion in water/glycol
soluble in water and can lead to corrosion of mild mixtures which are yet to be discovered.
steel. Higher molecular weight organic acids are not
water soluble, but are typically soluble in the oil phase 2.25.2.6.8 Effect of condensation in
and pose a corrosion threat at higher temperatures in wet gas flow
the refineries. Acetic acid CH3COOH (denoted as When transporting humid natural gas, due to the cool-
HAc in the text below) is the most prevalent low ing of the stream, condensation of water vapor occurs
molecular weight organic acid found in brines. on the internal pipe wall. The condensed water is pure
Other acids typically found in the brine are propionic, and, due to dissolved CO2, typically has a pH < 4.
formic, etc.; however, their behavior and corrosiveness This leads to the so-called top-of-the-line corrosion
is very similar to that of HAc and therefore HAc can (TLC) scenario. If the rate of condensation is high,
be used as a ‘surrogate’ for all the organic acids found plenty of acidic water flows down the internal pipe
in the brine. HAc is a weak acid; however, it is stronger walls leading to a very corrosive situation. If the con-
than H2CO3 (pKa 4.76 vs. 6.35 at 25  C), and it is the densation rate is low, the water film is not renewed and
main source of Hþ ions when the two acid concentra- flows down very slowly and the corrosion process can
tions are similar. The effect of HAc is particularly release enough Fe2+ to raise the local pH and saturate
pronounced at higher temperatures and low pH the solution, leading to the formation of protective
when the abundance of undissociated HAc can ferrous carbonate layer. The layer is often protective;
increase the CO2 corrosion rate dramatically as seen however, incidents of localized attack in TLC were
in Figure 10. Solid iron acetate does not precipitate in reported.21 Either way, the stratified or stratified-wavy
the pH range of interest since its solubility is much flow regime, typical for TLC, does not lead to a good
higher than that of ferrous carbonate. There are some opportunity for inhibitors to reach the upper portion
indications that the presence of organic acids impairs of the internal pipe wall and protect it. A very limited
1286 Liquid Corrosion Environments

range of corrosion management options for TLC pressure of CO2, etc. It seems that localized attack
exists. To qualitatively and quantitatively describe the occurs when the conditions are such that partially
phenomenon of corrosion occurring at the top of protective ferrous carbonate layers form. It is well
the line, a deep insight into the combined effect known that when fully protective ferrous carbonate
of the chemistry, hydrodynamics, thermodynamics, forms, low general corrosion rates are obtained and
and heat and mass transfer in the condensed water is vice versa: when no protective layers form, a high rate
needed. A full description exceeds the scope of this of general corrosion is seen. It is when the corrosive
review, and the interested reader is directed to some environment is ‘in between,’ in the so-called ‘gray
recent articles published on this topic.21,22 zone,’ that localized attack can be initiated most
often by some extreme flow conditions. There are
2.25.2.6.9 Nonideal solutions and gases many combinations of environmental and metallur-
In many cases produced, water has very high dissolved gical parameters that define the grey zone, making
solids content (>10 wt%). At such high concentra- this sound like a difficult proposal. However, there is
tions, the infinite dilution theory used above does not a single parameter which is easy to calculate: ferrous
hold, and corrections need to be made to account for carbonate supersaturation, SSðFeCO3 Þ (see eqn [19]
solution nonideality. A simple way to account for the above), which can be successfully used as a good
effect on nonideal homogenous water chemistry is delineator for the gray zone and as such as a predictor
to correct the equilibrium constants by using the con- for the probability for localized attack. When bulk
cept of ionic strength as indicated above. This ferrous carbonate supersaturation is in the range
approach seems to work well only for moderately 0.5 < SSðFeCO3 Þ <2 there is a risk of localized attack.
concentrated solution (up to a few weight percentage The further away the solution is from these bound-
of dissolved solids). For more concentrated solutions, a aries, the lower the risk. The scaling tendency ST
more accurate way is to use activity coefficients as (see equation [22] above) is conceptually even better
described by Anderko et al.23 The effect of concen- suited as a predictor of localized corrosion risk, how-
trated solutions on heterogeneous reactions such as ever, its calculation is much more difficult and uncer-
precipitation of ferrous carbonate and other layers is tain as it involves calculation of both the uniform
still largely unknown. Furthermore, it is unclear how corrosion rate and the precipitation rate.
the highly concentrated solutions affect surface elec- Based on mostly anecdotal evidence (field experi-
trochemistry. Some experience suggests that corrosion ence), the presence of H2S and HAc is related to the
rates can be dramatically reduced in very concentrated onset of localized attack, however, little is understood
brines; nevertheless a more systematic study is needed. about how and when this may happen.
At very high total pressure, the gas–liquid equili-
bria cannot be accounted for by Henry’s law. A simple
correction can be made by using a fugacity coeffi- 2.25.3 Aqueous H2S Corrosion of
cient, which accounts for nonideality of the CO2/ Mild Steel
natural gas mixture24 and can be obtained by solving
the equation of state for the gas mixture. Those cases, Corrosion of mild steel in the presence of hydrogen
in which critical point for CO2 is approached or sulfide (H2S) also represents a significant problem for
exceeded, warrant a separate analysis and are not the oil and gas industry.27–33 Many fields have been
covered by the considerations discussed above. developed that in addition to CO2 have high concen-
trations of H2S. In CO2/H2S corrosion of mild steel,
thermodynamic considerations suggest that both fer-
2.25.2.7 Localized CO2 Corrosion of Mild
rous carbonate and ferrous sulfide layers could theoret-
Steel in Aqueous Solutions
ically form on the steel surface. However, studies have
As illustrated above, significant progress has been demonstrated that the formation of the ferrous sulfide
achieved in understanding uniform CO2 corrosion, layers is dominant and presents one of the most impor-
without or with protective layers, and hence success- tant factors governing the H2S corrosion rate. The
ful uniform corrosion models can be built. However, sulfide layer growth depends primarily on the kinetics
much less is known about localized CO2 corrosion. It of the corrosion process as it is described below.
is thought that one of the main factors that ‘triggers’ Despite the relative abundance of experimental
localized attack is flow, tempered by other environ- data on H2S corrosion of steel, most of the literature
mental variables such as pH, temperature, partial is still confusing and somewhat contradictory.
Corrosion in Acid Gas Solutions 1287

