You are on page 1of 13

A Proposed Mechanism for Corrosion in Slightly Sour Oil and

Gas Production

Stephen N. Smith

Exxon Production Research Co.


P.O. Box 2189
Houston, Texas 77001

Abstract

A corrosion model is proposed to explain corrosion in oil and gas production environments that contain
CO2 and small quantities of H2S. Although a number of studies have been conducted into the
mechanism of corrosion by CO2 and there has been some study of the corrosion mechanism by H2S,
information about the transition between these corrosion mechanisms has not received much attention.
The proposed model is based upon the formation of a meta-stable form of iron sulfide, mackinawite,
which has been observed as a corrosion product both in the laboratory and in the field. Mackinawite
has been observed as the corrosion product that forms when small concentrations of H2S are present in
the corrosive fluids. A simple partial pressure or solution concentration guideline for the amount of
H2S required to form mackinawite on the iron surface cannot be developed since factors such as
temperature, H2S activity, pH and the ionic strength of the solution are all involved in the determination
of the conditions required for formation. Mackinawite is believed to form as a surface reaction between
the H2S in solution and the iron rather than as a precipitation reaction of ferrous sulfide from solution.
Therefore, even when the conditions for mackinawite formation are favorable on the iron surface,
mackinawite can remain soluble in the bulk solution. This means that as mackinawite forms on the
surface, the mackinawite immediately begins to dissolve. However, since the formation reaction has
been shown to be much more rapid than the dissolution reaction, a mackinawite film will form and
influence the rate of further corrosion.

The interface between the mackinawite region and the CO2 corrosion region is believed to be
independent of the CO2 corrosion reaction and is defined solely by the chemistry of mackinawite
formation. As the concentration of H2S is increased, an upper H2S boundary to the mackinawite region
is reached. This boundary is defined by the conditions where the solubility product for iron sulfide is
exceeded. Under these conditions the thermodynamically stable iron sulfide, pyrrhotite, precipitates on
the metal surface. Once pyrrhotite forms, any remaining mackinawite rapidly dissolves or transforms to
the slower growing pyrrhotite.

Key terms: Mackinawite, Sour Corrosion, H2S Corrosion

Introduction

The formation of an iron sulfide corrosion product at low H2S concentrations can have many practical
implications to oilfield corrosion control. Several situations have been encountered in which the
change from a CO2 controlled reaction to H2S controlled corrosion altered the corrosion rate and the
physical location of the most corrosive point within the production system. In some cases, the
changeover increased corrosion and in others the change significantly reduced the amount of corrosion
to levels below what was expected. The morphology of corrosion is also generally different between
CO2 and H2S control. Where CO2 corrosion often produces either general or mesa corrosion, H2S
control is typified by discrete local pitting attack. Thus, the change can have an important effect

upon design factors such as corrosion allowance considerations or in the selection and location of the
most appropriate corrosion monitoring system.

Although a number of studies of the CO2 corrosion mechanism have been conducted and there has been
some study of the corrosion mechanism by H2S, the transition region between these two corrosion
mechanisms in environments with low to moderate levels of CO2 and H2S has not received much
attention. This is especially true in the oil and gas industry where the majority of the effort has
concentrated on defining the mechanism of CO2 corrosion and the remaining research has looked at the
mechanism of corrosion in high concentrations of H2S and/or elemental sulfur.

For corrosive conditions where CO2 is present with virtually no H2S, the current theories state that the
corrosion mechanism will be determined by the fate of the Fe2 ion corrosion product. If the Fe2 remains
in solution, the corrosion mechanism will follow the chemistry originally described in deWaard and
Milliamsl, 2. If conditions conducive to FeCO3 formation are favorable, further corrosion is determined
by the quantity and quality of the FeCO3 corrosion product3, 4. When significant concentrations of H2S
are present in the system, the corrosion mechanism is controlled by the formation of the pyrrhotite form
of FeS5-7. None of these mechanisms explains the formation of the mackinawite form of iron sulfide
that was originally named kansite by Prange8. The practical existence of mackinawite as a corrosion
product has subsequently received extensive discussions.

