Вы находитесь на странице: 1из 83

SECTION 2

CORROSION and DAMAGE


MECHANISMS
CALTEX REFINERY MATERIALS MANUAL October 1999

SECTION 2
CORROSION AND DAMAGE MECHANISMS

TABLE OF CONTENTS

I. 0 ABSTRACT ................................................................
................................................................................................
...............................................
............... 1

2.0 INTRODUCTION ................................................................


................................................................................................
......................................
...... 1

3.0 CORROSION PRINCIPLES


PRINCIPLES ................................................................
..........................................................................................
.......................... 3
3.1 Low Temperature Refinery Corrosion ............................................................................................3
3.2 High Temperature Refinery Corrosion ...........................................................................................5

4.0 PROCESS CORROSION


CORROSION ................................................................
.............................................................................................
............................. 9
4.1 Low Temperature Conditions........................................................................................................9
4.2 High Temperature Conditions.....................................................................................................10

5.0 FORMS AND CAUSES ................................................................


.............................................................................................
............................. 11
5.1 Forms Of Metal Loss Due To General And/Or Localized Corrosion ..............................................11
5.1.1 Crevice Corrosion ...............................................................................................................11
5.1.2 Erosion Corrosion................................................................................................................12
5.1.3 Fuel Ash Corrosion ..............................................................................................................12
5.1.4 Galvanic Corrosion..............................................................................................................13
5.1.4.1 Oxidation/Scaling Resistance ........................................................................................................ 15
5.1.5 High Temperature Sulfidation (With and Without Hydrogen) ..............................................15
5.1.6 Intergranular Corrosion.......................................................................................................18
5.1.7 Microbiologically Influenced Corrosion (MIC) .....................................................................18
5.1.8 Oxidation ...........................................................................................................................20
5.1.9 Pitting.................................................................................................................................21
5.2 Specific Causes of General/Localized Corrosion...........................................................................27
5.2.1 Aluminum Chloride.............................................................................................................27
5.2.2 Amine ................................................................................................................................27
5.2.3 Ammonium Bisulfide (NH4HS).............................................................................................27
5.2.4 Atmospheric Corrosion.......................................................................................................28
5.2.5 Boiler Feedwater .................................................................................................................29
5.2.6 Cooling Water Corrosion ....................................................................................................29
5.2.7 Carbon Dioxide (CO2).........................................................................................................30
5.2.8 Corrosion Under Insulation .................................................................................................31
5.2.9 Hydrogen Chloride .............................................................................................................31
5.2.10 Hydrofluoric Acid................................................................................................................32
5.2.11 Naphthenic Acid ................................................................................................................33
5.2.12 Organic Chlorides ...........................................................................................................34
5.2.13 Phenol (Carbolic Acid).......................................................................................................35
5.2.14 Phosphoric Acid .................................................................................................................36
5.2.15 Process Chemicals ...........................................................................................................36
5.2.16 Soil Corrosion....................................................................................................................36
5.2.17 Steam Condensate ..........................................................................................................37
5.2.18 Sulfuric Acid ......................................................................................................................37
5.3 STRESS CORROSION CRACKING (SCC) ......................................................................................38
CALTEX REFINERY MATERIALS MANUAL October 1999

5.3.1 Chloride Stress Corrosion Cracking .....................................................................................40


5.3.2 Caustic Stress Corrosion Cracking .......................................................................................42
5.3.3 Carbonic Acid .....................................................................................................................43
5.3.4 Polythionic Acid Stress Corrosion Cracking (PASCC)............................................................43

Table of Tables
Table 2-1 Galvanic Series of Metals And Alloys in Sea Water ..............................................................14
Table 2-2 Maximum Skin Temperatures ............................................................................................21
Table 2-3 Corrosives Found in Many Refining Processes.....................................................................23
Table 2-4 Alloy Systems Subject to Stress Corrosion Cracking............................................................40
Table 2-5 Operating Conditions for C-O.5 Mo Steels........................................................................51
Table 2-6 Minimum Temperature for Creep ......................................................................................73

Table of Figures
Figure 2-1 Corrosion of Carbon Steel by Sulfuric Acid.38
Figure 2-2 Average Corrosion Rate McConomy Curves24
Figure 2-3 Cooper-Gorman Curves.25
Figure 2-4 Conditions Requiring Stress Relief of Carbon Steel in Caustic Service...41
Figure 2-5 Time-Temperature-Sensitization (TTS) Curves for the Commonly-Used
300-Series Stainless Steel...45
Figure 2-6 Operating Limits for Steel in Hydrogen Service to Avoid Decarburization & Fissuring55
Figure 2-7 Time For Incipient Attack Of Carbon Steel In Hydrogen Service.56
Figure 2-8 Experience with C-0.5Mo & Mn-0.5Mo Steel in High Temperature Hydrogen Service..57
Figure 2-9 Time for Incipient Attack Of 0.5Mo Steels in High Temperature Hydrogen Service....58
Figure 2-10 Microstructure Replicas for Creep-Life Estimation.....73
Figure 2-11 Relative 100,000 Hour Rupture Strength74
Figure 2-12 Variation of Larson-Miller Creep Rate Parameter with Stress for 5Cr - 1/2Mo Steel...75
Figure 2-13 5Cr - 1/2 Mo Steel ASTM A213 T5, A 335 P5, A200 T5.76
CALTEX REFINERY MATERIALS MANUAL October 1999

I. 0 ABSTRACT
Corrosion and other damage often leads to failures in refining processes at high and low
temperature conditions, the type and mechanism depending on the process conditions and
contaminants in the process. This section provides a brief overview of some of the types of
corrosion and other damage mechanisms active in typical refinery processes. It includes a short
description of low and high temperature corrosion principles and many of the process conditions
that drive corrosion.

2.0 INTRODUCTION
Refinery corrosion can be split into low temperature corrosion and high temperature corrosion.

Low temperature corrosion occurs at temperatures below 500oF (260oC), usually in the presence
of water. It is often also called aqueous corrosion, wet corrosion or electrochemical corrosion.
Low temperature corrosion requires the presence of an aqueous solution, even in very small
amounts, and an electrolyte in a hydrocarbon stream. In vapor streams, low temperature
corrosion is often found under dew point conditions where water collects.

High temperature corrosion occurs at temperatures above 500oF (260oC) with no water present.
Using oxygen as an example, corrosion is the result of a reaction between a metal and oxygen.
Sulfur compounds, steam, carbon containing gases, fuel ash components and naphthenic acids
are other chemicals which can cause severe high temperature corrosion in refinery process
equipment.

In refining applications, the material/environment interactions are extremely varied; many


refineries contain over 15 different process units, each having its own combination of corrosive
process streams and temperature/pressure conditions.