Therefore, the mechanism of H2S corrosion remains where KH2 S is the solubility constant of H2S in
much less understood when compared to that of CO2 mol l1 bar1:
corrosion. This uncertainty makes it more difficult to cH S
develop a model to predict the corrosion rate of mild KsolðH2 SÞ ¼ 2 ½63
pH2 S
steel in H2S saturated aqueous solution.
and can be found from34

ð634:27þ0:2709TK 0:11132103 TK2 ð16719=TK Þ


2.25.3.1 Chemistry of H2S Saturated KsolðH2 SÞ ¼ 10 261:9logTK Þ ½64
Aqueous Solutions – Equilibrium
Considerations As shown in Figure 11, the solubility of H2S decreases
Similar to CO2 discussed above, H2S gas is also with temperature, as it is observed for CO2. However,
soluble in water: for the same partial pressure and temperature, the
concentration of dissolved H2S actually exceeds that
KH2 S in the gas phase as shown in Figure 12.
H2 SðgÞ , H2 S ½62
Aqueous H2S is another weak acid which partly
dissociates in two steps:
0.20 Khs
H2 S , Hþ þ HS ½65
Species concentration (mol l–1)

 Kbs
0.15 HS , Hþ þ S2 ½66

where Khs is the dissociation constant of H2S:


0.10
cHþ cHS
H2S Khs ¼ ½67
0.05 cH2 S
CO2
and can be calculated as35
0.00
0 20 40 60 80 100
T (⬚C) 782:43945þ0:361261TK 1:6722104 TK2 ð20565:7315=TK Þ
Khs ¼ 10 142:741722lnTK ½68
Figure 11 Calculated solubility of H2S and CO2 as a
function of temperature; 25  C, pH2 S ¼ 1 bar, pCO2 ¼ 1 bar.
and Kbs is the dissociation constant of HS:

1.E+00

1.E–01
Species concentration (mol l–1)

1.E–02

1.E–03
H2S
1.E–04

H2S(g)
1.E–05

HS–
1.E–06

1.E–07
2 3 4 5 6 7
pH
Figure 12 Calculated sulfide species concentrations as a function of pH for an H2S saturated aqueous solution at
pH2 S ¼ 1 mbar, 25  C, 1 wt% NaCl.
1288 Liquid Corrosion Environments

ferrous sulfide – mackinawite is known as a func-


c þ c 2 tion of temperature36,37
Kbs ¼ H S ½69
cHS
There is a very large discrepancy in the reported
mackin
KspðFeSÞ ¼ 10ð2848:779Þ=Tk 6:347 ½71
values for Kbs, varying from 1:0  1019 to
1:1  1012 kmol m3 at room temperature (seven For other ferrous sulfides, only the values at room
orders of magnitude). In addition, these values are temperature are known, as listed in Table 1 below.
very small compared with other equilibrium con- It is convenient to show various ferrous sulfide solu-
stants, all suggesting that using Kbs to calculate the bilities in terms of an equilibrium concentration of
concentration of sulfide species, cS2 and further to the Fe2+ as a function of pH at a given H2S partial
predict the solubility product constants for ferrous pressure (concentration). An example is presented in
sulfides should be avoided. Figure 13 where it can be seen that the much less
Given the same gaseous concentrations of soluble pyrrhotite and troilite are thermodynami-
H2S and CO2, one obtains a similar aqueous concen- cally more stable forms compared to mackinawite
tration of dissolved H2S and CO2 (see Figure 11) and and amorphous ferrous sulfide. For a typical ferrous
the resulting pH is within 0.1 pH unit, therefore, ion concentration of cFe2þ ¼ 1 ppm, the saturation
values shown in Figure 1 for CO2 can be used for with respect to troilite and pyrrhotite is reached
H2S as the first approximation. The equilibrium dis- already at pH 5.4, while for mackinawite it is pH 6
tribution of sulfide species as a function of pH for an and for amorphous ferrous sulfide pH 6.7. Keeping in
open system is shown in Figure 12. The concentra- mind that the concentration of Fe2þ at a corroding
tion of bisulfide ion, cHS , becomes significant only steel surface can easily be much higher than in the
above pH 4, while the concentration of the sulfide bulk (e.g., 10 ppm or even higher) and that the pH is
ion, cS2 , is not even shown as it is very low and also higher at the surface than in the bulk (typically
unreliable to calculate. above pH 6), using Figure 13 one can expect a whole
Many types of iron sulfides, such as amorphous range of different ferrous sulfides to form on a cor-
ferrous sulfide (FeS), mackinawite (Fe1þxS), cubic roding steel surface at this H2S concentration at
ferrous sulfide (FeS), troilite (FeS), pyrrhotite different points in time.
(Fe1xS or FeS1þx), smythite (Fe3þxS4), greigite SEM images of a ferrous sulfide surface layer
(Fe3S4), and pyrite (FeS2) occur. Studies have sug- formed on mild steel after a week long exposure are
gested that some of these are stoichiometric such shown in Figure 14. The layered structure of the
as cubic ferrous sulfide, troilite, greigite, and sulfide is prominent, and it can be identified as mack-
pyrite, while others such as mackinawite, pyrrho- inawite. In longer exposures, the ferrous sulfide layer
tite, and smythite are not. Some are electrically thickens and eventually becomes more protective. An
nonconductive, others apparently behave as semi- image of a ferrous sulfide layer after a month long
conductors. However, there is no consensus on exposure is shown in Figure 15. The composition of
these issues and the interested reader is directed the layer is a mixture of mackinawite and pyrrhotite.
to the vast literature on iron sulfides for a more Another layered structure composed of a mixture of
in-depth treatment. The thermodynamics of these ferrous carbonate and ferrous sulfide is shown in
systems is very complicated; depending on envi- Figure 16.
ronmental conditions and time, transformation
from one type of ferrous sulfide into the other
occurs. Limited information exists on aqueous
solubility of the various sulfides. Avoiding the
Table 1 Solubility product constants for various ferrous
usage of the sulfide ion concentration, cS2 , one sulfides at 25  C38
can write a general equation for precipitation of
ferrous sulfide as Type of ferrous sulfide log Ksp(FeS)