In work performed at the Canadian AEC10, 11-20, a mechanism for the formation of a mackinawite scale
was reported for the Girdler-Sulfide method for concentrating heavy water, which utilizes H2S. The
Girdler-Sulfide H2S concentrations studied were significantly lower than those studied by
Ramanarayanan5,6 and Vedage7. The mackinawite corrosion product may therefore exist somewhere
between the two CO2 corrosion regions (Fe2 and FeCO3) and the pyrrhotite corrosion product region, as
illustrated in Figure 1.

This paper will utilize the work of the Canadian AEC and others10-24, to propose a mechanism for
mackinawite formation under oilfield production conditions. The paper will also describe the proposed
chemistry required for mackinawite formation. Methods for predicting the boundary conditions that
separate the mackinawite region from the other three corrosion product regimes are also discussed
along with suggestions for further laboratory work to verify the model.

Mackinawite

Mackinawite was first identified as a corrosion product by Prange8, who named it kansite. Prange
identified mackinawite as the corrosion product that had been observed in a number of sour oil
production systems in Kansas and West Texas. The official mineralogical name, Mackinawite, was
adopted many years later in 196421. Early corrosion literature citations of kansite are therefore usually
describing mackinawite.

Mackinawite is a thermodynamically meta-stable form of iron sulfide. Mackinawite will decompose to


the pyrrhotite form of iron sulfide. However, provided favorable conditions are maintained, this
decomposition may not occur for weeks or months. Like pyrrhotite, mackinawite is ferromagnetic and
can be highly reactive when fine powders are exposed to oxygen.
Mackinawite exists as the non-stoichiometric form of iron sulfide with the imbalance toward the
ferrous ion side. Although the non-stoichiometry has not been proven to be an excess of ferrous ions,
as opposed to sulfide ion vacancies, some literature seems to favor the excess ferrous ion optionl2, 13
where the composition would be expressed as Fe(1+X)S. Other literature22 presents evidence that
mackinawite may exist as the sole example of a sulfur deficient metal sulfide. In this case the
composition would be expressed as FeS(1-X)l

Mackinawite Corrosion Mechanism

Thermodynamics favors either pyrrhotite or pyrite as the iron sulfide corrosion product over
mackinawite. However, the rapid formation kinetics of mackinawite appears to give it the edge as the
initial iron sulfide corrosion product. Mackinawite can be observed to form rapidly (virtually
instantaneously) on the surface of carbon steel coupons dipped into H2S saturated deionized water at
room temperature. Shoesmithl5 proposed that "the first stage in the corrosion of carbon steel in aqueous
H2S is the formation of a mackinawite layer, via a solid-state mechanism."

An explanation for the preference for solid-state mackinawite formation over pyrrhotite is that the
equilibrium Fe2 and H2S concentrations in the bulk solution are not conducive to pyrrhotite
precipitation as determined by the pyrrhotite solubility product. However, the activity of Fe2 on a
carbon steel surface approaches 1.0 since the Fe activity on the solid steel surface is defined as 1.0. If
the combination of the increased Fe2 ion activity and the local H2S activity exceeds the equilibrium
constant for mackinawite formation on the metal surface, a mackinawite tarnish will form as a surface
reaction on the steel. Rickard23 studied kinetics of this reaction and found them to be extremely fast.
Mackinawite was observed after exposures of less than 0.1 seconds in solutions of 10-3 M Na2S at pH
7.0. By comparison, pyrrhotite formation kinetics appear to be quite slow. Shoesmithl6 and Wikjord20
both reported observing troilite (stoichiometric FeS) and pyrrhotite only after exposures greater than 5
hours. Mackinawite formation would therefore be favored kinetically over pyrrhotite.

However, if the Fe2 activity of the bulk solution is below the solubility product for mackinawite, the
mackinawite is still soluble and will immediately begin to dissolve as the Fe2 ions diffuse away from
the steel surface. If the Fe2 activity of the bulk solution is above the solubility product for the more
stable FeS, pyrrhotite, FeS will start to precipitate from solution. The driving force for mackinawiite
dissolution will also be reduced as the Fe2 ion concentration increases. This combination provides the
conditions that are conducive to pyrrhotite formation.

Since mackinawite formation kinetics are so fast, Tewari10 found that the rate of corrosion on a
mackinawite covered surface is determined by the combined FeS dissolution and corrosion product
diffusion reactions. In fact, Tewari hypothesized the existence of an Fe(HS)+ complex based upon his
corrosion product diffusivity values. Shoesmithl6, Taylorl7 and Ogundele24 also have discussed the
existence of an Fe(HS)+ complex based upon FeS dissolution.

