Você está na página 1de 11

Corros Rev 29 (2011): 123133 2011 by Walter de Gruyter Berlin Boston. DOI 10.1515/CORRREV.2011.

021

Sulfidic corrosion in refineries a review

Raul B. Rebak
GE Global Research, Schenectady, NY, USA,
e-mail: rebak@ge.com

Abstract
Sulfidic corrosion of steels in refineries is a prevalent
phenomenon that occurs in oil containing sulfur species
between 230C and 425C. There are several internal and
external variables controlling the occurrence of sulfidic corrosion. The most important external factors are temperature,
concentration and type of sulfur species, and presence of
naphthenic acid. The most important internal or metallurgical
factor to control sulfidic corrosion is the amount of chromium
in the steel. The refinery industry relies today in a vast industrial experience on the variables affecting sulfidic corrosion
but very little is known on the basic mechanism of attack.
There is ample room for research and the basic understanding
of this phenomenon.
Keywords: corrosion; refinery; sulfidic; temperature.

1. Introduction
Iron (Fe) or steel reacts readily with hydrogen sulfide (H2S)
to form iron sulfide (FeS or FexSy). Since H2S is a ubiquitous compound in the oil and gas industry, sulfidation of steel
occurs under upstream (exploration and production) and
downstream (refinery) conditions. Reaction of Fe with sulfur (S) also occurs in power generation during the burning of
fossil fuels. Table 1 shows three areas of sulfidation corrosion
of steels.
In the temperature range from ambient to approximately
200C, the sulfidation of carbon steel is a common occurrence
under upstream aqueous conditions (Kane, 2006). Several
types of FeS were identified; including amorphous, pyrrhotite, mackinawite, troilite, cubic, greigite, marcasite and
pyrite (Vedage, Ramanarayanan, Mumford, & Smith, 1993;
Harmandas & Koutsoukos, 1996; Sun & Nesic, 2007; Smith,
Brown, & Sun, 2011). The establishment of the various stoichiometric and non-stoichiometric forms of FexSy may be
influenced by factors such as the partial pressure of H2S, the
pH, and the temperature (Smith et al., 2011). For example, at
ambient temperature and at pH 4 mostly mackinawite forms
at all concentrations of H2S; however when the pH is 7, the
formation of pyrrhotite is favored for the higher concentrations of H2S (Smith et al., 2011).
At temperatures higher than 538C the sulfidation mode of
attack of steel by sulfur compounds may change from mostly
an external uniform corrosion reaction of a component to an

internal localized attack, when sulfur diffuses inside the bulk


metal (Table 1). The higher temperature sulfidation attack
generally occurs in gases, for example, in the gasification
of coal, in the refining industry and in gases that result from
burning fossil fuels (gas, liquid and coal) (Lai, 2007). In other
cases steel could be attacked by the combined presence of S
and O2, such as in the phenomenon known as hot corrosion
(Type I and Type II) (Rapp, 2002). The presence of molten
alkali metal sulfate, sulfur trioxide as well as vanadium pentoxide may destroy a protective oxide film on the metal surface accelerating corrosion locally. The hot corrosion process
may also happen in refineries in the fire side of heater tubes
(Wen & Mucek, 2011).
The higher (above 500C gaseous) and lower (upstream
wet) temperature sulfidation issues are not part of this review.
Only the sulfidation corrosion in refineries, often called sulfidic corrosion, is reviewed here. Sulfidic corrosion is the
corrosion process of engineering alloys (mainly steels) in
presence of hydrocarbons containing sulfur species, mainly
H2S. The temperature range in which the phenomenon of
sulfidic corrosion occurs is 232427C (Table 1). Several
subgroups of sulfidic corrosion in refineries may be listed,
including presence of H2S, presence of H2S plus S and/
or S-containing compounds (mercaptans), presence of H2S
plus H2, and presence of H2S plus naphthenic acids, and the
combination of all of above.
Typical sulfidic corrosion in refineries can occur in a large
number of components such as fittings and pressure vessels but it seems more prominent in piping. Sulfidic corrosion manifests itself as more or less uniform thinning of the
wall of the component, but in horizontal pipes the 12 oclock
position may be preferentially attacked (Niccolls, Gallon, &
Yamamoto, 2008). The mechanism of sulfidic attack is by
direct reaction of the sulfur with the metal.
Several review documents on sulfidic corrosion have
been published (NACE, 2004; API, 2008) to capture the
current understanding and consensus on the sulfidic phenomenon in refineries. However, very little is found in the
literature regarding fundamental research of sulfidic corrosion including kinetics and thermodynamics (Farrell &
Roberts, 2010).
The main objective of the current review is to analyze the
most recent (newer than 2006) published data on sulfidic
corrosion and to highlight incomplete and contradictory data.
1.1. Crude oil in a refinery

A crude oil refinery contains more than a thousand components housing nearly a hundred internal environments
(Jenkins, 1998). In such a refinery crude oil or petroleum is
converted into end products such as gasoline, kerosene, diesel oil, etc. Crude oil is a complex mixture of several 100
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

124 R.B. Rebak: Sulfidic corrosion in refineries a review

Table 1

Sulfidation corrosion of steels.

Application

Temperature range

Corrosion characteristics

Upstream oil and gas exploration


and production, geothermal wells

Ambient to 230C

Uniform type of corrosion in presence of water. Different type of iron


sulfide scales may form depending on the environmental characteristics
including oxidizing vs. reducing environments, pH, bacterial activity, etc.

Downstream oil refineries

232 427C

Sulfidic corrosion could be in the presence of liquid or vapor phase oil


containing sulfur species. The attack is generally uniform.

Fossil fuel power production

538 1100C

Sulfidation attack could be localized. It is generally a gaseous phase


corrosion. There are several types of corrosion environments, including
gasifiers (syngas), oxidizing conditions (SO2), reducing conditions (H2S),
hot corrosion, coal ash corrosion, etc. (Lai, 2007)

hydrocarbons containing approximately 84% carbon, 14%


H and 13% S and <1% each of metals, salts and oxygen
(Ruschau & Al-Anezi, 2001; Guedes Soares, Garbatov,
Zayed, & Wang, 2008). There are more than 150 varieties of
crudes produced worldwide (Bacon & Tordo, 2005). Crude
oils may be classified in different ways, for example: (1) as
paraffinic, naphthenic or aromatic based on the predominant
type of hydrocarbon molecule (Ruschau & Al-Anezi, 2001;
Guedes Soares et al., 2008); (2) as sour or sweet depending on
the amount and type of reactive sulfur species, such as H2S.
Sour oils contain more H2S; (3) as light or heavy depending
on the API gravity number (oils with a high API number and
high proportion of H are considered light) and (4) as acidic or
not acidic depending of the total acid number (TAN). A TAN
value higher than 1 can be considered acidic (Bacon & Tordo,
2005). The acidity in the oil is given by the amount and type
of carboxylic organic acids. A crude with a TAN number
lower than 0.5 will be less corrosive to a refinery plant (Kane
& Cayard, 2002).
Refineries may use blends of different varieties of crude
and at the same time there is variation in the oil that is being
processed (distilled) in one refinery from point to point of
the flow. Similarly, the contaminants that the oil carries are
different from different blends that come into the refinery
and for a typical blend, the amount and type of species also
change from sector to sector in the refinery. During the refining process, other substances may be added to the oil, such
as hot water (hydro treating) or emulsifiers to remove salts
and metals (desalting). In other treatments, the oils may be
sweetened by neutralizing H2S using amines or other caustics
(Jayaraman & Saxena, 1995).