Amorphous (FeS) 2.95


KspðFeSÞ
Fe2þ þ H2 S , FeSðsÞ þ 2Hþ ½70 Mackinawite (Fe1þxS) 3.6
Pyrrhotite (Fe1  xS or FeS1þx) 5.19
where the solubility constant for one type of Troilite (FeS) 5.31
Corrosion in Acid Gas Solutions 1289

1.E+04

M
ac
ki
Fe2+ concentration (ppm)

na
1.E+03

w
ite
Py
1.E+02

rrh
ot

Am
ite

or
ph
1.E+01

or
us
Tr
oi
lit
e
1.E+00
3 3.5 4 4.5 5 5.5 6 6.5 7
pH
Figure 13 Calculated solubility of various iron sulfides as a function of pH shown in terms of the equilibrium concentration
of Fe2þ, pH2 S ¼ 1 mbar, 25  C, 1 wt% NaCl.

Ferrous sulfide

Ferrous sulfide
Mild steel
Mild steel

Acc.V Spot Magn Det WD 200 µm Acc.V Spot Magn WD 100 µm


20.0 kV 5.0 100x SE 10.3 15.0 kV 3.0 250x 11.9

Figure 15 SEM images showing a cross-section view of a


ferrous sulfide layer formed on mild steel; 60  C, pH 6,
pCO2 ¼ 7.7 bar, pH2 S ¼ 0.25 mbar, 1 m s1 single phase flow
in a 100 mm ID pipe, 30 day exposure.

2.25.3.2 Mild Steel Corrosion in H2S and


Mixed H2S/CO2 Saturated Aqueous
Solutions
As aqueous H2S is another weak acid, it can be seen
as an additional reservoir of Hþ ions according to
Acc.V Spot Magn Det WD 200 µm reaction [65], similar to H2CO3. Therefore, stimula-
20.0 kV 5.0 100 x SE 10.3
tion of the hydrogen evolution reaction could also be
expected in the presence of H2S. Using the analogy
Figure 14 SEM images showing a cross-section
and a top view of a ferrous sulfide layer formed on mild with CO2 corrosion, one must also allow the possi-
steel; 60  C, pH 6, pCO2 ¼ 7.7 bar, pH2 S ¼ 0.25 mbar, 1 m s1 bility of direct reduction of H2S, that is, that the
single phase flow in a 100 mm ID pipe, 7 days exposure. H2S molecule can be adsorbed at the steel surface,
1290 Liquid Corrosion Environments

of the solution is below saturation level, the outer


sulfide layer will undergo a process of chemical
dissolution. Conversely, when the saturation is
exceeded, ferrous sulfide precipitation from the
bulk is possible. Eventually, the amount and protec-
Ferrous sulfide tiveness of the outer sulfide layer is determined by
the balance of the various formation and removal
Ferrous carbonate processes.39
The transformation of mackinawite into other
forms of less soluble and more stable ferrous sulfide
Mild steel (pyrrhotite and troilite, see Figure 13) may happen
over time. Among the various ferrous sulfides, mack-
Figure 16 SEM images showing a cross-section view of a
inawite is the prevalent ferrous sulfide that forms
mixed ferrous carbonate and ferrous sulfide layer formed on
mild steel; 60  C, pH 6, pCO2 ¼ 7.7 bar, pH2 S ¼ 1.2 mbar, in the corrosion of mild steel at low H2S concentra-
1 m s1 single phase flow in a 100 mm ID pipe, 25 day tion and low temperature. At increased levels of
exposure. H2S, mackinawite is less prevalent and pyrrhotite
is the main corrosion product. At very high H2S con-
followed by a reduction of the Hþ and oxidation of centrations, pyrite and elemental sulfur appear. While
iron in the steel. One can write the overall corrosion thermodynamics of ferrous sulfides may favor other
reaction as types of sulfide over mackinawite as the corrosion
product, the rapid kinetics of mackinawite formation
FeðsÞ þ H2 S ! FeSðsÞ þ H2 ½72 favors it as the initial corrosion product seen in most
situations. Overall, however, there is currently no
As solid ferrous sulfide (mackinawite) is always found clearly defined relationship between the nature of
on the corroding steel surface in the presence of H2S, the sulfide layer and the underlying corrosion process.
even below the solubility limit, this can been referred It is generally thought that all types of ferrous sulfide
to as a direct ‘solid state’ reaction pathway as both the layers offer some degree of corrosion protection for
initial and final state of Fe are solid(s).39 mild steel.
Experimental evidence suggests that corrosion of At very high H2S concentrations, elemental sulfur
mild steel by H2S initially proceeds by adsorption of can appear and lead to severe localized corrosion.
H2S to the steel surface followed by a very fast redox Large amounts of elemental sulfur can precipitate
reaction at the steel surface to form an adherent out of the gas stream and can even block the line,
mackinawite film (much like a tarnish). This initial due to the changes in pressure and temperature.
mackinawite film is very thin (1 mm) but appar- Alternatively, when there is O2 present, the most
ently rather dense and acts as a solid state diffusion likely pathways for formation of elemental sulfur
barrier for the species involved in the corrosion reac- are as follows:
tion. Therefore, this thin mackinawite film is one of
the most important factors governing the corrosion ferrous sulfide reacts with O2 and converts to iron
rate in H2S corrosion. It also impedes the mobility of oxide forming elemental sulfur probably via:
other species in reaching the steel surface and there-
fore corrosion rates due to CO2 are affected even if 3FeS þ 2O2 ! Fe3 O4 þ 3S ½73
very small amounts of H2S are present in the gas
phase (as little as 10–5 bar). at very high H2S concentration, the following
The thin mackinawite film continuously goes reaction can occur to yield elemental sulfur:
through a cyclic process of growth, internal stress
growth, cracking, and delamination that generates 2H2 S þ O2 ! 2H2 O þ 2S ½74
an outer sulfide layer, which thickens over time At very high temperatures, an alternative pathway is
(typically
1 mm) and forms an additional diffusion
barrier. However, this outer sulfide layer is very
H2 S ! H2 þ S ½75
porous and rather loosely attached to the steel sur-
face. Over time it cracks, peels, and spalls, a Localized corrosion by elemental sulfur occurs via a
process accelerated by turbulent flow. If the pH reaction with the iron in the steel, represented by the
Corrosion in Acid Gas Solutions 1291