Once equipment is placed in service, it is subject to operating, upset, and/or downtime


conditions that can cause deterioration or damage to equipment from mechanisms other than
corrosion such as creep, fatigue, erosion and embrittlement. The refining industry has equipment
designed for 100,000 hours, many of which have successfully operated for over 450,000 hours,
because many of these mechanisms operate very slowly.

Changes in process temperature or pressure, upsets, overfiring of furnaces to increase


throughput, instrument failures or exposure to fire can occur. These conditions can produce
metallurgical failures when changes in the microstructure and/or chemistry of original material of
construction occurs. For example, furnace tubes start to sag or bulge, vessel walls become
distorted and develop cracks or blisters, and piping becomes embrittled. Metal deterioration may
result in rather serious consequences. Failures are often accelerated by cyclic changes, including
periodic shutdowns.

Process industry equipment is designed to withstand a set of process conditions (pressure,


temperature, fluids, etc.) for a defined period of time, often 100,000 hours. To ensure that the
equipment has been correctly designed and fabricated, major equipment items are subjected to
inspection and testing before being placed into service. However, smaller equipment items,
especially such commodity items as valves and bolts, are often not subject to the same inspection
and testing requirements. Material mixing can easily occur with these items. Failures are
frequently reported which have occurred due to the installation of incorrect materials.

Page 2-1
CALTEX REFINERY MATERIALS MANUAL October 1999

Equipment is also subject to mechanical damage either during installation or in operation.


Examples of this might include mechanical overload of structural members and over-tightening
of bolts. Other types of mechanical damage which might be encountered include: brittle
fracture, fatigue of vibrating or rotating parts, and failure due to accidental over pressurization.

Page 2-2
CALTEX REFINERY MATERIALS MANUAL October 1999

3.0 CORROSION
CORROSION PRINCIPLES

3.1 LOW TEMPERATURE REFINERY CORROSION


Low temperature (aqueous) corrosion obeys electrochemical laws but is often controlled by
diffusion processes. Except for gold and platinum, metals and alloys are not inherently corrosion
resistant. During initial exposure to the corrosive environment, stable films form which reduce or
prevent further attack. Corrosion of a metal consists of two or more partial reactions of oxidation
and reduction and requires the presence of an electrolyte, such as water. Oxidation occurs at
anodic sites on the metal, reduction at the cathodic sites. The anodic reaction in every corrosion
process is the oxidation of a metal to its ionic form:

M MC+n + ne -

Typical cathodic reactions are the following:

2H+ + 2e- H2 (gas) hydrogen evolution

O2 + 4H+ + 4e- 4OH- oxygen reduction in acid solutions

O2 + 2H2O + 4e- 4OH- oxygen reduction in neutral or basic solutions

M+3 + e- M+2 metal ion reduction

M+ + e- M metal deposition (plating)

Hydrogen evolution and oxygen reduction are among the more common cathodic reactions. In
refinery equipment, the bisulfide reduction is also common and important:

2HS-+ 2e- H2 (gas)+ 2S-2

When iron or steel is in contact with water and exposed to the atmosphere, corrosion occurs. The
anodic reaction is:

Fe Fe+2 + 2e-

Since the water contains dissolved oxygen from air, the cathodic reaction is:

O2 + 2H2O + 4e- 4OH-

The overall corrosion reaction is obtained by addition of the anodic and cathodic reactions:

2Fe + 2H2O + 02 2Fe+2 + 4OH- 2Fe(OH)2 (solid ferrous hydroxide)

Ferrous hydroxide, which precipitates from solution, is oxidized to ferric hydroxide, as follows:

2Fe(OH)2 + H2O + 1/2 02 2Fe(OH)2 (solid ferric hydroxide)

Page 2-3
CALTEX REFINERY MATERIALS MANUAL October 1999

Ferric hydroxide is known as rust.

In practice, corrosion reactions are considerably more complex. When an alloy corrodes, its
elements go into solution as their own respective ions. Several cathodic reactions can occur at
the same time. If the water in the above example had been acidified, hydrogen evolution would
have taken place, in addition to oxygen reduction. The rates of anodic and cathodic reactions
must be equal. Therefore, two or more cathodic reactions result in greater electron consumption
and accelerate the anodic reaction.

Corrosion rates are measured as weight loss per unit area and are expressed in mils (0.001 inch)
of penetration per year (mpy). Corrosion rates below about 5 to 10mpy are generally considered
acceptable for long term service, but higher rates are routinely experienced and can be safely
tolerated if an effective corrosion monitoring program is in place.

Corrosion can be decreased by reducing the rate of either the anodic or cathodic reaction or
both. For example, iron will not corrode in deaerated water simply because oxygen reduction
can not take place. Corrosion inhibitors are formulated to retard the anodic or cathodic reaction.
Other corrosion inhibitors form a protective, non-conducting film on the metal surface. Paints
and protective coatings prevent corrosion in a similar manner.

Polarization retards the kinetics of electrochemical reactions including corrosion. Activation


polarization occurs when corrosion is controlled by the reaction sequence at the metal surface.
For example, hydrogen ions must be absorbed on the corroding surface before hydrogen
reduction can take place. Electron transfer must occur next forming atomic hydrogen. Two
hydrogen atoms then combine to form hydrogen gas which bubbles off the metal surface. If
hydrogen reduction is controlled by the slowest of these several reaction steps, corrosion is said
to be activation polarized. Corrosion in concentrated acids is usually controlled by one or more
reaction steps at the metal surface. Hydrogen blistering in vapor recovery units is controlled by
activation polarization in the presence of hydrogen sulfide.

Concentration polarization occurs when corrosion is controlled by diffusion in the corrosive


environment. With hydrogen evolution, corrosion is concentration polarized if hydrogen ion
diffusion becomes the rate controlling step. Corrosion in very dilute acids usually depends on ion
diffusion. Cooling water and steam condensate corrosion are controlled by concentration
polarization. Process changes will produce different results depending on the type of polarization
that controls the reactions. For example, lowering flow velocity will decrease corrosion only if the
cathodic reaction is controlled by concentration polarization.

Passivity refers to the increase in corrosion resistance of certain metals and alloys as a result of the
formation of protective surface films. Passivity occurs with most common engineering and
structural metals. If the protective films are destroyed, corrosion rates can easily increase many
thousand times, and the metal is said to have become active. The protective films usually are
stable over a wide range of conditions but are damaged or destroyed in highly reducing or
oxidizing environments. Active ions, such as chlorides, can undermine the integrity of surface
films. As a result, materials can suffer pitting corrosion and/or stress corrosion cracking (SCC).
For this reason, refineries are reluctant to use austenitic stainless steels in aqueous services.

Oxygen content is an important variable because metals and alloys that form protective oxide
films require a sufficiently oxidizing environment to maintain passivity. Usually, there is enough
dissolved oxygen in refinery service water to maintain the passivity of stainless steel or titanium
but not enough to passivate carbon steel. In contrast, chromates make effective cooling water
inhibitors because they readily oxidize and passivate carbon steel surfaces. However, chromates

Page 2-4
CALTEX REFINERY MATERIALS MANUAL October 1999

have been eliminated from cooling water systems due to environmental concerns with heavy
metals.