Therefore, mackinawite would be expected to form under conditions where the combined Fe2 and H2S
bulk solution activities are inadequate to precipitate FeS, but where the H2S activity is sufficient to
form mackinawite if the Fe2 activity on the steel surface is assumed to be 1.0. As illustrated in Figure 2,
the H2S diffuses to the metal surface and reacts with the steel surface to first form adsorbed molecular
FeS which then combines to form mackinawite. However, since the FeS is still soluble in the bulk
solution, the mackinawite immediately begins to dissolve to form Fe(HS)+, probably through a reaction
sequence such as:
FeS + H2S Fe(HS)2 Fe(HS)+ + HS- (1)

As soon as the Fe(HS)+ has had an opportunity to diffuse away, a fresh steel surface is exposed which
can immediately react with the H2S to form more mackinawite. The mechanism therefore assumes the
following:

FeS(pyrrhotite) is saturated in solution,


H2S will diffuse toward the Fe surface,
A reaction between Fe and H2S occurs on the surface to form FeS,
FeS begins to dissolve to form Fe(HS)+ and HS-,
The Fe(HS)+ diffuses away from the metal surface, and
More H2S immediately moves in to react with the exposed Fe.

This produces a very thin tarnish of mackinawite that is continually forming and dissolving. This is
consistent with observations where mackinawite is never observed as the thick scale that often forms
when pyrrhotite is the corrosion product. This also explains the poor x-ray diffraction patterns
frequently encountered with mackinawite. The thin FeS film, combined with the extremely fine crystal
size found in a tarnish layer, produces a very broad set of diffraction peaks, as illustrated in Figure 3.
This occurs even for the (001) crystal lattice reflection that produces the 100% I peak for
mackinawite2l. This often results in mackinawite not being detected by routine, computerized x-ray
diffraction analyses as the peak is too broad for the computer to identify. The problem is further
complicated by the low angle diffraction produced by the 5.03 spacing of the (001) plane, which can
disappear into the background noise produced by the commonly used Cu-K, excitation signal.

Definition of Corrosion Product Boundaries

Assuming that this explanation for mackinawite formation is correct, the boundary conditions between
the mackinawite corrosion product region and the other corrosion products can be defined.

Boundary with Fe2 Region

Since most oilfield waters have pH values in range of 3 to 6, the majority of the aqueous sulfide found
in oilfield waters exists as H2S rather than as HS- or S-2 ions. The equilibrium reaction between the
corrosion product regions for Fe2 and mackinav;ite can be expressed by the reaction-.

Fe2 + H2S FeSmackinawite + 2H+ (2)

Although equilibrium constant information for mackinawite formation as a function of temperature is


not known to be available, Bemer25 reported the standard free energy of formation for mackinawite and
MacDonaldl4 described a procedure to estimate mackinawite free energy as a function of temperature.
The calculated free energy can then be used to develop a value for Keq as a function of T.

For temperature T and assuming a 1.0 molal activity of all reactants and products,

GTreaction = (GTmackinawite + 2*GTH+) - (GTFe2 + GTH2S) (3)

Correcting for the fact that the reactants and products are normally not present at unit activity:

GTreaction = GTreaction - RT In KFeS (4)


where the equilibrium constant for reaction described by equation 2 is:

K FeS / Fe2 =
(a mackinawit e (aH )2 ) +
(5)
(aFe 2 a H2Saq )
where amackinawite is the mackinawite activity on the tarnish surface,
aH+ is the H+ activity in the solution,
aH2Saq is the activity of the H2S dissolved in the solution,
and aFe2 is the Fe2 activity in the solution.

The H2S value that defines the mackinawite/Fe2 boundary can then be calculated by equation 5.
However, the activity of the H2S dissolved in solution is generally not a readily available value. This
can be corrected by determining the aqueous H2S activity from the H2S fugacity in the gas phase.