overall reaction By eliminating the unknown interfacial concen-


trations coðH2 SÞ and ciðH2 SÞ from eqns [77] to [79], the
Fe þ S ! FeS ½76
following equation is obtained for the flux (corrosion
It is not very clear at this stage what the detailed rate) due to H2S:
mechanism of this reaction is. It appears that rapid !
attack is seen only when direct contact of sulfur with dos 1
cH2 S  FluxH2 S þ
the steel is achieved in the presence of water. A more DH2 S ec kmðH2 SÞ
in-depth discussion about the corrosion mechanisms FluxH2 S ¼ AH2 S ln
csðH2 SÞ
of mild steel involving elemental sulfur exceeds the
scope of this review. ½81

This is an algebraic nonlinear equation with respect


2.25.3.3 Calculation of Mild Steel to FluxH2 S, which does not have an explicit solution
H2S Corrosion Rate but can be solved by using a simple numerical algo-
rithm such as the interval halving method or similar
Due to the complexity of the underlying processes methods. These are available as ready-made routines
and a lack of mechanistic understanding, predictive in spreadsheet applications or in any common com-
models of H2S corrosion were not readily available puter programming language. The prediction for
until recently. One approach40 which has the capabil- FluxH2 S depends on a number of constants used in
ity to address a few simple H2S corrosion scenarios is the model which can be either found in handbooks
presented below. A pure H2S corrosion environment (such as DH2 S ), calculated from the established theory
is described first followed by a mixed H2S/CO2 (e.g., kmðH2 SÞ ) or are determined from experiments
corrosion scenario. (e.g., AH2 S ; csðH2 SÞ ). The unknown thickness of the
outer sulfide layer change with time and need to be
2.25.3.3.1 Pure H2S aqueous environment calculated as described below.
Due to the presence of the inner mackinawite film It is assumed that the amount of layer retained on
and the outer porous sulfide layer, it is assumed that the metal surface at any point in time depends on the
the corrosion rate of steel in H2S solutions is always balance of:
under mass transfer control. One can then write the layer formation kinetics (as the layer is generated by
flux of H2S due to: spalling of the thin mackinawite film underneath it
convective diffusion through the mass transfer and by the precipitation from the solution), and
boundary layer as layer damage kinetics (as the layer is damaged by
  intrinsic or hydrodynamic stresses and/or by
FluxH2 S ¼ kmðH2 SÞ cH2 S  coðH2 SÞ ½77 chemical dissolution):
¼  SDR
molecular diffusion through the liquid in the SRR SFR
½82
g

Sulfide layer Sulfide layer Sulfide layer


porous outer sulfide layer as retension rate formation rate damage rate

DH2 S ec   where all the terms are expressed in kmol m2 s1. In
FluxH2 S ¼ coðH2 SÞ  ciðH2 SÞ ½78 order to simplify the calculations, it can be assumed
dos
that in the typical range of application (4 < pH < 7),
solid state diffusion through the inner mackinawite precipitation and dissolution of ferrous sulfide layer
film as do not play a significant role and so it can be written
  !
BH2 S ciðH2 SÞ SRR ¼ CR  SDRm ½83
FluxH2 S ¼ AH2 S exp  ln ½79
RTk csðH2 SÞ
Some experiments involving mackinawite have shown
that even in stagnant conditions about half of the outer
In a steady state, the three fluxes are equal to each
sulfide layer that forms is lost from the steel surface
other and are equivalent to the corrosion rate as
due to intrinsic growth stresses by internal cracking
CRH2 S ¼ FluxH2 S MFe =rFe ½80 and spalling, that is, SDRm 0:5CR, so one obtains:

further corrected for appropriate corrosion rate unit. SRR ¼ 0:5CR ½84
1292 Liquid Corrosion Environments