2. The sulfidation corrosion process


The actual sulfidic corrosion mechanism of steel is not known
(Niccolls, 2005). It is generally accepted that the corrosion
progresses in the steels via a film on the surface. This film is
formed by metal sulfides and it is pseudo-passive, i.e., it
is semi-protective and not tenacious (Jayaraman & Saxena,
1995).
The following steps could lead to the formation of an FeS
film
FeFe2++2e-

(1)

H2SHS-+H+

(2)

as iron oxidizes, the hydrogen cation gets reduced and then it


may get dissolved in the metal or it may evolve as molecular
hydrogen gas in the stream
H++e-Ho

(3)

o
Ho Hdiss

(4)

Ho+HoH2

(5)

If Equation 4 is dominant over Equation 5, the dissolved


hydrogen in the steel may cause hydrogen embrittlement,
especially in welds. FeS forms on the surface by the reactions
in Equations 6 and 7
HS-+Fe2+FeS+H+

(6)

or it may form non-stoichiometric sulfide products such as


yHS-+xFe2+FexSy+yH+

(7)

The stability of the film will depend among other variables on the presence of flow or turbulence. As turbulence
increases, the pseudo-passive film becomes minimal. The
transport of sulfur through the pseudo-passive metal sulfide
layer is confirmed by a parabolic decrease in the film thickness as a function of time suggesting a diffusion controlled
mechanism (Qu, Zheng, Jing, Yao, & Ke, 2006).
If the temperature is high enough (500C and higher) it
is possible that besides forming a metal sulfide scale on the
surface, sulfur may diffuse inside the metal (mainly though
grain boundaries) and react with the metal forming internal
sulfides. The presence of sulfur inside the metal may interfere
with the dissolved and carbide precipitated carbon. The metal
carbides disappear where the sulfides form and the liberated carbon diffuses deeper into the metal (Hucinska, 2006).
This process of carbide decomposition is accelerated if H2 is
present in the environment (Hucinska, 2006).

3. Previous reviews on sulfidic corrosion


Sulfidic corrosion has occurred since modern refineries were
in operation and it is still pervasive today. However, refineries are familiar with sulfidation and know how to control
it by alloy selection and corrosion and materials monitoring
and management. Niccolls et al. (2008) state that refining is a
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

R.B. Rebak: Sulfidic corrosion in refineries a review

mature industry and therefore almost all of the types of degradation modes are well known to seasoned materials engineers
and inspectors. Shut downs or failures due to sulfidic corrosion are currently rare. The previous knowledge on sulfidic
corrosion has been summarized in a series of comprehensive
reports that were prepared by the experts in the field. These
reports include the NACE 34103 Overview and the API
939-C Guidelines.
3.1. NACE 34103 overview of sulfidic corrosion
in petroleum refining

The NACE International Specific Technology Group (STG)


34 on Petroleum Refining and Gas Processing formed the task
group TG176 called Prediction Tools for Sulfidic Corrosion.
This task group issued a consensus document in February
2004. This document states that little fundamental research
has been published studying the mechanism, thermodynamics
and kinetics of sulfidic corrosion in the temperature range of
interest for the refining industry. Most of the current understanding on the corrosion behavior of materials is based on
plant experience.
Some basic facts from the NACE 34103 document:
1. Carbon steels form an iron deficient sulfide scale (Fe1-xS)
on the surface that, as it grows, slows down the corrosion
process as a function of time.
2. Steels that contain chromium form a two-layer scale the
inner layer is a sulfo-spinel (FeCr2S4) and the outer layer
is the Fe1-xS. It is generally accepted that the sulfo-spinel
layer is more protective than the Fe1-xS.
3. Several steps may be involved in the corrosion process in
presence of sulfur compounds chromium (Cr) may poison a critical decomposition step of the sulfur compounds
previous to the incorporation of sulfur to the scale and
H2 may promote the decomposition of these sulfur compounds (therefore H2 may counteract the beneficial effect
of Cr).
4. Carbon steels can be used until 260C. The corrosion
resistance of carbon steel is marginal at 316C.
5. Lower than 5% Cr steels are not currently used in refineries,
5Cr steels are used up to 343C and 9Cr steels are used up
to 400C. Austenitic type 18/8 stainless steels have excellent resistance to sulfidic corrosion even under long-term
use at high temperature like in furnace tubes and furnace
transfer lines. Sensitization of the stainless steel does not
decrease its resistance to sulfidic general corrosion. If H2
is present in the system the austenitic 18/8 stainless steels
may need to be used above 260C.
6. The mechanism of increased aggressiveness due to the
presence of H2 has not been established yet. Some postulated that H2 does not allow for beneficial formation of
coke, others argue that H2 promotes the decomposition of
other sulfur bearing compounds into H2S, thus increasing
the aggressiveness of the environment.
7. The FeS scale that forms on the surface generally contains
cracks, fissures and spalls that may provide avenues for
sulfur ingress and promote more corrosion. Some argue

125

that the presence of coke may seal these paths for sulfur
ingress. That is, when coking starts occurring, the corrosion attack by sulfidation generally decreases.
3.2. API RP 939-C guidelines for avoiding sulfidic
corrosion

In January 2008 the API Subcommittee on Corrosion and


Materials issued the Version 5.0 of the RP 939-C Guideline
for Avoiding Sulfidation Corrosion (API, 2008). RP stands
for recommended practice to provide a practical guidance to
corrosion engineers and other personnel, such as inspectors,
project and maintenance engineers on how to address sulfidic
corrosion. The Subcommittee on Corrosion and Materials is
part of the API Committee on Refinery Equipment (CRE).
The mission of the CRE is to promulgate safe and proven
engineering practices for the design, fabrication, installation,
inspection, and use of materials and equipment in refineries and
related processing facilities. The Subcommittee on Corrosion
& Materials deals with issues, such as: (1) Fabrication requirements; (2) Corrosion mechanisms; (3) Equipment reliability;
(4) Refractory systems; and (5) Reducing capital and maintenance costs. The RP 939-C document is applicable to hydrocarbon process streams containing sulfur compounds, with
and without the presence of H2, which operate at temperatures
above approximately 230C up to about 540C. A threshold
limit for sulfur content is not provided because within the past
decade significant corrosion has occurred in the reboiler/fractionator sections of some hydroprocessing units at sulfur or
H2S levels as low as 1 ppm.
The API RP 939-C document captures the state of the art
consensus in the industry of the state of knowledge dealing
with sulfidation corrosion. That is, any significant contributions on the understanding of variables that affect sulfidation
corrosion are captured in this document. The information in
RP 939-C has been mined from published technical papers,
information exchanges at the API and NACE levels and also
from refinery owners and operators. This does not mean that
some findings previous to 2008 that may also be relevant may
not be captured in this document since they may be proprietary or because consensus has not been reached in the matter.
The following basic findings can be summarized from the
API RP 939-C document:
1. The sulfidic corrosion rate increases with the temperature
from 230C until 425C and then it decreases as the temperature increases until 540C. The rate of increase in the
corrosion rate between 230C and 425C is faster as the
temperature is increased.
2. Chromium is a beneficial alloying element protecting
against sulfidation corrosion, i.e., the corrosion rate of steel
with 5% Cr is lower than the corrosion rate of carbon steel.
3. The corrosion rate increases with the amount of sulfur in
the stream. The influence of sulfur on the corrosion rate is
less important than the effect of the temperature.
4. The presence of H2 tends to accelerate the corrosion rate.
The beneficial effect of Cr in the steel may be less effective when H2 is present in the oil mixture.
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