Most refinery streams contain hydrogen sulfide which produces a highly reducing environment.
Air contamination, as might occur from leakage on the suction side of pumps, reacts with
hydrogen sulfide, forming elemental sulfur and can result in sulfur plugging. In this case, air
contamination has no effect on corrosion unless large air pockets are formed. However, air
contamination of boiler feed water for steam generation can have serious consequences in boiler
tubes. For this reason, boiler feed water is chemically deaerated by treating with hydrazine or
sodium sulfite.

Flow velocity changes have no effect on corrosion processes which are controlled by activation
polarization. Corrosion processes which are controlled by concentration polarization will proceed
at a greater rate when flow velocity is increased. Extensive use of carbon steel in refineries is only
possible because of the protective sulfide scale that forms in the presence of aqueous hydrogen
sulfide. When exposed to high flow velocities, this sulfide scale can be damaged and removed,
resulting in severe, localized attack of the underlying metal. Sulfuric acid corrosion of carbon steel
is highly sensitive to velocity effects because resistance depends on a protective surface film that
exists on the steels surface. This should be kept in mind when operational changes are
contemplated for alkylation units and other systems handling sulfuric acid.

As with almost all chemical reactions, temperature increases corrosion rates. Another temperature
effect also needs to be considered in refinery operations. Increased temperatures may increase
the amount of water in liquid hydrocarbon and vapor streams. This means that more water is
likely to condense out in downstream distillation towers or in overhead condensing systems. As a
result, corrosion can occur in equipment that was thought to be dry.

Corrosion by strong acids, such as concentrated sulfuric acid in alklylation units, is highly
dependent on temperature. Carbon steel can be used for these units primarily because process
temperatures are relatively low. Stainless steels can exhibit a drastic change in corrosion
resistance as temperature reaches a certain level. This can manifest itself as sudden loss of
passivity, causing the corrosion rate to increase by a significant factor.

Concentration increases in the corrosive environment generally increase corrosion rates.


However, corrosion in concentrated acids often is minimal because water is absent. When a
material, such as stainless steel, is passive in a given environment, relatively large changes in
concentration are required to produce significant changes in corrosion. In refinery streams, the
concentration of a corrosive constituent in a hydrocarbon stream must be considered in addition
to the amount of associated water present. For example, carbon steel is severely attacked by
dilute sulfuric acid. Very little corrosion occurs, however, if droplets of this dilute acid are
dispersed in large volumes of liquid hydrocarbon.

3.2 HIGH TEMPERATURE REFINERY CORROSION


Oxidation was used to describe metal dissolution during low temperature corrosion. This
example will again be used to describe the reaction between a metal and oxygen at high
temperatures [above 500oF (260oC) with no water present]. For this example, oxygen causes
severe scaling and cracking of carbon steel above 1000oF (538oC) and ultimately converts it into
a brittle mass of iron oxide.

Page 2-5
CALTEX REFINERY MATERIALS MANUAL October 1999

The mechanisms, scale-morphology and kinetics of various types of high temperature corrosion
are quite similar, so discussion of these aspects will cover primarily reactions with oxygen. In
some cases, however, no scale is formed.

High temperature corrosion, as in the case with low temperature corrosion discussed earlier, is an
electrochemical process which consists of two or more partial reactions. Typically, these involve
oxidation and reduction reactions. In the case of exposure to air, metal is oxidized to an ion at
the metal scale interface:

M M+n+ ne-

At the same time, oxygen is reduced at the scale surface:

1/2 O2 + 2e- O-2

The overall corrosion reaction is obtained by addition:

M + 1/2 O2 MO

Metal oxides serve a number of functions analogous to those in low temperature corrosion:

They must be able to conduct ions.


They must be able to conduct electrons.
They must serve as an electrode for oxygen reduction.

Electronic conductivity of most oxides is much greater than their ionic conductivity. This means
that the reaction rate depends on the diffusion rates of either metal ions or oxygen ions, or both.
It should be obvious that oxidation can be controlled if these diffusion rates could be reduced in
some fashion, however, no practical methods of achieving this have been found. Instead,
oxidation resistance is improved by alloying so that more protective oxides are formed in the
scale.

Scale invariably consists of a number of different stable components. For example, when carbon
steel is oxidized, layers of FeO, Fe3O4 and Fe2O3 are formed in sequence. The layer containing the
highest proportion of oxygen (Fe2O3) will be found at the outer scale surface; the layer with the
highest proportion of iron (FeO) will be at the steel/scale interface. The relative thickness of each
oxide layer depends on the rates of ion diffusion through that layer.

Oxide scales grow primarily at the scale surface by outward diffusion of metal ions. It has been
proposed that some scales grow by dissociation of inner oxide layers, sending metal ions outward
and oxygen molecules inward. Such scales grow both at the metal/scale interface and at the scale
surface. In actuality, morphological aspects of scale formation are far more complex than
described above. Dissolution of oxygen atoms in some metals, low melting points and high
volatility of some oxides, and the existence of grain boundaries within the metal and the scale
complicate interpretation of oxidation mechanisms. The same applies to high temperature
corrosion by other chemicals, except that the reactions become even more complex.

From a practical point of view, the rate of high temperature corrosion is an important parameter.
Scale usually adheres to metal surfaces thus rates are measured and expressed in terms of weight
gain per unit area (rather than weight loss per unit area which is used to measure low
temperature corrosion). High temperature corrosion of common refinery metals obeys one of
two empirical rate laws, as follows:

Page 2-6
CALTEX REFINERY MATERIALS MANUAL October 1999

Linear kinetics apply when weight gain due to corrosion scaling is proportional to exposure
time. Cracked or porous scales are formed which do not prevent diffusion of metal or
reactant. Scaling is controlled by the rate of molecular dissociation (or some other reaction
step) at the metal/scale interface or scale surface. For example, linear kinetics apply to
naphthenic acid corrosion.

Parabolic kinetics yield a straight line when weight gain data are squared and plotted versus
exposure time. Parabolic scaling rates are controlled by ion diffusion through a scale layer
which is continuously increasing in thickness. In general, high temperature corrosion of many
engineering alloys obeys parabolic kinetics, including oxidation and sulfidation. Parabolic
kinetics apply to the external scaling of carbon steel and low alloy furnace tubes.

There are other rate laws, but these apply to unusual metals and exposure conditions. Obviously,
linear kinetics are less desirable than parabolic kinetics for metals and alloys in high temperature
applications.

