Вы находитесь на странице: 1из 109

400 Characteristic Corrosion

Phenomena

Author: B.L. (Brian) Jack

Abstract
This section discusses the theory and mechanisms of several corrosion phenomena
that affect both upstream and downstream units. It includes alternatives for material
selection to remedy commonly encountered corrosion problems.

Contents Page

410 Intergranular Corrosion and Polythionic Acid Stress Corrosion Cracking 400-3
411 Appearance of Intergranular Corrosion
412 Theory of Intergranular Corrosion
413 Factors Affecting Susceptibility to Intergranular Attack
414 Preventing Sensitization
415 Welding
416 Media (Chemicals) That Cause Intergranular Corrosion
417 Laboratory Tests For Detecting Sensitization
418 Polythionic Acid Stress Corrosion Cracking
420 Stress Corrosion Cracking 400-21
421 Stress Corrosion CrackingGeneral Description
422 Stress Corrosion Cracking Variables
423 General Preventive Measures
424 Oil and Chemical Industry Alloys
430 High-temperature Oxidation 400-39
431 Theory of Oxidation
432 Special Forms of Oxidation
433 Effects of Alloying Elements in Steels

Chevron Corporation 400-1 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

434 Oxidation of Carbon and Low-alloy Steels


435 Oxidation of High-alloy Steels
436 Oxidation of Other Metals
437 Oxidation-resistant Coatings
440 High-temperature Hydrogen Attack 400-55
441 Mechanism of Hydrogen Attack
442 Prevention of High-temperature Hydrogen Attack
443 Incubation Period for Hydrogen Attack
450 Hydrogen Damage Due to Wet H2S (Sour Service) or HF Acid 400-62
451 Mechanism of Hydrogen Blistering
452 Mechanism of Sulfide Stress Cracking (SSC) in Welds or High Strength
Steels
453 Mechanisms of Hydrogen Induced Cracking (HIC) and Stress Oriented
Hydrogen Induced Cracking (SOHIC)
454 Occurrence of Hydrogen Blistering and Cracking in Wet H2S Service
455 Inspection for Wet H2S Damage
456 Disposition and Repair of Wet H2S Damage
457 Prevention of Blistering and Cracking Due to Wet H2S
458 Hydrogen-Induced Cracking (HIC) of Line Pipe
459 Sour Service Considerations for Upstream Equipment
460 Hydrogen Embrittlement and Delayed Fracture 400-99
461 Mechanism of Hydrogen Embrittlement
462 Conditions Leading to Hydrogen Embrittlement
463 Hydrogen Outgassing of Reactors
464 Effects of Hydrogen Embrittlement on Fatigue
470 References 400-106

January 2001 400-2 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

410 Intergranular Corrosion and Polythionic Acid Stress Corrosion


Cracking
Introduction
Section 410 discusses intergranular corrosion and polythionic acid stress corrosion
cracking (For more information on polythionic and other corrosion cracking, see
Section 418). Polythionic cracking is a form of stress corrosion cracking (occurring
mainly in austenitic stainless steel) caused by the same factors that lead to intergran-
ular corrosion in stainless steels. For a discussion of other forms of stress corrosion
cracking, see Section 420.
Intergranular corrosion is a form of corrosive attack which proceeds selectively
along the boundaries between the grains of a metals structure. Many alloys are
susceptible to intergranular attack in specific environments; however, intergranular
corrosion is most prevalent in the 300Series austenitic stainless steels.
Discussion will be centered on the austenitic stainless steels, although Alloy 800
(Incoloy 800) and Alloy 20 are also subject to sensitization and subsequent inter-
granular attack. Discussion of intergranular corrosion in other alloys can be found in
the general references listed at the end of this section.

Definitions
Austenite is a form of iron in which the atoms in the crystal structure are arranged
in a repetitive pattern called a face-centered cubic unit cell (having an atom at each
corner and each face of the cube). While the austenitic structure is characteristic of
all steels at high temperatures, austenitic steels are those which do not transform to
other metallurgical structures during cooling. Although other factors can influence
whether or not a steel transforms, chemical composition is the most important.
An austenitic stainless steel contains 16%28% chromium and 6%22% nickel as
principal alloying elements. In general, corrosion resistance depends on chromium
content, and increased chromium content improves corrosion and oxidation
resistance.
A carbide is a compound of carbon with one or more metallic elements. In austen-
itic stainless steels, carbon normally is completely dissolved at high temperatures
(i.e., above 1700F), and rapid cooling can keep the carbon in solution. If the stain-
less steel is cooled slowly, carbon migrates to grain boundaries and forms chro-
mium carbide precipitates.
Sensitization is the harmful precipitation of chromium carbides such as (Cr,Fe) 23
C6, in an almost continuous network around the metal grains of an austenitic stain-
less steel. The formation of chromium carbides leaves a chromium-depleted zone at
the grain boundaries and thus renders the alloy susceptible to intergranular corro-
sion. The temperature range in which sensitization occurs is about 750F to 1550F.
Sensitization has a time-temperature dependence and is most rapid at about 1250F
to 1350F.

Chevron Corporation 400-3 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Stabilization is the addition to a stainless steel of an alloying element that has a


greater affinity for carbon than does chromium. The stabilizing elements form
carbides and leave the alloy without local depletion of chromium. Titanium and
columbium (niobium) are the elements most frequently used to stabilize stainless
steels. Stabilized alloys generally are not susceptible to intergranular corrosion, but
they can become so under certain conditions. Stabilization can also refer to a heat
treatment intended to cause the preferential precipitation of stabilizing element
carbides.
Extra-low Carbon Stainless Steel. An ELC or L-grade stainless steel has a
reduced carbon content of 0.03% or less to make it immune to sensitization during
welding. Type 304L and 316L are steels of this type. Another consequence of low
carbon content is a substantial reduction in strength.

411 Appearance of Intergranular Corrosion


Figure 400-1 shows the surface appearances of two Type 316 parts that have experi-
enced intergranular attack. In the coarse-grained impeller casting (left photograph),
the grains stand out in relief because of preferential attack along the grain bound-
aries. The surface of the fine-grained wrought pipe (right photograph) also exhibits
the typical roughened appearance frequently called sugaring. When the surface is
magnified many times, as in Figure 400-2, the grain boundaries and intergranular
attack are readily apparent. In instances of severe intergranular attack, pieces of the
metal may look like sandstone and individual grains can be separated easily.

Fig. 400-1 Surface Appearance of Equipment that Failed by Intergranular Corrosion in Warm Sulfuric Acid. Note
large grains in cast Type 316 pump impeller in the left picture that were revealed by accelerated grain
boundary attack. Roughened, sugared appearance of interior surface of Type 316 pipe in the right
picture is characteristic of wrought material. Magnification is approximately 1.3 X.

January 2001 400-4 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-2 Microstructure of Intergranularly Corroded Type 316 Pipe Shown in Figure 400-1. The condition at the
inner surface is shown in the left picture at a magnification of 150X. Attack occurred preferentially
through the grain boundaries, and grains literally fell out. At 500X magnification, the right picture shows
the attacked grain boundaries more clearly.

412 Theory of Intergranular Corrosion


The almost universally accepted theory for intergranular corrosion is based on
depletion of chromium in the grain boundary areas. The addition of chromium to
ordinary steel imparts corrosion resistance to the steel in many environments.
Generally, at least 12% chromium is needed to make a stainless steel. If the chro-
mium content is effectively lowered by carbide formation, the relatively poor corro-
sion resistance of ordinary steel is approached.
In the sensitization temperature range (750F to 1550F), carbon and carbides such
as Cr23C6 are virtually insoluble and precipitate out of solid solution if the carbon
content is about 0.02% or higher. Chromium and carbon are thereby removed from
solid solution, forming Cr-carbides along the grain boundaries. The result is metal
with lowered chromium content in the area adjacent to the grain boundaries. The
chromium carbide in the grain boundary is not attacked, but the chromium-depleted
zone near the grain boundary is corroded because it does not contain sufficient chro-
mium to resist attack in many corrosive environments.

Chevron Corporation 400-5 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Common 18-8 stainless steel, Type 304, usually contains from 0.06% to 0.08%
carbon, so that excess carbon is available to combine with the chromium and precip-
itate the carbide. This situation is shown schematically in Figure 400-3. Carbon
diffuses towards the grain boundary quite readily at sensitizing temperatures, but
chromium is much less mobile. Therefore, the chromium depleted when the
carbides form is not quickly replenished by diffusion of chromium from the centers
of the grains.

Fig. 400-3 Chromium Concentration Near a Chromium Carbide Particle Precipitated in a


Grain Boundary.

There is some evidence to indicate that chromium content at the boundary can be
reduced to a very low level or zero. This not only lowers corrosion resistance, but
also brings two dissimilar metal compositions into contact, creating a large, unfa-
vorable anode/cathode area ratio. The depleted area (anode) protects the grains
(cathode). The net effect is rapid attack in the depleted grain boundary area, with
little or no attack on the grains themselves.

January 2001 400-6 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

413 Factors Affecting Susceptibility to Intergranular Attack


An alloy is to intergranular corrosion if the following occurs:
The alloy is sensitized.
A corrosive environment exists.
Sensitized stainless steel can be used if a corrosive medium is not present.
The three major factors which influence the susceptibility of austenitic stainless
steels to intergranular attack are carbon content, grain size, and degree of exposure
(time and temperature).

Carbon Content
The amount of carbon in a stainless steel determines whether it will be susceptible
to intergranular attack. The effect of carbon content can be separated into three
composition ranges:
From 0% to about 0.016% carbon, all carbon remains dissolved regardless of
heat treatment. Alloys in this range are not commercially available. However,
they would be immune to intergranular attack.
From 0.016% to about 0.03% carbon, chromium carbides will precipitate in
the grain boundaries if the alloy is heated to between 750F and 1550F.
However, for comparatively short heating times, the extent of precipitation is
insufficient to form the continuous network necessary for serious intergranular
corrosion. Extra-low carbon grades of stainless (0.03% maximum carbon) can
therefore be used as welded. However, immunity is lost if extra-low carbon
steels are heated for long times in the sensitizing range. Therefore, the
Company generally limits extra-low carbon grades to a maximum service
temperature of 700F.
Above a nominal 0.03% carbon, alloys are readily susceptible to intergran-
ular attack if heated for a short time as during welding.

Grain Size
As has been stated, chromium carbide precipitation in unstabilized austenitic stain-
less steels occurs mainly at grain boundaries. Both carbon content and the amount
of grain boundary area available for carbides will govern the density of the precipi-
tate and hence the degree of sensitization. Because the grain boundary area is
inversely proportional to grain size, the boundary area of large-grained material is
reduced as compared to small-grained. For a given carbon content and heat treat-
ment, a small grain size causes the carbides to be more sparsely distributed over the
greater available grain boundary area. Therefore, the degree of sensitization is
lower.
In wrought products, the range of grain sizes encountered is usually relatively small,
and this variable exerts only a minor influence. (For Alloy 800, a small grain size
must be specified if that is what is desired, since Alloy 800 is generally furnished
with a large grain size for creep strength.)

Chevron Corporation 400-7 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

On the other hand, grain size varies widely in castings, depending chiefly on the
cooling rate immediately after the casting is poured. Grains with diameters up to
1 inches are not uncommon in large castings. Even with low carbon contents, such
large-grained material is easy to sensitize. Figure 400-4 shows a large Type 316L
valve seat that failed from intergranular attack by ammonium sulfate. The carbon
content was 0.026% and there was no record of improper heat treatment. With the
extremely large grain size evident in the picture, the casting was sensitized by
(1) the comparatively short time in the sensitizing zone during cooling from the
solution heat treatment, and (2) the heating resulting from the hard-surfacing of the
adjacent seating surface.

Fig. 400-4 Section of Intergranularly Corroded Cast Type 316L Seat from Isophthalic Plant
Crystallizer Dump Valve. Note the large grains. Although the carbon content was
0.026%, a sufficient density of carbides precipitated in the grain boundaries
during hard surfacing or heat treatment to sensitize the material. Magnification
approximately 0.6X.

Although the rate of sensitization is higher in a coarse-grained casting, the same


degree of susceptibility may be reached in a fine-grained wrought product if heating
time is sufficiently long. This is illustrated in Figure 400-5.

Time and Temperature of Exposure


Precipitation of chromium carbides in an austenitic stainless steel is dependent on
carbon diffusion and on solubility of carbon in the austenite. Each of these factors is
strongly temperature dependent; both diffusion rate and carbon solubility increase
with increasing temperature.
Below the sensitizing range (750F to 1550F) precipitation is too slow to be of any
importance. Above the sensitizing temperature range, the carbon remains largely

January 2001 400-8 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-5 Effect of Grain Size on Development of Sensitivity in an Austenitic Stainless


Steel Alloy.

dissolved in the austenite. Because atom mobility increases with temperature,


precipitation is more rapid in the upper end of the sensitizing range than in the
lower. Two minutes at 1450F might be sufficient to cause a small but measurable
susceptibility to intergranular attack, while a minimum of an hour at 1000F would
be needed to achieve the same susceptibility.
This temperature effect on precipitation is illustrated in Figure 400-6, which shows
the microstructures of Type 304 specimens after heating for 16 hours at 875F,
1000F, and 1100F. Virtually no carbides are precipitated at 875F, and the spec-
imen was not susceptible in corrosion tests. A much greater degree of sensitization
is evident in the 1100F specimen than in the 1000F specimen. However, if the
1000F specimen had been heated for a longer period, it would have become as
sensitized as the 1100F specimen.
The time effect is demonstrated in Figure 400-7, which shows the same Type 304
material heated at 1000F for up to 10,000 hours. The increase in precipitation as a
function of time and temperature is evident not only in the increasing grai boundary
attack but also in the eventual appearance of carbides at less preferred precipitation
sites such as substructure boundaries within the grains.
Although precipitation begins more rapidly at higher temperatures, the eventual
severity of sensitization is much greater at lower temperatures. Figure 400-8 shows
that maximum sensitivity for an 0.038% carbon Type 304 steel is reached in the
1000F to 1250F range. Each of the different austenitic stainless steels has slightly

Chevron Corporation 400-9 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-6 Effect of Temperature on the Degree of Sensitization in Type 304 Stainless Steel Following 16-Hour Expo-
sures. The specimen received from the mill and the 875 F specimen are essentially free of sensitization,
while the 1000 F and 1100 F specimens are sensitized enough to be attacked by acidified copper
sulfate test solution. Magnification is 500X.

a. As-received specimen b. 16 hours at 875F

c. 16 hours at 1000F d. 16 hours at 1100F

January 2001 400-10 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-7 Effect of Time at 1000F on the Sensitization of Type 304 Stainless Steel. Note increasing amount of grain
boundary attack with time. Specimens heated 16 hours or longer were found to corrode intergranularly
in acidified copper sulfate test solution. Magnification is 500X.

a. As-received specimen b. 1 hour at 1000F

c. 16 hours at 1000F d. 100 hours at 1000F

e. 1000 hours at 1000F f. 10,000 hours at 1000F

Chevron Corporation 400-11 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-8 Effect of Temperature and Time on the Severity of Sensitization of Type 304
Stainless Steel.

different sensitivity characteristics. The temperature that produces a maximum


susceptibility may vary from one alloy to another.
A sensitized stainless steel will finally heal itself if held in the sensitizing range
for sufficient time. This occurs because chromium will eventually diffuse into the
depleted zones at the grain boundaries from the adjacent chromium-rich regions.
Once the chromium at the grain boundaries has been replenished, the alloy will no
longer be sensitive to intergranular attack. The time required for healing will depend
on temperature but can be 1000 hours or more.

414 Preventing Sensitization


Three methods are used to control or minimize sensitization of austenitic stainless
steels:
High-temperature solution heat treatment, commonly called solution-
annealing
Lowering carbon content to below 0.03%
Adding stabilizing elements that are stronger carbide formers than chromium

January 2001 400-12 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Solution-annealing
Solution-annealing is accomplished by holding the stainless steel at a temperature
sufficiently high (1800F2100F depending on grade) to dissolve any existing
chromium carbides. Subsequently cooling the steel rapidly through the sensitizing
temperature range (preferably by water quenching) prevents precipitation of the
carbides. Standard stainless steel (Type 304 or Type 316) so treated is subject to
sensitization if subsequently heated or exposed to temperatures above 750F in
service. Mill products and castings of austenitic stainless steel should be purchased
in the solution-annealed (and pickled) condition.

Lowering Carbon Content


Carbon content of the alloy can be reduced to less than 0.03% so that damaging
amounts of carbides are not precipitated during normal fabrication. Type 304L and
316L are examples of low carbon (or L-grade) stainless steels.
Figure 400-9 shows the microstructures of Type 304 and Type 304L test specimens
after a two-hour stress relief at 1550F. The 304 alloy is thoroughly sensitized,
while the 304L alloy is not. However, when the same two steels are exposed for
10,000 hours at 875F, there is a change: the 304 alloy is sensitized, of course, but
now so is the 304L. See Figure 400-10.

Fig. 400-9 Microstructures of Types 304 (left) and 304L (right) Stainless Steel Stress Relieved for Two Hours at 1550-
1600F and Air Cooled. Note extensive grain boundary attack in Type 304 after exposure to an acid solu-
tion, and the absence of attack in 304L. This explains why it is possible to stress-relieve 304Lbut not
304without subsequently risking intergranular corrosion in corrosive services. Magnification is 500X.

Chevron Corporation 400-13 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-10 Microstructures of Types 304 (left) and 304L (right) Stainless Steel After 10,000 Hours at 875F. A suffi-
cient density of chromium carbides has precipitated in the grain boundaries to make both alloys suscep-
tible to intergranular corrosion. As a result of the extended time heating, Type 304L has lost its corrosion
resistance. Magnification is 500X.

While the low carbon grades resist sensitization during ordinary heating during
fabrication, welding, and stress relieving, they are not immune to sensitization from
long-term exposure to elevated temperatures. Thus, the low carbon grades of
stainless steel should not be used if, during fabrication or at any time in service,
they will be subjected to prolonged heating in the sensitizing range.
Only low carbon grade steels (304L, 316L) should be used for welded construction.
This is because even the short heating times obtained during welding will sensitize
Types 304 and 316 stainless steels. If operating temperatures are above 700F, use
Type 321 or 347 stainless steel.

Stabilization
Stabilization is the addition of an alloying element which has a greater affinity for
carbon than does chromium. The chromium is left in solution in the austenite grains,
and corrosion resistance remains unimpaired. The commonly used stabilizing
elements, columbium (or columbium plus tantalum) and titanium are used to
produce Type 347 and Type 321 stainless steels, respectively. These elements are
added in sufficient quantity to combine with all of the available carbon in the alloy.

January 2001 400-14 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Columbium and titanium carbides are highly stable at elevated temperatures. For
optimum resistance to sensitization, heat Type 347 and Type 321 with stabilizing
heat for several hours at about 1650F. This thermal stabilization results in the
formation of stabilizing-element carbides rather than chromium carbides. The sepa-
rate heat treatment is seldom needed, however, since the formation of the colum-
bium and titanium carbides is largely completed during the hot work of the original
ingot.
The stabilizing-element carbides are randomly scattered throughout the grains, in
marked contrast to the chromium carbides, which precipitate largely in the grain
boundaries of the unstabilized grades. This difference in microstructure is shown in
Figure 400-11 for Types 304, 304L and 347. After a 100-hour exposure at 1100F in
laboratory tests, both Type 304 and Type 304L were in a susceptible condition,
while Type 347 was immune to intergranular attack.
Stabilized grades of austenitic stainless steels can be used at higher temperatures for
long-term exposure.
Type 321 stainless steel can be operated indefinitely at temperatures up to
850F with negligible sensitization
Type 347 stainless steel can be operated indefinitely at temperatures up to
900F with negligible sensitization
Many years of experience show that these temperatures are acceptable for Types
321 and 347 stainless steel even without thermal stabilization. However, for equip-
ment operating for long periods above 900F (such as furnace tubes), we recom-
mend using only the Type 347 stainless, and that it be thermally stabilized for at
least four hours at 1650F +/- 25F.

415 Welding
Titanium is reactive at very high temperatures and can be lost to oxidation if used as
the stabilizing agent in welding electrodes. Such a large proportion of titanium is
lost in transfer across the welding arc that a weld deposit made with Type 321 wire
would not be stabilized. Columbium is less reactive, and therefore stabilized elec-
trodes are made from Type 347. For similar reasons Type 347 (or its cast equivalent
CF-8C) is used for castings.

Weld Decay
If equipment is fabricated from unstabilized or non-L-grade austenitic stainless
steel, intergranular corrosion can occur in a band adjacent to welds, where local
sensitization has resulted from the heat of welding. This specific case of sensitiza-
tion is called weld decay. An example is shown in Figure 400-12. One solution is to
use Type 321 or Type 347 base material and Type 347 weld rod.
A stabilized electrode will not eliminate sensitization in the heat-affected zones of
unstabilized grades of base metal. The area of failure from weld sensitization is in
the base metal, and little or no diffusion of columbium from weld metal to base
metal takes place. If the environment will attack sensitized stainless, all stainless in

Chevron Corporation 400-15 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

a welded structure must be either stabilized, L-grade, or solution-annealed after all


welding and other heat treatment take place.

Knife Line Attack


In media having a strong tendency to cause intergranular corrosion (for example,
nitric acid), an attack of stabilized grades can occur in a thin, sharply defined line
adjacent to the weld metal. This is referred to as knife line attack.
Both Types 321 and 347 are susceptible to knife line attack; however, except for
nitric acid environments, knife line attack has generally not been a problem in the
oil industry.
During welding, the base metal in this narrow zone is heated to a temperature high
enough to dissolve most of the columbium and titanium carbides, which then remain
in solution when the weld metal cools rapidly. If this zone is reheated to about
1200F during, for example, stress relief, chromium carbides precipitate in the grain
boundaries, and the stabilizing elements remain in solution. These zones are thereby
sensitized and susceptible to intergranular attack.
However, if the zones are reheated after welding to a temperature of approximately
1600F, stabilizing element carbides will precipitate in preference to chromium
carbides, and the material is again in a stabilized state. This occurs because at this
higher temperature the solubility of stabilizing element carbides is lower than that of
the chromium carbides.

January 2001 400-16 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-11 Microstructure of Types 304, 304L and 347 Stainless Steel Comparing the
Amount of Grain Boundary Carbides Precipitated During 100 Hours at 1100F.
The precipitate is continuous in 304, heavy but
discontinuous in 304L, and essentially nonexistent in 347. Only the Type 304
specimen was attacked intergranularly in acidfied copper sulfate test solution.
Magnification is 500X.

Type 304 after 100 hours at 1100F

Type 304L after 100 hours at 1100F

Type 347 after 100 hours at 1100F

Chevron Corporation 400-17 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-12 Intergranular Corrosion in the Heat-Affected Zone (Weld Decay) of Type 304 Stainless Steel Welded with
Type 308. Sensitization of the 304 resulted from the heat of welding. Specimen exposed to
nitric-hydrofluoric acid. Magnification of weld is approximately 3.5X, the microstructure is 250X.