126 R.B. Rebak: Sulfidic corrosion in refineries a review

5. Carbon steels containing <0.1% silicon (Si) have a faster


rate of sulfidic corrosion than carbon steels with higher
contents of Si.
6. Other variables that may affect the corrosion rate include
flow velocity, coking, and presence of steam (stripping).
3.3. McConomy and Couper-Gorman curves

McConomy and Couper-Gorman curves are prediction


tools based on survey data from refinery experience that
are used to estimate the corrosion rate of different steels
in crude oil as a function of the temperature (API, 2008).
The McConomy curves developed by API in 1963 are
used in the absence of dissolved hydrogen and the CouperGorman curves developed in 1971 are used when H2
and H2S are present (Qu et al., 2006; API, 2008; Farrell
& Roberts, 2010). Neither the McConomy nor the CouperGorman curves take into consideration the effect of flow
velocity. The McConomy curves were later modified in
1986 since the original ones were too conservative, i.e.,
the original curves predicted higher sulfidic corrosion rates
than the actually observed in refineries (API, 2008; Farrell
& Roberts, 2010). The modified McConomy curves were
developed from empirical data at a total sulfur concentration of 0.6 wt% in the crude oil. A correction factor may
be applied for higher sulfur levels (API, 2008; Farrell &
Roberts, 2010). The McConomy curves (Figure 1) show that
the corrosion rate of the different steels increases monotonically as the temperature increases. The highest corrosion
rates are for carbon steel and the lowest are for type 18/8
stainless steels. The curves do not cross each other in the
entire range of temperature reported from 260C to 400C.
The predictions from both types of curves (McConomy and
Couper-Gorman) are still conservative compared to industry experience (Riley, 2005). Materials behave differently
in H2S environment and in H2S+H2 environments, since the
beneficial effect of chromium for the H2S environment is
less effective in the H2S+H2 environment (Setterlund, 1991;
Niccolls, 2005).

4. Variables affecting sulfidic corrosion


As stated in the NACE 34103 document Overview of
Sulfidic Corrosion in Petroleum Refining very little basic
research has been dedicated to the understanding of the thermodynamics and kinetics of sulfidic corrosion. The little
interest in research may have resulted from the 40 years plant
experience that led to the preparation of guidelines to deal
with materials selection and plant performance in the field
(Chambers & Kane, 2008). Materials engineers may have felt
comfortable predicting the lifetime of several types of steels
according to the temperature and sulfur content of the process
and replacing the part at scheduled shut downs rather than
trying to fully understand the sulfidic corrosion mechanisms.
Table 2 shows a few of the internal and external variables that
influence sulfidic corrosion in refinery applications. Internal
variables pertain to the metallic component (pipe), such as
composition and phase distribution, heat treatment, etc.
External variables include temperature, type of crude, sulfur
content in the crude, flow velocity, etc.
4.1. Effect of the temperature

It is accepted that the most important variable affecting corrosion of a component in a refinery is the temperature. The
temperatures of interest for sulfidic corrosion are in the range
230425C. The temperature generally defines what material
is used in that application. For example, depending on the
particular application in the refinery and crude blend, carbon
steel may be used up to 260288C. Similarly, the upper limit
for the 5Cr steel could be 329343C and for temperatures
up to 400C 9Cr steel may be used. In general, for all materials, the corrosion rate increases as the temperature increases;
however, the relationship between corrosion rate and temperature may not be the same in the entire range of temperature.
As the temperature increases the activation energy value
seems to increase (API, 2008). For temperatures higher than
427C, sulfidic corrosion may actually decrease due to additional protection from coking (Gutzeit, 1986).

1.000

Carbon steel

0.100
1-3Cr
4-6Cr
7Cr
9Cr

0.010

Sultur content WT%

Corrosion rate (mm/year)

10.000

12Cr

18Cr/8

0.001
250

270

290

310

330

350

370

10
1
0.1
0.01
0.4

0.8

1.2

1.6

Corrosion rate multiplier

390

410

Temperature (C)

Figure 1

Modified McConomy curves to predict sulfidic corrosion.


- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

R.B. Rebak: Sulfidic corrosion in refineries a review

Table 2 Variables influencing sulfidic corrosion.


Internal variables

External variables

Type of steel
Heat treatment amount and
distribution of pearlite colonies
Cr content
Si content
Mo content

Temperature
Type and concentration of
sulfur species in the oil stream
Naphthenic (carboxylic) acids
Fluid velocity
Hydrogen in oil stream
Steam
Amines and other chemicals

Activation energy values for sulfidic corrosion are


scarce. No data were found for pure sulfidic corrosion in
contact with crude oil in the temperature range 230425C.
The activation energy was measured for the sulfidation
of Fe exposed to 10 mm Hg sulfur vapor (gas phase) for
up to 10 h in the temperature range from 250C to 500C
(Foroulis, 1978). It was reported that the formation of the
sulfide scale on the Fe surface followed an Arrhenius relationship between the parabolic rate constant and the inverse
of the absolute temperature. However, two ranges were
found; the activation energy was 59 kJ/mol in the higher
temperature range from 370C to 500C and 113 kJ/mol in
the lower temperature range (250370C) (Foroulis, 1978).
For the lower temperature range the scale consisted in an
inner layer of pyrrhotite and a thin external layer of pyrite;
however, for the higher temperature range the scale was
only pyrrhotite (Foroulis, 1978). The activation energy values reported by Foroulis (1978) may not be fully relevant
to oil refineries since it was in presence of sulfur vapor and
for pure Fe instead of steel.
For naphthenic acid corrosion Gutzeit reported that for
both carbon steel and 410SS the Arrhenius activation energy
was approximately 69 kJ/mol at temperatures higher than
288C (Gutzeit, 1977). More recently, an Arrhenius relationship was also reported for the sulfidic corrosion rate of carbon
steel between 210C and 300C in oil containing two types
of naphthenic acid. Values of activation energy from 23.8 to
31.8 kJ/mol were reported (Slavcheva, Shone, & Turnbull,
1998).
Two steels (carbon steel and 5Cr0.5Mo) were tested for
their response to sulfidic corrosion in the temperature range
between 230C and 270C (Qu et al., 2006). The carbon steel
was mostly a ferrite phase and the 5Cr0.5Mo steel was mostly
a pearlite phase. The tests were performed in a static autoclave for up to 65 h of testing time. As expected, for both
steels, the corrosion rate increased linearly with the temperature; however, the corrosion rate was higher for the carbon
steel than for the 5Cr0.5Mo steel. For the carbon steel the
corrosion rate increased from approximately 0.4 mm/year
at 230C to approximately 2 mm/year at 270C (Qu et al.,
2006). However Qu et al. did not report the value of activation energy for these experiments (Qu et al., 2006).
It seems a little surprising that activation energy for the
pure or single sulfidic corrosion mechanism were not found
in the literature for carbon or alloy steels. A few activation
energy values are published for the combined mechanism of

127

sulfidic corrosion and naphthenic acid corrosion. In general


the activation energy values available for steels in presence
of sulfur species and naphthenic acid seem to suggest that the
activation energy above 300C could be approximately twofold higher than the activation energy below 300C.
4.2. Sulfur species