Page 2-7
CALTEX REFINERY MATERIALS MANUAL October 1999

Page 2-8
CALTEX REFINERY MATERIALS MANUAL October 1999

4.0 PROCESS CORROSION

4.1 LOW TEMPERATURE CONDITIONS


Most corrosion problems in refineries are not caused by hydrocarbons being processed but by
various inorganic compounds, such as water, hydrogen sulfide, hydrogen chloride, sulfuric acid,
carbon dioxide and others. Table 2-1 lists corrosives found in many refining processes. The two
general sources of these compounds are crude oil contaminants and process chemicals. In
addition, corrosion problems are caused by the atmosphere, cooling water, boiler feed water,
steam condensate and soil.

Crude oil contaminants are the major cause of low temperature refinery corrosion. Most are
present in crude oil as it is produced. Some are removed during preliminary treatment in the oil
fields. The remainder end up in refinery tankage, along with contaminants picked up in pipelines
or marine tankers. In most cases, the actual corrosives are formed during initial refinery
operations. For example, highly corrosive hydrochloric acid evolves in crude oil furnaces from
relatively harmless calcium and magnesium chlorides. Other corrosives can form from corrosion
products after exposure to air during shutdowns. The following discussion will highlight some of
the most important crude oil contaminants that cause refinery corrosion problems.

Water is found in all crude oils and is difficult to remove completely. Water is not only an
electrolyte but also hydrolyzes some inorganic chlorides to hydrogen chloride. Water is primarily
responsible for various forms of corrosion in distillation tower overhead systems. In general,
whenever equipment can be kept dry through suitable process changes, corrosion problems will
be minimized. The addition of air can be especially detrimental. Moisture and air drawn into
storage tanks during normal 'breathing' as a result of temperature changes and transfers, directly
relates to the amount of tank corrosion experienced.

Crude and heavy oils form a somewhat protective oil film on the working areas of a tank shell.
Corrosion in tanks handling these stocks is generally limited to the top shell ring and the
underside of the roof where protective oil films are minimal if they are not normally in contact
with the oil. Tank bottom corrosion occurs mostly with crude oil tankage and is caused by
separated water and salt entrained in the crude oil. A layer of water usually settles out on the tank
bottom and becomes highly corrosive. Cracks in the mill scale on tank bottoms form anodic areas
which pit while the remaining attached scale acts as a large cathode.

Corrosion in tanks which handle gasoline and other light stocks occurs primarily at the middle
shell rings because these see more wetting and drying cycles than other areas. Light stocks do
not form protective oil films. Pits become so numerous that metal loss may appear to be uniform
corrosion. The rate of corrosion is proportional to the water and air content of light stocks and
chloride and hydrogen sulfide contamination accelerate attack.

Refinery equipment can be exposed to moisture and air which can be pulled into the suction side
of pumps if seals or connections are not tight. Air and moisture can also be dissolved in
hydrocarbons that are stored in tanks where air and moisture were accessible. In general, air
contamination of hydrocarbon streams can be more detrimental with regard to fouling than
corrosion. Cooling water from open recirculating cooling towers is saturated with air and can be
corrosive to carbon steel heat exchanger tubing.

Hydrogen sulfide is present in sour crude oils and gases handled by most refineries. During
processing at elevated temperatures, hydrogen sulfide is also formed by decomposition of
organic sulfur compounds. It causes aqueous sulfide corrosion in overhead systems of various

Page 2-9
CALTEX REFINERY MATERIALS MANUAL October 1999

distillation towers, hydroprocessing effluent streams, in the light ends recovery section of FCC
units and cokers, in sour water strippers, and in amine plants. Considerable corrosion is also
caused by hydrogen sulfide in storage tanks containing sour crude oils. At temperatures above
450oF (232oC), hydrogen sulfide in combination with organic sulfur compounds or hydrogen,
causes high temperature sulfidation and hydrogen damage, to be discussed later.

Generally, carbon steel has fairly good resistance to aqueous sulfide corrosion because a
protective iron sulfide film is formed on the steel. Where excessive corrosion has been
experienced, resistant alloys have been used successfully. Localized corrosion in vessels has been
mitigated by selective lining with these alloys. Alloy tubes have been used as replacements for
carbon steel in condensers and coolers at a number of units where this type of corrosion is a
problem. Alloys such as titanium, duplex stainless steel, super austenitic stainless steel and
austenitic stainless steel have been used in these applications.

4.2 HIGH TEMPERATURE CONDITIONS


High temperature corrosion problems in refineries are of considerable importance. Equipment
failures can have serious consequences because processes at high temperatures may involve high
pressures as well. With hydrocarbon streams, there is always the danger of fire, if leaks or ruptures
occur.

High temperature corrosion depends on the nature of the scale that is formed. Uniform scale
reflects uniform attack, pitting occurs where scale has been locally damaged, and intergranular
attack occurs when grain boundaries corrode in preference to the grains. Many refinery processes
at elevated temperatures involve vapor or mixed vapor/liquid streams at high flow velocities. It is
not surprising to find, therefore, that high temperature corrosion often results in fatigue, erosion
and cavitation damage.

Attack by naphthenic acids differs from most other types of high temperature corrosion in that
no protective scale is formed. Damage appears as localized areas of uniform attack on carbon
steels, low alloy steels and ferritic or martensitic stainless steels containing 12% chromium. In
contrast, naphthenic acids cause pitting of austenitic stainless steels, such as Types 304 or 316,
due to breakdown of the passive oxide film which normally protects these alloys from corrosion.

On a more positive note, high temperature refinery corrosion is caused primarily by various sulfur
compounds. Over the years, extensive research has been done to establish the mechanisms of
various forms of high temperature sulfide corrosion. Fortunately, corrosion rate correlations are
available so that equipment life can be predicted with some degree of reliability. (refer Section 2-
5.1.5)

Page 2-10
CALTEX REFINERY MATERIALS MANUAL October 1999

5.0 FORMS AND CAUSES


In general, the following types of damage are encountered in refining equipment:

1. Corrosion - general / localized


2. Stress Corrosion Cracking
3. High Temperature Hydrogen Attack
4. Metallurgical changes (embrittlement)
5. Mechanical failures
6. Other failures

Each of these general types of damage are caused by one or more of many specific damage
mechanisms.

Examples of each type are:


corrosion (naphthenic acid corrosion of carbon steel)
stress corrosion cracking (polythionic acid stress corrosion cracking of stainless steel)
embrittlement (temper embrittlement of 2-1/4 Cr - 1 Mo alloy steel)

Each of these damage mechanisms occur under very specific combinations of materials and
environment/operating conditions. A general discussion of specific damage mechanisms and
refining process conditions conducive to each mechanism and damage mechanisms for
commonly used alloys in the refining industry follows.

The various types of corrosion can be conveniently classified by the appearance of the corroded
metal. In investigating corrosion problems, it is important to examine the equipment prior to
cleaning and repairing it so that the appearance of corrosion characteristics is not lost. Usually,
various types of corrosion are interrelated.