H2Svap H2Saq (6)

a H2Saq
K H2Saq / H2Svap = (7)
a H2Svap

Since amackinawite 1.0 and aFe2 is assumed equal to 1.0 on the surface, equations 5 and 7 can be
combined and reduced to:

(aH )2
+
K FeS / Fe2 = (8)
K H2Saq / H2Svap a H2Svap

It is interesting to note that critical H2S activity for mackinawite formation is highly dependent upon
pH. In fact, H+ activity is a second order function and is therefore far more important than H2S activity
in defining the boundary. Attempts to use a single H2S value, such as a gas phase concentration of 50
ppm, would therefore appear to be misguided without first specifying a pH.

Although this equation appears to be fairly simple, the fact that the H2S values appear as activities, as
opposed to an aqueous H2S concentration or a gas phase partial pressure, makes application of this
relationship difficult. If the aqueous phase H2S solution concentration is known, the activity coefficient
for dissolved H2S, which often deviates substantially from 1.0, must be known before the boundary can
be defined to any useful degree of accuracy. If a gas phase partial pressure of H2S is known, then the
fugacity coefficient must be defined and the gas and aqueous phases must be assumed to be in
equilibrium before an aqueous phase H2S activity can be determined.

Boundary with FeCO3 Region

FeCO3 is known to be the corrosion product that is favored at elevated temperatures when a mildly
acidic aqueous solution containing dissolved CO2 is the corrosive. Similar to the mackinawite/Fe2
boundary, the boundary between the FeCO3 and mackinawite regions can be defined by the reaction:

FeCO3 + H2S FeSmackinawite + H2CO3 (9)


Tabulated Keq values as a function of temperature are available for FeCO3 and using the thermodynamic
data for mackinawite that was described above, an equilibrium constant for this reaction can be
determined by:

(amackinawit e aH CO ) 2 3
K FeS / FeCO3 = (10)
a FeCO3 a H2Saq

where amackinawite is the mackinawite activity on the surface,


aH2CO3 is the solution activity of H2CO3,
aH2Saq is the solution activity of H2S, and
aFeCO3 is the FeCO3 activity on the corrosion product surface.

However since asolids 1.0, equation 10 becomes:

a H2CO3
K FeS / FeCO3 = (11)
a H2Saq

Since the vapor phase CO2 activity is related to the H2CO3 solution activity through the following
reactions, the usefulness of this expression can be enhanced by making the following substitutions for a
H2CO3 and aH2Saq.

H2CO3 H2O + CO2aq (12)

(aH 2O
a CO 2aq )
K H2CO3 / CO 2aq = (13)
a H2CO3aq

CO2vap CO2aq (14)

a CO2aq
K CO 2aq / CO2 vap = (15)
a CO 2vap

Rearranging terms and substituting expressions 7, 13 and 15 for aH2C03 and aH2Saq into equilibrium
expression 11, the equation becomes:

(aH )
K CO 2aq / CO2 vap a CO 2vap
2O
K FeS / FeCO3 = (16)
(
K H2CO3aq K H2Saq / H2Svap a H2Svap )
This can be further simplified by combining the equilibrium constants into a single term and making the
simplifying assumption that the activity of water is 1.0 to produce:

a CO2 vap
K FeS / FeCO3 = C (17)
a H2Svap
If the oversimplifying assumption is made that activity equals partial pressure for CO2 and H2S in the
gas phase, the form of this equation is consistent with the concept of critical CO2/H2S ratios that have
been reported for various fields26,27.

Mackinawite Upper Temperature Boundary

Since mackinawite is a meta-stable phase of FeS, it is intuitively satisfying that there should be an
upper temperature limit to the phase's meta-stability. Takeno28,29 synthesized mackinawite by reacting
iron with a 1 atm solution of H2S in distilled water at 50oC. He then heated the mackinawite under
vacuum to temperatures ranging from room temperature to 300oC for periods up to 34 days. X-ray
diffraction analyses were then conducted on the thermally treated FeS samples to determine the
stability region for mackinawite.

Takeno found that mackinawite went through a solid state transformation to pyrrhotite at temperatures
between 170o and 200oC. This was illustrated by the sample baked at 170oC for 30 days, which showed
signs of both pyrrhotite and mackinawite whereas the 200oC sample was completely transformed to
pyrrhotite. Taylorl7 stated that the upper limit of thermal stability is 130oC. Either way, it is not known
whether this limit represents a thermodynamic limit or is somehow tied to the sluggish kinetics of the
mackinawite/pyrrhotite transformation at this low temperature.