The second materials selection criteria are the presence in the


stream of H2S and other sulfur containing species. As the concentration of sulfur species increase the corrosion rate of refinery plants components increase. For example, it was reported
that the corrosion rate of carbon steel exposed to a 329C oil
stream in a hydro-processing distillation column increased
practically 10-fold from 5 mpy to 45 mpy when the H2S concentration increased approximately three-fold from 800 ppm
to 2500 ppm (Niccolls, 2005). Sulfur is present in crude oil as
H2S, as thiols, mercaptans, sulfides, benzothiophenes, polysulfides, or as elemental sulfur (de Jong et al., 2007; API, 2008;
Guedes Soares et al., 2008). Sulfur becomes aggressive to steel
if its proportion in the crude oil is 0.2% or higher (Ruschau &
Al-Anezi, 2001). At temperatures higher than 230C, Fe reacts
with S to form FeS. In general, in the refineries crudes are classified as sweet (<1% S) and as sour (more than 0.5% S) (Bota,
Qu, Nesic, & Wolf, 2010). Sour crudes are blamed for sulfidic
corrosion in refineries. During refining some crude oils may
be treated with a caustic wash to remove sulfur in a so-called
sweetening process. Since the reactivity or corrosivity of these
different forms of sulfur varies from molecule to molecule, the
corrosion rate of steel generally is not proportional to the total
sulfur content in the oil stream (Bota et al., 2010). Smaller
molecules of sulfur compounds tend to be more corrosive than
the large ones (Setterlund, 1991). It was also reported that
sulfur species that have the ability to decompose into H2S at
the exposure temperature would have an impact on the corrosion rate of the steel (Kane & Cayard, 2002). Sometimes it is
reported that mercaptans are even more aggressive than H2S in
affecting sulfidic corrosion (Niccolls, 2005).
A study was carried out to determine the corrosion behavior of 1018 carbon steel in presence of mercaptans in crude
oil in the liquid phase in the temperature range 200300C
(de Jong et al., 2007). Coupons of carbon steel were exposed
to crude containing four different mercaptans at different concentrations ranging from 100 ppm to 3000 ppm (de Jong et
al., 2007). It was reported that at 275C the corrosion rate of
steel increased significantly with the concentration of mercaptans up to 1000 ppm and more slowly for the higher mercaptan concentrations. The highest corrosion rate at 275C
corresponded to the higher molecular weight mercaptan but
an explanation was not given for this behavior (de Jong et
al., 2007). It was also reported that the corrosion rate did
not increase monotonically as the temperature increased but
peaked at approximately 280290C and then decreased for
the higher temperatures (de Jong et al., 2007).
In general the information in the published literature regarding the effect of sulfur species on sulfidic corrosion is highly
contradictory, some argue that thiols and mercaptans are more
corrosive than H2S and some argue just the opposite.
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

128 R.B. Rebak: Sulfidic corrosion in refineries a review

4.3. Effect of naphthenic acid

Fe+2RCOOHFe(RCOO)2+H2

(8)

Fe(RCOO)2+H2SFeS+2RCOOH

(9)

The corrosion of iron by naphthenic acid is given by Equation


8. Iron naphthenate is soluble in oil, but if enough sulfur is
present in the oil, it may react with iron naphthenate to reform
the FeS layer on the surface of the component (Equation 9).
However, this reaction regenerates naphthenic acid in the system. Therefore, it is claimed that crude with high naphthenic
acid and low sulfur may be more corrosive than crude with
a similar naphthenic acid content with a higher sulfur level
(Turnbull et al., 1998; Laredo et al., 2004).
Two type of steels (carbon steel and 5Cr0.5Mo) were
tested for 24 h in presence of two naphthenic acids at a concentration of 0.25 mol/L, and in presence of the same naphthenic acid plus 0.1% H2S (in argon) at 275C in two types
of oil (heavy vacuum gas oil and mineral oil) (Slavcheva et
al., 1998). Figure 2 shows that the presence of H2S reduced
the corrosion rate of the steels promoted by naphthenic acids.
The inhibitive effect also depended on the type of naphthenic
acid and the type of steel. Figure 2 also shows that in the
presence of the naphthenic acid mixture the corrosion rate of

10

8
Corrosion rate (mm/year)

The third important criteria for materials selection in refineries is the total acid number (TAN), which is an indication of
the relative amount of carboxylic acids or naphthenic acid in
the oil. Naphthenic acid corrosion (NAC) is most prevalent
in the temperature range 200400C and it seems to peak at
approximately 370C (Jayaraman & Saxena, 1995). The corrosion caused by naphthenic acid is more important for TAN
higher than 0.5 (Laredo, Lopez, Alvarez, & Cano, 2004).
There is also a complex relationship between sulfidic corrosion and naphthenic acid corrosion. For example, it has been
claimed that the naphthenic acid and sulfidic mechanisms
act synergistically, i.e., the presence of sulfur may accelerate naphthenic acid corrosion, but the opposite has also been
defended, i.e., that the presence of sulfur may inhibit naphthenic corrosion (Slavcheva et al., 1998; Tebbal, 1999; Kane
& Cayard, 2002; Kanukuntla, 2008; OKane, Rudd, Cooke,
Dean, & Powell, 2010). It is claimed sometimes that there is
a continuum between naphthenic acid corrosion and sulfidic
corrosion (Kane & Cayard, 2002). Other researchers state
that the relationship between the two modes of corrosion may
change depending on the levels of sulfur and naphthenic acid
in the system (Messer, Tarleton, Beaton, & Phillips, 2004;
Chambers & Kane, 2008). Since FeS is insoluble in oil, this
may protect the steel from attack by naphthenic acid (Piehl,
1988; Turnbull, Slavcheva, & Shone, 1998; Kanukuntla,
2008; Bota et al., 2010). However, since iron naphthenate is
soluble in oil, the presence of naphthenic acid may weaken
the protectiveness of FeS on the surface promoting scale debonding and favoring more corrosion. In a corroded refinery
component, the evidence of naphthenic acid corrosion may be
supported by the absence of a surface scale on the corroded
component (Gutzeit, 1977).
The following reactions have been proposed to explain the
interaction between naphthenic acid (RCOOH) and FexSy

Inhibitive effect of H2S on


naphthenic acid corrosion
mineral oil, 24 h, 275C
Cyclohexane carboxilic acid
Cyclohexane carboxilic acid+H2S
Naphthenic acid mixture
Naphthenic acid mixture+H2S

0
Carbon steel 1018

5Cr 0.5Mo steel

Figure 2 Inhibiting effect of hydrogen sulfide on the corrosion caused by naphthenic acid. Plot prepared from table 9 data in
Slavcheva et al. (1998).