416 Media (Chemicals) That Cause Intergranular Corrosion


Intergranular corrosion failures in service environments can result from exposure to
acids, salts that hydrolyze to acids, or solutions that have been acidified. Serious
intergranular attack has occurred in the following environments:
Organic sulfurous acid
Inorganic sulfurous acid
Sulfuric acid
Acetic acid
Phosphoric acid
Nitric acid
Hydrofluoric acid
Lactic acid
Maleic acid
Hydrocyanic acid
Napthenic acid
Carbolic acid
Ammonium nitrate
Calcium nitrate

January 2001 400-18 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Aluminum sulfate
Ammonium sulfate
Chlorinated solvents1
Hydrochloric acid solutions1

417 Laboratory Tests For Detecting Sensitization


A number of laboratory tests are available for detecting sensitization. They are
sometimes needed to assure a manufactured component is free of sensitization.
They are also used as tools for failure analysis and materials selection.
Sensitization detection tests are detailed in ASTM A262. There are six different
ASTM tests available as follows:
Oxalic Acid Etch Test for Classification of Etch Structures of Austenitic Stain-
less Steels (Practice A).
Ferric Sulfate-Sulfuric Acid Test for Detecting Susceptibility to Intergranular
Attack in Austenitic Stainless Steels (Practice B).
Nitric Acid Test for Detecting Susceptibility to Intergranular Attack in Austen-
itic Stainless Steels (Practice C). This is also called the Huey test. It is useful
because it gives a quantitative measure of the degree of sensitization.
Nitric-Hydrofluoric Acid Test for Detecting Susceptibility to Intergranular
Attack in Molybdenum-Bearing Austenitic Stainless Steels (Practice D).
Copper-Copper Sulfate-Sulfuric Acid Test for Detecting Susceptibility to Inter-
granular Attack in Austenitic Stainless Steels (Practice E). Also called the
Strauss test. This test provides a straight forward, qualitative method of
detecting sensitization.
Copper-Copper Sulfate-50% Sulfuric Acid Test for Detecting Susceptibility to
Intergranular Attack in Molybdenum-Bearing Cast Austenitic Stainless Steels
(Practice F).

418 Polythionic Acid Stress Corrosion Cracking


Partially oxidized sulfur acids, commonly known as polythionic acids, are not
normally found in refinery process units during operation but develop during shut-
down periods. They are formed by oxidation of iron sulfide in the presence of mois-
ture and oxygen. Iron sulfide is a corrosion product in many processes, such as
crude distillation and the first stages of catalytic reforming and hydroprocessing.
These processes may also require austenitic stainless steel in some sections for
corrosion protection. During shutdowns, this stainless equipment and its sulfide
scale may be exposed to moisture either from steaming to remove hydrocarbons or
from the atmosphere during inspection. The polythionic acids that are formed will

1. Solution concentrations have ranged from very dilute to very strong. Temperatures have ranged from ambient to
the boiling point of the solution involved.

Chevron Corporation 400-19 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

cause intergranular corrosion and intergranular cracking of sensitized austenitic


stainless steel.
The formation of polythionic acids is best explained by analogy to sulfurous acid.

Fe + H2S FeS + H2 (in service)


(Eq. 400-1)

FeS + O2 Fe2O3 + SO2 (during shutdowns)


(Eq. 400-2)

SO2 + H2O H2SO3 (during shutdowns)


(Eq. 400-3)
What actually forms in equipment is H 2SO3 plus a number of other partially
oxidized acids, including the polythionics (H2SxO6, where x 3). Because partially
oxidized sulfur acids are not stable, a number of them will typically coexist in a
given solution. The presence of H2SO3 is an indicator that the environment is
capable of causing polythionic acid stress corrosion cracking.
The stress required for polythionic stress corrosion cracking (SCC) of a sensitized
stainless steel can be either applied or due to residual stresses from fabrication.
Because the stress level necessary for cracking is very low, stress relief is not an
effective method of preventing polythionic SCC. If conditions are right, cracking
can be severe and occur in less than an hour. A typical failure is shown in
Figure 400-13.

Fig. 400-13 Failure of Type 304 Stainless Steel Catalytic Reformer Line at a Thermowell Boss Weld. This failure is an
example of polythionic acid stress corrosion cracking of a sensitized alloy. Note the intergranular char-
acter of the cracks. Magnification of left photograph is 2X; right is 150X.

January 2001 400-20 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

The earlier discussion of sensitization also applies to polythionic acid SCC. If the
stainless becomes sensitized, either during fabrication or during service, it will be
subject to cracking if exposed to polythionic acids. Nonsensitized stainless will not
crack.

Prevention of Polythionic Acid Stress Corrosion Cracking


Prevention of stress corrosion cracking from polythionic acids is usually accom-
plished in two ways:
Avoiding sensitization. Use stabilized grades such as 321 or 347, or an extra
low carbon grade such as 304L (see Section 414). Such alloys are initially
immune to stress corrosion cracking by polythionic acids, but will become
susceptible if exposed long enough at temperatures in the sensitizing range
Avoiding formation of polythionic acids. This is required in plants that do not
have sensitization-resistant materials, but is often done in other plants for addi-
tional protection
To avoid polythionic acid formation, equipment should, when possible, be filled
with nitrogen during shutdowns to prevent entry of oxygen and moisture. If the
equipment must be opened during shutdown, it should first be washed with a
sodium carbonate (soda ash) solution. Remember that protection by soda ash
washing lasts only a few days. Recommendations on nitrogen purging and soda ash
washing are given in NACE Recommended Practice RP-01-70. Information on
when to use soda ash wash in specific plants and equipment is provided in Volume 2
of this manual.

TTS Curves
Figure 400-14 shows time-temperature-sensitization (TTS) curves for the
commonly used 300Series stainless steels. The exact shape and position of a TTS
curve for a given alloy will vary with chemical composition. The curves in
Figure 400-14 are average, typical curves for the alloys indicated.1
Each alloy on the curve has a separate sensitization loop. Within the loop the alloy
will be sensitized. Outside the loop it will not be. The curves are based on time-
temperature conditions that produce a degree of sensitization resulting in a 500-mil-
per-year corrosion rate in the ASTM A262 Practice C test (Huey test). This degree
of sensitization results in susceptibility to intergranular corrosion and polythionic
acid stress corrosion cracking.

420 Stress Corrosion Cracking


Many failures of piping, equipment and vessels are attributed to stress corrosion
cracking. These failures occur with little warning, and repair is difficult because the
cracks are numerous and branchy. This section covers the fundamentals of stress

1. The 321 and 347 SS curves in Figure 400-14 are without thermal stabilization (see Section 414). With thermal
stabilization, the 321 SS loop shrinks dramatically, and for practical purposes, the 347 SS loop appears to
disappear.

Chevron Corporation 400-21 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-14 Time-Temperature-Sensitization (TTS) Curves for the Commonly-Used 300-Series Stainless Steels.
Curves are based on 500 mpy corrosion rates in the Huey test, which corresponds to susceptibility to
intergranular corrosion and polythionic acid SCC. For conditions within the loop for a given alloy, it will
be sensitized. Outside its loop, the alloy will not be sensitized. These are average, typical curves to be
used for guideline purposes only.

corrosion cracking and provides a detailed look at specific metals and environments
in the oil and chemical industries where stress corrosion cracking is a concern. The
purpose is to help a field engineer recognize specific combinations of metal and
environment that can cause cracking and offer some recommendations to prevent its
occurrence.

421 Stress Corrosion CrackingGeneral Description


Stress corrosion cracking (SCC) is the failure of a metal through the combined
action of stress and corrosion. For cracks to occur, the metal must be exposed to
both a unique environment and tensile stress.
Environment, as it relates to stress corrosion cracking, refers to the specific
service variables, for example, the chemical species, pH, and temperature that are in
contact with a metal.
Specific metals are only susceptible to cracking in unique environments. For
example, 130F, 50 Baum caustic can crack stressed carbon steel but not nickel
alloys. Also, changing the environmental conditions for carbon steel by either
decreasing the temperature to 100F or decreasing the caustic strength to 20 Baum
will prevent cracking.
Figure 400-15 shows the susceptibility of common metals to cracking by various
corrosives.
Stress corrosion cracks are either intergranular (occurring along metal grain bound-
aries) or transgranular (occurring across metal grains) depending upon the specific
corrodent and metal. See Figures 400-16 and 400-17. Both intergranular and trans-
granular cracks rarely occur in the same metal/environment/stress combination.

January 2001 400-22 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Laboratory studies and failure experience are the only reliable ways to predict
which metals and environmental conditions are necessary to produce stress corro-
sion cracks. Corrosivity alone is not an indicator. Solutions that are highly corrosive
to a material almost never promote stress corrosion cracking.
The two most accepted theories for the cause of stress corrosion cracking are anodic
dissolution and stress-sorption cracking. However, neither theory can account for all
observed characteristics.
Anodic dissolution involves selective corrosion (or oxidation) of local anodic areas.
Grain boundaries and slip planes (locations where deformation occurs) are often
anodic to the surrounding metal because they have higher potential energy. Local
electrochemical action allows cracking to grow by corrosion of these anodic areas.
A tensile stress breaks any protective film formed by the corrosion process,
allowing it to continue. See Section 100 for an explanation of electrochemical
corrosion.
Stress-sorption cracking theory involves the surface energy of metal. Chemical
species in the solution adsorb on the metal surface and decrease its surface energy
enough that a tensile stress causes the surface layer to crack. The degree of adsorp-
tion depends upon electrical potential. Critical cracking potential is defined as the
potential above which adsorption takes place and below which desorption occurs.
Only specific adsorbents effectively decrease the surface energy of a particular
metal.

Chevron Corporation 400-23 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-15 Susceptibility of Common Metals to Cracking in Various Environments(1) (1 of 2)


Carbon/Low
Alloy/ Ferritic Austenitic Copper Nickel
Corrosive Stainless Stainless Steel Alloys Alloys
Aluminum chloride x
Aluminum sulfate x
Ammonia x x
Ammonium biphosphate x
Ammonium chloride x
Ammonium hydroxide x
Ammonium nitrate x
Aniline x
Calcium bromide x
Calcium chloride x
Calcium nitrate x
Carbon tetrachloride x
Chlorine x
Chromic acid x x x
Dichlorophenol x
Diethanolamine x
Ethylamine x x
Ethyl chloride x
Ferric chloride x x
Ferrous chloride x
Glycerol x
Hexachloroethane x
Hydrofluoric acid vapors x x
Hydrofluoric acid (aerated) x x x
Hydrofluoric acid (no air) x x x
Hydrogen chloride x
Hydrogen sulfide (wet) x x
Lithium chloride x
Magnesium chloride x x x
Manganese chloride x

January 2001 400-24 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-15 Susceptibility of Common Metals to Cracking in Various Environments(1) (2 of 2)


Carbon/Low
Alloy/ Ferritic Austenitic Copper Nickel
Corrosive Stainless Stainless Steel Alloys Alloys
Monoethanolamine x x
Nickel chloride x
Nickel nitrate x
Nitric acid x x
Pentachloroethane x
Phosphoric acid x
Phosphoric acid (aerated) x
Polythionic acid x
Potassium carbonate x
Potassium chloride x x
Potassium chromate x
Potassium hydroxide x x x
Silver Nitrate x
Sodium bisulfate x
Sodium bisulfite x
Sodium carbonate x x
Sodium chloride x
Sodium fluoride x x
Sodium hydroxide x x x
Sodium nitrate x x
Sodium phosphate (tribasic) x
Sodium sulfate x
Sodium sulfide x
Sulfate liquor x x
Sulfonated oil x
Sulfur x
Sulfurous acid x
Toluene x
Zinc chloride x
(1) This table is for guidance only. For questions on specific environments and metals contact CRTC Materials and Equipment Engineering.

Chevron Corporation 400-25 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-16 Intergranular Cracking Adjacent to Welds. Stress corrosion cracks in carbon steel pipe exposed to 170F
lean diethylamine (DEA). Cracks occurred opposite an externally welded pipe support.

Fig. 400-17 Transgranular Cracking of Nonsensitized Austenitic Steel. Failure of crude unit furnace control valve
stem. Exposure was hot emulsion of crude oil and saltwater. Branching transgranular cracking in the
right photograph is characteristic of chloride cracking of nonsensitized austenitic stainless steel.
Magnification of the left photograph is approximately 2.5X; right is 500X.

January 2001 400-26 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

422 Stress Corrosion Cracking Variables


The principal variables of both anodic dissolution and the stress-sorption cracking
mechanisms are:
Tensile stress
Service temperature
Solution environment
Duration of exposure
Metal properties
All these variables are interrelated, and cracking may occur under numerous condi-
tions. The general influence of these variables on cracking is described as follows.
Specific environment/metal combinations are discussed in Section 424.

Tensile Stress
Tensile stress consists of residual stress from welding and fabrication, plus in-
service stresses. Increasing the total tensile stress increases stress corrosion
cracking. Residual weld stresses often cause failure. They are significantly higher
(near yield stress levels) than in-service stresses unless welds are stress-relieved by
heat treatment.
The amount of stress necessary to cause cracking for a given set of conditions is
called the threshold stress level. At a stress below the threshold stress level cracking
will not occur. Threshold stress values are unique and vary considerably, depending
upon the environment. For instance, stainless steel has cracked at stress levels as
low as 500 psi under severe chloride solution test conditions. For sulfide cracking,
stress levels have been as high as 100 ksi before cracking high strength carbon steel.
The threshold stress levels determine the acceptable level of stress in service or after
fabrication.

Service Temperature
The influence of temperature can only be described in general. In most environ-
ments, increased temperature causes more severe cracking, and there is usually a
threshold temperature below which cracking will not occur.
Temperature plays an important role in concentrating the chemical species that
causes cracking. High temperatures may evaporate liquids and concentrate crack-
producing species. Even though the species may be below the threshold value for
cracking in the bulk fluid, the species may concentrate enough to exceed threshold
level. Evaporation and concentration in boilers and heat exchangers are a special
concern.

Solution Environment
The solution environment significantly influences SCC. The pH of the solution,
concentration of crack-producing species, and presence of oxygen all play impor-
tant roles. If the crack-producing species is not present above threshold level,
cracking will not occur.

Chevron Corporation 400-27 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Dissolved gases may influence the tendency for cracking by changing the threshold
level of other variables. An example is austenitic stainless steel in aqueous chlo-
rides, where oxygen greatly increases cracking susceptibility.
Figure 400-15 lists environments and materials found in our industry in which stress
corrosion cracking can occur.

Duration of Exposure
The duration of exposure is a variable not well defined. Usually cracking times are
extremely short. Typically, when failure occurs after several years, it is because of a
change in solution conditions that made the metal susceptible to cracking.
Little literature is available on the threshold time for cracking, although everyone
agrees that there must be a threshold and that it varies with the conditions and envi-
ronment. Because of the uncertainty of threshold times, decreasing the amount of
time a metal is exposed to a solution is not a practical way to control cracking.

Metal Properties
A metals properties can affect its susceptibility to SCC. Some metals must undergo
metallurgical change to be susceptible. Certain metallurgical phases are more prone
to cracking than others in the same environment. In other environments, the strength
of the material must exceed a threshold value for the material to be susceptible to
cracking.

423 General Preventive Measures


The incidence of stress corrosion cracking is specific for a given metal, and is deter-
mined by a combination of the following variables: time, temperature, solution,
environment, and stress conditions. These variables are not always known or
defined. A change in one variable that places it outside the critical range can elimi-
nate cracking just as easily as placing it in the critical range can cause cracking. The
variables we can most easily change to prevent cracking are:
Type of alloy
Amount of stress
As has been said, controlling duration of exposure is not practical. Also, it is usually
impossible to change the process variables (e.g., temperature, concentration of crit-
ical chemical species, and stream composition).
The best insurance against cracking is to choose a material resistant to cracking in
the given environment. Another practical measure in well-known cracking environ-
ments is stress relief heat treatment, which reduces the stress level below threshold
stress. However, in some environments, the threshold stress is either unknown or so
low that stresses can not be reduced adequately to prevent cracking.
Stress corrosion cracking often occurs in locations where residual fabrication
stresses are highestfor example, welds. Fabrication stresses can usually be
controlled by stress relief heat treatment. Fabricated equipment designed for
services where stress corrosion cracking may be encountered should be heat-treated

January 2001 400-28 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

to minimize residual stresses. External attachment welds, for example, those associ-
ated with pipe supports and anchors, should be stress relieved as well since these
external welds can create through-wall stresses.
Heat treatment should not introduce undesirable side-effects through metallurgical
changes or distortion. Effective stress relief requires control of heating and cooling
rates, uniformity of temperature, and proper mechanical support of equipment.
Otherwise, the heat treatment itself will impart residual stresses.
The temperatures for stress relief of common materials are given in the Welding
Manual. To obtain the greatest reduction in the residual stress, stress relief tempera-
tures should be on the high side of the temperature range for a given material.
Adding corrosion inhibitors to process streams is not effective in controlling stress
corrosion cracking. For example, laboratory tests show that the usual filming amine
inhibitors are useless in preventing cracking of stainless steel by chloride-ion-
contaminated phenol. In certain instances, amine inhibitors increased rather than
decreased the stress corrosion cracking susceptibility in ammonium chloride.

424 Oil and Chemical Industry Alloys


This section addresses common metal/environment combinations encountered in the
oil and chemical industry where stress corrosion cracking is possible, and offers
ways to prevent or solve cracking problems. It also lists specific variables that
promote cracking for each environment and describes the crack features. If appro-
priate, some details are provided on where these environments are commonly
located. See also Figure 400-15.

Carbon/Low Alloy Steels


Carbon and low alloy steels are susceptible to SCC in numerous environments. The
most common are as follows:
Carbonate Alkaline Environments. Carbonate stress corrosion cracking has been
reported in CO2 removal plants using hot potassium carbonate (Benfield and
Catacarb) and most recently in FCC overhead streams. Cracks are typically inter-
granular, and somewhat branchy. The overall crack growth direction is relatively
straight.
Some CO2 absorbers in hot potassium carbonate service cracked at repair welds that
were not stress relieved. Originally the vessels were stress relieved, but excessive
corrosion necessitated weld build-up with carbon steel and strip-lining with stain-
less steel. The carbonate solution leaked behind the strip lining and cracked the
weld build-up.
The specific environment necessary for cracking is not well understood, primarily
because the electrochemical potential of the steel in the carbonate solution must be
in a relatively narrow range for cracking to occur. This potential range changes with
chemical species, temperature, and pH of the solution. Stress relief of welds is a
good preventive measure against cracking; however, it may not be a guarantee if the
applied stress exceeds the threshold stress for SCC.

Chevron Corporation 400-29 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Amine Environments. Amines currently used in CO2 and H2S removal plants have
caused intergranular stress corrosion cracking of carbon steel piping and vessels.
Some of the more common amine solutions are MEA (monoethanolamine), DEA
(diethanolamine) and MDEA (methyl diethanolamine). Cracks have occurred in
both lean and rich amine service. Temperature of the amine is probably the most
influential variable. Cracks occur readily at temperatures exceeding 200F and
decrease in severity as the temperature decreases. However cracks have been
reported in MEA equipment operating as low as 100F.
Stress relief is recommended to prevent cracking. Our current guidelines are to
stress relieve all new vessels and exchangers in MEA, DEA, and MDEA service
regardless of temperature. All new piping above 100F in MEA service should be
stress relieved. The same is true for piping in DEA and MDEA, except stress relief
is not required for socket and seal welds 1 inches and smaller. The stress relief
guidelines are relaxed for DEA and MDEA because they are less prone than MEA
to cause stress corrosion cracking of carbon steel.
For existing equipment that has not been stress relieved increased inspection
frequency is recommended. Inspection should be performed with the wet fluores-
cent magnetic particle technique on clean steel surfaces to locate any stress corro-
sion cracks. Consult the CRTC Materials and Equipment Engineering for more
specific information on inspection and inspection frequency.
Ammonia Environments. Ammonia stress corrosion cracking has occurred in
anhydrous ammonia storage vessels. For cracking to occur, oxygen must be present
at levels greater than about 1 ppm.
Most failures and cracking of spheres have occurred outside the United States and
with foreign steels. The foreign steels typically obtain their strength from small
alloying additions. This is in contrast to domestic steels, which increase strength
through carbon and manganese additions.
The best preventive measure is thermal stress relief of welded vessels. Following
are some other potential actions to minimize cracking:
Ensure that weld hardness does not exceed Brinell 225.
Use low strength carbon steel.
Add 0.2% minimum water to the ammonia.
Operate at temperatures below -33F.
Keep oxygen levels as low as possible.
Hydrofluoric Acid (HF) Environments. Hydrofluoric acid has caused cracking of
high strength carbon steel. The cracks are typically transgranular. The mechanism is
believed related to hydrogen embrittlement from the hydrogen generated in the
corrosion reaction. This mechanism is the same as described for sulfide cracking.
In HF alkylation plants originally built to Chevron specifications, carbon steel welds
should be stress relieved to prevent cracking. HF alkylation plants built by Gulf
followed the process licenser (Phillips) recommendation to control weld hardness
by specifying steel chemistry limits. Both approaches have been successful in
preventing SCC. The practice originally established for a plant should be continued.

January 2001 400-30 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Bolts should be ASTM B7M because the higher strength B7 bolts could crack in
service.
Hydroxide EnvironmentsCaustic. Sodium hydroxide can cause intergranular
cracking in contact with carbon steel at elevated temperatures. This is commonly
referred to as caustic cracking and caustic embrittlement. While sodium hydroxide
causes most of this cracking, other hydroxides such as potassium hydroxide do so as
well. Figure 400-18 shows the concentration of caustic versus metal temperature
where carbon steel is susceptible to cracking.
Stress relief is the most effective way to prevent caustic cracking. We recommend
stress relief of all carbon steel welds when the metal temperature exceeds 140F for
less than 30 wt.% (36 Baum) and 110F when the concentration exceeds 30 wt.%
(36 Baum). The same holds true for cold-formed parts like vessel heads and U-tube
exchanger tubes. A location often overlooked is steam traced lines where local
metal surface temperatures exceed the temperature guidelines and make the steel
susceptible to cracking.
Note that general corrosion increases rapidly with increasing temperatures. For
Area C in Figure 400-18, nickel alloys should be used instead of carbon steel.

Fig. 400-18 Recommended Limits of Carbon Steel in Sodium Hydroxide Service

Nitrate Environments. In what are primarily fertilizer industry services, cracking


of carbon steel has been documented in ammonium nitrate, sodium nitrate, potas-
sium nitrate, calcium nitrate, and their mixtures. The stress corrosion cracks are
intergranular and can occur at ambient temperature. Stress relief is most effective in
preventing cracking in this situation.

Chevron Corporation 400-31 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

The Company has experienced cracking of small, nonstress-relieved field storage


tanks of calcium-ammonium nitrate fertilizer solution. Interestingly, the failures
were highly time-temperature dependent. During the first year the tanks were
installed, failures occurred only in areas with high ambient temperatures in Arizona
and Southern California. The second year failures occurred in the slightly cooler
San Joaquin Valley in California. In Northern California, there were no failures for
several years until, at Richmond, a highly restrained (i.e., highly stressed) lap joint
failed. Figure 400-19 shows stress corrosion cracks next to a weld exposed to
calcium-ammonium nitrate solution.

Fig. 400-19 Stress Corrosion Cracks Adjacent to Welds. Carbon steel test specimens exposed to calcium-ammo-
nium nitrate fertilizer solution at 140F. Note intergranular pattern of cracks. Magnification of left photo-
graph is 1.5X; right is 500X.

Sulfide Environments. Wet hydrogen sulfide can cause cracking of steel. This is
discussed in detail in Section 450.