5.1 FORMS OF METAL LOSS


LOSS DUE TO GENERAL AND/OR LOCALIZED CORROSION

5.1.1 Crevice Corrosion


Crevice corrosion is associated with stagnant solutions in crevices, such as under bolt heads,
gaskets, washers and in threaded and lap joints. It also occurs in rolled tube-to-tubesheet joints,
under wet packing or insulation and under corrosion products. In the latter case, crevice
corrosion is also referred to as 'underdeposit attack'. Stainless steels are especially susceptible to
crevice corrosion in hot sea water environments. In refineries, crevice corrosion of carbon steel is
often seen under various deposits, and to a lesser extent, at gasket connections.

For crevice corrosion to occur, the crevice must be wide enough to allow liquid to enter and
narrow enough to maintain a stagnant condition. This means that crevice corrosion is usually
limited to openings which are less than a few mils wide. The mechanism of crevice corrosion is
similar to that of pitting corrosion, with the crevice acting as a relative large pit. At one time it
was thought that crevice corrosion was caused by differences in metal ion or oxygen
concentration between the crevice and the bulk liquid. This lead to the term 'concentration cell
corrosion' which has been erroneously used to describe this type of attack. Like pitting, crevice
corrosion usually occurs in the presence of chlorides, has an incubation period and, once started,
becomes an auto-catalytic process.

There are a number of ways in which crevice corrosion can be avoided:

Page 2-11
CALTEX REFINERY MATERIALS MANUAL October 1999

Equipment should be designed for proper drainage during downtime.


Solids deposition should be minimized by frequent cleaning, by-passing equipment if
necessary to keep a unit on stream. Welded connections are less likely to suffer crevice
corrosion than are flanged and bolted connections.
Wet packing should be taken out of critical equipment during long shutdown periods.
Low chloride insulation should be specified and insulation must be kept dry by proper
wrapping and caulking.

5.1.2 Erosion Corrosion


Erosion corrosion is accelerated attack due to flow velocity and mechanical factors in addition to
corrosion. Abrasion and mechanical wear accelerates the corrosion rate, or from another
viewpoint, corrosion aggravates abrasion. Typically, damage is in the form of grooves, gullies,
elongated holes and valleys having a directional pattern. Depending on the flow regime, damage
areas may be smooth or sharp edged.

Erosion corrosion occurs when protective surface films are damaged or worn away so that fresh
metal is continuously exposed to corrosion. For this reason, alloys of aluminum, chromium steels,
and stainless steels are especially subject to attack since they depend on a surface film for their
resistance to corrosion. Bends, elbows and tees of piping, pump cases and impellers, compressor
blades, valve internals, agitators, baffles, thermowells, and orifice plates are subject to various
forms of erosion corrosion. In general, any increase in velocity will increase erosion corrosion,
especially if suspended solids are involved. Flow turbulence at the inlet of heat exchanger tubes
can result in rapid corrosion of the first several inches of tubing.

Erosion corrosion due to droplets of liquid suspended in a vapor stream is a real problem in many
refinery applications. Known as 'impingement corrosion', this type of erosion corrosion occurs in
overhead piping and condensers of distillation towers when vapor velocities exceed 25 ft/sec (8
m/sec). The usual cause is water droplets containing dissolved hydrogen sulfide and hydrochloric
acid. Areas most likely to be attacked are elbows in overhead piping, condenser shell inlet
nozzles, and condenser upper tube rows. Condensed hydrocarbons or hydrocarbon carryover
from damaged trays can cause similar damage.

Corrosion inhibitors are usually ineffective as far as erosion corrosion is concerned. However,
there are a number of other ways in which erosion corrosion problems can be minimized:

Increase metal thickness to provide greater corrosion allowance. Install sacrificial


impingement baffles.
Streamline bends and remove any obstructions to smooth flow, such as rough weld crowns,
and use larger diameter pipe and fittings.
Install protective ferrules in tube inlet ends of heat exchanger bundles.
Rotate tube bundles from time to time to distribute impingement damage and maximize
bundle life.
Install a lining in the corroded areas with alloys that resist the corrosion.
Use titanium or other alloy heat exchanger tubes which are highly resistant to impingement
corrosion.

5.1.3 Fuel Ash Corrosion


Fuel ash corrosion has been a problem in heaters and boilers when burning high sulfur fuels with
50 ppm or more vanadium. The sulfates and vanadates formed during combustion combine to
form low melting compounds that are liquid and very corrosive above about 1000F (538oC).

Page 2-12
CALTEX REFINERY MATERIALS MANUAL October 1999

The corrosive liquids flux the normally protective scales and can rapidly attack the non-cooled
heater and boiler parts. A simplified mechanism is as follows:

1. The normally protective oxide film is fluxed by the molten salt


2. Subsurface penetration and attack by a sulfur species to form chromium sulfides and deplete
the area of chromium
3. The chromium depleted areas oxidize. After the fluxing of the protective oxide film, wastage
continues due to alternate sulfidation and oxidation. High levels of sodium and lead
compounds can dramatically increase wastage rates.

The 50-50 and 60-40 Cr-Ni alloys offer generally acceptable corrosion resistance in fuel ash
environments. Additions of aluminum and magnesium oxides to the fuel modify the low melting
liquids by raising the melting point of the sulfate-vanadate compounds so that the fluxing is
minimized. Some refractories have been effective in protecting non-cooled furnace/boiler parts.
Some large power station boilers have run at very low-excess air during combustion which can
minimize the problem. However, process furnaces are usually difficult to run at the very low
excess air needed for minimum corrosion. Environmental regulations limiting sulfur emissions in
the U.S. and stricter regulations worldwide will probably be the most effective method of
minimizing fuel ash corrosion.

5.1.4 Galvanic Corrosion


Galvanic corrosion can occur when two metals or alloys are coupled (joined electrically) and
exposed to an electrolyte. The more active and the less active alloy form a corrosion cell, often
called a galvanic cell. Corrosion of the more active metal or anode increases, while that for the
less active metal or cathode decreases. This is known as galvanic corrosion. Galvanic corrosion
can be a major problem in sea water service, such as in cooling water heat exchangers, but is of
lesser concern in most refinery services. However, the explanation of some unexpected and
peculiar problems may involve galvanic corrosion.

Based on galvanic corrosion tests and electrical potential measurements in seawater, various
metals and alloys can be ranked in the form of a galvanic series as seen in Table 2-1. Metals near
the top of the table become anodic or active and corrode when in contact with a metal nearer
the bottom of the list. Certain alloys, such as austenitic stainless steel, are shown in two positions
depending on whether they are in the active or passive state. The further apart two metals are in
the series, the more likely the less noble metal in the couple will experience galvanic corrosion.
Although, strictly speaking, the galvanic series applies only to sea water corrosion, it serves as a
rough guide to how metal couples behave in other aqueous series environments.