the 5Cr0.5Mo steel was higher than the corrosion rate of the
carbon steel. The latter result is one of the surprising findings
in which a 5Cr steel is found more prone to corrosion than
plain carbon steel. The authors claim that the inhibitive effect
of sulfur also depended of the type of carrier oil (not shown in
Figure 2) (Slavcheva et al., 1998).
Two steels (5Cr and 9Cr) were tested for resistance to
impingement corrosion in Tufflo 1200 oil at 343C using
flow velocities of 1697 m/s. The tests were performed in
oil containing only naphthenic acid (up to TAN 3.5) and in
oil containing naphthenic acid plus two different levels of
H2S (0.2 and 0.45 psia) (Kane & Cayard, 2002). For the pure
naphthenic acid environment, the impingement corrosion
was evident for TAN values higher than 1.5. Both the 5Cr
and the 9Cr steel had similar degradation rate. As the TAN
value increased, the velocity to onset impingement corrosion
decreased. When the 0.2 psia level of H2S was added it caused
inhibition of impingement corrosion in both steels. When the
H2S was increased to the higher level (0.45 psia) the 5Cr steel
started to corrode but the 9Cr still maintained the inhibition
(Kane & Cayard, 2002).
The interaction between naphthenic acid and sulfidic corrosion was also investigated at 270C using 5Cr0.5Mo steel
(Qu et al., 2006). For a solution containing a TAN between 6
and 16, the corrosion rate was inhibited by adding 1% S (Qu
et al., 2006). However, then the TAN was 32, the addition of
1% S was detrimental (increased the corrosion rate promoted
by naphthenic acid). For the carbon steel, the addition of 1% S
always increased the corrosion rate by naphthenic acid. Sulfur
inhibited corrosion caused by naphthenic acid only in the
5Cr0.5Mo steel (Qu et al., 2006). It was more recently claimed
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

R.B. Rebak: Sulfidic corrosion in refineries a review

that no fundamental studies were carried out to investigate the


mechanism of interaction between the sulfide layer and the
naphthenic acid in solution (El Kamel et al., 2010b). Coupons
of four different alloys (carbon steel, 2.25Cr1Mo, 5Cr0.5Mo
and 304L) were pre-sulfidized in 150 mbar of pure H2S at
300C. The FeS that formed on the surface was pyrrhotite.
Then the coupons were exposed to white oil containing naphthenic acid to a total TAN=4 at 260C for different amount of
times and the effect of the naphthenic acid on the pre-existing
H2S scale was monitored (El Kamel et al., 2010b). It was
reported that the naphthenic acid attacked locally the sulfide
scale on the coupons causing eventual detachment of the scale
from the surface of the coupons. Figure 3 shows a representation of the required testing time to produce detachment of the
FeS scale from the coupons (El Kamel et al., 2010b). Scale
detachment from the 304L steel (18Cr) did not occur even
after 9 h of exposure to the oil. Figure 3 also shows an exponential fit of the data for the first three points (carbon steel,
2.25Cr1Mo and 5Cr0.5Mo). An extrapolation of these results
show that detachment may have occurred for the 304L steel
only after more than 100 h exposure (Figure 3).
Recent laboratory results show the intricacy of the relationships between steel compositions, temperature, sulfur species,
naphthenic acid, etc. Current laboratory data do not contradict
findings from the plants but at the same time do not provide
too much insight on the corrosion mechanism or mechanisms.
Interestingly, the interaction between sulfidic corrosion and
naphthenic acid seems to be one of the most widely tested
phenomena in the laboratory.
4.4. Effect of composition of the steel

The selection of materials in a refinery is not only based on


the ability to resist corrosion but also on price, availability and
ability to weld. Table 3 shows some regular materials used in
1000

Sulfide scale detachment time (min)

Exponential fit, R2=0.97

304L

100
5Cr

2.25Cr

Carbon steel

Sulfided coupons
white oil, TAN=4, 260C

10
0

12

16

20

Weight % Cr

Figure 3 Detachment time for a sulfide scale when exposed to


white oil containing naphthenic acid TAN=4 at 260C (plotted from
data by El Kamel et al., 2010b).

129

the petroleum refining industry. For carbon steels it is known


that steels containing <0.1% Si corrode faster than steels with
higher Si content (API, 2008). The presence of Si in the steel
may help to form a more adherent and stable sulfide scale
on the surface (API, 2008). The modified McConomy curves
(Figure 1) do not differentiate between low and high Si steel.
These curves show that at each temperature the corrosion rate
is practically reduced about half its value for series carbon
steel >2.25Cr steel >5Cr steel >9Cr steel >12Cr >18/8 steel. It
has also been reported that the corrosion rate can be reduced
10-fold when ferritic 9Cr steel is used instead of carbon steel
(Hucinska, 2006). In general the resistance of the steels to
sulfidic corrosion increases according to the following order
(Farraro & Stellina, 1996; Qu et al., 2006): Carbon steel,
Carbon steel +0.5Mo, 5Cr+0.5Mo, 9Cr+1Mo, 12Cr (410),
17Cr (430), 304SS, 316SS, and 317SS.
Stainless steels are used to resist high temperature sulfidic
corrosion (Farraro & Stellina, 1996). Stainless steels containing molybdenum are used to combat corrosion mainly by
naphthenic acid (Farraro & Stellina, 1996). The effect of Cr
to protect against sulfidic corrosion may be more important
under flow conditions (Qu et al., 2006). The beneficial effect
of Cr may originate of its ability to poison the decomposition
of sulfur compounds (Farrell & Roberts, 2010).
Two steels (carbon steel and 5Cr0.5Mo) were tested for
their response to naphthenic acid corrosion and sulfidic
corrosion (Qu et al., 2006). The carbon steel was mostly a
ferrite phase and the 5Cr0.5Mo steel was mostly a pearlite
phase. The tests were performed in a static autoclave for up
to 65 h of testing time. Coupons were exposed to the liquid
and vapor phase inside the vessel. Testing temperature was
from 220C to 320C (at 20C intervals). The base fluid (carrier) was transformer oil to which (1) naphthenic acid to a
TAN=2 to 14.51 and (2) dimethyl disulfide with [S]=1% were
added separately (Qu et al., 2006). It was reported that in
the naphthenic acid environment with TAN=2 at 270C, the
carbon steel was found more resistant to corrosion than the
5Cr0.5Mo steel. However, in the dimethyl disulfide solution
([S]=1%) the 5Cr0.5Mo steel was found more resistant to
corrosion. Cross sections of the scale formed on the testing
coupons showed that on the 5Cr0.5Mo steel the scale consisted of two layers while in the carbon steel the scale consisted of only one layer (Qu et al., 2006). Laboratory tests and
field results may suggest that the presence of Cr in the steel
may not be beneficial to protect against naphthenic acid corrosion (Qu et al., 2006).
In the case when hydrogen is present, the Couper-Gorman
curves do not predict a large decline of the corrosion rates
between carbon steel and 5Cr0.5Mo steel. Only when the 9Cr
steel was used there is modest decrease in the corrosion rate,
and a distinctive improvement was noticed when the 18%Cr
austenitic steel was used (Hucinska, 2006).
Coupons of three steels (carbon steel, P5 and 304L) were
exposed to flowing high sulfur crude oil at 300C and 35 bar
pressure for times as long as 98 h (El Kamel et al., 2010a).
Mass losses after cleaning the sulfide layer were transformed
to pyrrhotite layer thickness. For carbon steel and P5 steel
(5Cr0.5Mo) the same thicknesses were found, indicating
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

130 R.B. Rebak: Sulfidic corrosion in refineries a review

Table 3 Typical materials in sulfidic corrosion applications.