Ferritic Stainless Steels


Ferritic stainless steels, like carbon steel, are susceptible to stress corrosion cracking
in caustic environments and wet H2S environments. See the discussion of caustic
cracking of carbon steel at the beginning of this section. Section 460 discusses
sulfide cracking.

January 2001 400-32 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Austenitic Stainless Steels


Austenitic stainless steels are particularly susceptible to stress corrosion cracking.
They rely on the formation of a passive chromium oxide film for corrosion protec-
tion. If the film does not form or breaks down in local areas, fresh metal ripe for
corrosion is exposed. Most failures of austenitic stainless steel result from contact
with one of the following:
Aqueous solutions containing chloride ions
Aqueous solutions containing hydroxyl ions
Polythionic or sulfurous acids
Chloride Environments. Stress corrosion cracking of austenitic stainless steels in
aqueous chloride ion solutions occurs at temperatures above 150F at relatively low
threshold stress levels. Below 150F, cracking has occurred only with sensitized
stainless steel in unusually aggressive environments of high stress and high chlo-
ride concentrations.
Chloride concentrations as low as 100 ppm can cause stress corrosion cracking,
especially at pH levels below 8. With more alkaline solutions, a higher chloride
concentration can be tolerated without SCC. Complete resistance to chloride
cracking can only be assured by complete absence of chloride ions.
Cracks are generally transgranular and very branchy. Figure 400-17 shows a control
valve stem that cracked when exposed to an emulsion of hot crude oil and salt
water. If the stainless steel is sensitized cracks will be predominantly intergranular
(see Section 410 for a complete definition of sensitization).
Insulated lines are often overlooked as locations for stress corrosion cracking of hot
stainless steel. The concern is chloride cracking from the external environment.
Insulation materials should be purchased with low chloride content. When insula-
tion becomes wet, chlorides can leach out of the insulation and concentrate on the
surface. Or the chlorides may come from the atmosphere, especially at locations
near the ocean.
Preventive measures against chloride cracking are as follows:
Use of an alloy resistant to chloride cracking. Alloys with high nickel
content, such as Alloy 20 and Incoloy 800, resist cracking. Complete immunity
occurs at about 42% nickel. Figure 400-20 shows the effect of nickel content on
susceptibility to chloride cracking in extremely severe boiling magnesium chlo-
ride tests.
Use of duplex alloys. Duplex refers to a dual-phase matrix containing ferrite
and austenite. The ferrite phase is more resistant to chloride cracking than
austenite. These alloys can be cost-effective, depending on the severity of the
environment and the alloy content necessary to resist cracking. Company expe-
rience with duplex stainless steels is limited.
Thermal stress relief. Use thermally stress-relieved Type 304L, Type 316L,
Type 321, or Type 347 stainless steel. Thermal stress relief of Type 304 or Type
316 can sensitize these grades, making them less resistant to chloride cracking.

Chevron Corporation 400-33 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-20 Effect of Nickel Content on Susceptibility of Cr-Ni Alloys. [1] 18% to 20% chromium content; exposure in
boiling 42% magnesium chloride. From Copsons data. Several commercial alloys are shown at right;
probable resistance to cracking increases with nickel content. (From Copson, H.R., Physical Metallurgy
of Stress Corrosion Fracture, Interscience Publishers, p 2247 (1959).

Figure 400-21 shows a section of oil line that cracked at a nonstress-relieved


temporary weld attachment. It illustrates the danger of weld repairs and tempo-
rary weld attachments on already stress-relieved equipment in a stress corro-
sion cracking environment
Use of insulation materials with low chloride content (less than 50 ppm).
The following are general guidelines for stainless steel exposed to chloride
solutions:
At less than 50 ppm chlorides, stainless steel is considered safe from cracking.
Above 50 ppm, stress relief is recommended if the temperature exceeds 150F.
Most failures occur under an evaporative heat flux, where chlorides can
concentrate by evaporation of the liquid. Even though a process stream contains
less than 50 ppm in chlorides, in an evaporating environment the chloride levels
can exceed the threshold value and crack stainless steel.

January 2001 400-34 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Hydroxyl IonsCaustic. Stainless steels have cracked in hydroxyl ion solutions


such as NaOH, KOH, and fused NaOH. Their susceptibility to cracking generally
follows the same temperature limits as that of carbon steel in caustic service.
Figure 400-22 shows some conservative limits for use of austenitic stainless steels
in caustic service. Few failures resulting from this cracking mechanism have been
reported, although studies indicate that the cracks are transgranular, intergranular, or
a combination of the two. Cracking tends to be predominantly intergranular if the
stainless steel is sensitized. Figure 400-23 shows typical stress corrosion cracks
produced in a 4% NaOH solution between 300F and 400F.

Fig. 400-21 Exterior of a Section of Type 304 Waste Oil Fig. 400-22 Stress Corrosion Cracking of Austenitic
Line Cracked from Exposure to Salt Water. Stainless Steels in Commercial Grade
The causative stresses were introduced by Sodium Hydroxide. This plot illustrates the
the short length of the weld in the center of time-temperature-concentration interde-
the photograph. The weld apparently pendence of this type of cracking.
temporarily held a clip or welding ground to
the line. Random or nonstress-relieved
welds in equipment subject to stress corro-
sion are a hazard. Magnification is approxi-
mately 2X.

At temperatures above 300F, the corrosion rate of stainless steel can be high, and
upgrading materials to a nickel base alloy may be the most effective way to prevent
stress corrosion cracking failures and excessive general corrosion.

Chevron Corporation 400-35 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-23 Stress Corrosion Cracking of Type 304 Exposed to 4% NaOH Solution at 300F to 400F. In the left
photograph, the chromium carbides are in solution (the material is not sensitized) and the cracking is
characteristically transgranular. In the right photograph, the alloy has been sensitized by heating at
1200F to precipitate chromium carbides in the grain boundaries, and the cracking is intergranular.
Magnification is 150X.

Polythionic Acid Stress Corrosion Cracking. Polythionic acid stress corrosion


cracking is detailed in Section 410. Following are the key variables for cracking
susceptibility:
The stainless steel must be sensitized.
The stainless steel must have a sulfide scale on the surface.
Water and oxygen must be present.
Most failures occur during shutdowns because proper precautions are not taken to
prevent exposure of the equipment to water and oxygen. Because threshold stress
levels are low, stress relief is not a good preventive measure. The best preventive
measures center on ensuring that none of the key variables exist at the same time.

Copper Alloys
Copper alloys are susceptible to cracking in two environments commonly found in
process streams: ammonia solutions and sulfate solutions.

January 2001 400-36 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Ammonia Environments. All copper alloys used in plants can fail by stress corro-
sion in a moist ammonia atmosphere. Ammonia may be present in the process or
ammonia will be added for pH control. Figure 400-24 shows a failed admiralty
brass tube. The transgranular, branchy, crack pattern is typical of stress corrosion of
brass, but failure may not always be preceded by the pitting obvious in this case.
Failures are most often associated with admiralty brass heat exchanger tubes. Even
though it is common to use annealed tubing, residual stresses can develop from
rolling into the tubesheet or from vibration in U-bend bundles. Residual stress
should be limited by annealing, and operational stress limited by use of additional
baffles. However, these measures will not ensure complete immunity to cracking.
The best method for preventing failures is to use a material resistant to ammonia
stress corrosion cracking. Relative resistance of several common copper alloys from
most to least resistant is as follows:
Copper
9010 copper-nickel
7030 copper-nickel
Silicon bronze
Aluminum bronze
Admiralty brass
Aluminum brass
Sulfate Solutions. Copper alloys are also susceptible to stress corrosion cracking in
sulfate solutions. This form of cracking is not commonly identified, partly because
it is nearly impossible to differentiate between cracking due to ammonia solutions
and sulfate solutions. However, regardless of the exact chemical species respon-
sible for the cracking, the solution to the problem is the same.

Nickel Alloys
Ammonium Bisulfide and Sea Water. K-Monel studs have undergone cracking
failures in ammonium bisulfide and in sea water environments as a result of
hydrogen embrittlement. The metal becomes charged with hydrogen from the corro-
sion reaction and cracks from applied stress. Threshold stress for embrittlement is
quite high since the failures most often occur in overtorqued bolts. The best preven-
tive measure is to not exceed the recommended stress level on the bolts.
Hydrofluoric Acid Environments. Stress corrosion cracking of nickel-base alloys is
primarily limited to hydrofluoric acid alkylation plants, and HF acid is the primary
culprit for failures in Monel, K-Monel, and Inconel. Cracks can be both intergran-
ular and transgranular. The vapor phase is more likely to produce cracking than the
liquid phase. Figure 400-25 shows cracking failure of a K-Monel stud in an HF
alkylation plant heat exchanger. Stress relief of all Monel and other nickel-base
alloys does away with the problem.

Chevron Corporation 400-37 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-24 Stress Corrosion Cracking in an Admiralty Fig. 400-25 Intergranular Stress Corrosion Cracking of
Brass Tube. From an alkylation plant isobu- K-Monel. K-Monel stud from floating head
tanizer overhead condenser. Pitting of HF alkylation plant exchanger. The
preceded transgranular cracking. Tube cracks resulted from exposure to hydroflu-
was exposed to triethanolamine and SO2, oric acid vapors. Magnification is 150X.
both of which can crack brass.
Magnification is 200X.

Monel K-500 bolts in HF service should have the following characteristics:


Manufactured from cold-drawn bar stock
Bolts annealed after thread cutting but before precipitation hardening
heat treatment
100 ksi minimum yield
140 ksi minimum tensile
Rc 30 maximum hardness
These requirements come from HF alkylation plant recommendations.

January 2001 400-38 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

430 High-temperature Oxidation


All common metals and alloys are unstable in air or oxygen and react to form oxides
even at room temperature. This chemical reaction between metal and oxygen is
called oxidation.
Unlike most chemical reactions, oxidation of many important alloys creates an
oxide film that, as it thickens, forms an increasingly effective barrier between the
metal and the gas.
The rate of oxidation is controlled not by the chemical reaction but by diffusion of
metal outward or oxygen inward through the oxide. This diffusion is dependent on
temperature, temperature fluctuations, the integrity of the oxide layer, and the pres-
ence of other gases in the atmosphere.
High-temperature oxidation is a separate problem of particular interest to process
plant engineers and warrants separate treatment. It can result in excessive corrosion
or scaling and becomes a concern at approximately 1000F.
Our discussion of high-temperature oxidation is confined principally to commonly-
used iron-base and nickel-base alloys, since they are the most widely used in
refinery applications. Fortunately, the capabilities and limitations of most of these
alloys have been thoroughly explored, so unexpected failures from simple oxida-
tion are rare. Failures that do occur usually result from an unforeseen upset in oper-
ating conditions.
A summary of the salient facts concerning the high-temperature oxidation of alloys
commonly used in refinery process units follows:
The maximum temperature at which carbon steel has adequate, long-term resis-
tance to scaling in an oxidizing atmosphere is about 1050F.
Oxidation rates are often not constant with time. The oxide scale can act as a
diffusion barrier, so that the oxidation rate progressively drops as the scale layer
builds up.
Temperature cycling substantially increases the scaling rate of many high-
temperature, heat-resistant alloys because of spalling of the scale.
The oxidation resistance of steels is approximately proportional to the chro-
mium content, but the resistance of a chromium-bearing steel is further
enhanced by small additions of silicon, aluminum, titanium and columbium.
Traces of sulfur gases in the high-temperature environment may increase the
scaling rate of both low- and high-alloy steels. High-nickel, heat-resistant steels
and nickel-base alloys should be used with caution in hot gases containing
appreciable quantities of sulfur.
Oxidation resistance is only one of several criteria to be used in selecting mate-
rials for high-temperature applications. Good high-temperature mechanical
properties, freedom from development of brittle metallurgical constituents, and
resistance to process-side corrosion also are usually necessary.

Chevron Corporation 400-39 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

431 Theory of Oxidation


High-temperature oxidation of metals follows either a Parabolic Rate Law or a
Linear Rate Law, depending on whether the metal oxide layer is protective or
nonprotective. Our discussion of these laws and related theory is necessarily brief.
References [2] and [3] contain extensive and excellent treatments of oxidation
theory.

Parabolic Rate Law


The Parabolic Law of Oxidation (Parabolic Rate Law) applies when formation of a
protective oxide layer provides a continuous barrier between oxygen and metal and
inhibits further oxidation. The protective effect is directly proportional to the layer
thickness. In other words, the rate of further growth is inversely proportional to
layer thickness. The Law of Oxidation is expressed as follows:

W2 = K p t
(Eq. 400-4)
where:
W = weight gain from oxidation per unit area
Kp = parabolic oxidation constant
t = elapsed time
Figure 400-26 illustrates the Parabolic Rate Law. This equation predicts a rate of
oxidation that, initially, is relatively high but continuously decreases according to a
parabolic function.
Most metals and alloys, including carbon steels and low-alloy steels, follow a para-
bolic law of oxidation. In the early stages of film formation, rate of growth is
controlled by surface reactions, first at the metal/oxygen interface and later, as the
film achieves finite thickness, at the metal/oxide and the oxide/oxygen interfaces.
Later, when the film has become appreciably thicker, the controlling factor becomes
the diffusion of metal or oxygen through the oxide layer. See Figure 400-26.
The transport of material through the oxide layer occurs either as movement of
metal cations and electrons to the oxide/oxygen interface, or by oxygen anions to
the metal/oxide interface with an electron flow in the opposite direction. Metal
diffusion occurs more frequently, possibly because the metal cation is generally
smaller than the oxygen anion and passes more readily through the oxide. These two
cases [3] are illustrated schematically in Figure 400-27.

Linear Rate Law


The Linear Law of Oxidation (Linear Rate Law) applies when a nonprotective oxide
layer allows continuous, steady access of oxygen to the metal. Rate of growth of the
oxide therefore is independent of thickness, and the thickness of the layer increases
in a linear manner with time. At high temperatures and long time intervals, the
metal will completely oxidize because the oxidation rate never slows down. The
Linear Law of Oxidation is expressed as follows:

January 2001 400-40 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

W = K1 t
(Eq. 400-5)
where:
W = weight gain from oxidation per unit area
K1 = linear oxidation constant
t = elapsed time

Fig. 400-26 Illustration of Parabolic Law of Oxidation. Fig. 400-27 Two Ways by Which an Oxide Layer on a
Oxide is protective and rate of formation Metal Thickens. Cationic (metal) diffusion
decreases parabolically as oxide layer (A), anionic (oxygen) diffusion (B). After
thickens.(Courtesy of Butterworth Heine- Shreir [3]. (Courtesy of Butterworth Heine-
mann Publishers) mann Publishers)

Figure 400-28 illustrates the Linear Rate Law. Linear oxidation may result when the
oxygen supply is insufficient to maintain the rate at which the oxide layer is capable
of thickening. In short, the oxygen replenishment rate controls the oxidation
process. Linear oxidation may occur in an atmosphere where the oxygen content is
very low, or as a result of the cracking and spalling of the oxide layer. If the oxide
forms in a stressed condition (e.g., with volume changes in the conversion of metal
to oxide) then at some critical thickness the stresses may become large enough to
fracture the layer. In the cracked condition, the oxide layer is nonprotective and the
oxidation rate becomes very high for a short period, gradually reducing as build-up
takes place. This cycle repeats, and, if the critical film thickness is relatively small,
the measured oxidation rate appears roughly constant.

Chevron Corporation 400-41 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

At the start, oxidation rarely follows the Linear Rate Law. A brief initial period
exists in which the rate changes, followed by a period in which the rate is constant.
The result is two oxide layers with dissimilar properties. The first layer forms
during the initial nonlinear period as a thin, continuous film adjacent to the metal.
The thickening rate is controlled by diffusion through this film rather than the
oxygen reaction, so that the rate slows as the film thickens.
At some point during oxidation, the oxide layer changes from a thin continuous film
to a nonprotective porous scale. As already described, the oxide layer may crack and
spall. Oxidation now follows the Linear Rate Law if the thickening rate of the
porous layer equals the rate at which it cracks. The thin inner layer remains of
constant thickness so that it appears that the oxidation rate is roughly constant.
Among the metals that oxidize in this manner are molybdenum, columbium,
tantalum, titanium, zirconium, thorium, and tungsten.

432 Special Forms of Oxidation


Oxidation at Grain Boundaries
In contrast to general oxidation, which is relatively uniform, preferential oxidation
may occur along grain boundaries (intergranular oxidation), particularly at high
temperatures. Figure 400-29 shows a section from a failed carbon steel tube which
overheated in a phthalic anhydride converter. This selective penetration at grain
boundaries can occur only by a movement of oxygen inward through the surface
oxide layer. Penetration is possible only if the layer is permeable.

Catastrophic Oxidation
Catastrophic oxidation occurs when a metal is exposed to vapors of a low-melting-
point oxide or when metal undergoing oxidation contains an element forming a low-
melting-point oxide. Elements causing catastrophic oxidation include vanadium,
bismuth, tungsten, molybdenum, and lead. These liquid oxides attack the metal,
destroying the adhesiveness of the protective oxide layer.
Molybdenum-bearing stainless steels can undergo oxidation at rates ten times higher
than normal. This catastrophic oxidation has been observed chiefly in the austenitic
grades, Types 316 and 317, which contain 2% to 4% molybdenum. At furnace
temperatures of 1600F to 2000F, the most stable oxide of molybdenum is volatile
and present as a vapor. Typically, where the flow of furnace gases sweeps the
molybdic oxide away as rapidly as it forms, oxidation proceeds in a normal fashion.
However, in crevices or protected areas where the atmosphere is stagnant, the MoO2
accumulates and enriches the scale forming a loose and nonprotective scale. The
resulting oxidation and metal loss may be extremely rapid.

Subsurface Oxidation
At temperatures over about 1700F, stainless steels and high-nickel alloys may
suffer subsurface oxidation. This occurs in low-oxygen environments when condi-
tions are oxidizing to some elements in an alloy but not to others. For example,
oxygen atoms may diffuse through austenite grains in a stainless steel without

January 2001 400-42 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-28 Illustration of Linear Law of Oxidation. Fig. 400-29 Section From Failed Carbon Steel Tube
Oxide is porous and nonprotective. Rate of From Phthalic Anhydride Reactor Over-
thickening is constant. Nonlinear part of heated in Service. Note intergranular
curve b may be result of formation of a very penetration of oxide. Magnification is 300X.
thin protective oxide layer adjacent to the
metal surface. (Courtesy of Butterworth
Heinemann Publishers)

reacting, then combine with the chromium in a chromium carbide precipitate. The
ultimate effect is selective oxidation of one or more elements within the body of the
alloy rather than at the surface. In Fe-Ni-Cr alloys, an oxide mass containing grains
of almost pure nickel is often found. Fortunately, subsurface oxidation is rare.
Within the Company, it has only been observed in very high temperature furnaces,
such as steam-methane reformers and ethylene plants, and then only when run hotter
than normal and, presumably, with low excess oxygen.

Vanadium Pentoxide Corrosion


When vanadium-containing fuels are burned in a furnace firebox, ash deposits
containing vanadium pentoxide (V2O5) form. If metal temperatures are high
enough, the deposits melt and the resulting slag dissolves the protective oxide film
from the metal, promoting rapid oxidation. This type of corrosion leaves rough,
irregular surfaces without oxide scale. Corrosion commonly occurs on tube hangers
rather than tubes because, normally, tube wall temperatures arent hot enough to
melt the ash deposits.

Oxidation of Iron
Commercially pure iron follows the Parabolic Rate Law. Three oxides are formed:
FeO (wustite), Fe3O4 (magnetite), and Fe2O3 (hematite). These oxides can coexist
over a wide temperature range in varying proportions. At comparatively low
temperatures only magnetite and hematite are stable, and the scale is usually

Chevron Corporation 400-43 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

predominately magnetite. Wustite is stable in air at temperatures above about


1050F, and exposure above this temperature produces a three-layer scale, most of
which is wustite. Figure 400-30 illustrates oxides formed on iron at temperatures

Fig. 400-30 Effect of Temperature on Oxides Formed on Iron. Oxide layers formed at temper-
atures 1050F or higher (left), and below 1050F (right). After Schreir [3]. (Cour-
tesy of Butterworth Heinemann Publishers)

above and below 1050F. Kubaschewski and Hopkins [4] believe that wustite grows
primarily by iron cation (Fe++) diffusion, hematite by oxygen anion (O ) diffu-
sion, and magnetite largely by anion diffusion but also by minor cation diffusion. A
magnetite oxide layer on the inner surface of a boiler steam generating tube is
shown in Figure 400-31.

433 Effects of Alloying Elements in Steels


Among the elements intentionally added to steel or present as residuals, carbon,
manganese, sulfur, and phosphorus do not help the steel resist high-temperature
oxidation. Chromium, silicon, titanium, columbium, and aluminum markedly
improve resistance to high-temperature oxidation. The roles of molybdenum and
nickel are controversial.

Chromium
Chromium is the most important element added to steel to impart oxidation resis-
tance. Chromium creates this resistance by formation of a chromium-rich oxide that
slows down oxygen diffusion and hence protects the metal underneath from attack.
Chromium oxidizes preferentially to iron, and when sufficient chromium is present,
the protective scale is essentially chromic oxide.

January 2001 400-44 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Although even small additions of chromium are beneficial, larger amounts allow a
higher temperature range. The temperature at which an alloy may be used safely
without excessive scaling increases as chromium content increases. Maximum
effect is achieved at about 27% as illustrated in Figure 400-32. The use of chro-
mium has made possible a whole series of alloy steels with a wide range of oxida-
tion resistance.

Fig. 400-31 Oxide on Inner Surface of Carbon Steel in Fig. 400-32 Effect of Chromium in Steel on Maximum
Air at 1290F to 1300F. The oxide is prima- Service Temperature Without Excessive
rily Fe3O4, magnetite. Scaling [5] and on Metal Loss by Oxidation
in Air at 1800F [6]. (Data courtesy of U.S.
Steel & Babcock & Wilcox)

Silicon
Silicon is even more effective than chromium in reducing oxidation of steels.
Figure 400-33 compares the effects of silicon and chromium on the high-
temperature oxidation of steel in air.
Unfortunately, silicon adversely affects the mechanical properties of steels, making
it brittle at room temperature and reducing creep strength at high temperatures.
Therefore, silicon additions are limited to a maximum of about 3% in high-alloy
heat-resisting castings and about 1% in low-alloy wrought materials. Even with
this limitation, the effects of silicon on oxidation are appreciable. The addition of
1% of silicon to a standard 5 Cr-1 Mo steel raises its safe scaling temperature
from about 1200F to 1300F.

Chevron Corporation 400-45 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-33 Effects of Silicon and Chromium on the Oxidation of Steel in Air at 1290F to
1300F.

For reasons which are not completely clear, silicon does not have the same benefi-
cial effect in steam as it does in air. In both 1 Cr- Mo and 5 Cr- Mo steels,
silicon actually reduces oxidation resistance in steam atmospheres.

Aluminum
When added to steel, aluminum promotes formation of tightly adhering scales and
markedly improves resistance to scaling. However, it is difficult to add appreciable
quantities of aluminum to steel without producing certain undesirable effects, partic-
ularly a loss of toughness and ductility.

Molybdenum
Rahmel et al. report that molybdenum steels with 2% molybdenum markedly reduce
the oxidation rate of iron in the range of 900F to 1800F. Other investigators [5]
have found similar effects: a 5% molybdenum steel oxidized at about one-fifth the
rate of a plain carbon steel at 1100F. As mentioned earlier, molybdenum can cause
catastrophic oxidation under certain circumstances; however, this has been prima-
rily associated with the austenitic stainless steels. Although molybdenum is
frequently present in high-temperature alloys, its chief function is to impart creep
resistance, rather than oxidation resistance.