Page 2-13
CALTEX REFINERY MATERIALS MANUAL October 1999

TABLE 2-1 GALVANIC SERIES OF METALS AND ALLOYS IN SEA WATER

Corroded End - Anodic - More Active - Less Noble

Magnesium
Magnesium Alloys
Zinc
Aluminum
Aluminum Alloys
Steel
Cast Iron
Type 410 Stainless Steel (active state)
Ni- Resist
Type 304 Stainless Steel (active state)
Type 316 Stainless Steel (active state)
Lead
Tin
Nickel (active state)
Brass
Copper
Bronze
Copper-Nickel
Monel
Nickel (passive state)
Type 410 Stainless Steel (passive state)
Type 304 Stainless Steel (passive state)
Type 316 Stainless Steel (passive state)
Titanium
Graphite
Gold
Platinum

Protected end - Cathodic - Less Active - More Noble

It is important to note that the galvanic series only compares corrosion tendencies and tells
nothing about the corrosion rate to be expected, either for the metals or alloys acting separately
or electrically connected together. If there is an acidic constituent (i.e., CO2) or dissolved oxygen
present to provide a strong cathodic reaction, the cell can be quite vigorous. In operation of the
cell, the more active alloy (the anode) corrodes faster than it does when uncoupled, and the less
active alloy (the cathode) corrodes more slowly. Also, if there is a large ratio of anode area to
cathode area, the cathode will effectively be protected and the galvanic effect minimized. If, on
the other hand, there is a very small anode area and a large cathode area, the smaller anode will
tend to corrode rapidly.

Page 2-14
CALTEX REFINERY MATERIALS MANUAL October 1999

Small anode - large cathode effects are often not significant. Consider the case of a steel water
pipe coupled to a brass fitting. From the galvanic series, it is seen that the steel is more active
than the brass. The steel is the anode and the brass is the cathode. Near the point of contact, the
steel will corrode faster than normal, while the brass will corrode more slowly. The area of steel
affected and the intensity of corrosion will depend upon the relative size of the brass component,
geometry of the coupled parts, availability of dissolved oxygen, pH, and the resistivity of the
water. Depending on the influence of these variables, the steel pipe corrosion pattern can range
from localized knife-line attack to broad, general wastage.

Metals and alloys need not be grossly dissimilar to produce galvanic action when they are
coupled. Weld metal and the heat affected zone of welds can be sufficiently dissimilar from the
parent metal to cause galvanic cells. If the weld or heat affected zone is anodic to the parent
metal, a highly unfavorable situation of small anodic area and large cathodic area exists. In some
refinery environments this can lead to preferential weld corrosion, sometimes described as weld
decay - ie. Carbon steel ERW welded pipe can suffer rapid corrosion in raw water in spite of
correct post weld heat treatment. An elevated temperature heat treatment has sometimes been
successfully used to combat preferential weld corrosion by producing a more homogeneous weld
and base metal microstructure.

To minimize galvanic corrosion problems, the following things must be kept in mind:

Corrosion is more severe near the junction of two dissimilar metals with attack decreasing
with increasing distance from that point.

The severity of corrosion is related to the electrical conductivity of the solution. Galvanic
corrosion does not occur in hydrocarbon or vapor systems unless free water is present.

The area of the more anodic metal should be as large as possible compared to that of the
cathodic material.

Dissimilar metals should be electrically insulated wherever practical. Insulation must be


complete, otherwise, corrosion can actually be accelerated.

Painting or coating, when used, must be done to the entire assembly or at least the less
active, cathodic member. If only the anode is coated, breaks in the coating can cause the
exposed area to corrode very rapidly.

Corrosion inhibitors may be used to reduce galvanic effects in many refinery aqueous
environments. Sacrificial anodes along with paint/coatings may be used to reduce galvanic
effects.

5.1.4.1 Oxidation/Scaling Resistance


Table 2-2 shows the approximate temperature to which materials can be exposed without
significant scaling in steam and flue gas. These temperatures should be considered approximate
and may change significantly depending upon the specific environment.

5.1.5 High Temperature Sulfidation (With and Without Hydrogen)


(i) WITHOUT hydrogen
High temperature sulfur corrosion (without hydrogen present) becomes a problem with
hydrogen sulfide and various sulfur compounds above a temperature of about 260oC (500oF).
The corrosion depends on the concentration and type of sulfur compounds involved. Five
reactive groups of sulfur compounds that cause sulfur corrosion are:

Page 2-15
CALTEX REFINERY MATERIALS MANUAL October 1999

Elemental sulfur
Polysulfides
Hydrogen sulfide
Aliphatic sulfides
Aliphatic disulfides

Hydrogen sulfide is the most active of these from a corrosion standpoint. In fact, most of the
other compounds are considered inert, as far as corrosion is concerned, until the petroleum
reaches the refinery and is heated to elevated temperatures. Even then there is some question as
to whether the resulting corrosion is caused by the complex chemical forms themselves or
whether the corrosive attack results from conversion of the sulfur compounds into H2S.

Aliphatic sulfur compounds are generally more corrosive than aromatic or heterocyclic types.
They break down more easily to form H2S. Tertiary mercaptans are more corrosive than
secondary or primary mercaptans. The corrosion mechanism apparently proceeds by conversion
of the sulfur from its original form to hydrogen sulfide, followed by reaction of the H2S with the
steel.

High temperature corrosion problems related to H2S first appeared in the early 1930's in
refineries when the new thermal cracking processes resulted in higher operating temperatures. It
was quickly discovered that at temperatures above 260oC (500oF), the addition of small amounts
of chromium would reduce the corrosion associated with sulfur on steel. The degree of
improvement was related to the amount of chromium added. A typical curve relating corrosion
rates, temperature, and sulfur content is shown in Figure 2-2. There is a rapid increase in
corrosion rate above 260oC (500oF), especially for carbon steel. Velocity plays a part, as well as
temperature, in determining the corrosion rate at a given sulfur content. The effect of velocity is
not shown on the curve.

As a rule, sulfur corrosion in crude distillation units, coking units, FCC units, FCC feed
hydrotreaters, hydrodesulfurizers and hydrocrackers (ahead of the hydrogen injection point)
follow the general pattern described above. The McConomy Curves (Figure 2-2) is a set of data
useful for materials selection and prediction of the relative corrosivity of crude oils and their
various fractions. Although developed primarily for naphtha desulfurizers, these curves form a
reasonable design basis for hydrogen free sulfide corrosion service. Hydroprocessing fractionation
columns may not follow the corrosion rates predicted by the McConomy Curves. As the rate of
sulfur corrosion starts to decrease as the temperature exceeds 454oC (850oF). The most likely
reason for the rate to drop off is coke forming a protective layer on the exposed surface. There is
also a time at temperature factor in that thermal breaking up of the sulfur compounds may result
in less H2S being produced the further into the processing sequence the particular oil stock goes.