that is, about half of the iron corroded is found on the The flux FluxHþ is directly related to the corrosion
steel surface in the form of mackinawite. It is not rate by Hþ ions: FluxHþ MFe
known if and how this ratio is different when other CRHþ ¼ ½87
2 rFe
types of ferrous sulfide layers form, for example, the
more adherent and protective pyrrhotite. Moreover, further adjusted for the appropriate corrosion
additional experimentation is required to determine rate unit.
how the mechanical layer damage is affected by hydro- By solving eqns [81] and [86] sequentially in time,
dynamic forces. the total corrosion rate in mixed pure H2S aqueous
Once the layer retention rate SRR is known, the environments can be calculated as
change in mass of the outer sulfide layer can be easily CR ¼ CRH2 S þ CRHþ ½88
calculated as
Dmos ¼ SRRMFeS ADt ½85
2.25.3.3.2 Mixed CO2/H2S environments
The porosity of the outer sulfide layer was deter-
For mild steel corrosion in mixed CO2/
mined to be very high (e 0:9) by comparing the
H2S containing environments, one can account for
weight of the layer with the cross-sectional SEM
the effect of CO2 by assuming that the rate
images showing its thickness. On the other hand,
controlling step in this additional process is the dif-
this layer has proven to be rather protective (i.e.,
fusion of CO2 through the ferrous sulfide layers.
impermeable to diffusion) which can only be
Then a similar expression can be obtained for the
explained by its low tortuosity arising from its lay-
corrosion rate due to CO2:
ered structure. By comparing the measured and  
calculated corrosion rates in the presence of the dos 1
cCO2 FluxCO2 þ
outer sulfide layer, the tortuosity factor was calcu- DCO2 ec kmðCO2 Þ
lated to be c ¼ 0:003. FluxCO2 ¼ ACO2 ln
csðCO2 Þ
A time-marching explicit solution procedure
could now be established where ½89
The flux FluxCO2 is equivalent to the corrosion rate
1. the corrosion rate FluxH2 S in the absence of outer by CO2:
sulfide layer can be calculated by using eqn [81],
and assuming dos ¼ 0; FluxCO2 MFe
CRCO2 ¼ ½90
2. the amount of sulfide layer dmos formed over a 2 rFe
time interval Dt is calculated by using eqn [85]; further adjusted for appropriate corrosion rate unit.
3. the new corrosion rate FluxH2 S in the presence of By solving eqns [81], [86], and [89], the total
sulfide layer can be recalculated by using eqn [81]; corrosion rate in mixed CO2/H2S environments can
4. a new time interval Dt is set and steps 2 and 3 be calculated as
repeated.
CR ¼ CRH2 S þ CRHþ þ CRCO2 ½91
At very low H2S gas concentrations (ppmw range),
there is very little dissolved H2S and the corrosion
rate is directly affected by pH. A mackinawite layer
still forms and controls the corrosion rate; however, 2.25.3.4 Limitations of Modeling of
the corrosion process is largely driven by the reduc- Aqueous H2S Corrosion of Mild Steel
tion of Hþ ions, rather than of H2S. By analogy with
The calculation model presented above covers
the approach laid out above, the following expression
uniform H2S and CO2/H2S corrosion. There are
is obtained for the flux of Hþ ions controlled by the
numerous limitations:
presence of the ferrous sulfide layers:
  It does not predict localized corrosion in either
dos 1
cHþ  FluxHþ þ environment.
DHþ ec kmðHþ Þ While it covers a very broad range of H2S partial
FluxHþ ¼ AHþ ln
csðHþ Þ pressures, it is not recommended to use this model
½86 below pH2 S ¼ 0.01 mbar or above pH2 S ¼ 10 bar.
Similar limits apply to the CO2 partial pressure.
Corrosion in Acid Gas Solutions 1293

This leaves a very broad area of applicability for main underlying assumptions about the formation
the present model. and protective nature of a mackinawite layer are
This H2S model does not account for any precipi- correct. The comparison of the performance of this
tation of ferrous sulfide, ferrous carbonate, or any model with experimental data is given in the follow-
other scale; therefore, in cases where this is ing section, which covers the main factors affecting
deemed important for corrosion, the model should CO2/H2S corrosion of mild steel.
be used with caution. The model also does not
account for various transformations of sulfide
layer from one type to another which are known 2.25.3.5 Key Factors Affecting Aqueous
to happen over time. H2S Corrosion of Mild Steel
The present model does not account for dissolu-
Some of the key factors affecting aqueous H2S corro-
tion of the sulfide layer that may occur at very low
sion are discussed in these sections, and the experimen-
pH. Therefore, the use of this model at pH < 3 is
tal results are compared with the model described
not recommended. Similarly, the model should be
above.
used with caution for pH > 7 where it has not been
tested.
2.25.3.5.1 Effect of H2S partial pressure
The model in its present state does not cover the
Corrosion rate of mild steel at extremely low
effect of organic acids on mixed H2S and CO2/
H2S partial pressures is seen in Figure 17 wherein
H2S corrosion, and therefore it should not be used
atmospheric glass cell experiments pH2 S ranged from
when organic acids are present in the system.
0.0013–0.32 mbar, corresponding to 1–250 ppmm in the
A practical threshold for the validity of the present
gas phase at 1 bar CO2. Clearly, this is a CO2 dominated
model is <1 ppm of organic acids in the brine.
corrosion scenario (pCO2 /pH2 S ratio is in the range
The model does not account for the effect of
103–106); however, the presence of H2S controls the
high chloride concentrations, oxygen, elemental
corrosion rate. Even when present in such minute
sulfur or any other unspecified condition, which
amounts, H2S reduces the pure CO2 (H2S-free) corro-
is known to affect the corrosion rate and is not
sion rate by 3–10 times due to the formation of a thin
explicitly covered in the theoretical underpinnings
mackinawite film. The model presented above success-
discussed above.
fully captures this effect as shown in Figure 17.
While this calculation model is clearly not inclu- At higher H2S partial pressures, the same effect is
sive of all the important processes in aqueous observed as shown in Figure 18, which shows results
H2S corrosion of mild steel, it is believed that the from autoclave experiments conducted at a very high

H2S gas concetration (ppmm)


0 50 100 150 200 250
10
Mod.
Exp.
Corrosion rate (mm year–1)

Pure CO2 corrosion rate


1

0.1

0.01
0 0.1 0.2 0.3
H2S partial pressure (mbar)

Figure 17 The corrosion rate vs. partial pressure of H2S; experimental data (exp.) shown as points, model predictions
(mod.) shown as lines; conditions: total pressure p ¼ 1 bar, pCO2 ¼ 1 bar, pH2 S ¼ 0.0013–0.32 mbar, T ¼ 20  C, reaction time
24 h, pH 5, 1000 rpm. For reference: pure CO2 corrosion rate is measured to be 1 mm year1. Data taken from Lee.41
1294 Liquid Corrosion Environments

total pressure (p ¼ 138 bar) and a high CO2 partial ratio of 3500 (pCO2 ¼ 13.8 bar, pH2 S ¼ 40 mbar), CO2
pressure (pCO2 ¼13.8 bar). When comparing the pre- accounts for 70% of the corrosion rate and 30%
dictions with the experimental results, it can be seen can be ascribed to H2S. At the lowest pCO2 /pH2 S ratio
that the model underpredicts the observed rate of 1180 (pCO2 ¼ 13.8 bar, pH2 S ¼ 116 mbar), CO2
of steel corrosion by approximately a factor of 2. accounts for 57% of the corrosion rate and 43%
However, when this is compared with a pure CO2 can be ascribed to H2S.
(H2S-free) corrosion rate under the same conditions Corrosion rates of mild steel at very high partial
(which is not reported but can be predicted to be pressures of H2S (pH2 S ¼ 3–20 bar) and CO2
almost 20 mm year1), the accuracy of the model can (pCO2 ¼ 3–12.8 bar) for exposures lasting up to 4 days
be considered as reasonable. At the highest pCO2 /pH2 S are shown in Figure 19. This is a situation where the