Nickel
Nickel has little or no direct effect on the oxidation of iron. Nickel steels with 35%
nickel oxidize at about the same rate as carbon steel. At the 8-10% Ni level of auste-

January 2001 400-46 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

nitic stainless steels, the nickel does not noticeably improve scaling resistance under
conditions of constant temperature. If temperature cycling exists, nickel is defi-
nitely beneficial because it minimizes spalling of the scale. It accomplishes this by
reducing differential thermal expansion between metal and oxide, thereby reducing
stress at the interface.

434 Oxidation of Carbon and Low-alloy Steels


Detailed studies conducted by the Company have shown that the high-temperature
oxidation of carbon, carbon-moly, and the common austenitic stainless steels
follows a parabolic rather than a linear rate law. Equation 400-4 expresses the Para-
bolic Rate Law in terms of weight gain as a function of time. So long as the scale
remains on the metal, weight is gained during the oxidation reaction because the
metal oxide weighs more than the metal consumed in its formation. However, the
Parabolic Rate Law can also be written in terms of weight loss (the amount of metal
converted to metal oxide). The equations are the same; only the numerical value of
the constant Kp changes. Therefore, instead of writing the law as:

W2 = K p t
(Eq. 400-4)

the law can be written as:

P2 = Kt
(Eq. 400-6)
Then, if R equals corrosion rate (in mils per year),

R = P/t = (K/t)1/2 = K1/(t)1/2


(Eq. 400-7)
where:
P = penetration, in mils
t = time, in hours
K, K1 = new constants, proportional to Kp but numerically different
In other words, the corrosion rate of carbon steel and the common alloy steels is
inversely proportional to the square root of time, and the oxidation rate slows down
as time passes.
One result of this phenomenon is that published oxidation rate data, usually based
on 100- or 1000-hour tests, show corrosion rates that are unrealistically high for
long-term applications. The Parabolic Rate Law gives us a technique for converting
such data into long-term rate predictions.

Chevron Corporation 400-47 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Oxidation rates decrease with time as the growing oxide scale increasingly prevents
oxygen diffusion. In practice, however, the oxide scale cannot continue to grow
indefinitely. At intervals it may spall off owing to temperature cycles, plant startups
or shutdowns, or the high stresses which develop in the oxide scale as it becomes
thicker. When this happens, fresh metal is exposed and the oxidation rate again
increases to the high initial rate experienced by a clean metal surface. This rate will
slowly decrease with time as the oxide scale redevelops. Over time, the metal may
experience cycles of scale growth and spalling, as shown in Figure 400-34. The
average long-term corrosion rate will then be the value determined by the Parabolic
Rate Law, in which the time, t, is the length of one growth-spall cycle.

Fig. 400-34 Oxidation by Parabolic Rate Law, With and Without Periodic Spalling of the
Oxide Scale

January 2001 400-48 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Predicting Oxidation Rates of Carbon and Low-alloy Steels


To predict long-term oxidation rates one must assume a value for the frequency at
which the oxide scale is likely to be disrupted. At present this determination is an art
rather than a science, but we are learning by experience. Initially, the practice has
been to assume a period of one year. This was based on the logic that scale spalling
would most likely occur due to the temperature and environmental changes during
plant startups and shutdowns, and that one year was about the minimum shutdown
frequency for most process plants. It was thought that this would give a reasonable,
conservative basis for oxidation rate predictions. Experience indicates that a one-
year time factor gives accurate-to-slightly-conservative rate predictions, and this is
recommended for most materials selection purposes. Rates based on a one-year time
factor are shown in Figure 400-35.

Fig. 400-35 Effect of Temperature on Oxidation Rate of Steels in Air

Chevron Corporation 400-49 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Predicting Rates Under Cyclic Conditions. The Parabolic Rate Law also allows
prediction of oxidation rates under cyclic conditions, by assuming that the scale
spalls off every temperature cycle. This assumption may be valid if the oxide scale
builds up to a reasonable thickness between cycles. However, very thin scale layers
dont develop a high internal stress, and they resist spalling during heating and
cooling cycles. A limited amount of experimental evidence, developed on super-
heaters of gas turbine-powered ships, indicates that oxide scale resists spalling, even
under severe thermal cycling, if less than about 20 mils (0.020 inch) thick. This is
equivalent to about 5 mils metal penetration.
Maximum Temperatures of Steels in Oxidizing Atmospheres. Low-alloy steels,
including alloys with up to 9% chromium, are suitable for temperatures up to about
1300F. Carbon steel can be used up to temperatures of 1050F. Generally, as
Cr content increases the maximum temperature for use increases. The limit for each
alloy is controlled by the alloy content and by conditions of the particular applica-
tion. Chromium, silicon, aluminum, and titanium are the alloying elements of major
importance. Changes in oxidation characteristics brought about by each as single
additions have been discussed in Section 433. However, these elements are usually
present in combinations, and the effects may be compounded. Figure 400-36 gives
the maximum temperature limit for long life under oxidizing conditions for some of
the more common alloys. However, these limits apply to oxidation resistance only.
Factors such as embrittlement or lack of high-temperature strength may limit the
usefulness of these alloys to temperatures lower than the tabulated values.
Much of the available oxidation data on steels has been gathered in short-time labo-
ratory tests in pure air. Industrial atmospheres are much more complex, and the
effects of the constituent gases are sometimes subtle. As a result, observed oxida-
tion behavior in service frequently deviates from what might be predicted in
Figure 400-36.

Fig. 400-36 Maximum Temperatures for Use of Steels in Air [5], [6] (Data courtesy of U.S.
Steel & Babcock & Wilcox)
Alloy Temperature, F
Carbon Steel 1050
C - Mo 1050
1 Cr - Mo 1085
1 Cr - Mo 1100
2 Cr - 1 Mo 1175
3 Cr - 1 Mo 1175
5 Cr - Mo 1200
5 Cr - Mo - 1Si 1300
7 Cr - Mo 1250
9 Cr - 1 Mo 1300
Type 410 (12 Cr) 1400
Note that these limits apply to oxidation limits only. Factors such as embrittlement or lack of high temperature
strength may limit the usefulness of these alloys to temperatures lower than the tabulated values.

January 2001 400-50 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-36 Maximum Temperatures for Use of Steels in Air [5], [6] (Data courtesy of U.S.
Steel & Babcock & Wilcox)
Alloy Temperature, F
Type 405 (12 Cr - Al) 1400
Type 430 (17 Cr) 1525
Type 446 (27 Cr) 2050
Type 304 (18 Cr - 8 Ni) 1600
Type 316 (18 Cr - 8 Ni - 3 Mo) 1600
Type 321 (18 Cr - 8 Ni - Ti) 1600
Type 347 (18 Cr - 8 Ni 1 Cb) 1600
Type 309 (25 Cr - 12 Ni) 1900
Type 310 (25 Cr - 20 Ni) 2050
Note that these limits apply to oxidation limits only. Factors such as embrittlement or lack of high temperature
strength may limit the usefulness of these alloys to temperatures lower than the tabulated values.

Oxidation Rates for FCC Regenerators. Experience has shown that the oxidation
of carbon steel and low alloys in FCC regenerators follows the Parabolic Rate Law.
When using promoted catalyst, oxygen levels in regenerators are comparatively
high (several percent), and oxidation rates are essentially the same as in air at the
same temperature. Rates can be predicted with good accuracy by using the one-year
time factor.
Oxidation Rates in Flue Gas Environments. In many applications, such as power
plant boilers and process unit heaters, the tubes, tube supports and other firebox
hardware are exposed to flue gases rather than to air or oxygen. Flue gas is a
mixture of combustion products and usually consists of CO, CO2, O2, H2 and water
vapor, in addition to N2. Depending on whether the unit is fired with an excess or
deficiency of air, the flue gas may either be oxidizing because it contains free
oxygen or reducing by virtue of the carbon monoxide present. Because most fuels
also contain some sulfur, the combustion gases are likely to contain SO2, SO3, and
perhaps H2S. Oxidation rates of carbon and low-alloy steels are appreciably higher
when sulfurous gases are present in the flue gas, particularly if CO is also present. A
quarter of one percent SO2 is sufficient to double the oxidation rate of carbon steel
in flue gases at 1200F. Additions of chromium improve resistance to attack by
SO2-bearing flue gas more or less in proportion to the amount added. At the
9% chromium level, rates may be as little as one-tenth that of plain carbon steel.
Oxidation Rates in Steam Environments. Also of practical interest is the oxida-
tion of steels by steam. Generating tubes in a boiler must resist not only the attack of
flue gases on their exterior (fired) side, but also that of steam on the inner surfaces.
While some investigators have shown that there are differences between steam and
air oxidation rates, these differences are not of great practical significance. It is
generally assumed that oxidation rates in steam and air are the same.

Chevron Corporation 400-51 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

435 Oxidation of High-alloy Steels


The basic guidelines for producing oxidation resistance in highly alloyed steels were
worked out at least 50 years ago. Modern heat-resistant alloys are designed on these
general precepts to give the optimum combination of scaling resistance, high-
temperature strength, cost, and ease of fabrication. Most standard heat-resistant
steels rely on chromium for scaling resistance, supplemented in certain instances by
silicon. Nickel improves the adherence of the scale under cyclic temperature condi-
tions, but high-nickel alloys do poorly in atmospheres having a high sulfur content.
Figure 400-36 shows the maximum temperatures at which wrought high-alloy steels
may be used without excessive scaling in oxidizing atmospheres. Generally, these
limits also apply to cast alloys, although the cast equivalent of a wrought alloy may
resist oxidation slightly better because of the higher silicon content.
Common cast high alloy austenitic stainless steels include HK-40, which is the cast
equivalent of the wrought alloy 310. This alloy is a common material for furnace
tubes, where high temperature oxidation resistance is important. Other cast high
alloy stainless steelsthe HP-modified alloys, such as Manurite 36Xhave even
better oxidation resistance then HK-40.

436 Oxidation of Other Metals


Nickel and Nickel Alloys
Compared to metals such as iron and copper, nickel resists oxidation at high temper-
atures well. However, it is inferior to chromium. For example, in oxygen at 1300F,
nickel may scale at a rate 1/50 that of iron, but twice that of chromium. The only
oxide formed is NiO, which forms a closely adhering and protective film that
follows the Parabolic Rate Law. The resistance of nickel to oxidation may be modi-
fied considerably by alloying.
The alloying element with the greatest influence on the oxidation behavior of nickel
is chromium. Additions up to about 7% markedly increase the rate of oxidation. As
increasing amounts of chromium are added above 7%, the oxidation rate decreases
and above about 10% chromium, the oxidation resistance of nickel is substantially
improved. Figure 400-37 shows how chromium content affects the oxidation rate of
nickel-chromium alloys at 2000F.
High-nickel, low-chromium alloys such as Inconel 600 show increased suscepti-
bility to subsurface oxidation, as described in Section 432. Such alloys are also
sensitive to attack in atmospheres containing sulfur compounds, such as flue gas.
One of the more complex nickel base alloys, Hastelloy X, has found limited refinery
application in protective sheaths for furnace tube skin temperature thermocouples
and supports for platinum gauze in nitric acid manufacture. In addition to about
21% chromium, it also contains approximately 19% iron, 9% molybdenum, 2%
cobalt, and 1% tungsten. Hastelloy X has excellent high-temperature mechanical
properties, and it also forms a tight, spall-resistant scale that makes it particularly
suitable for oxidizing atmospheres where temperature cycling is expected. Its

January 2001 400-52 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

performance at 2000F is compared to that of Type 309 stainless steel and a 75-15
Ni-Cr alloy in Figure 400-38.

Fig. 400-37 Effect of Chromium Content on Oxidation Fig. 400-38 Oxidation of Several Alloys Containing
Rate of Nickel-Chromium Alloys. Exposure Nickel. Intermittent exposure in dry air at
at 2000F. Calculated from Zima [7]. 2000F.

Copper and Copper Alloys


Although copper and copper alloys are ordinarily not selected for applications
requiring high-temperature oxidation resistance, they are mentioned briefly because
of their widespread use in other applications. Pure copper follows the Parabolic Rate
Law at a rate somewhat higher than iron at any given temperature. The limiting
maximum temperature for extended exposure without excessive scaling is about
200F lower than iron (850F versus 1050F). The oxide formed in air is primarily
cuprous oxide, Cu2O. Copper-zinc alloys and copper-aluminum alloys are more
resistant to oxidation than pure copper. Dunn [8] found that small additions of
silicon, nickel, or arsenic had little or no effect on oxidation resistance.

437 Oxidation-resistant Coatings


The oxidation resistance of carbon steel can be greatly improved by application of
diffusion coatings of aluminum, chromium, or silicon. These are not coatings in the
usual sense, but rather are high-alloy surface layers obtained by diffusion of the
alloying element.
Calorizing is a trade name for a comparatively old process for alloying the surface
of carbon steel with aluminum. Aluminum is applied to the part to be coated either
by spraying or by packing a retort with powdered aluminum and ammonium chlo-
ride. Subsequent heating for 12 to 48 hours at 1600F diffuses the aluminum into

Chevron Corporation 400-53 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

the steel to a depth of about 0.015 to 0.025 inch. A similar process frequently
mentioned in industry publications is called alonizing, a proprietary name of Alon
Processing, Inc., U.S.A. The Alonizing process diffuses aluminum into steel to a
depth of 0.010 to 0.040 inch by heating the part in a retort containing a mixture of
aluminum powders. Parts of almost any shape tubing, piping, castings, or struc-
tural shapes can be alonized on one or more surfaces provided they are of a size
that fits into a retort. The surface layer produced is an aluminum-iron alloy that
forms a protective aluminum oxide film at high temperatures. Carbon steel thus
protected can be operated satisfactorily in an oxidizing atmosphere as high as
1700F.
Diffusion of the aluminum does not improve the high-temperature strength of the
steel. If high-temperature strength is needed as well as oxidation resistance, the
calorizing or alonizing processes can be applied to the low-alloy chromium-
molybdenum steels.
Also, the aluminum-iron alloy formed by diffusion of the aluminum into the steel is
relatively brittle. This is significant when the cross-section of the base metal is
small, as in wire screens; it limits the minimum size of wire that can be aluminized
and the depth of diffusion permissible on small wires. On heat exchanger tubes, the
brittleness of the aluminum layer may result in its cracking during tube rolling. As a
result, exchanger tube ends are often masked off during the aluminizing, so that the
section to be rolled is not aluminized.
Steels with high-aluminum surface layers have good resistance to attack by SO2,
SO3, and H2S in high-temperature environments above the dew point of the gas. In
the early days of catalytic reforming, calorized piping was used in considerable
quantity in process units to resist high-temperature H 2S attack. While some installa-
tions proved satisfactory, others did not, principally because the diffused layer was
not pore-free, and localized attack occurred. For this reason, this material has not
been widely used in refinery applications in recent years. Improved processing tech-
niques may significantly reduce porosity, and increase the use of this material in the
future.
A common application of alonized stainless steel is in the catalyst support screens
for VGO, VRDS, and isocracker first stage reactors. Austenitic stainless steels are
subject to corrosion and scaling under the conditions encountered in these reactors
(850F, 4050 psi hydrogen sulfide partial pressure, 3,000 psi total pressure). While
the corrosion rate of a Type 347 stainless steel under these conditions is not exces-
sive for most equipment, the sulfide scale formed occupies approximately four
times the volume of the metal lost by corrosion and would plug the screens in a rela-
tively short time. In pilot plant tests, the corrosion rate of alonized Type 347 steel
was negligible and no scaling was observed. On the basis of these tests, this mate-
rial was selected for reactor screens and has performed well for many years.
Because of the small diameter of the screen wire and the brittle nature of the alumi-
nized layer, it is necessary to limit the depth of the aluminized layer to 0.003 to
0.006 inch. Please refer to Specification EG-2663 for details.
Not all diffusion coatings are intended to improve oxidation resistance. Some, like
nitriding and the proprietary TMT-5 (Turbine Metals Technology, Inc.), are prima-

January 2001 400-54 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

rily wear-resistant materials. Nitriding, in fact, reduces oxidation resistance by tying


up chromium in the alloy as chromium nitride, contributing nothing to oxidation
resistance.

440 High-temperature Hydrogen Attack


High-temperature hydrogen attack can degrade carbon and low alloy steels oper-
ating in high-temperature hydrogen environments. Hydroprocessors, hydrotreaters,
naphtha hydrotreaters, catalytic reformers, and hydrogen manufacturing plants, for
example, are all exposed to conditions promoting hydrogen attack. These condi-
tions are not found in upstream operations, and this phenomenon is primarily a
downstream concern.
Proper design, alloy selection, and operation can generally prevent hydrogen attack
in process equipment. Figure 400-39 covers high-temperature hydrogen attack and
compares it to other hydrogen damage mechanisms.

Chevron Corporation 400-55 January 2001


Fig. 400-39 Summary of Hydrogen Damage Mechanisms. The definitions listed in this table are the preferred Chevron definitions. Some other definitions are
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


used throughout the industry. (1 of 2)
Temperature Related Hydrogen Damage Mechanisms in Steel Plate and Piping
Type Mechanism Factors Appearance & Detection Source of Hydrogen Control Methods

High-Temperature Hydrogen H diffuses into steel, reacts Occurs in carbon and C-1/2 Detected with UT or at high At high temperature and high Controlled by upgrading
Attack (Section 440) with carbon, forms methane, Mo steels above 450F and magnification under a micro- pressure, hydrogen has an material as specified in API
and creates an internal pres- 100 psig H2. Cr-Mo (low alloy) scope prior to failure. equilibrium between H2 and 941. Cr and Mo additions to
sure which causes cracks steels have greater resis- H. As H is formed some steel increase resistance.
and fissures. tance. diffuses into the steel.

Defect Related Hydrogen Damage Mechanisms in Steel Plate


Type Mechanism Factors Appearance & Detection Source of Hydrogen Control Methods

Hydrogen Blistering H formed during corrosion The presence and density of Forms circular blisters on ID H forms on surface by Can be eliminated in H2S
(Section 451). diffuses to internal disconti- non-metallic plate-like inclu- of vessels. Detected inter- corrosion, in the presence of service by coating with
nuities (non-metallic inclu- sions. nally with visual inspection hydrogen recombination stainless steel. Performance
sions), recombines to H2, or externally by UT thick- poison (H2S or HF acid), and can be improved by using
creates internal pressure, ness gauge. diffuses into steel. It is low sulfur steel or corrosion
400-56

and forms a blister. accelerated by cyanides or controls (water wash or


ammonia. APS).

Hydrogen Induced Cracking An advanced degree of The presence and density of Surface breaking cracks at Same as hydrogen blisters. Same as hydrogen blisters.
(HIC) or Stepwise Cracking hydrogen blistering, in which non-metallic inclusions and center or edge of blisters,
(Section 452). pressure in blisters causes blisters. can be detected with WFMT
cracks to grow in the or visually, if severe. Subsur-
through wall direction face cracks can be detected

400 Characteristic Corrosion Phenomena


between internal discontinu- with shear wave UT.
ities.

Stress Oriented Hydrogen Similar to HIC, but weld Residual stresses in the HAZ Surface breaking cracks in Same as hydrogen blisters. Same as hydrogen blisters,
Induced Cracking (SOHIC) residual stresses cause and the presence of non- the weld HAZ, can be plus stress relief will improve
(Section 452). crack growth in the through metallic inclusions. detected with WFMT. resistance.
wall direction between
internal discontinuities.

Note H2 - Molecular Hydrogen, H - Atomic Hydrogen, UT - Ultrasonic Testing, WFMT - Wet Fluorescent Magnetic Particle Testing,
HAZ - Heat Affected Zone, Rc - Rockwell C Hardness, BHN - Brinell Hardness
January 2001
Fig. 400-39 Summary of Hydrogen Damage Mechanisms. The definitions listed in this table are the preferred Chevron definitions. Some other definitions are
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


used throughout the industry. (2 of 2)
Property Related Hydrogen Damage Mechanisms Plate and Pipe
Type Mechanism Factors Appearance & Detection Source of Hydrogen Control Methods

Sulfide Stress Cracking H diffuses into a region of High strength & hardness Surface-breaking cracks, H forms on surface by Avoid the use of overly
(SSC) (Section 452). high stress (crack tip or (>Rc 22 in HAZ or >BHN 200 often in the weld or HAZ. Can corrosion, in the presence of strong or hard materials.
internal flaw) in hard steel for the weld metal for steel). be detected with internal H2S, and diffuses into steel. Upgrade to materials listed
and causes crack growth. WFMT. Can also crack steel in NACE MR-01-75. PWHT
Cracking can occur quickly. machinery. improves weldment perfor-
mance.

Hydrogen Stress Cracking Same as sulfide stress Same as sulfide stress Same as sulfide stress H forms on surface by Same as sulfide stress
(HSC) cracking. cracking. cracking. corrosion, in the presence of cracking. In HF service
(Section 452). hydrogen recombination Monel is a common upgrade,
poison other than H2S (HF but is susceptible to
acid), and diffuses into steel. cracking.

Hydrogen Embrittlement or H which enters at high High strength and high hard- No change in appearance H can come from any Avoid high stresses and
400-57

Hydrogen Assisted Cracking temperatures and high ness steels are affected. The until failure. Mechanical source including electro- stress concentrations while
(Section 460). hydrogen partial pressures, degradation of the proper- property tests while the plating, corrosion, or high material is hydrogen
from corrosion, or from any ties is transitory. The normal material is hydrogen temperature hydrogen and charged. Bake out hydrogen
source reduces the mate- properties return if the charged are needed for rapid cooling. at 600F to restore proper-
rial's ductility and resis- hydrogen diffuses out. detection. ties.
tance to crack growth.

Hydrogen Cracking of Welds H diffuses to a region of High strength & hardness Cracking will occur at the H diffuses to the HAZ from Apply proper preheat. Clean

400 Characteristic Corrosion Phenomena


(Delayed Cracking, Cold high stress (crack tip or (>Rc 22 in HAZ or >BHN 200 weld toe after the weld has the base metal (if the vessel and degrease basemetals
Cracking, or Underbead internal flaw) and causes for the weld metal). cooled, up to 48 hours after has been in corrosive before welding. Bake out
Cracking) (Welding Manual crack growth in the HAZ. welding. Can be found using service and not baked-out metal before welding. Use
Section 140). WFMT. prior to welding) or during clean, dry low hydrogen
welding. electrodes.