Material

Designation

Typical composition, weight %


Single figures are maximum

A53

Grade B ASTM A53

Fe, 0.3C, 0.4Cr, 1.2Mn, 0.045S, 0.05P, 0.4Cu, 0.4Ni, 0.15Mo,


0.08V, (Cu+Ni+ Cr+Mo +V= max 1.0 %)

A106

Grade B ASTM A106


(K03006)

Fe, 0.3C, 0.4Cr, 0.29 1.06Mn, 0.1Si, 0.035S, 0.035P, 0.4Ni,


0.4Cu, 0.08V

5Cr 0.5Mo (pipe)

ASTM A335 P5

Fe, 0.15C, 4 6Cr, 0.45 0.65Mo, 0.3 0.6Mn, 0.5Si,


0.025S+ 0.025P

9Cr 1Mo (pipe)

ASTM A335 P91

Fe, 0.08 0.12C, 8 9.5Cr, 0.85 1.05Mo, 0.3 0.6Mn, 0.2 0.5Si,
0.18 0.25V, 0.01S, 0.02P

12Cr 410SS

UNS S41000

Fe, 0.15C, 11.5 13Cr, 1.00Mn, 0.040P, 0.030S, 1.00Si

316LSS (pipe)

ASTM A312 (S31603)

Fe, 0.03C, 16 18Cr, 10 14Ni, 2 3Mo, 2Mn, 1Si, 0.03S, 0.045P

317SS (pipe)

ASTM A312 (S31700)

Fe, 0.08C, 18 20Cr, 1115Ni, 3 4Mo, 2Mn, 1Si, 0.03S, 0.045P

little beneficial effect of the presence of 5% Cr in the steel.


However, there was a strong benefit with the 18% Cr since
they could not find weight change for the 304L coupons (El
Kamel et al., 2010a).
Current laboratory findings and some plant operation
experience may not support the use of 5Cr steel in refineries since in many applications its corrosion behavior cannot
be fully differentiated from the behavior of carbon steels.
Moreover, some current commercial 5Cr steel pipes now
contain <4.5% Cr. In plant applications when there are limitations to the performance of carbon steel, some plant engineers may find it appropriate to upgrade the pipe material
directly to 9Cr steel, skipping the recommendation of 5Cr
steel.
4.5. Effects of velocity, hydrogen and steam

Fluid velocity and turbulence are important factors affecting sulfidic attack (Kane & Cayard, 2002). It is accepted
that the lowest corrosion is found when the surface of the
metal is completely wetted with the hydrocarbon under low
flow (Gutzeit, 1986). Fluid velocity up to 60 m/s may hamper the ability of the steel to form a semi protective sulfide
film on the surface and therefore high velocity may accelerate corrosion (API, 2008). An adherent sulfide scale may
control further attack of the steel either by more sulfidation or by naphthenic acid attack. However, if the sulfide
scale is removed by the shear stress resulting from fluid
flow, the attack of the underlying steel may be accelerated
(Kane & Cayard, 2002; Qu, Liu, Jiang, Lan, & Shan, 2011).
This is especially true in presence of naphthenic acids (Bota
et al., 2010; Qu et al., 2011). It has also been argued that
too little flow may also be detrimental since more H2S may
be allowed to evolve and concentrate in certain pipe areas
(API, 2008).
Materials behave differently in H2S environment than in
H2S+H2 environments, since the beneficial effect of chromium for the H2S environment may seem less effective
in the H2S+H2 environments (Setterlund, 1991; Niccolls,

2005). It has been argued that the presence of H2 may be


detrimental for the corrosion resistance of the steels since
it inhibits the formation of semi-protective coke on the surface (NACE, 2004). Another explanation for the hydrogen
effect is that it reacts with less corrosive sulfur containing
species and forming more H2S and therefore increasing the
corrosiveness of the system (Gutzeit, 1986; NACE, 2004).
In general, if H2 is present in the stream above 260C it is
recommended to use 18Cr steel (e.g., type 316SS) and avoid
carbon steel and lower Cr steels altogether (Gutzeit, 1986;
NACE, 2004).
The presence of vaporization and steam may yield higher
corrosion rates in steels. The most detrimental corrosion may
happen within a spray flow with vapor loads higher than
60% containing liquid droplets that may destroy the sulfide
scale via impingement (Gutzeit, 1986). It was reported that
in sulfidic environments the corrosion rate is approximately
six-fold higher when the metal surface is exposed to vapor
vs. liquid (McLaughlin, 2005). Carbon steel and 5Cr0.5Mo
steels were tested for resistance to sulfidic corrosion at
230C, 250C and 270C (Qu et al., 2006). In the sulfidic
solution ([S]=1%), the corrosion rate of both steels increased
monotonically as the temperature increased both for the
liquid and vapor phases. For both steels the corrosion rate
was the same in the liquid and vapor phases (Qu et al., 2006).
However, the corrosion rate was higher for the carbon steel
than for the 5Cr0.5Mo steel (Qu et al., 2006). The effect
of the sulfur concentration on the corrosion rate was also
studied at 270C. When the sulfur concentration increased
from 0.5 to 1.25, the corrosion rate increased up to [S]=1
and then decreased from 1 to 1.25. The corrosion rate for
both steels was higher in the vapor phase than in the liquid
phase and the corrosion rate in both phases of the carbon
steel was higher than the corrosion rate of the 5Cr0.5Mo.
When mixtures of naphthenic acid with sulfur compounds
were tested, a complex relationship of the variables involved
was reported (Qu et al., 2006). Nevertheless, the corrosion
rate of the 5Cr0.5Mo steel was lower than the corrosion rate
of the carbon steel (Qu et al., 2006).
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

R.B. Rebak: Sulfidic corrosion in refineries a review

4.6. Effect of other chemical species, such as amines


and ammonium disulfide

Crude oil may contain many other chemicals, such as amines


added to control wax deposition, enhance flow characteristics, aid in water separation, etc. (Kapusta, van den Berg,
Daane, & Place, 2003). For example, amines such as monoethanolamine, di-ethanolamine, aminoethoxyethanol, methyl
di-ethanolamine and di-isopropanolamine are added at different points in the refinery to remove H2S, mercaptans and
carbon dioxide from the process stream (Shahid & Faisal,
2009; Lagad, Cayard, & Srinivasan, 2010). The presence of
amines and other chemicals add another degree of complexity
to the sulfidic corrosion mechanism. Little or no information
is available in this area.
Some crude oils may also contain species such as cyanide ions (CN-), which may destroy the protective layers of
FexSy on the surface and promote hydrogen ingress into the
steel what eventually may cause hydrogen induced cracking
(Groysman, Feldman, Kaufman, & Balali, 2011).
4.7. Effect of testing time

The sulfide scale that forms on the test metallic coupons


could be semi-protective depending on the testing conditions. That is, as the time increases the corrosion rate may
decrease. When coupons of carbon steel and 5Cr0.5Mo steel
were tested in the liquid and vapor phase of transformer oil
containing 1% [S] as dimethyl disulfide at 270C, the corrosion rate initially increased (up to 8 h) and then decreased
as the time increased up to 65 h (Qu et al., 2006). These
results suggest that diffusion of sulfur through the sulfide
scale is the rate limiting step in the corrosion rate (Qu et al.,
2006). Since the parabolic law rate constant for the carbon
steel was higher than for the 5Cr0.5Mo steel, it was concluded that under the tested conditions the 5Cr0.5Mo steel
was more resistant to sulfidic corrosion than the carbon steel
(Qu et al., 2006).
4.8. Plant experience