H2S gas concentration (ppmm)


0 100 200 300 400 500 600 700
20
Mod.
Pure CO2 corrosion rate Exp.
Corrosion rate (mm year –1)

15

10

0
0 20 40 60 80 100 110 120
H2S partial pressure (mbar)

Figure 18 The corrosion rate vs. H2S partial pressure; experimental data (exp.) shown as points, model predictions
(mod.) shown as lines; conditions: total pressure p ¼ 137.9 bar, pCO2 ¼13.8 bar, pH2 S ¼ 40–120 mbar, T ¼ 50  C, experiment
duration 3 days, pH 4.0–6.2, stagnant. Experimental data taken from Smith and Pacheco et al.31

40
Test A and B mod. Test A and B exp.
35 Test C mod. Test C exp.
Test D mod. Test D exp.
Corrosion rate (mm year –1)

30 Test E mod. Test E exp.


Test F mod. Test F exp.

25

20

15

10

0
0 20 40 60 80 100
Time (h)
Figure 19 The corrosion rate vs. time; experimental data (exp.) shown as points, model predictions (mod.) shown as lines;
Test A and B: p ¼ 8.3 bar, pCO2 ¼ 5.3 bar, pH2 S =3 bar, T ¼ 60  C, 71 h (a) and 91 h (b); Test C: p ¼ 24 bar, pCO2 ¼ 4 bar,
pH2 S ¼20 bar, T ¼ 70  C, 91 h; Test D: p ¼ 15.7 bar, pCO2 ¼ 3.5 bar, pH2 S ¼12.2 bar, T ¼ 65  C, 69 h; Test E: p ¼ 20.8 bar,
pCO2 ¼12.8 bar, pH2 S ¼8 bar, T ¼ 65  C, 91 h; Test F: p ¼ 7.2 bar, pCO2 ¼3 bar, pH2 S ¼4.2 bar, T ¼ 65  C, 63 h; experimental data
taken from Bich and Goerz.43
Corrosion in Acid Gas Solutions 1295

H2S was the dominant corrosive species. At the high- in the less extreme experiments 1 and 2 (pCO2 ¼ 3.3
est pCO2 /pH2 S ratio of 1.8 (pCO2 ¼ 5.3 bar, pH2 S ¼ 3 bar), bar; pH2 S ¼ 10 bar) both at low (25  C) and high tem-
H2S generated 86% of the corrosion rate. At the perature (80  C). In experiment 3 which was con-
lowest pCO2 /pH2 S ratio of 0.2 (pCO2 ¼ 4 bar, pH2 S ¼ 20 ducted at the most extreme set of conditions
bar), H2S generated 97% of the overall corrosion rate. (pCO2 ¼ 10 bar; pH2 S ¼ 30 bar) and high temperature
It is also noted that the model predictions show that (80  C) the model overpredicts the corrosion rate by
the corrosion rate in the first reaction hour is on a factor of 2.5. In all three experiments reported by
average 20 mm year1 with an initial corrosion rate Omar et al.,44 the pCO2 /pH2 S ratio was about 0.3, that is,
of 60 mm year1 and a final corrosion rate of 10 mm the corrosion process and corrosion rate were
year1. The pitting corrosion rate was reported to be completely dominated by H2S, which contributed
30 mm year1 in a field case with similar conditions, 95% of the corrosion rate.
which is related to the very high, H2S-driven corro-
sion seen at the beginning of experiments before a
2.25.3.5.3 Effect of time
thick protective ferrous sulfide film forms.
A marked decrease of corrosion rate with time was
seen in autoclave tests as reported in Figure 19 above;
2.25.3.5.2 Effect of flow the same was observed in stratified pipe flow experi-
The effect of flow velocity in H2S corrosion is shown ments where pure CO2 corrosion rate decreased with
in Figure 20 for the three long-term experiments time due to the presence of H2S, as shown in Figure 21
reported by Omar et al.44 Flow loop experiments below. The latter is also a mixed CO2/H2S corrosion
lasting 15–21 days were conducted at severe condi- scenario. At a pCO2 /pH2 S ratio of 200 (pCO2 ¼ 2 bar,
tions: high partial pressure of H2S (pH2 S ¼10–30 bar), pH2 S ¼ 4 mbar), the CO2 contribution to the corro-
high partial pressure of CO2 (pCO2 ¼3.3–10 bar) and sion rate is 75% with most of the balance provided
low pH 2.9–3.2. No effect of velocity on the uniform by H2S. At the pCO2 /pH2 S ratio of 28 (pCO2 ¼ 2 bar,
corrosion rate could be observed in these long-term pH2 S ¼70 mbar), both CO2 and H2S account for
exposures, which is due to the build-up of a thick 50% of the overall corrosion rate.
protective sulfide layer. The model predictions also Corrosion experiments at high temperature
shown in Figure 20 confirm this trend and show a (120  C), high partial pressures of CO2 (pCO2 ¼ 6.9
remarkable agreement with the experimental results bar), and H2S (pH2 S ¼ 1.38–4.14 bar) in exposures