It is therefore possible to conclude that an upper boundary exists and that it appears to be somewhere in
the range of 130o to 170oC. The transition temperature may be toward the lower end of the range if
exposure times longer than Takeno used are considered.

Pyrrhotite Boundary

Studies of the mechanism that defines the low temperature boundary between the two iron sulfides,
mackinawite and pyrrhotite, have not been reported. Extension of the mechanism proposed in this work
to the pyrrhotite/mackinawite boundary presumes that the boundary is defined by the saturation of
pyrrhotite in the bulk solution. If the solution is undersaturated with respect to pyrrhotite, then
mackinawite tarnish will be the kinetically preferred iron sulfide corrosion product. As the solution
activity of H2S and Fe2 ion (including Fe(HS)+ and any other Fe/S complex ions that may exist)
approaches the saturation value for pyrrhotite, the mackinawite dissolution reaction may slow or stop as
the effects of the increased bulk Fe2 activity reduce the Fe2 ion flux from the surface. This provides the
time required for pyrrhotite nucleation kinetics to form pyrrhotite crystals on the surface. The
pyrrhotite crystals then grow, cover the surface and thereby gain control of the corrosion mechanism, as
described by Ramanarayanan.

Therefore, the boundary between the mackinawite and pyrrhotite regions is defined by pyrrhotite
saturation in the bulk solution. The boundary chemistry is determined in the same overall manner as
the Fe/mackinawite boundary, except that aFe2 is no longer defined as 1.0 but is defined by the Fe2 ion
activity in solution and the Keq is based upon the free energy of formation of pyrrhotite rather than
mackinawite.

K pyrr / Fe2 =
(a pyrrhotite (a H )2 )
+
(18)
(aFe 2 a H2Saq )
since apyrrhotite 1.0,
(aH )2
+
K pyrr / Fe2 = (19)
(aFe 2 aH2Saq )

As with other activity values, extreme care must be taken when converting from solution concentrations
to activity values. This is especially true for divalent ions like Fe2. A modified Bromley formalism for
calculating activity coefficients was used to estimate the activity coefficient for Fe2 in a 3 wt% NaCl
solution at 50oC. The activity coefficient was determined to be less than 0.2. Using the simplifying
assumption that the activity coefficient is equal to 1.0 would therefore result in the under-estimation of
the critical Fe2 content for pyrrhotite formation by a factor of five times.

The complexing nature of Fe2 ions in aqueous chloride media also complicates the matter of
determining the critical Fe2 level to form pyrrhotite. Oilfield produced water analysis procedures
intentionally acidize the water to yield a total iron ion value and this acidification breaks down most of
the iron complexes that would be found in service. Only ferrous ions that exist as free Fe2 or as
Fe(HS)yx complexes are probably available for the pyrrhotite formation reaction. If significant
quantities of iron/chloride, iron/oxy-hydroxide and/or iron bicarbonate type complexes exist, they may
reduce aFe2 sufficiently to influence the location of the pyrrhotite/mackinawite boundary.

Suggestions for Further Work to Verify Model

There is surprising little corrosion laboratory data reported about the formation of mackinawite. The
vast majority of the work that is reported was conducted in water saturated with H2S at atmospheric
pressure and at temperatures significantly less than 100oC. The numerous difficulties in conducting
carefully controlled experiments in this area certainly explains the limited volume of data. Besides the
safety problems related to working with H2S, the problems include:

the meta-stability of mackinawite,


the complication of the corrosion reaction consuming H2S, which must be available in ppm
level concentrations to limit pyrrhotite formation,
changing Fe2 ion buildup in solution through the experiment,
the complication that the corrosion reaction increases the pH of the bulk solution by consuming
H+ and releasing HS-, which decreases the calculated amount of H2S required for mackinawite
formation, and
the physical difficulty of conducting experiments at the elevated temperatures and pressures
representative of oilfield production conditions.

The answer to conducting this type of experiment is the use of a continuously replenished flow system
that is capable of handling H2S containing salt solutions at temperatures up to 150oC and 150
atmospheres. Such systems are complex, expensive to build and can be difficult to operate safely.