Corrosion issues in crude refineries are generally solved


by alloy selection based on experience accumulated by the
industry in many decades. Trade documents cited before
(API RP 939-C, NACE 34103) provide a guide for alloy
selection under the different conditions found in the plants.
Some refineries may use additional measures, such as inhibition or neutralization (Tuttle, 2005). Continuous monitoring
is essential in many plants to measure the degradation rate
of some components and to determine if these components
need replacement in the next scheduled shut down. Most of
the recent failures due to sulfidic corrosion were traced to the
use of carbon steel pipes with insufficient amount of silicon
(API, 2008).
An unusual high temperature sulfidic corrosion has been
reported in the catalytic refining unit in a plant in Indiana
(Wilks, 2000). The failure occurred in a hot dip aluminized
steel pipe elbow that was exposed to a turbulent two-phase

131

flow. Failure occurred after the aluminized layer was corroded or eroded (Wilks, 2000). This pipe operated above
316C. Some of the corrosion attack progressed under the
aluminized layer lifting it away into the flowing stream. The
leaking failure was ductile overload due to the thinning of the
pipe wall. As a consequence of the reported failure, all piping in the area operating above 260C has been upgraded to
5Cr0.5Mo (Wilks, 2000).
A corrosion study was conducted for 20 days at a refinery
during a sour operation when a blend of high sulfur crudes
were processed (Farrell & Roberts, 2010). The total amount
of sulfur in the blend was 1.80.2 wt% and the TAN number was relatively low (0.24). Weight loss coupons of five
different steels were exposed to the oil stream in piping at
two points in the plant, (1) in the heavy atmospheric gas
oil (HAGO) with a flow of 1.5 m/s at 354C and (2) in the
light vacuum gas oil (LVGO) with a higher flow of 2.4 m/s
but at a lower temperature of 204C. The tested materials
were: (a) carbon steel, (b) 5Cr, (c) 9Cr, (d) 410SS 12Cr and
(e) cast CA6NM 12Cr4Ni0.5Mo (Farrell & Roberts, 2010).
For the coupons exposed to the LVGO (204C), the carbon
steel, the 5Cr and the 9Cr steel performed well with corrosion rates below the detection limit of 0.1 mpy (<0.0025
mm/year). This agrees well with the common knowledge
that sulfidic corrosion is becoming an important issue at
temperatures above 232C, which is above the temperature
in the LVGO. Surprisingly it was reported that in the LVGO
both the 410SS and the CA6NM coupons showed a measurable corrosion rate of 1.8 mpy (0.05 mm/year) for the 410SS
and 0.2 mpy (0.005 mm/year) for the CA6NM steel (Farrell
& Roberts, 2010). Figure 4 is a graphic representation of
results reported by Farrell and Roberts in their Table 2 for
the HAGO system at 354C. Figure 4 shows that the corrosion rates of carbon steel, 5Cr and 9Cr materials were higher
than 10 mpy (>0.254 mm/year). The highest corrosion rate
was for the 5Cr steel at 29.1 mpy (0.74 mm/year) (Farrell
& Roberts, 2010). It may seem unanticipated that under the
tested conditions the corrosion rate of the 5Cr steel was
higher than the corrosion rate of the carbon steel. Figure 4
also shows a sharp decline in the corrosion rates between
the 9Cr and the 12Cr steels, since for the latter materials
the corrosion rate was <1 mpy (<0.025 mm/year). Other
results from Farrell and Roberts show a significant effect
of the temperature on the sulfidic rate of the steels between
204C and 354C, since the corrosion rate increased more
than two orders of magnitudes between these temperatures.
Under the tested conditions, the effect of fluid flow was not
significant since for the higher velocity of 2.4 m/s at 204C
the corrosion rate was approximately two orders of magnitude lower than for the lower velocity of 1.5 m/s at 354C
(Farrell & Roberts, 2010). These results seem to indicate that
the temperature is a more important factor controlling
corrosion than the fluid velocity. It was also noted by Farrell
and Roberts that, after the in-situ plant tests, the carbon steel
coupons yielded corrosion rates lower than the values predicted from the McConomy curves, while the 5Cr and 9Cr
steel yielded higher corrosion rates than the McConomy
curves predicted values (Farrell & Roberts, 2010).
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

132 R.B. Rebak: Sulfidic corrosion in refineries a review

Coupons in plant piping


sour crude ~2% [S], 20 days, 354C

100

10
0.1

mm/year

Corrosion rate (mpy)

1
0.01

0.1

Carbon
steel

5Cr

9Cr

410SS CA6NM

Figure 4 Corrosion behavior of engineering alloys coupons


exposed to a plant stream containing sulfur (plotted from tabular data
by Farrell & Roberts, 2010).

5. Summary and conclusions


Sulfidic corrosion is a complex mechanism of steel degradation in crude refineries occurring between 232C and 427C.
Internal and external factors affect the sulfidic corrosion
degradation rate including alloy composition, temperature,
total sulfur content in the oil, and presence of naphthenic
acid. Two review consensus documents have been issued by
NACE International in 2004 and API in 2008 capturing the
state of knowledge in plant experience of how the environmental variables affect the sulfidic corrosion performance of
the engineering steels. Most of the process of material selection to replace degraded parts or for new refineries is based on
almost a century of data from plant experience.
The main aim of the current review was to bring together
the most recent research results from laboratory testing and
evaluate their findings in perspective of the plant experience.
Based on the reviewed literature it seems apparent that the
basic mechanism of sulfidic corrosion of carbon steel and
alloy steels is not fully understood. Little or no systematic
research has been carried out in laboratory or in plant to
measure, for example, kinetics of sulfide scale growth, the
mechanical properties and adherence of the scales, the activation energy for scale formation and its dependence of temperature ranges, alloy composition, sulfur species, etc. Little
or no information exists to determine why the presence of
Cr in the steel is beneficial for sulfidic corrosion resistance.
There is ample room for systematic laboratory testing on the
sulfidic corrosion of engineering alloys in simulated and plant
refinery environments, mainly on the effect of temperature,
sulfur species, alloy composition, and microstructure.

References
API. RP 939-C Guidelines for avoiding sulfidation (sulfidic) corrosion railures in oil refineries. API Subcommittee on Corrosion &