Note H2 - Molecular Hydrogen, H - Atomic Hydrogen, UT - Ultrasonic Testing, WFMT - Wet Fluorescent Magnetic Particle Testing,
HAZ - Heat Affected Zone, Rc - Rockwell C Hardness, BHN - Brinell Hardness
January 2001
400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

441 Mechanism of Hydrogen Attack


Hydrogen attack requires the presence of atomic hydrogen. Atomic hydrogen is
normally unstable at low temperatures, and two hydrogen atoms will combine spon-
taneously to form a hydrogen molecule. At high temperatures, however, molecular
hydrogen exhibits a definite dissociation into hydrogen atoms according to the
following reaction:

H2 2Ho
(Eq. 400-8)
where:

2 24 ,000
--------------------
( PH ) T
-------------- = e
PH2

and
P H = partial pressure of atomic hydrogen
P H 2 = partial pressure of molecular hydrogen
T = temperature, K
This is an equilibrium reaction, dependent only on temperature. In other words, at a
given temperature a fixed percentage of the hydrogen will exist in the atomic state.
As the temperature is changed, the percentage of hydrogen will vary. In a hot
hydrogen environment, atomic hydrogen always exists and will diffuse into and
through the walls of the container regardless of the metal.
As in low temperature blistering (see Section 460), atomic hydrogen tends to
combine into the molecular form within voids or discontinuities in the metal.
However, this cannot cause appreciable damage at high temperatures, since the
hydrogen pressure within the void cannot exceed the hydrogen pressure in the bulk
fluid, owing to the equilibrium between atomic and molecular hydrogen. Hydrogen
attack damage, therefore, results from other reactions.

Methane Fissuring and Internal Decarburization


At temperatures above about 450F and hydrogen partial pressures above 100 psi,
atomic hydrogen can chemically react with the carbon compounds in steel to form
methane gas according to the reaction:

Fe3C + 2H2 3Fe + CH4


(Eq. 400-9)
(Hydrogen can also react with impurities such as oxides or sulfides, but these reac-
tions are of less importance.)

January 2001 400-58 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Since carbon acts as the major strengthening agent in steel, the removal of carbon
(decarburization) by the reaction with atomic hydrogen causes a loss of strength.
The methane formed by this reaction cannot diffuse out of the steel, but is trapped
inside.
As attack continues and more methane is formed, high internal pressures are devel-
oped within the steel, which causes small bubbles at grain boundaries and nonme-
tallic inclusions. As the attack progresses, the bubbles link up to form fissures.
These fissures expand progressively due to the high internal gas pressure. This
formation of methane bubbles and fissures and/or the decarburization of the steel is
called high-temperature hydrogen attack.
The metallurgical structure of carbon steel following severe attack is shown in
Figure 400-40.

Fig. 400-40 Severe Fissuring and Decarburization of Carbon Steel From High-temperature Hydrogen Attack.
In the right photograph, notice the progressive change in carbon content, varying from the normal
structure at the top to the completely decarburized structure near the fissures.

In addition to bubbles and fissures that occur within the steel and cannot be seen on
the surface, surface blisters also may be found. Blisters can be seen on the surface of
the steel. These blisters can form at laminations in the steel or by the joining of a
number of separate fissures. They differ from low temperature blisters in that they
contain methane rather than hydrogen and behave in a ductile rather than brittle
manner. The low temperature blisters usually crack when they rise to a height of
perhaps 10% of the blister diameter. High temperature blisters from methane forma-
tion seldom crack, even though they often attain an almost hemispherical shape.

Chevron Corporation 400-59 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

High-temperature hydrogen attack also causes the metal to lose ductility. This does
not represent any inherent embrittlement of the steel. It is due to the stresses being
localized by the methane fissures. Actually, the steel remains ductile but failures
appear brittle because of the short fracture path between fissures.
Hydrogen attack also reduces Charpy impact toughness, and therefore brittle frac-
ture resistance.

Surface Decarburization
When carbon steel is exposed to high-temperature, low-pressure hydrogen (above
1050F and below 200 psi) surface decarburization occurs, with a resultant loss in
strength and an increase in ductility. Most likely this is caused by a continuous
migration of carbon to the surface, where it is removed as methane (or carbon
monoxide if carbon dioxide is present.)

Effect of Temperature and Pressure


Whether steel deteriorates by surface decarburization or by methane fissuring and
internal decarburization depends on several variables, primarily hydrogen tempera-
ture and pressure, and alloy content of the steel. At relatively high temperatures and
low pressures, surface decarburization proceeds faster than internal attack, and a
steel may become completely decarburized before methane fissuring occurs. At
relatively high pressures and low temperatures, on the other hand, carbon mobility
is decreased, and internal attack may occur without a significant amount of surface
decarburization. When both pressure and temperature are sufficiently high, both
phenomena can occur simultaneously [9].

442 Prevention of High-temperature Hydrogen Attack


In ordinary carbon steel, carbon takes the form of iron carbides. When hydrogen
diffuses into the steel at high temperature, these carbides are chemically reduced to
form metallic iron and methane. Therefore, hydrogen attack can be prevented or at
least reduced by adding an alloying element to the steel that preferentially combines
with the carbon to form a more stable carbide. The important carbide-stabilizing
elements are chromium, molybdenum, columbium, titanium, vanadium and tung-
sten. Elements such as nickel and silicon that do not form stable carbides have no
effect on hydrogen attack resistance.

Nelson Curves
The alloys most frequently used commercially to resist hydrogen attack include the
chromium-molybdenum low-alloy steels and the austenitic stainless steels. The
limits for hydrogen attack resistance of carbon steel and certain low-alloy steels are
shown in Figure 1 of API 941. Hydrogen attack can be expected for steel exposed to
temperature-hydrogen partial pressure conditions above and to the right of its repre-
sentative curve.
Figure 1 of API 941 was developed by G.A. Nelson of Shell Development Company
using the service experience of many companies including Chevron, and is often
referred to as the Nelson curves. The Nelson curves are now sponsored by the

January 2001 400-60 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

American Petroleum Institute, and issued as API 941, Steels for Hydrogen Service
at Elevated Temperatures and Pressures in Petroleum Refineries and Petrochemical
Plants, a copy of which is included in this manual. They are widely used and
accepted as providing the most reliable information on the resistance of iron-base
alloys to high-temperature hydrogen attack.
For equipment operating in high-temperature hydrogen service, Chevron requires
materials selections to be based on a safety margin of 50F and 50 psi below the
Nelson curve for the particular steel. Austenitic alloys such as 18% chromium 8%
nickel stainless steels are not shown on Nelson curves; they are considered immune
to high temperature hydrogen attack under all conditions.

Clad Steel
Use of adequately resistant alloys is the only effective way to prevent high-tempera-
ture hydrogen attack. A resistant alloy lining or cladding does not necessarily
protect the backing material from attack. Atomic hydrogen readily diffuses through
a lining and attacks susceptible backing material deficient in alloy content.
However, alloy cladding reduces the concentration of atomic hydrogen in the
backing material by lowering the effective hydrogen partial pressure at the cladding-
backing interface. Therefore, strictly speaking, Nelson curves represent the resis-
tance to hydrogen attack of unclad material only. Theoretically the curves shift to
higher temperatures and higher hydrogen partial pressures for clad materials [10].
Although stainless steel cladding is commonly used to resist corrosion in equip-
ment that is also subject to hydrogen attack, current Company practice is to disre-
gard the potentially beneficial effects of alloy cladding in the design and operation
of equipment to resist hydrogen attack. We use backing materials that have adequate
hydrogen attack resistance with or without cladding. However, we sometimes take
advantage of the beneficial effect of stainless cladding in determining inspection
requirements for specific pieces of equipment.

Carbon - 0.5% Molybdenum Steel


From the time they were first issued by API in 1970, the Nelson curves have been
periodically updated with new data. For example, the API 941 (April 1990) does not
include carbon-% molybdenum steel in Figure 1. The curve has been moved to a
separate figure to show the industry reported incidences of hydrogen attack of
C- Mo operating below the Nelson curve. These reports indicate that some heats
of C- Mo steel have no better hydrogen attack resistance than carbon steel. For
new or replacement equipment, the Company no longer recommends C- Mo steel
to prevent hydrogen attack, though it may be used for other reasons. Rather, where
you need better hydrogen attack resistance than carbon steel can provide,
1% chromium - % molybdenum steel is recommended as the next step up.
Because a number of companies including Chevron have experienced failures by
hydrogen attack of C- Mo equipment operating below the 1983 C- Mo Nelson
curve, Chevron has developed a policy to deal with existing C- Mo equipment in
high temperature hydrogen service. The report Recommendations for Hydrogen
Attack Inspection of Carbon- Moly Equipment was issued in 1992 and revised in
1993. It gives recommended inspection procedures, intervals, and priorities for

Chevron Corporation 400-61 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

specific equipment items and piping lines. Check with CRTC Materials and Equip-
ment Engineering for further information.

443 Incubation Period for Hydrogen Attack


Damage to steels by high-temperature, high-pressure hydrogen is preceded by a
period where no noticeable change in properties can be detected. After this period
has elapsed, decarburization and/or methane bubble formation and fissuring will
occur. The length of time before attack begins has been termed the incubation
period and varies with temperature, pressure, and alloy content of the steel.
The length of the incubation period is of considerable practical importance, because
damage is reversible only during this period, at which time the steel can be
considered to maintain its initial mechanical properties. After this period has
elapsed, permanent degradation proceeds. Incubation times have been established
and are given in chart form in API Publication 941. It must be emphasized that these
curves serve only as a guide for determining approximate safe operating times for
steels exposed to conditions above their prescribed limits. Materials selection should
never be based on these incubation curves. Company policy is to use incubation
data only for evaluating the condition of steels exposed to high temperature or pres-
sure conditions during operational excursions.

450 Hydrogen Damage Due to Wet H2S (Sour Service) or HF Acid


The forms of hydrogen damage covered in this section are hydrogen blistering and
three forms of cracking:
Sulfide Stress Cracking (SSC)
Hydrogen Induced Cracking (HIC)
Stress Oriented Hydrogen Induced Cracking (SOHIC)
All of these damage mechanisms can be caused by hydrogen charging due to expo-
sure to wet H2S conditions. Because HF acid also causes hydrogen charging and
similar damage, this discussion and the subsequent sections also apply to HF acid
services. Figure 400-39 in Section 441 covers the hydrogen damage which occurs in
wet H2S environments and compares it to other hydrogen damage mechanisms.

Scope of Wet H2S Environments and Damage


Since 1989, Chevron has performed over 1100 inspections for wet H2S cracking and
blistering. Wet H2S cracks deeper than 3/8" or over 50% through-wall have been
found in forty-eight vessels. Over 140 vessels have required repair or replacement
due to wet H2S cracking and/or fabrication defects.
Hydrogen blistering and cracking are common to refinery and upstream processing
plant equipment that contains >50 ppm H2S in water, between ambient tempera-
tures and 300F. Longitudinally or spiral welded pipe is also susceptible in these
environments. Neither seamless pipe, forgings, nor castings usually crack or blister
in wet H2S, provided hardness controls are maintained on weldments.

January 2001 400-62 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Mechanisms of Hydrogen Charging in Wet H2S


Hydrogen damage in wet H2S services is caused by the generation of atomic
hydrogen as a byproduct of the corrosion reaction, and the subsequent diffusion of
that atomic hydrogen into the steel. Atomic hydrogen is produced in many corro-
sive environments as follows:

Fe + 2H+ >> Fe++ + 2H


(Acidic Environments)
(Eq. 400-10)

Fe + 2OH- >> FeO2= + 2H


(Alkaline Environments)
(Eq. 400-11)
In either case, hydrogen is first formed in the atomic state (Ho), then two atoms
subsequently combine to form a molecule of hydrogen gas:
2Ho >> H2 (FAST!)
However, certain compounds such as sulfide, cyanides (e.g. HCN), and arsenates,
called recombination poisons or catalyst poisons, retard the conversion of atomic
hydrogen to molecular hydrogen. In the presence of a catalyst poison, the surface
concentration of atomic hydrogen rises, and a corresponding increase occurs in the
amount of hydrogen diffusing into the metal. Atomic hydrogen can diffuse through
solid steel at rates of several cubic centimeters per square centimeter per day.
This elevated concentration of atomic hydrogen can affect the steel in several ways:
1. At laminations or inclusions, the hydrogen atoms may recombine to form
molecular hydrogen, which is then too large to diffuse further through the steel
and is trapped. If laminations are large enough, the internal hydrogen pressure
may become sufficient to cause distortion and formation of a bulge on the
surface (blistering).
2. The high concentration of atomic hydrogen can result directly in embrittlement
and cracking of the steel, particularly high strength or high hardness steels. This
often includes the heat affected zones in low strength steels that have not been
PWHT (SSC).
3. A combination of the two effects may occur, wherein laminations on parallel
planes are linked by cracks in the through-thickness direction (HIC), or where a
stacked array of short laminations appears ahead of a sulfide stress crack and
enables its further propagation (SOHIC).
Thus, any of several forms of wet H2S damage can occur if all three of the
following are present:
a corrosion reaction that generates hydrogen atoms
a chemical environment to enhance atomic hydrogen absorption into the steel
(commonly sulfides, with cyanides being particularly potent)

Chevron Corporation 400-63 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

a susceptible microstructure
The various wet H2S damage mechanisms are discussed in greater detail in the
following sections.

451 Mechanism of Hydrogen Blistering


Hydrogen blisters are subsurface planar cavities in steel, caused by the buildup of
internal hydrogen pressure. Blistering is caused by atomic hydrogen diffusing into a
steel and being trapped at voids, laminations, or non-metallic inclusions as shown in
Figure 400-41 (Section 450 discusses sources of atomic hydrogen). As hydrogen
atoms enter these sites, they combine to form molecular hydrogen, which cannot
escape by outward diffusion. If inclusions or laminations are large, internal pressure
may become sufficient to cause distortion and formation of a blister on the surface
(Figure 400-42). These blisters may appear on either or both surfaces of the plate, or
on top of one another, depending on the location of the original laminations. The
size and appearance of blisters vary from small protrusions to swellings several feet
or more in diameter.
The most common sites for blistering are manganese sulfide inclusions. During the
rolling of steel plates, manganese sulfide inclusions get flattened like pancakes in
the plane of the plate. Thus, their shape is favorable for both intercepting atomic
hydrogen that is diffusing through the steel, and for initiating blistering damage. In
contrast, the manufacturing process for forgings and for seamless pipe causes less
flattening of MnS inclusions. Consequently, seamless pipe and forgings are highly
resistant to hydrogen blistering. Welded pipe, which is made from rolled plate, is
equally as susceptible as pressure vessel plate steel.
Once a blister has formed, any further increase in internal pressure may cause it to
rupture, either by cracking at the center of the blister (Figure 400-43) or by tearing
at the edges (Figure 400-44). In both cases, the development and propagation of this
crack is aided by hydrogen embrittlement of the steel (see Section 460).
Uncracked blisters have little effect on the pressure-retaining capability of a vessel.
However, if blisters have produced surface bulges on the steel, then it is not possible
to determine whether blisters will crack in the future. Such areas are typically evalu-
ated as both a local thin area (as though the bulged material was missing) and as a
crack. See Section 456 for more information on disposition of blistered equipment.

January 2001 400-64 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-41 Formation of Hydrogen Blisters by H2S corrosion

452 Mechanism of Sulfide Stress Cracking (SSC) in Welds or High Strength


Steels
Note Section 459 specifically addresses sour service considerations for upstream
equipment, which are predominantly related to SSC.
Sulfide stress cracking (SSC) is a form of hydrogen embrittlement cracking that
occurs in high strength steels, hard welds, and hard weld heat affected zones
(HAZs) that are subjected to sour environments at temperatures below about 200F.
It requires a tensile stress, which can be either externally applied or residual
(e.g., from welding). SSC can occur rapidly in susceptible steels, and it results in
cracks that have a macroscopically brittle appearance and little or no ductility.

Chevron Corporation 400-65 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-42 Hydrogen Blistering of Carbon Steel Plate. Fig. 400-43 Cross-section of Hydrogen Blister
Note cracks at either top or edge of which has Cracked Through the
blisters. Surface of the Plate at the Apex of the
Blister. Note second blister beneath
the first.

Fig. 400-44 Cross-section of Hydrogen Blister which has Cracked


at the Edge Rather than Across the Center

January 2001 400-66 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Carbon steels with a hardness above Rockwell C 22 or BHN 241 are considered
susceptible, although steel composition and microstructure can influence the
threshold hardness for susceptibility. Low strength carbon steels used in pressure
vessel construction are immune to this form of cracking, except at weldments.
Common locations for SSC failures in sour refinery services include the following:
High strength 12 Cr (Type 410 SS) valve trim
Compressor shafts, sleeves, and other high strength machinery parts exposed to
sour gas
Bolts (including common B7, especially in floating head closures)
Alloy steel relief valve springs that have not been A1 plated or isolated from
sour relief gas by bellows designs
Hard welds and weld heat affected zones, most commonly at low heat input
fillet welds (see further discussion below). Hard welds have sometimes resulted
from material mix-ups and the use of alloy filler metals in welds that should
have been carbon steel.
Weld metal hardness of most refinery steels is commonly limited to BHN 200,
which is the normal state of E6010 or E7018 (and similar) weldments with no
special treatment. This is sufficient to prevent SSC in bulk weldmetal.
Weld heat affected zones are a common location for SSC, even in nominally low
strength steel. This is particularly true for single pass fillet weld attachments, which
are usually made with low heat input and without preheating of the basemetal.
Welding heats a narrow band of basemetal into the austenitizing temperature range,
and the subsequent rapid cooling can leave a hard zone adjacent to the weld. Weld
heat input, steel composition, thickness, preheat temperature, and other factors
affect whether this will occur on any given weldment. On multipass welds, the root
passes and fill passes are tempered by the heat of the subsequent passes, leaving
only the HAZ of the cap pass in the untempered and therefore most vulnerable
condition. Post weld heat treatment reduces weld HAZ hardness and virtually elimi-
nates SSC in most refinery steels. However, microalloying elements such as Nb or
V can cause steels to be resistant to tempering and to retain high HAZ hardness
even after PWHT.
NACE Standard MR0175 details the material requirements for preventing SSC in
upstream environments. Compliance with MR0175 is sufficient to prevent SSC in
most refinery environments as well, but it does not address the other potentially
serious forms of wet H2S damage such as HIC, SOHIC, and blistering.

453 Mechanisms of Hydrogen Induced Cracking (HIC) and Stress Oriented


Hydrogen Induced Cracking (SOHIC)
Hydrogen induced cracking (HIC) and stress oriented hydrogen induced cracking
(SOHIC) refer to cracking in the through-thickness direction of the plate or pipe.

Chevron Corporation 400-67 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

HIC is the cracking that links hydrogen blisters or laminations at different planes in
the metal either to each other or to the surface of the steel. HIC cracking requires no
externally applied stress. Instead, these cracks are driven by stresses from the
internal buildup of hydrogen at blisters. HIC is sometimes called step-wise cracking,
in reference to its appearance. Figure 400-45 shows a crack with a stepwise appear-
ance that is typical of HIC.
Since HIC is closely associated with blistering, it occurs most frequently in steels
with high levels of sulfide inclusions. It is also associated with steels in which
manganese segregation and a banded microstructure is observed. The resistance of a
steel to HIC can be greatly increased by reducing sulfur levels, and by controlling
the shape of sulfide inclusions through the use of calcium or rare earth additions to
produce spherical inclusions rather than flattened ones.
SOHIC refers to cracks in the through-thickness direction that are formed by the
link-up of a stacked array of small HIC cracks. SOHIC is aligned perpendicular to
an applied or residual stress. It is most commonly found in the heat affected zone of
welds, where it typically initiates from other cracks or defects (e.g., a sulfide crack
at a hard weld HAZ). See Figure 400-46.

Fig. 400-45 HIC in a A516-70 Hydrogen Reformer Amine Fig. 400-46 SOHIC Crack (at 50X) from a Sphere
Contactor/Water Wash Tower From NACE Containing Sour LPG Note the stepwise
Standard RP0296-96 appearance common to both SOHIC
and HIC.

In contrast to HIC behavior, the resistance of a steel to SOHIC seems to be reduced


by calcium or rare earth additions intended to control inclusion shape [11]. The
reason for this isnt clear, but one theory holds that the in-plane delaminations that
occur in conventional steels may blunt SOHIC cracks and retard their progress.
PWHT has a net positive effect on reducing susceptibility to SOHIC. It reduces
weld HAZ hardness, which in turn reduces the likelihood of sulfide stress cracks
that can trigger SOHIC. Also, by reducing weld residual stress it reduces the driving
force for SOHIC. These benefits are partially offset by the slight tempering effect of
PWHT on basemetal strength, which in turn has a slightly adverse effect on HIC
and SOHIC resistance.

January 2001 400-68 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Figure 400-39 summarizes HIC and SOHIC and compares them to other hydrogen
damage mechanisms.

454 Occurrence of Hydrogen Blistering and Cracking in Wet H2S Service


Wet H2S cracking of any degree is a serious concern, but first priority must be given
to cracking that has the highest potential to threaten the pressure integrity of equip-
ment. Experience has shown that this is most likely under the following
circumstances:
The equipment has a history of blistering.
Significant non-PWHT repairs or alterations have been made, particularly if
they were in response to wet H2S damage. Circumstances will vary, but
significant generally means welds greater than 50% of wall thickness, or "
in depth, or greater than a few inches in length. Butt patches and replacements
of nozzles, shell courses, or heads are at the top of the list.
Within any piece of affected equipment, the highest priority must be given to
inspection of the pressure containing welds (seams and nozzles). Figure 400-47
shows that deep cracks (> 3/8" deep or 50% of wall thickness) most often occur at
nozzles and seams, and only infrequently at attachments.

Fig. 400-47 Percent of Vessels in Wet H2S Service with Cracks Greater than 3/8" Deep or
50% of Wall Thickness
3.5%
% OF VESSELS WITH DEEP CRACKS

3.0%
RESULTS OF 1100 INSPECTIONS
2.5%
Fab Defects
2.0%
Wet H2S Cracks
1.5%

1.0%

0.5%

0.0%
Nozzles Seams Attachments Not Reported
CRACK LOCATION

It is common to find more than one cracking mechanism active in a given service.
Blistering often results in HIC at the edges or center of the blister. Sulfide stress
cracks, which are discussed in Section 452 can initiate SOHIC [12] [13] [14].
All of these cracking mechanisms are the result of hydrogen that forms during
corrosion on the steel surface. Corrosion rates are highest on a clean steel surface.
Higher corrosion rates mean that more atomic hydrogen is being generated, and it is
available to diffuse into the steel. With time, an iron sulfide (FeS) scale forms on the
surface and slows the corrosion rate. Therefore, blistering or cracking is likely to
occur soon after a vessel is put into service or following process upsets, and then
decrease with time.

Chevron Corporation 400-69 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Inspection results indicate that certain wet H2S services are more likely to cause
blistering and cracking. Though a limited degree of cracking has been found in
equipment containing less than 50 ppm H2S in liquid water, 50 ppm H2S is the
minimum H2S level Chevron uses for wet H2S inspections. Consider upset condi-
tions when evaluating the H2S levels [15].
Blistering occurs most frequently in services which contain H2S and cyanides or
ammonia. The acidity or alkalinity of the system influences the rate of damage, with
blistering generally minimized by an approximately neutral pH. Small amounts of
cyanide greatly increase corrosion rates and hydrogen entry into carbon steel. Most
references indicate that cyanides do not enter directly into the corrosion reaction,
but increase corrosion and atomic hydrogen generation by removing the protective
FeS scale.
Hydrocarbon vapor and LPG services are more likely to cause cracking and blis-
tering than the heavy crude services. Water is carried with the hydrocarbon vapor
while the crudes inhibit the corrosion of the steel and less atomic hydrogen is
formed.
These are the most likely places to find blistering and cracking:
LPG and hydrocarbon vapor services
FCC plants and gas recovery systems
amine plant absorbers, regenerators, and regenerator reboilers
hydrocracker and hydrotreater cold high and low pressure separators
services with pH above 8.0 and greater than 2,000 ppm H2S
services where cyanides are present
Heat exchanger shells, vessels, and columns manufactured from plate are all suscep-
tible to blistering, HIC, SOHIC, and SSC. Tanks and large diameter welded pipe
have no special inspection requirements but both are susceptible to blistering.
Piping welds with high hardness (greater than 200 BHN) are susceptible to SSC. If
problems are suspected in a certain piece of equipment, contact CRTC Materials and
Equipment Engineering for Company-wide inspection history.