An additional area that needs to be addressed is the question of activity coefficients, especially for Fe2,
H+, H2S, and CO2. Existing activity coefficient estimation procedures can be used, but their accuracy
for highly concentrated oilfield brine solutions is debatable. This has been a long standing physical
chemistry problem, so there is little hope for a theoretical breakthrough. However, collection of the
multi-component interaction parameters specific to oilfield waters for the Bromley, Meissner or Pitzer
activity coefficient formalisms might help to improve the accuracy of these predictions.
A final suggested area of research is the upper temperature meta-stability limit of mackinawite. Takeno
conducted studies that lasted up to 34 days. Longer studies at more temperatures than just 170o and
200oC would help to improve the definition of the upper temperature limit of mackinawite.

Summary

The existence of a corrosion product regime between the sweet corrosion products, Fe2 and FeCO3, and
the sour corrosion product pyrrhotite (FeS) has been postulated. The existence of this FeS corrosion
product, mackinawite, has been confirmed in both the field and the laboratory. Equilibrium
relationships between mackinawite and the other known corrosion products have been developed and
are illustrated in the graph shown in Figure 4. The numbers on the graph reference the appropriate
equations in the text.

However, application of these equilibrium relationships may be very difficult to achieve in the field.
The thermodynamically non-ideal nature of oilfield gas and aqueous phases means that the use of
solution concentrations and gas phase partial pressures for the activities of the various chemical species
can introduce substantial errors. Unfortunately, calculation of activity coefficients in many oilfield
brine systems is also not an accurate science with the current level of technology.

Therefore, extensive laboratory investigations will be required under carefully control conditions to
confirm the corrosion model proposed by this paper. Even more investigation will then be required
before the thermodynamic equilibrium relationships could be extended to the kinetic calculations that
will be required to predict practical corrosion rates.