Materials, Version 5.0, February 2008. New Orleans, LA: API,


2008.
Bacon R, Tordo S. Crude oil price differentials and differences in
oil qualities: a statistical analysis. ESMAP Technical paper 081,
October 2005. Washington, DC: ESMAP, 2005.
Bota GM, Qu D, Nesic S, Wolf HA. Naphthenic acid corrosion of
mild steel in the presence of sulfide scales formed in crude oil
fractions at high temperature. Corrosion/2010, Paper 10353.
Houston, TX: NACE International, 2010.
Chambers BD, Kane RD. Refining high acid crudes. PTQ Q4 2008;
137149.
De Jong J-P, Dowling N, Sargent M, Etheridge A, Saunders-Tack A,
Fort W. Effect of mercaptans and other organic sulfur species on
high temperature corrosion in crude oil and condensate distillation units. Corrosion/2007, Paper 07565. Houston, TX: NACE
International, 2007.
El Kamel M, Galtayries A, Vermaut P, Albinet B, Foulonneau G,
Roumeau X, Roncin B, Marcus P. Sulfidation kinetics of industrial steels in a refinery crude oil at 300C: reactivity at the nanometer scale. Surface Interface Anal 2010a; 42: 605609.
El Kamel M, Galtayries A, Foulonneau G, Roumeau X, Morel G,
Marcus P. High temperature corrosion of sulfided steels in naphthenic acid environments. Eurocorr 1317 September 2010,
Moscow. Paper 9377. 2010b.
Farraro T, Stellina RM Jr. Materials of construction for refinery applications. Corrosion/1996, Paper 614, Conference and Exposition.
Houston, TX: NACE International, 1996.
Farrell D, Roberts L. A study of high temperature sulfidation
under actual process conditions. Corrosion/2010, Paper 10358.
Houston, TX: NACE International, 2010.
Foroulis ZA. Kinetics and mechanism of the reaction of iron with sulfur vapor in the temperature range of 250 to 500C. Werkstoffe
Korros 1978; 29: 385393.
Groysman A, Feldman B, Kaufman A, Balali R. Hydrogen damage
and prevention in the oil refinery. Corrosion/2011, Paper 11295.
Houston, TX: NACE International, 2011.
Guedes Soares C, Garbatov Y, Zayed A, Wang G. Corrosion wastage
model for ship crude oil tanks. Corros Sci 2008; 50: 30953106.
Gutzeit J. Naphthenic acid corrosion in oil refineries. Mater Perform
1977; 16: 24.
Gutzeit J. Refinery corrosion overview. In: Process industry corrosion
the theory and practice. Houston, TX: NACE International,
1986. pp. 171189.
Harmandas NG, Koutsoukos PG. The formation of iron(II) sulfides
in aqueous solutions. J Crystal Growth 1996; 167: 719724.
Hucinska J. Influence of sulphur on high temperature degradation of
steel structures in the refinery industry. In: Advances in materials
science. Gdansk, Poland: Versita Open, 2006.
Jayaraman A, Saxena RC. Corrosion and its control in petroleum refineries a review. Corros Prevent Control 1995; 42:
123131.
Jenkins W. A refinery process equipment database. Corrosion/1998,
Paper 401. Houston, TX: NACE International, 1998.
Kane RD. Corrosion in petroleum production operations. In: ASM
metals handbook, Volume 13C Corrosion: environments and
industries. Metals Park, OH: ASM International, 2006. pp.
922966.
Kane RD, Cayard MS. A comprehensive study on naphthenic acid
corrosion. Corrosion/2002, Paper 02555. Houston, TX: NACE
International, 2002.
Kanukuntla V. Formation of sulfide scales and their role on the naphthenic acid corrosion of steels. Master of Science Thesis. Athens,
OH: Ohio University, 2008.
- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

R.B. Rebak: Sulfidic corrosion in refineries a review

Kapusta S, van den Berg F, Daane R, Place MC. The impact of oil
field chemicals on refinery corrosion problems. Corrosion/2003,
Paper 03649. Houston, TX: NACE International, 2003.
Lagad VV, Cayard MS, Srinivasan S. Prediction and assessment
of rich amine corrosion under simulated refinery conditions.
Corrosion/2010, Paper 10183. Houston, TX: NACE International,
2010.
Lai GY. (2007). Sulfidation. In: High-temperature corrosion and
materials applications. Metals Park, OH: ASM International,
2007. pp. 201234.
Laredo GC, Lopez CR, Alvarez RE, Cano JL. Naphthenic acids,
total acid number and sulfur content profile characterization in
Isthmus and Maya crude oils. Fuel 2004; 83: 16891695.
McLaughlin J. Sulfidation corrosion due to H2S/H2, effect of H2
partial pressure, effect of fluid flow. API Spring Conference, 19
April 2005. New Orleans, LA: API, 2005.
Messer B, Tarleton B, Beaton M, Phillips T. New theory for naphthenic
acid corrosivity of Athabasca oilsands crudes. Corrosion/2004,
Paper 04634, Conference and Exposition. Houston, TX: NACE
International, 2004.
NACE International. Overview of sulfidic corrosion in petroleum
refining. Publication 34103. Task Group 176 of Prediction Tools
for Sulfidic Corrosion. Houston, TX: NACE International,
2004.
Niccolls EH, Gallon AE, Yamamoto K. Systematic integration of
advanced NDE and corrosion monitoring for improved refinery reliability. Corrosion/2008, Paper 08280, Conference and
Exposition. Houston, TX: NACE International, 2008.
Niccolls N. Sulfidation: thoughts on theory and practice. Energy
Technology Company, ChevronTexaco. Augusta, GA: Doe
Hydrogen Pipeline Working Group, 2005.
OKane JM, Rudd TF, Cooke D, Dean FWH, Powell SW. Detection
and monitoring of naphthenic acid corrosion in a visbreaker
unit using hydrogen flux measurements. Corrosion/2010, Paper
10351. Houston, TX: NACE International, 2010.
Piehl RL. Naphthenic acid corrosion in crude distillation units. Mater
Perform 1988; 27: 3743.
Qu DR, Zheng YG, Jing HM, Yao ZM, Ke W. High temperature
naphthenic acid corrosion and sulphidic corrosion of Q235 and
5Cr0.5Mo steels in synthetic refining media. Corros Sci 2006;
48: 19601985.
Qu D, Liu X, Jiang X, Lan Z, Shan G. Setting critical operational
TAN and sulfur level for crude distillation units. Corrosion/

133

2011, Paper 11362. Houston, TX: NACE International,


2011.
Rapp RA. Hot corrosion of materials: a fluxing mechanism? Corros
Sci 2002; 44: 209221.
Riley J. Roundtable on Sulfidation. API Spring Conference, 19 April
2005. New Orleans, LA: API, 2005.
Ruschau GR, Al-Anezi MA. Petroleum refining: corrosion control
and prevention. In Corrosion costs and preventive strategies in
the United States, Appendix U. Washington, DC: US Department
of Transportation, 2001.
Setterlund RB. Selecting process piping materials. Hydrocarb
Process 1991; 70: 93100.
Shahid M, Faisal M. Effect of hydrogen sulfide gas concentration
on the corrosion behavior of ASTM A-106 Grade A carbon
steel in 14% diethanol amine solution. Arab J Sci Eng 2009; 34:
179186.
Slavcheva E, Shone B, Turnbull A. Factors controlling naphthenic
acid corrosion. Corrosion/1998, Paper 579. Houston, TX: NACE
International, 1998.
Smith SN, Brown B, Sun W. Corrosion at higher H2S concentrations and moderate temperatures. Corrosion/2011, Paper 11081.
Houston, TX: NACE International, 2011.
Sun W, Nesic S. A mechanistic model of H2S corrosion of mild steel.
Corrosion/2007, Paper 07655. Houston, TX: NACE International,
2007.
Tebbal S. Critical review of naphthenic acid corrosion. Paper 380.
NACE Corrosion Conference, 1999. Houston, TX: NACE
International, 1999.
Turnbull A, Slavcheva E, Shone B. Factors controlling naphthenic
acid corrosion. Corrosion 1998; 54: 922930.
Tuttle RN. Petroleum production and refining. In: Corrosion tests
and standards: application and interpretation, 2nd ed., Chapter
76. West Conshohocken, PA: ASTM International, 2005. pp.
812821.
Vedage H, Ramanarayanan TA, Mumford JD, Smith SN.
Electrochemical growth of iron sulfide films in H2S-saturated
chloride media. Corrosion 1993; 49: 114121.
Wen S, Mucek M. Low-temperature hot corrosion in the refining
industry. Corrosion/2011, Paper 11365. Houston, TX: NACE
International, 2011.
Wilks GW. Unusual aspects of corrosion failures in refinery
hydrotreater units. Corrosion/2000, Paper 00702, Conference
and Exposition.

- 10.1515/CORRREV.2011.021
Downloaded from PubFactory at 08/22/2016 04:09:50PM
via free access

Você também pode gostar