455 Inspection for Wet H2S Damage


The system described below is used for prioritizing and executing inspections for
wet H2S damage in Chevrons domestic refineries. It is applicable to all equipment
containing a liquid water phase with more than 50 ppm H2S. This policy is main-
tained and updated periodically in Chevrons Wet H2S Cracking Notebook [15], last
issued December 23, 1996. As of this writing, the information below is consistent
with that policy. Since this policy is continuously being reviewed and updated,
check with a CRTC materials engineer before implementing the policy.

January 2001 400-70 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

A distinction is made between the first inspection of an equipment item for wet H2S
cracking, and reinspections of the same equipment. For the first inspection, use a
prioritization system which takes into account the following three key factors:
Service severity
Susceptibility of the steel
Potential consequences of a leak
For all follow-up inspections, frequency and priority is primarily driven by the
results of the first inspection, as described below.

Required Inspection Schedule and Priority (First Inspection)


A five-step method is used to establish the schedule and priority of first-time inspec-
tions for wet H2S equipment on both new and old equipment.

Step 1. Assign a service severity factor.


a. Check the flow diagrams (Figures 400-49 to 400-56) for the severity
number for specific equipment. If no severity number is given for an indi-
vidual equipment item, use Part b.
b. For equipment not given a severity number on the flow diagrams in
Figures 400-49 to 400-56, use the following guidelines:
1. Use a severity number of 2 as a default for equipment in wet H2S
service.
2. Add 1 if the H2S content in water is greater than 2000 ppm.
3. Add 2 if the cyanides are present (generally FCCs and Cokers).
4. Add 2 if the equipment is in hydrocarbon vapor or LPG service.
Note The severity of the service indicated on the flow diagrams has been deter-
mined from historical data collected by Chevron since 1989, including over 1100
inspection reports submitted to CRTC. Where a significant number of vessels in a
given service have been inspected, and/or process conditions are well established, a
severity of service factor has been assigned to that vessel type. Upset conditions can
result in significant damage that would be unexpected based on the nominal oper-
ating conditions.

Chevron Corporation 400-71 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Step 2. Assign a steel susceptibility factor based on fabrication and repair


history as follows:
Susceptibility
History Factor
History of Cracking Requiring Weld Repair, no PWHT 16
History of Blistering With Linking (HIC or Stepwise Cracking) 14
History of Cracking Requiring Weld Repair, PWHT 12
History of Blistering 12
History of Cracking, No Weld Repair Required 10
Any Non-Original Welds or Welded Alterations 9
Conventional Steel, No WFMT History, No PWHT 6
Conventional Steel, No WFMT History, PWHT 5
(1)
HIC Resistant Steel , No WFMT History, No PWHT 4
(1)
HIC Resistant Steel , No WFMT History, PWHT 3
(1) HIC Resistant means steel was purchased to the severe sour service requirements of EG-4749, EG-
4750, or as stated in the Wet H2S Cracking Notebook (either current or previous editions).

Step 3. Establish a cracking factor by multiplying the service severity factor


times the steel susceptibility factor.

Step 4. Evaluate the relative consequences of a leak. Assign a fluid service


category for each piece of equipment as follows:
Category 1 (highest consequence): LPG, rich amine, vapor streams
containing 3 wt. % H2S, and all streams above 1500 psia operating
pressure.
Category 2 (moderate consequence): Hydrogen, fuel gas, natural gas, lean
amine, liquid streams that vaporize quickly upon release, and all streams
above 500 psig operating pressure.
Category 3 (lowest consequence): All other sour hydrocarbon and sour
water streams.
The consequences of a leak are difficult to predict, but are an important factor for
prioritizing inspections. For this purpose, the three process stream categories are
based on toxicity and explosivity. These categories are broadly similar to the fluid
service categories in API 570 for piping.

January 2001 400-72 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Step 5. Establish the required inspection schedule by using Figure 400-48.

Fig. 400-48 Determining Inspection Schedules

1
Inspect Next S/D

Inspect Within 10 Years


Service Category

No Special Inspection
2

0 10 20 30 40 50 60 70 80 90 100 110 120

Cracking Factor

Inspect Next S/D:


Fluid Category 1 with Cracking Factor 35
Fluid Category 2 with Cracking Factor 55
Inspect Within 10 Years:
Fluid Category 1 with Cracking Factor of 0 to 34
Fluid Category 2 with Cracking Factor of 10 to 54
Fluid Category 3 with Cracking Factor 20
No Special Inspection Required:
Fluid Category 2 with Cracking Factor of 0 to 9
Fluid Category 3 with Cracking Factor 0 to 20

Chevron Corporation 400-73 January 2001


Fig. 400-49 Amine Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


H2S to
Sulfur Plant
Sweet Gas to Fuel

Sweet Gas Amine


K. O. Drum Regenerator O/H Acid Gas
Severity = 4 Condenser KO Drum
Severity = 4 Severity=2
Amine
Regenerator
Reflux Drum
Severity = 2
Sour Water
H2S To
Absorber Amine
Severity = 8 Sump
Amine
Regenerator To Amine Relief
Ammonia Caustic
Severity = 8 Sump or Drain
Scrubber Scrubber
Column Severity = 8
Amine
400-74

Severity = 4
Cooler
Severity = 2
Amine
Regenerator
Reboiler
Severity = 8 Spent
Caustic

400 Characteristic Corrosion Phenomena


Offgas

Sour Gas
Rich/Lean Amine
K. O. Drum
Exchanger Sour water to Sour Water Stripper
Severity = 2
Severity = 8
Drain

Severity Numbers
Rich Amine Very Mild 1
Flash Drum Mild 2 All Amine
Severity = 2 Moderate
January 2001

4 Equipment
Severe 8 128/292 cracked Rev. 4 10/1/96
Fig. 400-50 Coker Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Debutanizer
Interstage O/H Condenser
Debutanizer
Cooler Severity=2
Reflux Drum
Severity=4 EK
Severity=4
2
Light

Fractionator O/H Gasoline

Condenser
Severity=4
Interstage KO Gasoline
Severity=8 Debutanizer
Splitter
Severity=2
Severity=2

Tailgas
Gasoline
Sponge
Splitter
Fractionator Absorber
Severity=2
O/H Drum Severity=4
Severity=8
400-75

Fractionator
Severity=2

Compressor
After Cooler

400 Characteristic Corrosion Phenomena


Severity=4 Intermediate
Gasoline
Splitter
Severity=2
Absorber
Severity=4 High Pressure
Separator Intermediate

Severity=4 Gasoline

To Coke Drums Severity Numbers


Very Mild 1
All Coker Mild 2
Equipment Moderate 4
January 2001

17/53 cracked Severe 8


Rev 4 10/1/96
Fig. 400-51 Crude Unit Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Atmospheric
Atmospheric O/H
O/H Cooler
Product Drum Stabilizer O/H
Severity=1 Stabilizer
Severity=1 Condenser
Severity=2

Atmospheric
Atmospheric Atmospheric
O/H Condenser
O/H Condenser Reflux Drum
Severity=4
Severity=2 Severity=4 Stabilizer Reflux
Drum
Severity=4 Dirty Water

Vacuum Ejector
Aftercondenser
Severity=1
Steam
Gasoline
Stabilizer
Vacuum O/H
Atmospheric Severity=2
Condenser
Column
Severity=1
Severity=2
400-76

Vacuum
Seal Drum
Severity=1

Storage/Splitter
Vacuum Vacuum
Column Column O/H
Drum Severity Numbers

400 Characteristic Corrosion Phenomena


Severity=1 Very Mild 1
Mild 2
Moderate 4
Severe 8

Desalter Desalter
Severity=1 Severity=1
All Crude Unit
Equipment
Resid 42/123 cracked
Crude
Feed
Dirty
Water
January 2001

Rev.4 10/1/96
Fig. 400-52 FCC Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Sour gas to H2S removal section

Interstage
Suction
Cooler Debutanizer Depropanizer
KO Drum
Severity = 8 Accumulator Accumulator
Severity = 8
Sponge Severity = 8 Severity = 4
Interstage Absorber Depropanizer
Debutanizer
KO Drum Severity = 8 O/H Condenser
O/H Condenser
Severity = 8 Severity = 4 Severity = 4

Water

Fractionator
Overhead
Condenser
Severity = 4 Water
Propane cut
to H2S plant
Deethanizer
Feed
Condenser
400-77

Severity = 8 Deethanizer Debutanizer


Depropanizer
Severity = 8 Severity = 4
Severity = 4

Fractionator
Reflux Drum
Severity = 8

400 Characteristic Corrosion Phenomena


Sour Deethanizer Feed
water Separator
Severity = 8
Main
Fractionator
Foul
Severity = 1
water

Severity Numbers Butane cut to Alky

Very Mild 1
Mild 2
Moderate 4 All FCC To gasoline splitter
Heavy cycle oil to tankage
Severe 8 Equipment
January 2001

101/210 cracked Rev. 4 10/1/96


Fig. 400-53 Hydrocracker and Heavy Hydrotreater Reaction Section Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Recycle
KO Drum Cold High
Severity = 8 Pressure
Separator
Severity = 8

Effluent
Air Cooler
Severity = 8
Cold Low
Pressure
Wash Separator
Water Severity = 8
Reactor
400-78

To Fractionator System
Hot High
Pressure
Separator

400 Characteristic Corrosion Phenomena


Severity Numbers
Very Mild 1
Mild 2
Moderate 4
Hot Low
Severe 8
Pressure
Separator
All Hydrocracker
Reaction Section
Equipment
31/53 cracked
January 2001

Rev. 4 10/1/96
Fig. 400-54 Hydrocracker Fractionation Section
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Compressor
After Cooler
Fractionator Severity = 2
Reflux Cooler
Severity = 2

Fractionator
Reflux Drum
Severity = 2

KO Drum
Stripper O/H To Deethanizer
Severity = 4
KO Drum Condenser feed cooler
Severity = 4 Severity = 2

Fractionator
Reflux Drum
Severity = 4 To Deethanizer
Fractionator
O/H Condenser feed drum
Fractionator Severity = 2
Stripper O/H
O/H Condenser Accumulator
Severity = 2 Severity = 4
Fractionator
400-79

Severity = 4 Stripper
Severity = 4
From
Reaction
Section

Fractionator
Severity = 2

400 Characteristic Corrosion Phenomena


All Hydrocracker All Hydrocracker Severity Numbers
Fractionator Section Fractionator Section Very Mild 1
From Reaction Section
Equipment with Equipment without Mild 2
H2S Stripper H2S Stripper Moderate 4
15/32 cracked 15 /32 cracked Severe 8
Rev. 4 10/1/96
January 2001
Fig. 400-55 Hydrocracker and Heavy Hydrotreaters Gas Recovery Units
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Depropanizer
Sponge Debutanizer O/H Condenser
Absorber O/H Condenser Severity = 2
Severity = 8 Severity = 2

Deethanizer
Feed Cooler
Severity = 4 Debutanizer
Reflux Drum Depropanizer
Severity = 4 O/H Accumulator
Severity = 4

Deethanizer Debutanizer
Severity = 8 Severity = 4
400-80

Deethanizer
Feed Drum Depropanizer
Severity = 8 Severity = 4
Deethanizer
Reboiler
Severity = 1

400 Characteristic Corrosion Phenomena


Debutanizer
Reboiler
Severity = 1

Severity Numbers All Hydrocracker &


Very Mild 1
Heavy Hydrotreaters
Mild 2
Gas Recovery Units
Moderate 4 9/18 cracked
January 2001

Severe 8
Rev. 4 10/1/96
Fig. 400-56 Naphtha, Jet, and Diesel Hydrotreater Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Severity Numbers
Reactor Charge Very Mild 1
Heater Mild 2 Desulfurizer O/H Sour Gas
Moderate 4 Condenser Treating

Reactor Severe 8 Severity = 2

Desulfurizer
Reflux Drum
Severity = 4 Foul
Desulfurizer Water
Severity = 1

Recycle Gas
KO Drum
Severity = 4
Water
Sour Gas
Drain Pot
Treating
Severity = 2
400-81

Foul
To Low Water
Pressure
Separator
Desulfurizer
Reboiler

High

400 Characteristic Corrosion Phenomena


Pressure
Separator
Severity = 8

Effluent Low
Desulfurizer
Cooler Pressure All Naphtha
Feed/Effluent
Separator Hydrotreater
Exchanger
Severity = 4 Equipment
Foul Severity = 2
Water 27/92 cracked
Foul Water Rev. 4 10/1/96
January 2001
Fig. 400-57 Sour Water Stripper Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


Stripper O/H
Condenser
Severity = 8
Vent to flare To Sulfur Plant

Sour Water
Feed Stripper Reflux
Water Drum
Degasser Severity = 8
Severity = 4

Sour Water
Stripper
400-82

Severity = 4

400 Characteristic Corrosion Phenomena


All Sour Water
Stripper Equipment
Stripper 10/30 cracked
Reboiler
Severity = 2 Severity Numbers
Very Mild 1
Mild 2
Moderate 4
Stripper Feed
Severe 8
Effluent Exchanger
Severity = 4
January 2001

Stripped Water
Rev. 4 10/1/96
Fig. 400-58 Waste Water Treater Equipment
Chevron Corporation

Corrosion Prevention and Metallurgy Manual


H2S to Sulfur Recovery Unit Ammonia Stripper
O/H Condenser
H2S Severity=8
Stripper
KO Drum
Severity=4

Ammonia
Stripper
Reflux
To Caustic
Drain Drum
Scrubber
Severity=8
To Disposal

H2S
Stripper Ammonia
Severity=4 Scrubber
Sour
Ammonia Severity=4
Water
400-83

Feed Water Stripper Cooler


Severity=4 Severity=4
H2S Stripper
Feed Preheater
Severity=4

Degasser
Severity=4

400 Characteristic Corrosion Phenomena


To Slop Oil
H2S
Stripper To Ammonia Stripper
Reboiler Ammonia Reflux Drum
Severity=2 Stripper
Reboiler
Severity=2 Severity Numbers
Very Mild 1
Mild 2
All Waste Water Moderate 4
Treater Equipment Severe 8
January 2001

22/54 cracked
Rev. 4 10/1/96
400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Required Inspection Method and Extent of Coverage


The extent of inspection coverage and methods required are as follows:
1. Due to the service severity and the failure consequences, the following equip-
ment requires WFMT inspection of 100% of the interior surfaces of nozzle
welds, 3050% of the pressure containing seam welds, and at least 20% of the
non-pressure-containing attachment welds:
FCC Sponge Absorber
FCC Deethanizer
FCC Interstage KO Drum
Hydrocracker/Isomax Deethanizer
Hydrocracker/Isomax HP Separator
Amine Unit H2S Absorber
Amine Regenerator
Amine Rich/Lean Exchanger
All equipment with a history of blistering
All field repair welds and field alterations of equipment in wet H2S service
Inspections should be extended if cracks deeper than 0.10" or 15% of the wall
thickness are found.
2. All other equipment not in listed in (1) above requires WFMT or UT shear
wave inspection of 100% of the nozzle welds and at least 1020% of the seam
welds. Again, inspections should be extended if cracks deeper than 0.10" or
15% of the wall thickness are found. Inspection of attachment welds is not
required for this equipment.

Reinspection Schedule, Based on Results of First Inspection


1. Follow up inspections for wet H2S damage may be extended to a maximum of
10 years from the previous inspection, provided ALL of the following criteria
are met:
All crack indications were removed
All weld repairs or alterations were PWHTd (note: temper bead or high
preheat procedures are not equivalent to PWHT)
No active blistering is known to be occurring in the equipment
The inspection for equipment meeting these criteria should be as stated in
Figure 400-59.

January 2001 400-84 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-59 Extent of Inspection for Wet H2S Damage


Equipment Item Method; Extent of Coverage
FCC Sponge Absorber WFMT; 30-50% of seams and 100% of nozzles
FCC Deethanizer WFMT; 30-50% of seams and 100% of nozzles
FCC Interstage KO Drum WFMT; 30-50% of seams and 100% of nozzles
Hydrocracker/Isomax WFMT; 30-50% of seams and 100% of nozzles
Deethanizer
Hydrocracker/Isomax HP WFMT; 30-50% of seams and 100% of nozzles
Separator
Amine Unit H2S Absorber WFMT; 30-50% of seams and 100% of nozzles
Amine Regenerator WFMT; 30-50% of seams and 100% of nozzles
Amine Rich/Lean Exchanger WFMT; 30-50% of seams and 100% of nozzles
All Other Equipment WFMT; or UTSW 10-20% of pressure seams and
100% of nozzles

2. For equipment not meeting the criteria outlined in (1) above, reinspection
should occur at the next scheduled plant turnaround, not to exceed five years
from the previous inspection when the damage was found.
The required inspection methods are the same as in (1) above. Reinspec-
tion must be by WFMT for the eight specific equipment items listed in
Figure 400-59, but may be by either WFMT or UTSW for all other
equipment.
Reinspection may be limited to the areas in which the prior damage, repair,
or alteration occurred.
Any reinspection that meets the scope and intent established in (1) above
may be used to re-set the 10-year reinspection frequency. Inspections of
lesser scope shall not be used to re-set the 10 year reinspection frequency,
even if no damage is found.

Discussion of Inspection Methods


There are two approved methods for inspecting equipment for wet H2S damage for
Chevron facilities:
Wet Fluorescent Magnetic Particle Testing (WFMT)
Ultrasonic Shear Wave (UTSW), either manual or automated
A third method, eddy current inspection, is currently under development and may be
available for future inspections. Contact CRTC for the latest status. Hydrogen
probes are also discussed below.
WFMT Inspection. WFMT is the most sensitive, non-destructive examination
(NDE) tool available for tight, surface-breaking cracks.
Surface preparation is critical to the sensitivity of the WFMT inspection. White
metal blast or near-white metal blast is necessary. The following surface

Chevron Corporation 400-85 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

preparation methods have been tried by Chevron Operating Companies. Only the
first two (grit blast and flapper wheel) are approved for wet H2S cracking
inspection.
Grit blast with Du-Pont Starblast is the recommended surface preparation method
because it removes the sulfide and mill scales without peening over crack indica-
tions. On rough or pitted surfaces, additional preparation by a flapper wheel to
smooth the surface may be required.
Flapper wheel surface preparation is also acceptable. It allows for the most sensi-
tive inspection. Also, the flapper wheel will remove shallow non-relevant indica-
tions. Flapper wheel surface preparation of part of each vessel is recommended
when inspecting for carbonate stress corrosion cracking in the FCC. For equipment
in other wet H2S service, spot checking areas with flapper wheel surface prepara-
tion is not a requirement.
High pressure hydroblasting is not recommended because it does not always remove
mill scale that masks significant indications. Flash rusting can also be a problem,
and improper use of corrosion inhibitor can mask indications.
High pressure hydroblasting with garnet is not a recommended surface preparation
method. It can peen over tight cracks, flash rusting can interfere with inspection,
and equipment damage has occasionally occurred due to the highly abrasive nature
of the process.
UTSW Inspection. This includes both automated (AUT) versions and manual tech-
niques. UTSW is an extremely valuable tool because it allows inspection without
entering the vessel. Operating equipment up to 300F can be scanned using AUT,
while manual techniques are limited to about 120-140F. Unfortunately, external
reinforcing pads at nozzles and various other construction details prevent access to
some welds. Also, internal attachment welds at tray supports, downcomers, etc. can
be difficult to locate, and the UT signals difficult to interpret.
AUT scanning is expensive and coverage is slow. Its advantages are the ability to
inspect higher temperature equipment and the availability of hard copy reports,
including C-scans for blistering and laminations, and B-scans for cracking. All
AUT reports of cracking need to be verified by manual UTSW. Also, sizing of
defects cannot be done by AUT, but must be done manually. A technician can
manipulate the transducer angle for optimum viewing of the crack indication by
UTSW, and can also apply other UT techniques to distinguish cracking from other
forms of sound reflectors.
Both manual and automatic UT testing are extremely sensitive to the equipment
being used, the testing and calibration parameters, and to the skill of the technician
(including basic NDE skill and familiarity with the types of damage that may be
present). Most UT practitioners are not able to pass screening tests established by
CRTC NDE specialists, despite being ASNT qualified. Thus, it is essential to use
only contractors and technicians individually certified by CRTC for this work.
Among qualified UTSW practitioners, detection of cracks deeper than 0.1" in butt
seams is achievable. Sizing accuracy is indeterminate at this time, and is the subject

January 2001 400-86 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

of ongoing research. Crack indications showing apparent sizes of 0.1" and greater
require follow-up internal inspection by WFMT. Nevertheless, UTSW using quali-
fied techniques and technicians can reliably screen vessel weld seams and verify the
absence of significant cracks.
Note that the lack of ability to inspect nozzles and certain other pressure seams
means that UTSW alone cannot satisfy the inspection coverage requirements stated
earlier and in the Wet H2S Cracking Notebook.
Eddy Current (EC) Inspection. Eddy current inspection tools are currently under
development and field testing, but are not yet ready for widespread application. If
successful, they may be able to detect and accurately size cracks without removing
the surface scale. EC probes have been adapted to inspect both fillet welds and butt
seams. This could potentially eliminate the need for grit blasting or flapper wheel
preparation of most welds, although vessel entry would still be required. Contact
CRTC for current recommendations on EC inspection.
Hydrogen Probes. Hydrogen probes are a method of detecting the hydrogen
activity which may lead to cracking and blistering (see Section 540). Hydrogen
probes measure the amount of hydrogen diffusing through a steel vessel. Signifi-
cant increases in the amount of hydrogen indicate an increase in the severity of the
service, perhaps due to a change in process conditions or to some difficulty with the
corrosion control system (e.g., polysulfide injection, wash water injection, etc.).
This may also indicate a need for follow-up inspection [16].

456 Disposition and Repair of Wet H2S Damage


Blisters
Blisters are typically found during internal visual inspections or during external
ultrasonic inspections. To size blisters, measure the diameter from the inside of the
vessel or estimate it by locating the perimeter with ultrasonic thickness gauging.
Determine the depth of a blister using ultrasonic thickness gauge measurements or
by removing the blister with a grinder and measuring the depth.
As the blister size and frequency increase, the probability of blisters linking up to
form HIC or stepwise cracks increases.
Blisters with no associated cracking and that do not extend below Tmin do not
require repair. Blisters that extend below Tmin or have associated cracks may be
evaluated per Chevrons Guide to Evaluating Equipment Integrity[17], NBIC, or
API 510. Materials Engineering, Fitness for Service Specialists are available to
provide additional guidance as needed.
The drilling or venting of blisters from inside a vessel is common, but this is not
necessary unless heat treatment is planned, in which case unvented blisters are
likely to grow or crack.