Bibliography

1. deWaard, C. and Milliams, D.E., "Prediction of Carbonic Acid Corrosion in Natural Gas
Pipelines," Proc. First International Conference on the Internal and External Protection of
Pipes. Paper F2. September 1975.
2. dewaard, C. and Milliams, D.E., "Carbonic Acid Corrosion of Steel," Corrosion, Vol. 31, No. 5,
pp. 177-181, 1975.
3. Schmitt, G., "CO2 Corrosion of Steels, An Attempt to Range Parameters and Their Effects,"
Advances in CO2 Corrosion, NACE, Houston, 1984.
4. Burke, P.A., "Synopsis: Recent Progress in the Understanding of CO2 Corrosion," Advances in
CO2 Corrosion," NACE, Houston, 1984.
5. Ramanarayanan, T.A. and Smith, S.N., "Scaling Phenomena in High Temperature Aqueous and
Gaseous Environments Containing Sulfur," Proc. International Symposium of Corrosion
Science and Engineering, Brussels, March 1989.
6. Ramanarayanan, T.A. and Smith, S.N., "Corrosion of Iron in Gaseous Environments and in Gas-
Saturated Aqueous Environments," Corrosion, Vol. 46, No. 1, pp. 66-74, 1990.
7. Vedage, H., Ramanarayanan, T.A., Mumford, J.D. and Smith, S.N., "Electrochemical Growth of
Iron Sulfide Films in H2S Saturated Chloride Media," Corrosion, Vol. 49, No. 2, pp. 114-
121, 1993.
8. Meyer, F.H., Riggs, O.L., McGlasson, R.L. and Sudbury, J.D., "Corrosion Products of Mild Steel
in Hydrogen Sulfide Environments," Corrosion, Vol. 14, No. 2, pp. 69t-115t, 1958.
9. Smith, J.S. and Miller, J.D.A., "Nature of Sulfides and their Corrosive Effect on Ferrous Metals: A
Review," Br. Corrosion J., Vol. 10, No. 3, pp. 136-143, 1975.
10. Tewari, P.H. and Campbell, A.B., "Dissolution of iron during the initial corrosion of carbon steel
in aqueous H2S solutions," Canadian J. Chemistry, Vol. 57, pp. 188-196, 1979.
11. Pankow, J.F. and Morgan, J.J., "Dissolution of Tetragonal Ferrous Sulfide (Mackinawite) in
Anoxic Aqueous Systems, 1. Dissolution Rate as a Function of pH, Temperature and Ionic
Strength," Environmental Science and Technology, Vol. 13, No. 10, pp. 1248-1255, 1979.
12. Pound, B.G., Abdurrahman, M.H., Glucina, M.P., Wright, G.A. and Sharp, R.M., "The Corrosion
of Carbon Steel and Stainless Steel in Simulated Geothermal Media," Australian J.
Chemistry, Vol. 38, pp. 1133-1140, 1985.
13. Pound, B.G., Wright, G.A. and Sharp, R.M., "The Anodic Behavior of Iron in Hydrogen Sulfide
Solutions," Corrosion, Vol. 45, No. 5, pp. 386-392, 1989.
14. MacDonald, D.D. and Hyne, J.B., "The Thermodynamics of the Iron/Sulfide/Water System,"
Atomic Energy of Canada Ltd, Report AECL-5811, 1979.
15. Shoesniith, D.W., "Formation, Transformation and Dissolution of Phases Formed on Surfaces,"
Lash Miller Award Address, Electrochemical Society Meeting, Ottawa, Nov. 27, 1981.
16. Shoesniith, D.W., Taylor, P., Bailey, M.G. and Owen, D.G., "The Formation of Ferrous
Monosulfide Polymorphs during the Corrosion of Iron by Aqueous Hydrogen Sulfide at
21oC, " J. Electrochemical Society, Vol. 127, No. 5, pp. 1007-1015, 1980.
17. Taylor, P., "The stereocheniistry of iron sulfides -- a structural rationale for the crystallization of
some metastable phases from aqueous solution," American Mineralogist, Vol. 65, pp. 1026-
1030, 1980.
18. Tewari, P.H., Wallace, G. and Campbell, A.B., "The Solubility of Iron Sulfides and Their Role in
Mass Transport in Girdler-Sulfide Heavy Water Plants," Canadian Atomic Energy
Commission, Report AECL5960, 1978.
19. Tewari, P.H., Bailey, M.G. and Campbell, A.B., "The Erosion-Corrosion of Carbon Steel in
Aqueous H2S Solutions Up to 120oC and 1.6 MPa Pressure," Corrosion Science, Vol. 19,
pp. 573-585, 1979.
20. Wikjord, A.G., Rummery, T.E., Doern, F.E. and Owen, D. G., "Corrosion and Deposition During
the Exposure of Carbon Steel to Hydrogen Sulfide-Water Solutions," Corrosion Science,
Vol. 20, pp 651-671, 1980.
21. Milton, C., "Kansite = Mackinawite, FeS", Corrosion, Vol. 22, No. 7, pp. 191-192, 1966.
22. Morse, J.W., Millerao, F.J., Cornwell, J.C. and Rickard, D., "The Chemistry of the Hydrogen
Sulfide and Iron Sulfide Systems in Natural Waters," Earth Science Reviews, Vol. 24, pp.
1-42, 1987.
23. Rickard, D., "Experimental concentration-time curves for the iron (II) sulfide precipitation process
in aqueous solutions and their interpretation," Chemical Geology, Vol. 78, pp. 315-324,
1989.
24. Ogundele, G.I. and White, W.E., "Some Observations on the Corrosion of Carbon Steel in Sour
Gas Environments: Effects of H2S and H2S/CO2/CH4/C3H8 Mixtures," Corrosion, Vol. 42,
No. 7, pp. 398-408, 1986.
25. Bemer, R.A., "Tetragonal Iron Sulfide," Science, Vol. 137, p. 669, 1962.
26. Dunlop, A.K., Hassell, H.L. and Rhodes, P.R., "Fundamental Considerations in Sweet Gas Well
Corrosion," Corrosion/83, Anaheim California, Paper 46.
27. Mlliams, D.E. and Kroese, C.J., "Aqueous Corrosion of Steel by H2S and H2S/CO2 Mixtures,"
Third International Conference on the Internal and External Protection of Pipes, London,
Paper H1, September 1979.
28. Takeno, S. Zoka, H. and Niihara, T., "Metastable Cubic Iron Sulfide -- With Special Reference to
Mackinawite," American Mineralogist, Vol. 55, pp. 1639-1649, 1970.
29. Takeno, S., "Thermal Studies on Mackinawitc," J. Science Hiroshima Univ., ser. C, 4, pp. 455-
478, 1965.
30. Zamaitis, J.F., Clark, D.M., Rafal, M. and Scrivner, N.C., Handbook of Aqueous Electrolyte
Thermodynamics, Am. Inst. of Chemical Engr., New York, 1986.

You might also like