Relatively small changes in temperature can have significant, unexpected affects on sulfur
corrosion rates. Convection section tubes in crude oil feed furnaces and fired heater reboilers
normally operate at low enough temperatures so that little corrosion occurs. However,
accelerated, localized attack may occur at points where convection section tubes pass through
tube supports because of higher heat flux and temperature at these points. Care must also be
taken in changing from plain to finned or studded heater tubes. Increased sulfidation will be
likely due to the localized increase in tube metal temperature which could be as much as 93oC
(200oF).

H2S and sulfur compounds that break down at elevated temperature to form H2S begin to cause
metal loss problems above 260oC (500oF), serious enough to consider alloy protection. The
addition of chromium to carbon steel increases its resistance to high temperature sulfide attack.

Page 2-16
CALTEX REFINERY MATERIALS MANUAL October 1999

Increasing chromium results in more protective sulfide scales. In order to select materials resistant
to sulfide corrosion, it is important to determine if the service environment has hydrogen present
or is hydrogen free. Hydrogen does not allow the lower chromium alloys to form protective
sulfide scales. Therefore, 5Cr-0.5Mo and 9Cr-1Mo alloys generally corrode at a rate too high to
be used in H2/H2S service. Usually 11-12Cr alloys are the lowest chromium level to give effective
protection in H2/H2S service.

The scale can be pyrophoric so some care must be taken when pulling exchanger bundles that
have been heavily scaled and in towers with scale buildup that are left open to the atmosphere
during shutdowns.

(ii) WITH hydrogen


High temperature sulfur corrosion (with hydrogen present) is more severe than the usual high
temperature sulfur corrosion without hydrogen present. Hydrogen converts organic sulfur
compounds to hydrogen sulfide and corrosion becomes a function of H2S concentration or
partial pressure. This type of attack occurs primarily in FCC feed hydrotreating units,
hydrodesulfurizers, and hydrocrackers downstream of the hydrogen injection point. Refinery
experience has shown that corrosion data based on McConomy Curves do not apply where
hydrogen is present. The problem has to do with the alloys ability to form a protective scale.
The 11-13%Cr level is needed to form protective sulfide scales in hydrogen service.

Reasonably reliable corrosion data for the prediction of H2S/H2 corrosion rates are based on
the Couper-Gorman Curves (Figure 2-3 pages 1 & 2) developed from a NACE field survey of
refiners. As an example, Figure 2-3/Figure A1 is a curve for carbon steel in naphtha
desulfurizer, hydrogen sulfide/hydrogen service. As shown by the iso-corrosion curves, the
mole percent H2S in the process stream and the operating temperature define the expected
corrosion rates. When the rate of corrosion is too high for carbon steel equipment to have
useful life, a more appropriate alloy can be selected.

There is little improvement in corrosion resistance of low alloy steels unless chromium content
exceeds 9%. H2S/H2 corrosion is generally more severe in gas oil desulfurizers than in naphtha
units. Above a certain temperature and at low hydrogen sulfide concentrations there is a zone of
no corrosion. This is because formation of iron sulfide is thermodynamically impossible under
these conditions. However, because corrosives other than H2S may still attack steels, caution is
advisable when selecting materials for this region. In particular, alloying is often required to resist
high temperature hydrogen attack.

The estimated rate, the service involved and the desired design life set the material and its
corrosion allowance. Rates of up to 10 mils per year are usually considered acceptable. However,
consideration must be given to potential downstream pressure drop problems from scale
buildup. Sulfide scale volume may exceed 7 times the metal loss. Higher corrosion rates can
usually be tolerated from a mechanical design standpoint but equipment fouling/pressure drop
problems must be considered for units with long anticipated run lengths.

Even when ferritic or martensitic stainless steel with 12% chromium is an acceptable material of
construction from a corrosion standpoint, scaling may be sufficient to plug reactor catalyst beds
and foul clearances in screens. In practice, austenitic stainless steels, such as Types 304L, 321, or
347 are used for most equipment operating above 260oC (500oF) in the presence of hydrogen
sulfide and hydrogen. Figure 2-3/Figure A5 is a corrosion rate curve showing the dramatic
improvement in corrosion resistance offered by austenitic stainless steel over other alloys,
including 12% chromium stainless.

Page 2-17
CALTEX REFINERY MATERIALS MANUAL October 1999

As with the McConomy Curves, there are refiners that have come up with their own curves or
revised the Couper-Gorman curves to better reflect their own H2/H2S sulfiding corrosion
experience. Most of these curves were developed about the same time as Couper-Gorman and
reflect the need for more realistic corrosion rates than were predicted by the early short time
laboratory tests which generally did not take a decrease in rates due to scale formation into
account.

5.1.6 Intergranular Corrosion


Intergranular corrosion is highly localized corrosion at and adjacent to grain boundaries. Since
there is relatively little corrosion of the grains, the alloy disintegrates by grain separation.
Intergranular attack is caused by the corrosive action of a specific chemical environment on the
metal grain boundaries that are susceptible to attack due to impurities or the enrichment or
depletion of one of the alloying elements at grain boundaries.

With austenitic stainless steels, intergranular attack can be caused by chromium depletion as a
result of sensitization. This occurs when the stainless steel alloy is held in or cooled slowly
through the temperature range of 371o to 816oC (700o to 1500oF). This can happen during
welding or while the equipment is in elevated temperature service. Carbon combines with
chromium and depletes the grain boundary region of chromium. The chromium depleted grain
boundary zone is then subject to preferential corrosion. The intergranular attack that results is
sometimes called 'weld decay'.

Intergranular attack of austenitic stainless steels can be minimized or prevented by one of the
following approaches:

Specify low carbon grades, such as Type 304L, 316L, or 317L. These contain insufficient
carbon for chromium carbide precipitation to be significant in most cases during welding.
However, the low carbon grades have lower high temperature allowable strengths.

Use chemically stabilized grades that have been properly stabilize annealed, such as Type 321
(titanium bearing) and Type 347 (niobium) wherein the alloying elements tie up carbon.
Austenitic stainless steels are supplied in the solution annealed condition but fabrication,
welding or heating can sensitize them.

Solution annealing the stainless steel by heating to 1093oC (2000oF) followed by water
quenching to redissolve any precipitated chromium carbide and uniformly distribute
chromium within the microstructure can eliminate the sensitized microstructure. Resistance
to sensitization can be further increased by heat treatment - i.e. stabilizing anneal. Under
some circumstances, the solution annealing can cause sensitization in Type 347.

In refineries, austenitic stainless steels are extensively used in hydroprocessing units and catalytic
cracking units where polythionic acids can form during shutdowns. Sensitized structures in
contact with polythionic acids can result in grain boundary stress corrosion cracking, which will
be discussed later. (refer Section 2 5.3.4) In most high temperature applications, sensitization
does not significantly affect high temperature performance. Sensitized stainless steel has
decreased resistance in aggressive acid situations such as dew point operation and shutdown
conditions and, therefore, exposure must be minimized.