Chevron Corporation 400-87 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Cracks
Most cracks found in vessels in wet H2S service are at blisters, in welds, or in weld
heat affected zones. Cracks may be evaluated per Chevrons Guide to Evaluating
Equipment Integrity [17], NBIC, or API 510. CRTCs Fitness for Service
Specialists are available to provide additional guidance.
Use the weld repair guidelines given below to minimize the occurrence of addi-
tional cracking after repair. Hydrogen bake-out prior to weld repairs and post weld
heat treatment improve the chances of making a sound repair.
For services which have severe or recurring cracking problems, non-porous, holiday
free stainless steel coatings will make welds resistant to SOHIC and SSC. Stainless
steel coatings can be applied by weld overlay. Thermal spray stainless steel and
nickel based coatings have been tried in several cases by Chevron, and are now
being evaluated further. Coatings and overlays are intended for welds with a history
of cracking, not for the entire vessel.

Weld Repair Guidelines for Equipment in Wet H2S Service


These guidelines present the best industry practice for repair of vessels in wet H2S
service. They are intended to enable the completion of crack-free repairs, minimize
the likelihood of future in-service cracking, and minimize the likelihood of equip-
ment failure if some cracking does occur. Use these guidelines after determining
that weld repair is required.
Recommended Weld Repair Method for Wet H2S Service:
1. Perform a hydrogen bake-out by heating the equipment to a minimum of
450 600F for 4 hours. Vessels with wall thickness greater than 1.5 inches
should be baked out at 600F.
Hydrogen bakeout is strongly recommended when cracking is severe or
several areas require weld repair. Bake-out is also recommended when
cracks appear to grow during grinding or when delayed hydrogen cracking
occurs after weld repair.
If visible blisters are present and if the repairs will be subsequently PWHT,
then a pre-repair hydrogen bake-out at full PWHT temperature is recom-
mended. This is because trapped molecular hydrogen in blisters will likely
cause additional cracking during PWHT. By heat treating both before and
after welding, only a single repair cycle is needed. Consider drilling large
blisters to minimize damage during heating.
When cracking and repairs are minor, the bake-out can be omitted. If the
bake-out is omitted, then preheating to a higher temperature may help
prevent cracking during and after welding.
2. Remove all crack indications (wet H2S or fabrication defects) by grinding or air
arc-gouging. Clean any arc gouge cavities with a grinding disk.
3. Verify complete removal of the crack indications using WFMT.

January 2001 400-88 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

4. Repair as follows:
Step 1: Preheat according to thickness
Thickness Preheat & Maintain Preheat of:
1" 100F if the metal temperature is less than 50F
1" 200F
Step 2: All surfaces shall be free of moisture.
Step 3: Use E7018 low hydrogen electrodes.
Step 4: Use PWHT at 1100F to 1200F in accordance with ASME
Section VIII Division 1 Table UCS-56.

5. Inspect completed repair using WFMT.


6. Verify that the weld hardness does not exceed 200 BHN (Brinell hardness
number).
7. Hydrotest at 1.5 times the ambient temperature maximum allowable working
pressure. Hydrotesting reduces the weldment residual stresses that can
contribute to crack propagation and rapid brittle fracture. The larger the repair,
the more important it is to hydrotest.
Alternate Weld Repair Methods for Wet H2S Service. Use great caution when
considering these alternative repair methods. They do not include PWHT, and hence
have two major drawbacks:
1. They do not reduce the residual stresses due to welding or fitup that can drive
large cracks to failure.
2. While both methods reduce heat affected zone hardness as compared to
conventional weld procedures, neither is as effective as PWHT. Consequently,
repairs made by these alternative methods will be more susceptible SSC
cracking in service than if the welds were PWHT.
The alternate repair methods should not be used when any of the following circum-
stances apply:
The repairs are major or extensive, such as butt patches, nozzle replace-
ments, or repair grooves greater than " in depth or 50% of wall thickness.
PWHT is required to comply with Code
The equipment is in an environmental cracking service that requires
PWHT to meet Company requirements (caustic, amine, potassium
carbonate, or carbonate in FCC plants).
Furthermore, the alternative repair methods should generally be avoided for Cate-
gory 1 fluid service (see Consequence Ratings), and for services with high hydrogen
charging (a high environmental severity factor per the flow diagrams, hydrogen
blistering, or extensive cracking).
Alternate Method A: Temper Bead Repair. Use the recommended weld repair
method except substitute the following for item 4 in the recommended method:

Chevron Corporation 400-89 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

4. Repair as follows:
Step 1: Preheat weld areas and maintain at a minimum of 200F during
welding. The maximum interpass temperature shall be 500F.
Step 2: Use E7018 low hydrogen electrodes.
Step 3: Perform vertical welding in the uphill direction.
Step 4: Run a tempering weld pass over the cap pass of the repair weld on
the base metal side, as shown in Figure 400-60. Remove the
tempering pass by grinding when weld repair is complete and
cooled. A similar procedure should be used for repairs transverse
to the weld or in the base metal.

Fig. 400-60 Temper Bead Repair Weld

Outer Surface
Original Weld

Repair
Weld Inner Surface

1/16 - 1/8 Tempering Weld Bead

Alternate Method B: High Preheat. Use the recommended weld repair method
except substitute the following for item 4 in the recommended method:
4. Repair as follows:
Step 1: Preheat weld areas and maintain at a minimum of 300F during
welding. The maximum interpass temperature shall be 500F.
Step 2: Use E7018 low hydrogen electrodes.
Step 3: Perform vertical welding in the uphill direction.

457 Prevention of Blistering and Cracking Due to Wet H2S


To prevent hydrogen blistering, hydrogen induced cracking (HIC), or stress oriented
hydrogen induced cracking (SOHIC), consider the conditions necessary for
cracking.
Corrosion
A chemical environment which slows the recombination of atomic hydrogen
Discontinuities in the steel
Stress

January 2001 400-90 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Corrosion Control
Atomic hydrogen will not be generated if corrosion is eliminated; therefore,
hydrogen will not diffuse into the steel and initiate cracking.
One method to prevent all forms of hydrogen induced cracking in wet H2S services
is the use of stainless steel clad plate. Follow Specification PVM-EG-1322, Stain-
less Steel Clad Pressure Vessels. For small vessels, weld overlay or thermal spray
stainless steel coatings of the welds also protect a vessel against hydrogen blistering
and cracking, assuming the coating is holiday free.
Inhibitors are also used to control corrosion. Organic corrosion inhibitors success-
fully prevent blistering for sections of vessels in continuous contact with liquid. In
columns or accumulators that do not operate full of liquid, inhibitors are relatively
ineffective in protecting surfaces exposed to vapors.

Hydrogen Recombination Poison Control


Hydrogen sulfide and cyanides are hydrogen recombination poisons, which allow
hydrogen to diffuse into the steel instead of recombining on the surface of the steel
to form molecular hydrogen. The poisons are usually water soluble and can be
removed or diluted by water washing. Good results have been obtained by injecting
water into piping and withdrawing it in a subsequent vessel, providing the addi-
tional water can be tolerated. Also, polysulfide is sometimes injected to control
cyanides in refinery plants. (See Section 630 for further discussion of polysulfide
injection.) The effectiveness of polysulfide additions should be monitored using
hydrogen probes (see Section 540). Monitoring determines the water injection rates
needed to minimize hydrogen generation and control hydrogen blistering and
cracking.

Steel Selection and Fabrication Guidelines


For new equipment fabrication for wet H2S services (> 50wppm dissolved in water
phase), Chevron uses the different specifications for steel selection, fabrication, and
testing shown in Figure 400-61.
The detailed requirements for each of the sour service classifications and more
specific guidance on their application is included in the Wet H2S Cracking Note-
book and in Corporate EG specifications. Many of the factors included, or consid-
ered for inclusion, in the guidelines are listed and discussed in Figure 400-62.
HIC Resistant Steels were widely touted during the 1980s as a solution to wet
H2S cracking problems. They were essentially very low sulfur steels with Ca treat-
ment. The Ca hardens the MnS and minimizes the flattening that otherwise would
occur during rolling. The performance test standard required of HIC resistant steels
was the NACE TM0284 test in the NACE TM0-177 acidified sour brine solution
(pH 3). This test measured cracking in both the rolling plane and the through thick-
ness direction of non-stressed specimens.
The technology grew from the pipeline industry where HIC and stepwise cracking
had caused failures. However, pipeline materials, construction, and service differ
from pressure vessels in several important ways.

Chevron Corporation 400-91 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-61 Steel Selection Specifications


Service Definition Requirements
Non-sour Service < 50 wppm H2S No special measures for H2S

Sour Service > 50 wppm H2S; service severity factor of 5 No special steel selection required, but
or less per the flow diagrams in the Wet PWHT and increased inspection are
H2S Cracking Notebook (Figures 400-49 to required, including WFMT and hardness
400-56 in this chapter). testing.
Severe Sour Service > 50 wppm H2S; service severity factor of 6 Special steel chemistry and fabrication
or greater per the flow diagrams in the Wet requirements apply, along with additional
H2S Cracking Notebook (Figures 400-49 to testing and inspection.
400-56 in this chapter).
Severe blistering or corrosion services; Stainless steel cladding
consult Materials Engineering Specialist

Fig. 400-62 Additional Considerations for New Construction (1 of 2)


Factor Consideration
PWHT Reduces weldment hardness and thereby prevents SSC
Reduces local residual stresses that can cause shallow sulfide cracks or HIC cracks
to propagate by SOHIC

Reduces global residual stresses and thereby reduces the likelihood of rupture or
fracture in the presence of large flaws
WFMT Inspection Ensures that no cracks are present. Pre-existing cracks are stress risers which may
help initiate wet H2S cracking in service.

Ensures that any defects subsequently found are due to in-service cracking. This is
important because hydrogen-related welding defects can be indistinguishable from
some forms of wet H2S damage.

UT Exam per SA578 Ensures minimal midwall laminations and inclusions are present. Note that SA578
Level 3 requires that any areas in which loss of backwall reflection exceeds 50%
must be limited in size to a 1" diameter circle.

Normalize & Temper Heat Refines grain size, which laboratory testing has shown to improve resistance to blis-
Treatment tering, HIC, and SOHIC. A finer grain size also improves toughness. Quenching and
tempering has similar effects, but also eliminates ferrite/pearlite banding and Mn
segregation. Also, N&T and Q&T heat treatments enable steels to meet minimum
tensile strength requirements at lower levels of carbon and carbon equivalent,
which in turn means these steels have lower weld HAZ hardnesses.

January 2001 400-92 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-62 Additional Considerations for New Construction (2 of 2)


Factor Consideration
Hardness Testing Hardness testing of finished welds ensures resistance to SSC, and serves as a
check against the use of improper welding consumables or improper heat treat-
ment. A laboratory microhardness traverse across the HAZ on production test plates
confirms that the weldment will respond to PWHT and soften sufficiently to be resis-
tant to SSC. The presence of microalloying elements such as V or Cb/Nb can cause
locally hard zones in the HAZ that resist tempering.

Chemistry Requirements Carbon equivalent limit - reduces the tendency of steels to form hard HAZs.

Cb + V limit - These elements increase a steels resistance to softening during PWHT.


Steels containing these microalloying elements can retain HAZ hardnesses high
enough for SSC to occur even after PWHT.

S limit - As discussed in Section 451, S is present in steel primarily as MnS inclu-


sions which are flattened during the steel rolling process and become ideal sites for
blistering and HIC cracking.

Ca additions - for inclusion shape control; reduces flattening of MnS during rolling,
and thereby eliminates sites for blistering and HIC cracking. Unfortunately, recent
testing has shown that excessive Ca treatment increases the susceptibility of steel
to the SOHIC form of cracking.

The final welding pass is always on the OD of a pipeline weld, and therefore
never in direct contact with the process fluid. The concern for SSC due to hard
weld HAZs is minimal because there are no internal attachment welds.
Higher strength pipeline steels are often fabricated using a controlled rolling
process at temperatures below the austenitizing temperature. This worsens the
flattening effect of rolling upon the MnS inclusions.
Due to the stress pattern of both applied and residual stresses, there was little
concern for SOHIC in pipelines.
Thus, while HIC resistant steels are giving good service in the pipeline industry,
process plant experience has been less successful. While blistering and HIC damage
is reduced, there appears to be a greater tendency for SSC and subsequent SOHIC
cracking, which may ultimately reduce the reliability of equipment. Many users,
including Chevron, are now eliminating or drastically curtailing their use of Ca-
treated HIC resistant steels in process plants.
A new class of steels manufactured to SA-841 have recently performed well in labo-
ratory tests for HIC and SOHIC resistance. These steels are made by thermo-
mechanically controlled processing (TMCP) or are quenched and tempered.

Chevron Corporation 400-93 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Although they have performed well in these tests, they are just beginning to be used
in process plant applications in the US.

458 Hydrogen-Induced Cracking (HIC) of Line Pipe


A phenomenon similar to the blistering of vessels can occur in pipelines (primarily
welded pipe) operating in sour service (sour gas or sour crude). This phenomenon
has been given several names in the literature, including hydrogen-induced cracking
(HIC), hydrogen pressure-induced cracking (HPIC), stepwise cracking (SWC),
hydrogen blister cracking, and others.
Like all forms of hydrogen damage, hydrogen blistering and cracking (HIC) of line
pipe is caused by hydrogen that diffuses into the steel. As with blistering, HIC does
not depend on the hardness of the steel. It occurs when hydrogen atoms accumulate
at nonmetallic inclusions in the steel, especially sulfides, and then recombine to
form H2 gas, with a resultant build-up of pressure at the inclusion sites. The result is
small blisters or internal fissures parallel to the pipe surface, which link up to cause
a leak path to the pipe surface.
The damage can appear as discrete cracks or as an array of cracks parallel to each
other, which are connected by cracks between them, as shown in Figure 400-63.

Fig. 400-63 Examples of HIC in Pipeline Steel

January 2001 400-94 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Blistering and HIC are primarily a problem with seam-welded pipe (as opposed to
seamless) because the sulfide inclusions are flattened out into pancake shapes
during plate rolling. Long, flat inclusions are much more detrimental than short,
round ones. Controlled rolled steels used for X52, X60, X65 and higher grades of
line pipe are particularly susceptible to HIC because the sulfide inclusions elongate
severely during the lower temperature rolling steps used in the plate rolling process.
Seamless pipe is less susceptible to blistering and HIC because higher forming
temperatures and less wall reduction are employed, so that the inclusions do not
elongate as severely. However, very few mills can make seamless pipe larger than
about 16-inch diameter, and it is much more expensive than welded pipe. Besides,
HIC can be prevented in welded pipe.

Prevention of HIC and Blistering


HIC and blistering can be prevented by reducing the sulfur content of the steel and
controlling the shape of the remaining sulfide inclusions. For sour service welded
pipe, the Company requires steel with a maximum sulfur content of 0.003% and
would prefer 0.001% maximum. We also specify that the steel be treated with
calcium, which controls the shape of the remaining sulfide inclusions. In order for
the shape control treatment to be effective, the steel must be fully killed. The
Company also limits impurities from other elements besides sulfur, specifically
phosphorous and nitrogen.
The Company also requires testing the HIC resistance of welded pipe according to
the NACE TM0284, using the test solution in NACE TM0177. In addition,
quenching and tempering or controlled rolling of the skelp is required for improved
toughness of the pipe.
Specification PPL-MS-4041, which can be used to purchase line pipe that will resist
HIC and blistering, can be found in the Pipeline Manual. Consult the Pipeline
Manual and Materials Engineering for guidance on whether to use HIC-resistant
steel for any particular project.

459 Sour Service Considerations for Upstream Equipment


Sour service damage in upstream equipment is similar to that in downstream equip-
ment. The hydrogen entry into the steel follows essentially the same process
(described in Section 450), and the blistering and cracking damage mechanisms are
similar (described in Sections 450453). However, several broad differences are
worth noting:
The predominant sour service concern in upstream is sulfide stress cracking
(SSC), due to the frequent use of high strength steels in casing, tubing, and
completion equipment.
Conformance to NACE Standard MR-01-75 is widely required.
Blistering and HIC damage is usually limited to longitudinally or spiral welded
linepipe and to surface processing equipment.

Chevron Corporation 400-95 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Examples of potential sulfide cracking failures in upstream operations include:


Cracking of high-strength tubing or casing, e.g., N-80 or P-110 in a sour oil or
gas well
Cracking of high-strength 12 Cr (Type 410 stainless steel) valve trim or other
12 Cr components in a sour well or process stream
Cracking of a hard piping or pipeline welds in a line carrying sour crude and
brine or wet sour gas
Cracking of high-strength drill pipe, tool joints, and other drilling tools exposed
to sour gas

NACE Standard MR-01-75


NACE (National Association of Corrosion Engineers) Standard MR-01-75 [24],
Sulfide Stress Cracking Resistant Metallic Materials for Oil Field Equipment, is the
industry standard which provides guidelines for the selection of materials for sour
environments. We are currently required by law to comply with MR-01-75 in Texas
wherever the Texas Railroad Commission's Rule 36 has jurisdiction. Compliance
with MR-01-75 is also required by federal law for all U.S. outer continental shelf
platforms, effective May 31, 1988. Unfortunately, MR-01-75 has many faults, and
many of its guidelines are controversial. In some cases, compliance with MR-01-75
may contradict sound engineering judgment. Improving MR-01-75 is an area of
ongoing concern and activity within NACE and the industry.
MR-01-75 is intended for oil field equipment only, with emphasis primarily on
downhole and wellhead equipment and valve materials. However, because there is
no analogous document for refinery plants, gas plants, or other operations, it has
begun to receive more widespread application.
Use caution when applying MR-01-75 outside its jurisdiction, especially in refinery
plants where experience may dictate H2S limits or material selections that conflict
with MR-01-75. While its treatment of sulfide stress cracking of carbon steels and
low-alloy steels is adequate for both upstream and downstream applications, it does
not address damage by other mechanisms such as HIC or SOHIC. Also, many of its
aspects, particularly those dealing with stainless steels and more highly alloyed
metals, are subject to debate.
MR-01-75 defines a sour gas environment follows:
The gas being handled is at a total pressure of 65 psia or greater.
The partial pressure of H2S in the gas is greater than 0.05 psia.
It also defines a sour oil or multiphase system as one which has a partial pressure of
H2S in the gas phase greater than 0.05 psia. However, in addition to this partial pres-
sure criterion, if an oil well meets all four of the following criteria, it is not consid-
ered sour.
The maximum gas/liquid ratio is 5000 SCF per bbl.
The gas phase contains a maximum of 15% H2S.

January 2001 400-96 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

The partial pressure of H2S in the gas phase is no higher than 10 psia.
The surface operating pressure is a maximum of 265 psia.
MR-01-75 contains curves that provide a convenient method of determining
whether or not a gas well or an oil well is sour. By knowing the total pressure and
the H2S concentration of the gas in the well, you can readily determine from the
graph if it falls within the sour (i.e., cracking susceptibility) region.
The definition of sour or nonsour, as outlined in MR-01-75, is intended only to
make clear where metals are subject to sulfide cracking at temperatures near
ambient. It does not address all aspects of materials selection for sour environ-
ments, such as weight loss corrosion, stress corrosion cracking, and hydriding by
H2S-chloride environments at elevated temperatures. Also, its definition of sour
does not consider H2S toxicity and personnel protection.

Preventing Sulfide Stress Cracking in Oilfield Services


Key points from MR-01-75 (and for sulfide cracking resistance in general) include
the following:
All carbon steels with hardness less than or equal to HRC 22 are suitable for sour
service. This includes all commonly used piping, plate, vessel and tank steels, and
valve body steels.
All low-alloy steels with hardness less than or equal to HRC 22, except those
containing 1% or more nickel, are suitable for sour service. This would include, for
example, AISI 4130 low-alloy steel (1% Cr - 0.2% Mo) used for wellheads,
christmas trees and valve bodies. The 1% Ni limitation is a controversial subject. In
areas not governed by MR-01-75, steels with greater than 1% Ni (added for
improved brittle fracture resistance) can safely be used if the heat treatment is care-
fully controlled.
12 Cr alloys (i.e., Type 410 stainless steel) have been used successfully in many
situations for valve trim and some other components when the hardness is held to a
maximum of HRC 22. However, since 12Cr alloys have rather poor inherent SSC
resistance, caution should be exercised in selecting 12Cr alloys for highly-stressed
components in severe H2S environments, i.e., a high-H2S gas well.
Stainless steels (i.e., Types 304 and 316) and more highly-alloyed materials are
generally very resistant to SSC. However, 300 Series stainless steels and some more
highly-alloyed stainless materials are susceptible to chloride stress corrosion
cracking in certain environments (see Section 420). For special applications such as
severe sour downhole applications, use of highly-alloyed materials is an area of
ongoing research within the industry. Because the technology is evolving,
MR-01-75 is inadequate. In special situations where alloys are being considered,
consult with the CRTC Materials and Equipment Engineering.
Service temperature has a significant effect on SSC resistance. Sulfide cracking
can only occur in the hydrogen embrittlement range of about -50F to 200F. It is
most severe at room temperature, and diminishes as the temperature either increases
or decreases. The greater the susceptibility of a steel to SSC, the wider its SSC

Chevron Corporation 400-97 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

temperature range within the -50F to 200F limit. For example, steel with a yield
strength of about 140 ksi is susceptible to cracking in about the 0F - 175F range.
A 175 ksi yield strength steel might still suffer SSC between 175F and 200F.
MR-01-75 makes use of this fact by permitting quenched and tempered API N-80
tubing and casing above 150F and API P-110 casing above 175F in sour service.
Neither of these grades of tubulars is resistant to sulfide cracking below the indi-
cated temperatures.
Care must be exercised in taking advantage of the temperature effect on SSC
susceptibility; all exposure conditions must be considered, including shut-in. For
example, a sour well with a flowing wellhead temperature of 175F may see the
same sour environment at 70F when the well is shut in. Similarly, a piece of equip-
ment operating in a sour environment above 200F may still be charged full of
hydrogen when it cools for a shutdown.
Welding standard carbon steel pipe (i.e., Grade B) or plate ordinarily should not
result in welds that are hard enough to be susceptible to SSC. Some precautions
should be taken, however, including: (1) requiring a hardness traverse across the
weld, both heat affected zones, and into the base metal on the welding procedure
qualification, (2) checking fabrication welds with a Telebrineller hardness tester to
be sure the welds do not exceed 200 BHN, per NACE Recommended Practice
RP-04-72 [25], and (3) checking weld procedure specifications to be sure active flux
is not used in submerged arc welds.
In some cases, stress relief of vessels used in sour service has been used to ensure
that both weld hardness and residual stresses are low, to minimize the risk of SSC.
Welding low-alloy steels, e.g., those containing chromium and molybdenum,
requires post-weld heat treatment in order to reduce heat-affected zone hardness.
For more information on welding requirements, refer to the Welding Manual, Piping
Manual, Pipeline Manual, and Pressure Vessel Manual.
Downhole tubulars of the following grades are suitable for sour service at all
temperatures:
API H-40
API J-55
API K-55
API C-75
API L-80
C-90SS
C-95SS
C-90SS and C-95SS refer to the Company's designations for manufacturer's special
proprietary, high-strength sour service grades of tubing and casing. These grades are
manufactured with careful control of chemistry and heat treatment in order to maxi-
mize their resistance to sulfide cracking. They retain their sulfide cracking resis-
tance at hardness above HRC 22, and MR-01-75 permits their use to HRC 26.