5.0
Exp. 1
Exp. 2
Exp. 3
4.0
Corrosion rate (mm year–1)

Mod. 1
Mod. 2
Mod. 3
3.0

2.0

1.0

0.0
0 1 2 3 4 5 6
Velocity (m s–1)
Figure 20 The corrosion rate vs. velocity; experimental data (exp.) shown as points, model predictions (mod.) shown as
lines; exp 1.: 19 days, p ¼ 40 bar, pCO2 ¼ 3.3 bar, pH2 S ¼ 10 bar, T ¼ 80  C, pH 3.1, v ¼ 1–5 m s1; exp 2.: 21 days, p ¼ 40 bar,
pCO2 ¼ 3.3 bar, pH2 S ¼10 bar, T ¼ 25  C, pH 3.2, v ¼ 1–5 m s1; exp 3.: 10 days, p ¼ 40 bar, pCO2 ¼ 10 bar, pH2 S ¼ 30 bar,
T ¼ 80  C, pH 2.9, v ¼ 1–5 m s1; experimental data taken from Omar et al.44
1296 Liquid Corrosion Environments

10 pH2S = 0 mbar (exp.)


pH2S = 4 mbar (exp.)
pH2S = 70 mbar (exp.)
Pure CO2 corrosion rate pH2S = 4 mbar (mod.)

Corrosion rate (mm year –1)


8 pH2S = 70 mbar (mod.)

0
0 100 200 300 400 500 600
Time (h)
Figure 21 The corrosion rate vs. time; experimental data (exp.) shown as points, model predictions (mod.) shown as lines;
conditions: total pressure p ¼ 3 bar, pCO2 ¼2 bar, pH2 S ¼ 3–70 mbar, T ¼ 70 C, experiment duration 2–21 days, pH 4.2–4.9,
liquid velocity 0.3 m s1. Experimental data taken from Singer et al.42

100
pH2S = 1.38 bar (exp.)

Pure CO2 corrosion rate pH2S = 2.76 bar (exp.)


pH2S = 3.45 bar (exp.)
Corrosion rate (mm year–1)

pH2S = 4.14 bar (exp.)


pH2S = 1.38 bar (mod.)
10 pH2S = 2.76 bar (mod.)
pH2S = 3.45 bar (mod.)
pH2S = 4.14 bar (mod.)

0
0 100 200 300 400 500 600
Time (h)
Figure 22 The corrosion rate vs. time; experimental data (exp.) shown as points, model predictions (mod.) shown as lines;
conditions: total pressure p ¼ 7 bar, pCO2 ¼ 6.9 bar, pH2 S ¼ 1.38–4.14 bar, T ¼ 120  C, experiment duration 1–16 days,
pH 3.95–4.96, liquid velocity 10 m s1. Experimental data taken from Kvarekval et al.45

lasting up to 16 days are shown in Figure 22. lowest pCO2 /pH2 S ratio of 1.67 (pCO2 ¼ 6.9 bar,
A steadily decreasing corrosion rate was observed pH2 S ¼ 4.14 bar), H2S generated 82% of the overall
due to build-up of a protective ferrous sulfide layer. corrosion rate.
The effect of pH2 S increase on corrosion rate was very The longest H2S containing corrosion experi-
small and practically vanished over time. Both these ments which are practically achievable in the lab are
effects were readily captured by the model with very of the order of a few weeks or at best a few months,
good accuracy as seen in Figure 22. In this case, the while predictions are meant to cover a period of
H2S is the dominant corrosive species. At the highest at least a decade, in order to be meaningful. With
pCO2 /pH2 S ratio of 5 (pCO2 ¼ 6.9 bar, pH2 S ¼ 1.38 bar), this in mind, it is interesting to take the experimental
H2S generated 70% of the corrosion rate. At the conditions above (pCO2 ¼ 6.9 bar, pH2 S ¼ 3.45 bar,
Corrosion in Acid Gas Solutions 1297

10.00
pH2S = 3.45 bar (mod.)
pH2S = 3.45 bar (exp)
Corrosion rate (mm year -1)

1.00

0.10

0.01
0.00 0.01 0.10 1.00 10.00 100.00
Time (year)
Figure 23 Extension of corrosion prediction to a 25-year lifetime; experimental (points), predicted (lines); conditions:
pCO2 ¼ 6.9 bar, pH2 S ¼ 3.45 bar, T ¼ 120  C, pH 4, liquid velocity 10 m s1. Data from Kvarekval et al.45