January 2001 400-98 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Because it is critical that these grades be properly manufactured to retain cracking


resistance at these high-strength levels, the Company has its own purchase specifi-
cation for C-90SS and C-95SS. Specification E-350.10 [26] is for casing and
E-350.16 [27] is for tubing. These specifications have requirements beyond those of
API C-90 and API T-95 (a sour service grade), and API C-95 (not a sour service
grade), and should be used for the purchase of C-90SS and C-95SS tubulars. For
more information refer to the Drilling Technology Centers Drilling Reference
Series [28], or contact the CRTC Materials and Equipment Engineering.
9Cr-1Mo, 13Cr, and duplex stainless steel alloy tubulars can be used in sour service
with a limited amount of H2S. Current guidelines are 1.5 psia maximum H2S partial
pressure for Grade L-80 tubing and 10 psia maximum H2S partial pressure for well-
heads. Higher-alloyed corrosion resistant alloys (CRAs) i.e., Sanicro 28, Alloy 825,
Hastelloy G-3, Hastelloy C-276, can be used in more severe environments with
higher concentrations of H2S. The use and limitations of all these alloys in sour
service is an area of ongoing research both within the Company and the industry.
When considering the use of alloys for downhole or wellhead completions, contact
the CRTC Materials and Equipment Engineering for guidance.

460 Hydrogen Embrittlement and Delayed Fracture


Hydrogen embrittlement is a deterioration of a metals mechanical properties caused
by the presence of atomic hydrogen within the metal. Hydrogen embrittlement is a
low temperature phenomenon seldom encountered above 200F; it primarily affects
high strength steels. It can occur without any apparent changes to the visual appear-
ance or the microscopic structure of the metal. See Figure 400-39 for a summary of
damage caused by hydrogen embrittlement and how it compares to other hydrogen
damage mechanisms.
Steel can be embrittled when hydrogen is present in the matrix, but the original
properties of the metal return when the hydrogen diffuses out. Hydrogen embrittle-
ment is the basic cause of several different types of failures such as hydrogen-
assisted cracking, hydrogen stress cracking, hydrogen embrittlement cracking,
hydrogen cracking of welds (delayed cracking, cold cracking, or underbead
cracking of welds), and sulfide stress cracking. The term used to describe the failure
mechanism depends on the characteristics of the metal and the hydrogen source.
Because sulfide stress cracking (SSC) is such a widespread concern to both
upstream and downstream operations, it is discussed separately in Sections 451
through 459. Hydrogen cracking of welds is discussed in the Welding Manual
Section 140. Other examples of hydrogen embrittlement failures include the
following:
high-strength steel bolts exposed to hydrofluoric acid
high-strength steel or K-Monel bolts exposed to sea water, particularly if
cathodic protection is applied
electroplated steel springs

Chevron Corporation 400-99 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

461 Mechanism of Hydrogen Embrittlement


Hydrogen embrittlement depends on the presence of atomic hydrogen in the metal.
In a hydrogen gas environment, hydrogen molecules have a finite tendency to disso-
ciate into the atomic form, so that some free hydrogen atoms are always present. At
ambient temperatures, the concentration of hydrogen atoms is insignificant unless
the hydrogen gas partial pressure is several thousand psi. At elevated temperatures,
however a significant concentration of hydrogen atoms exists at partial pressures as
low as a few hundred psi. Hydrogen embrittlement can result from rapid cooling of
hot, high pressure hydrogen equipment.
During aqueous corrosion, atomic hydrogen is formed, although it is in a transitory
and thermodynamically unstable state, i.e., it wants to recombine into molecular
hydrogen, H2. However, the presence of compounds such as sulfides and cyanides
can retard the recombination of atomic hydrogen into molecular hydrogen, thus
providing a ready supply of atomic hydrogen. (Additional discussion of recombina-
tion poisons is contained in Section 450.)
Atomic hydrogen can also be produced by cathodic protection currents or electro-
plating, or by the presence of moisture or oil during welding. The small size of the
hydrogen atom relative to other atoms and to the spacing between metallic atoms in
a metal allows hydrogen atoms to easily diffuse into metals.
Hydrogen embrittlement is a two-stage process. In the first stage the atomic
hydrogen is charged into the steel, but there is no damage. Tensile testing of
hydrogen-charged steel reveals a decrease in ductility (decrease in elongation and
reduction of area) and a decrease in stress to cause fracture. The fracture surface
also has a brittle appearance. However, if the atomic hydrogen is removed before
any cracking occurs, by heating for a short time in the absence of hydrogen to
between 450F and 600F, normal mechanical properties are restored. This can also
be accomplished by aging for a longer time at ambient temperature.
Cracking (i.e., permanent damage) can occur if the stress level on the hydrogen-
charged steel is high enough and the temperature is less than 200F. Hydrogen
cracking is the second stage in the hydrogen embrittlement process.
In higher-strength steels, the presence of hydrogen reduces the energy required for a
crack to grow and the stress at which unstable brittle fracture can occur [18]. It is
this mechanism that is of primary practical significance in oil industry equipment.
The mechanism of hydrogen embrittlement in high strength steels is still not fully
understood and is currently the subject of worldwide research. Mechanisms of
hydrogen embrittlement in other alloy systems are less understood still, and like-
wise the subject of research.

January 2001 400-100 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

462 Conditions Leading to Hydrogen Embrittlement


The presence of hydrogen in a metal does not necessarily mean a failure will occur.
For hydrogen cracking failure to occur, the following four conditions must be met
simultaneously:
The metal must be susceptible to hydrogen embrittlement.
The metal must be charged with hydrogen.
The temperature of the metal must be in the hydrogen embrittlement range
(about -50F to 200F, with the greatest susceptibility at around 70F).
A tensile stress (either applied or residual) above a threshold level must be
exerted on the metal.

Susceptibility of Metal to Hydrogen Embrittlement


Hydrogen has little or no effect on the properties or behavior of metals with a face-
centered cubic crystal structure (e.g., austenitic steels such as 18% chromium-8%
nickel stainless steels) except at very high hydrogen contents. On the other hand, all
ferritic steels (body-centered cubic crystal structure) are affected in varying degrees.
For ferritic steels the susceptibility to hydrogen embrittlement increases markedly
with increasing strength level. In high-strength steels, relatively small amounts of
hydrogen lead to large changes in mechanical properties; but in lower-strength
steels the effect of hydrogen decreases. The reasons for this are not well understood.

Applied Stress Versus Time to Fracture


Hydrogen assisted cracking normally exhibits subcritical growth. This means that
cracks grow at a finite rate that depends on the hydrogen concentration, steel type,
applied stress intensity, temperature, and other factors. The time during which
equipment is exposed to conditions causing crack growth may be quite limited (e.g.
reactors during shut down periods) or they may be continuous. In either case, the
time required for cracks to grow from any initial size (say, from a weld defect) to
the size required to trigger brittle fracture is currently the subject of active research.
Service failures from hydrogen embrittlement are generally delayed fractures. These
can occur at relatively low stress levels after the steel has been in service for a rela-
tively long time. This phenomenon is illustrated by Figure 400-64. The notched
tensile specimen is a tensile test specimen that has a sharp notch machined into the
gage section to act as a stress concentrator, as opposed to a smooth gage section
tensile specimen.
In Figure 400-64, there is a region in which the applied stress has very little effect
on the time-to-fracture. This stress is commonly called the upper critical stress and
may be much lower than the strength of a hydrogen-free notched specimen. At
stresses above the upper critical stress, fracture occurs without a time delay.
There is also a minimum stress below which fracture does not occur. It is commonly
called the lower critical stress. At stresses below the lower critical stress, hydrogen
is not damaging. The principal factors affecting the lower critical stress are
hydrogen concentration, stress concentration or notch severity, the strength of the

Chevron Corporation 400-101 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Fig. 400-64 Schematic Drawing of Characteristic Delayed Fracture Behavior of Notched


Tensile Specimens Containing Hydrogen (From J.W. Combs et al, Environ-
mental Behavior of 2-1/4 Cr - 1 Mo Hydrocracker Reactor and Piping Materials,
from paper given at API Refining Division meeting, May, 1972.)

steel, temperature, and the test duration. At intermediate stress levels between the
upper and lower critical stress, failure is inevitable provided the stress is maintained
long enough.
Fracture in hydrogen-charged specimens occurs as a series of events: incubation,
crack initiation, crack propagation and final fracture [19]. This behavior is repre-
sented in Figure 400-64 by two stress-time curves, one for crack initiation and one
for fracture. The crack initiation curve represents the incubation period for crack
formation, or the onset of irreversible damage. The fracture curve is simply the time
to complete rupture of the specimen.
An increase of hydrogen content in the steel, or an increase in notch severity (i.e.,
stress concentration) normally leads to a decrease in both the upper and lower crit-
ical stresses and shortens the incubation period. In general, increasing the strength
of the steel also decreases both the upper and lower critical stresses and shortens the
incubation period, but the importance of strength depends on the hydrogen content
of the steel. In severe hydrogen charging environments, such as aqueous solutions
of hydrogen sulfide, increasing the strength of the steel decreases both the time to
fracture and the lower critical stress.

January 2001 400-102 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

The combined effects of hydrogen concentration, notch severity, strength of the


steel, and temperature make it difficult to generalize about the magnitude of the
lower critical stress. However, ferritic steels with a yield strength of less than
90,000 psi, a tensile strength of less than about 110,000 psi, or a hardness of less
than about 22 HRC (all three properties are generally related) are virtually immune
to hydrogen cracking in most petroleum refining and producing services as long as
the yield strength of the steel is not exceeded.

463 Hydrogen Outgassing of Reactors


The equilibrium hydrogen content of a steel depends on temperature and the
external partial pressure of hydrogen. The equilibrium hydrogen content decreases
with temperature and hydrogen partial pressure, the effect of temperature being
much greater than hydrogen partial pressure. At ambient temperatures the equilib-
rium atomic hydrogen content of a steel is essentially zero unless the external
hydrogen partial pressure is several thousand psi. Therefore, hydrogen embrittle-
ment cannot occur unless the steel is supersaturated with hydrogen.
Supersaturation occurs when a steel is cooled too quickly to allow equilibrium by
diffusion of atomic hydrogen out of the steel. Allowing enough time for diffusion to
reduce the amount of supersaturated hydrogen to a safe level is called hydrogen
outgassing. The kinetics of hydrogen outgassing have been studied and outgassing
procedures described in a CRTC Materials and Equipment Engineering report, Safe
Hydrogen Levels for Shutdown of Hydrocrackers [20]. Failure to properly follow
these guidelines can result in an as-yet undefined degree of crack growth, up to and
including potential failure. Also, a significant degree of subcritical crack growth
could occur and remain undetected until a future inspection or excursion.
High-temperature, high-pressure hydrogen processes such as hydrofining and
hydrocracking charge an equilibrium amount of hydrogen into pressure vessels,
piping, and other process equipment by diffusion. At operating temperatures, this
thermally charged hydrogen is not damaging, provided the steel has sufficient alloy
content to resist high-temperature hydrogen attack (see Section 440). If sufficient
hydrogen is not allowed to diffuse out of the steel during the shutdown process to
maintain equilibrium as the metal temperature decreases, hydrogen cracking may
occur. As previously discussed, the susceptibility to hydrogen embrittlement, and
thus the need for hydrogen outgassing, depends on the strength of the steel, the
amount of hydrogen charged into the steel, and the applied or residual stress,
including the effects of stress concentration.
The amount of hydrogen thermally charged into steels in current hydrofining and
hydrocracking processes (825F to 850F maximum temperature) is normally insuf-
ficient to cause hydrogen embrittlement in steels with tensile strengths less than
100,000 psi (including welds and heat-affected zones). Thus, hydrogen outgassing is
normally not needed.

Chevron Corporation 400-103 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

464 Effects of Hydrogen Embrittlement on Fatigue


Hydrogen reduces the fatigue strength of steel [21] [22] [23]. While the effect is
more pronounced in high strength steel, even low strength steel is significantly
affected. Atomic hydrogen within the steel increases fatigue crack propagation rates
significantly, even though it apparently has little effect on the initiation of fatigue
cracks. The increase in crack propagation rate is believed to be a result of hydrogen
embrittlement.
Pressure swing absorber vessels have also been found to suffer hydrogen assisted
fatigue cracking, including through-wall cracks. This equipment cycles from about
400 psi to less than 50 psi on 20 minute intervals, in a predominantly hydrogen
atmosphere. The cyclic stress combined with stress concentrations at tray support
attachment welds to the shell resulted in rapid fatigue cracking. Figure 400-65
shows the detrimental effect of hydrogen, and the enormous acceleration of crack
growth rates in low strength pressure vessels steels such as SA-516 Gr. 70. Equip-
ment subjected to cyclic stress in hydrogen service requires special design precau-
tions to minimize stresses, stress concentrations, and weld defects which can
contribute to fatigue cracking.
Care must be taken in designing high-pressure hydrogen piping close to compres-
sors or other machinery capable of causing vibration. (For more information on
vibrating piping design, see the Piping Manual.) Fatigue failures have occurred
more frequently in hydrogen piping than in piping handling other gases. Small-
diameter pipe nipples used for branch connections between pieces of equipment
have been particularly troublesome. Hence, bridge welds are used around compres-
sors in high-pressure plants to minimize stress concentrations.
Investigators have found that hydrogen sulfide has a larger effect on fatigue strength
than does hydrogen gas. This is consistent with a hydrogen embrittlement mecha-
nism: at ambient temperatures moist hydrogen sulfide gas charges steel with more
atomic hydrogen than does hydrogen gas.

January 2001 400-104 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Fig. 400-65 Effect of Hydrogen Atmosphere on Fatigue Crack Growth Rates Under Conditions Typical of Pressure
Swing Absorbers (400 psia hydrogen) (Courtesy of UOP Equipment Systems Technology Group)

Chevron Corporation 400-105 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

470 References
1. Copson, H. R., Physical Metallurgy of Stress Corrosion Fracture, Interscience
Publishers, p. 2247 (1959).
2. Evans, U. R., The Corrosion and Oxidation of Metals. St. Martins Press Inc.,
1960.
3. Shreir, L. L., ed. Corrosion. John Wiley & Sons, Inc., 1963.
4. Kubaschewski, O., B. E. Hopkins., Oxidation of Metals and Alloys. Academic
Press Inc., 1962.
5. U.S. Steel Bulletin, Steels for Elevated Temperature Service, 4th Printing, 1961,
p. 15.
6. Babcock and Wilcox Technical Bulletin 6G, Properties of Carbon and Alloy
Steel Tubing for High Temperature and High Pressure Service.
7. Zima, G. E., Trans. American Society for Metals, Vol 49, p. 924, 1957.
8. Dunn, J. S., Proc. Jour, Inst. Met., Vol. 46, p. 25, 1931.
9. Nelson, G. A., Action of Hydrogen on Steel at High Temperature and High
Pressures Welding Research Council Bulletin No. 145, October 1969,
pp. 3342.
10. Anticipated Effect of Stainless Steel Cladding on Hydrogen Attack and Blis-
tering of Pressure Vessel Steels, Materials and Equipment Engineering Report,
File 45.55, July 29, 1974.
11. The Materials Properties Council, Inc. Evaluation of Advanced Steels for
SOHIC Resistance, Draft Topical Report on Task 1 of the Joint Industry Spon-
sored Research Program, Hydrogen Induced Cracking, Phase II, HIC-II-19,
October 1996.
12. Merrick, R. D., Refinery Experiences with Cracking in Wet H2S Environments.
Presented at NACE Corrosion/87, San Francisco, March 1987, Paper No. 190.
13. Merrick, R. D. and M. L. Bullen., Prevention of Cracking in Wet H2S
Environments. Presented at NACE Corrosion/89, New Orleans, April 1989,
Paper No. 269.
14. Failure Analysis of Amine-absorber Pressure Vessel, Materials Performance,
Vol. 26, No. 8, pp. 18-24 (1987) August. NACE International, Houston, Texas.
15. Wet H2S Cracking Flow Diagrams/Wet H2S Notebook, CRTC Materials and
Equipment Engineering File: 18.40, December 23, 1996.
16. Avoiding Environmental Cracking in Amine Units, API Recommended Practice
945, First Edition, August 1990, Washington, D.C.
17. Bizzel, K.P., W.J. Carter, R. Klehn, and E.H. Niccolls, Chevrons Guide to Eval-
uating Equipment Integrity. CRTC Materials and Equipment Engineering File:
54.21, April 10, 1995.

January 2001 400-106 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

18. Interrante, C. G., Interpretive Report on Effect of Hydrogen in Pressure Vessel


Steels - Basic and Research Aspects, Welding Research Council Bulletin
No. 145, October 1969, p. 1 - 30.
19. Coombs, J. W., J. G. Kerr, R. J. Olsen, and P. Purgalis., Environmental
Behavior of 2 Cr-1Mo Hydrocracker Reactor and Piping Materials, Preprint
No 56-72. Paper presented at the 37th Mid-year Meeting of the American
Petroleum Institute's Division of Refining, New York, May 11, 1972.
20. Outgassing of Chevron Hydrotreater and Isocracker Reactors. Materials and
Equipment Engineering Report, File 45.115, October 23, 1974.
21. Harrison, J. D. and G. C. Smith. Effects of Hydrogen on the Fatigue Behavior
of Mild Steel Bar and Weld Metal, British Welding Journal, Vol 14, No. 9,
pp. 493502, September 1967.
22. Prowse, R. L. and M. L. Wayman. Effect of Environment on the Fatigue
Behavior of a Medium Carbon Steel, Corrosion, Vol 30, No. 8, pp. 280284,
August 1974.
23. Effect of Hydrogen on the Fatigue Strength of Steels. Materials and Equipment
Engineering Report, File 45.25, December 12, 1972.
24. Sulfide Stress Cracking Resistant Metallic Materials for Oil Field Equipment,
NACE Standard MR-01-75, National Association of Corrosion Engineers,
Houston, TX.
25. Methods and Controls to Prevent In-Service Cracking of Carbon Steel Welds in
P-1 Materials in Corrosive Petroleum Refining Environments. NACE Standard
RP-04-72, National Association of Corrosion Engineers, Houston, TX.
26. Specification E350.10, Specification for C908S and C958S Richmond Grade
Casing for Sour Service; available from Materials and Equipment Engineering
at CRTC, Richmond.
27. Specification E350-16, Specification for C8555, C-9055, and C-9555 Grade
Tubing for Sour Service; available from Materials and Equipment Engineering
at CRTC, Richmond.
28. Drilling Reference Series. Available from the Drilling Technology Center,
Houston.

Resources
Andersson, B., Corrosion, Vol. 18, p. 425t (1962).
ASM Metals Handbook, Ninth Edition, Volume 13, 1987, p 123, p 239.
Baldy, M. F., and R. C. Bowden, Jr., Corrosion, Vol. 11, p. 417t (1955).
Bradley, B. W., and N. R. Dunne, Corrosion, Vol. 13, p. 239t (1957).
Cauchois, L., J. Didier, and E. Herzog, Corrosion, Vol. 13, p. 263t (1957).
Copson, H. R., and C. F. Cheng, Corrosion, Vol. 12, p. 647t (1956).

Chevron Corporation 400-107 January 2001


400 Characteristic Corrosion Phenomena Corrosion Prevention and Metallurgy Manual

Copson, H. R., The Influence of Corrosion on the Cracking of Pressure Vessels,


Welding Journal, Feb. 1953.
Dravnieks, A., H. J. McDonald. High Temperature Corrosion of Metals, Corro-
sion, Vol. 5, July (1949), pp. 227233.
Edeleanu, C. A., and P. P. Snowden, Jour. Iron & Steel Institute, Vol. 186, p. 406
(1957).
Edmunds, G., E. A. Anderson, and R. K. Waring, Symposium on Stress Corrosion
Cracking of Metals, ASME-AIME, p.7 (1944).
Effinger, R. T., M. L. Renquist, A. Wachter and J. G. Wilson., Hydrogen Attack in
Refinery Equipment, Proceedings of American Petroleum Institute, Vol 31M,
1951.
Fontana, M. G., Corrosion Engineering, Third Edition. McGraw-Hill, 1986.
Fontana, M. G., F. H. Beck, and J. W. Flowers, Metal Progress, Dec. 1961.
Fraser, J. P., and R. S. Treseder, Corrosion, Vol. 8, p. 342 (1952).
Fraser, J. P., and G. G. Eldredge, Corrosion, Vol. 14, p. 525t (1958).
Halsted, W. G., High-Temperature Oxidation in Catalytic Cracking Regenerators,
Materials and Equipment Engineering File 18.45, 3/8/79.
Heindhofer, K. B. M. Larsen, Rates of Scale Formation on Iron and a Few of its
Alloys, Transactions of the American Society for Steel Treating, Vol. 21, October
(1933), pp. 865895.
Hoar, T. P., Corrosion, Vol. 19, p. 331t (1963).
Hoar, T. P., and J. G. Hines, Stress Corrosion Cracking and Embrittlement, John
Wiley & Sons, p. 107 (1956).
Hoke, J., F. Eberle., American Society of Mechanical Engineers,
Paper No. 57-A-175, 1957.
Kirk, W. W., F. H. Beck, and M. G. Fontana, Physical Metallurgy of Stress Corro-
sion Fracture, Interscience Publishers, p. 227 (1959).
Lenhart, S. J., High Temperature Oxidation Relationship Between Time, Tempera-
ture, and Oxide Thickness, Materials and Equipment Engineering File 18.45, 8/1/79.
MacQuigg, C. E. 25 to 30 Per Cent Chromium-Iron Alloys, The Book of Stainless
Steels, ASM, 1935, p. 351
McGlasson, R. L., W. D. Greathouse, and C. M. Hudgins, Corrosion, Vol. 16,
p. 557t (1960).
Miley, H. A., Theory of Oxidation and Tarnishing of Metals, Transactions of the
Electrochemical Society, Vol. 81 (1942), pp 391411.

January 2001 400-108 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 400 Characteristic Corrosion Phenomena

Minutes of NACE Tech. Committee T-1F-2 Meeting, NACE South Central Region
Conference, Oklahoma City, October 2, 1957.
Morgan, E. R., V. F. Zackay., Ductile Iron-Aluminum Alloys, Metals Progress, 68,
p. 126, Oct. 1955.
Nielsen, N. A., Physical Metallurgy of Stress Corrosion Fracture, Interscience
Publishers, p. 121 (1959).
Piehl, R. L., Stress Corrosion Cracking in Sulfur Acids, API paper presented at
29th mid-year meeting, May 11, 1964; preprint no. 1664.
Priest, D. K., F. H. Beck, and M. G. Fontana, Trans. ASM, Vol. 47, p. 473 (1955).
Rahmel, A., W. Jaeger, and K. Becker., Archiv Eisenhuttenwesen, 30, p. 351, 1959.
Revised Guidelines for Carbon- Molybdenum Steel in Hydrogen Service,
Materials and Equipment Engineering Report, File 45.55, April 30,1987.
Scharfstein, L. R., and W. F. Brindley, Corrosion, Vol. 14, p. 588t (1958)
Scheil, M. A., Symposium on Stress Corrosion Cracking of Metals, ASTM-AIME,
p. 395 (1944).
Schuetz, A. E., and W. D. Robertson, Corrosion, Vol. 13, p. 437t (1957).
Snowden, P. P., and Jour, Iron and Steel Institute, Vol. 194, p. 181 (1960).
Uhlig, H. H. and R. W. Revie., Corrosion and Corrosion Control, Third Edition.
John Wiley & Sons, Inc., 1985.
Williams, W. L., and J. R. Eckel, Jour. Am. Soc. Naval Engineers, Vol. 68, (1956).

Chevron Corporation 400-109 January 2001

Вам также может понравиться