T ¼ 120  C, pH 4, v ¼ 10 m s1) and extend the simu- References


lation to 25 years. The result is shown in Figure 23.
The corrosion rate was predicted to start out rather 1. Bonis, M. R.; Crolet, J. L. In Corrosion/89; NACE
high as observed in the experiments; however, it was International: Houston, TX, 1989; Paper no. 466.
reduced to below 0.1 mm year1 after 2 years and 2. Oddo, J.; Tomson, M. In SPE of AIME; Society of
Petroleum Engineers: Richardson, TX, 1982; pp 1583–1590.
was as low as 0.03 mm year1 after 25 years. The aver- 3. Palmer, D. A.; van Eldik, R. Chem. Rev. 1983, 83, 651.
age corrosion rate over this period was only 0.06 mm 4. Sun, W.; Nešić, S.; Woollam, R. C. Corros. Sci. 2009,
year1, which amounts to a wall thickness loss of only 5.
51(6), 1273–1276.
Sun, W.; Nešić, S. Corrosion 2008, 64, 334.
1.5 mm over the 25 years, an acceptable amount 6. Drazic, D. M. In Aspects of Electrochemistry; Plenum
by any practical account. Actually, most of the other Press, 1989; Vol. 19, p 79.
7. Lorenz, W.; Heusler, K. In Corrosion Mechanisms;
conditions simulated have shown that rather low Mansfeld, F., Ed.; Marcel Dekker: New York, 1987.
H2S uniform corrosion rates are obtained for very 8. Nešić, S.; Thevenot, N.; Crolet, J. L. In Corrosion/96;
long exposures, which agrees with general field expe- NACE International: Houston, TX, 1996; Paper no. 3.
9. de Waard, C.; Milliams, D. E. Corrosion 1975, 31, 131.
rience as recently discussed by Bonis et al.33 Never- 10. Gray, L. G. S.; Anderson, B. G.; Danysh, M. J.;
theless, no quantitative long-term lab data are Tremaine, P. G. In Corrosion/89; NACE International:
currently available to back-up these long-term predic- Houston, TX, 1989; Paper no. 464.
11. Eriksrud, E.; Søntvedt, T. In Advances in CO2 Corrosion,
tions, and therefore they should be used with caution. Proceedings of the Corrosion/83 Symposium on CO2
Corrosion in the Oil and Gas Industry; Hausler, R. H.,
Goddard, H. P., Eds.; NACE: Houston, TX, 1984; Vol. 1, p 20.
2.25.3.6 Localized H2S Corrosion of Mild 12. Schmitt, G.; Rothman, B. Werkst. Korros. 1977, 28, 816.
Steel in Aqueous Solutions 13. Gray, L. G. S.; Anderson, B. G.; Danysh, M. J.;
Tremaine, P. R. In Corrosion/90; NACE International:
Localized H2S corrosion of mild steel is even less Houston, TX, 1990; Paper no. 40.
14. Delahay, P. J. Am. Chem. Soc. 1952, 74, 3497.
understood than its uniform counterpart. While it is 15. Nešić, S.; Postlethwaite, J.; Olsen, S. In Corrosion/95;
not very common, anecdotal evidence exists that has NACE International: Houston, TX, 1995; Paper no. 131.
linked localized H2S corrosion in aqueous environ- 16. Nešić, S.; Pots, B. F. M.; Postlethwaite, J.; Thevenot, N.
J. Corros. Sci. Eng. 1995, 1 Paper no. 3. Available at
ments to other factors such as high chloride content, http://www.cp.umist.ac.uk/JCSE/Vol1/PAPER3/
the presence of elemental sulfur and the transforma- V1_p3int.htm
tion of one type of sulfide into another. Intense research 17. Chokshi, K.; Sun, W.; Nešić, S. In Corrosion/05; NACE
International: Houston, TX, 2005; Paper no. 285.
of these topics is ongoing with a breakthrough in 18. Sun, Y.; Nešić, S. Corrosion/2004; NACE International:
understanding expected in the decade to come. Houston, TX, 2004; Paper no. 380.
1298 Liquid Corrosion Environments

19. Nešić, S.; G.Solvi, T.; Enerhaug, J. Corrosion 1995, 51, 32. Smith, S. N.; Joosten, M. In Corrosion/2006; NACE
p 773. International: Houston, TX, 2006; Paper no. 06115.
20. Sun, Y.; George, K.; Nešić, S. In Corrosion/2003; NACE 33. Bonis, M.; Girgis, M.; Goerz, K.; MacDonald, R. In
International: Houston, TX, 2003; Paper no. 3327. Corrosion/2006; NACE International: Houston, TX, 2006;
21. Gunaltun, Y. M.; Larrey, D. In Corrosion/2000; NACE Paper no. 06122.
International: Houston, TX, 2000; Paper no. 71. 34. Suleimenov, O. M.; Krupp, R. E. Geochim. Cosmochim.
22. Vitse, F.; Nešić, S.; Gunaltun, Y.; Larrey de Torreben, D.; Acta 1994, 58, 2433–2444.
Duchet-Suchaux, P. Corrosion 2003, 59, 1075. 35. Suleimenov, O. M.; Seward, T. M. Geochim. Cosmochim.
23. Anderko, A.; Young, R. In Corrosion/99; NACE Acta 1997, 61, 5187–5198.
International: Houston, TX, 1999; Paper no. 31. 36. Benning, L. G.; Wilkin, R. T.; Barnes, H. L. Chem. Geol.
24. de Waard, C.; Lotz, U. Corrosion/93; NACE International: 2000, 167, 25–51.
Houston, TX, 1993; Paper no. 69. 37. Sun, W.; Nesic, S.; Young, D.; Woollam, R. Ind. Eng.
25. Berger, F. P.; Hau, K.-F. F.-L. Int. J. Heat Mass Transfer Chem. Res. 2008, 47(5), 1738–1742.
1977, 20, 1185. 38. Criaud, A.; Fouillac, C.; Marty, B. Geothermics
26. Eisenberg, M.; Tobias, C. W.; Wilke, C. R. J. Electrochem. 1989, 18(5–6), 711–727.
Soc. 1954, 101, 306. 39. Sun, W.; Nesic, S.; Papavinasam, S. Corrosion 2008,
27. Shoesmith, D. W.; Taylor, P.; Bailey, M. G.; Owen, D. G. J. 64(7), 586–599.
Electrochem. Soc. 1980, 125, 1007–1015. 40. Sun, W.; Nesic, S. Corrosion 2009, 65(5), 291–307.
28. Shoesmith, D. W. Formation, transformation and 41. Lee, K. J. Ph.D. dissertation. Ohio University, 2004.
dissolution of phases formed on surfaces, In Lash Miller 42. Singer, M.; Brown, B.; Camacho, A.; Nešić, S.
Award Address, Electrochemical Society Meeting, In Corrosion/07; NACE International: Houston, TX, 2007;
Ottawa, 27 November 1981. Paper no. 07661.
29. Smith, S. N. In Twelfth International Corrosion 43. Bich, N. N.; Goerz, K. In Corrosion/1996; NACE
Congress, Houston, TX, 19–24 September 1993; Paper International: Houston, TX, 1996; Paper no. 26.
no. 385. 44. Omar, I. H.; Gunaltun, Y. M.; Kvarekval, J.; Dugstad, A. In
30. Smith, S. N.; Wright, E. J. In Corrosion/94; NACE Corrosion/05; NACE International: Houston, TX, 2005;
International: Houston, TX, 1994; Paper no. 11. Paper no. 05300.
31. Smith, S. N.; Pacheco, J. L. In Corrosion/2002; NACE 45. Kvarekval, J.; Nyborg, R.; Choi, H. In Corrosion/03; NACE
International: Houston, TX, 2002; Paper no. 02241. International: Houston, TX, 2003; Paper no. 03339.

You might also like