5.1.7 Microbiologically Influenced Corrosion (MIC)


Microbiologically Influenced Corrosion (MIC) has been documented for metals exposed to
seawater, fresh water, demineralized water, process chemicals, foods, soils, aircraft fuels, human

Page 2-18
CALTEX REFINERY MATERIALS MANUAL October 1999

plasma and sewage. MIC does not produce any unique type of localized corrosion. Instead, the
presence and activities of microorganisms can cause pitting, crevice corrosion, selective
dealloying and differential aeration cells in addition to enhanced galvanic and erosion corrosion.
Furthermore, there are no definite tests that can be used to detect MIC. Diagnosis requires
sophisticated microbiological, surface analytical, and electrochemical techniques. Often, the
most convincing evidence used to document MIC is the spatial relationship that exists between
specific physiological types of bacteria and the manifestation of localized corrosion. But, because
bacteria are ubiquitous and can be attracted to both cathodes and anodes, spatial relationships
can be easily overinterpreted or misinterpreted. As a consequence of the multidisciplinary nature
of MIC and the absence of definitive indicators, identification of MIC is often complicated.

Bacterial activity of interest to the corrosion engineer always involves several different types of
bacteria acting in concert to produce corrosive effects. These bacteria communities maintain
their spatial relationships by functioning within a matrix of biopolymers called a biofilm. The
environment within the biofilm is maintained by the metabolic activity of the bacteria and can be
very different from the chemistry of the bulk solution. The biofilm and its modified environment
will act as a further barrier to any biocides added to the system.

Bacteria can influence the corrosion process in several ways:

Utilization of Oxygen. When aerobic microorganisms colonize a surface, their metabolic


activity removes oxygen from the local environment and the oxygen deficient area becomes
the locus of concentration cell corrosion.
Utilization of Hydrogen. Many of the bacteria implicated in the corrosion process are able to
utilize hydrogen as an energy source. The resulting depolarization of the cathode is believed
to cause increased corrosion rates.
Production of Corrosive Metabolites. The end product of anaerobic metabolism is often
organic or inorganic acids. These products can accelerate corrosion and metal loss.
Production of Concentration Cells. The presence of the polymer matrix of the biofilm can
result in the concentration of specific ions at the metal surface. These localized differences in
concentration can shift the potential of the metal surface, resulting in the creation of a
corrosion cell.
Utilization of Protective Coatings. Some bacteria have the ability to degrade protective
coatings or the adhesives that bind them to the metal surface. A localized area where
coatings are removed will become a small anode electrically connected to a large cathode
beneath the remaining coating. Rapid metal loss can ensue.
Utilization of Inhibitors. Corrosion inhibitors and other products that are added to reduce
corrosion rates can be actively metabolized by certain microorganisms, resulting in loss of
protection.
Removal of Metal Ions. Metabolic by-products such as H2S can remove metal ions from a
solution that depolarizes the anodic process. The polymers in some biofilms have the ability
to bind metal ions and some bacteria also have the ability to produce proteins that can
aggressively remove metal atoms from their environment.
Precipitation of Metal Sulfides. Metal sulfides depolarize the cathode and improve the flow
of protons and electrons in the deposit.

Bacteria are not the only microorganisms contributing to MIC.

Bacteria exist in isolation from other organisms only in the laboratory. In natural environments,
they occur in association with fungi and microalgae. Fungi predominate in fuel water mixtures
and, because they produce copious amounts of organic acids, fungi are the causative agents for
corrosion in many fuel storage containers of carbon steel and aluminum. Fungi also predominate
in moist environments where organics are rich but water may be limited. Acid production by

Page 2-19
CALTEX REFINERY MATERIALS MANUAL October 1999

fungi growing in moist grain in a hold of a transport ship is responsible for the failure of the
carbon steel storage area. Fungi growing on wooden spools used to store grease-coated wire
rope are responsible for dissolution of the protective grease and localized corrosion of the wire
rope.

In fresh and marine environments, bacteria are found in association with microalgae that may
dominate the microflora. In the presence of light, algae produce oxygen and pH changes that
may influence corrosion reactions, but the extent that algae contribute to corrosion in the
absence of bacteria is not known.

MIC is not a unique type of localized corrosion. Instead, the presence and activities of
microorganisms on metal surfaces produce environments that can alter rates of partial reactions
in corrosion processes and shift corrosion mechanisms. The most severe MIC takes place in
natural environments where several physiological groups of aerobic and anaerobic
microorganism interact.

Fungi can produce MIC in humid environments. For carbon steel, copper and nickel alloys
exposed to sulfate-containing electrolytes, operating conditions that alternate between aerobic
and anaerobic increase corrosion caused by sulfate reducing bacteria. Microbe/metal
interactions are complex and reliable test methods for identifying or predicting MIC and cannot
be based on enumeration of organisms or their spatial relationship to corrosion products alone,
but must also consider those interactions.

5.1.8 Oxidation
Oxidation occurs when carbon steels, low alloy steels and stainless steels react at elevated
temperatures with oxygen in the surrounding air or flue gas and become scaled. Most theories
postulate a diffusion mechanism for protective scale growth which usually follows the parabolic
rate law. Nickel alloys also can become oxidized, especially if spalling of the scale occurs.
Oxidation of copper and aluminum alloys is not a typical refinery problem since they are rarely
used where operating temperatures exceed 260oC (500oF). Thermal cycling, applied stresses,
moisture and sulfur-bearing gases all tend to decrease scaling resistance.

In refineries, high temperature oxidation is primarily limited to the outside of furnace tubes, to
furnace tube hangers, and other internal furnace components which are exposed to combustion
gases containing excess air. Table 2-2 lists the maximum metal temperatures for various refinery
metals which will result in acceptable scaling rates in the presence of air. Acceptable scaling rate
represents a weight gain of less than 0.002 grams per square inch per hour.

Alloying steel with both chromium and nickel increases scaling resistance. Stainless steels or
nickel alloys are required to provide satisfactory oxidation resistance at temperatures above
704oC (1300oF). Silicon is also effective in steels, even when present in relatively small quantities.
Alloys with added silicon and aluminum may be very difficult to fabricate.

Aluminum, appropriately applied to the surface of steels, also improves oxidation resistance. It
can be applied by spraying, dipping or cementation. This practice is particularly applicable to low
alloy steels as a means for improving oxidation resistance at moderate temperatures. it is also
useful at high temperatures but a diffusion treatment must be included in the application process
to create a sturdy alloy bond and also eliminate the presence of metallic aluminum on the
surface. "GPS-A9 Selection of Metallic Materials" recommends that this treatment not be applied
to pressure containing parts, especially piping. This is because of the problems of ensuring
proper coverage and subsequent long-term inspection problems.

Page 2-20
CALTEX REFINERY MATERIALS MANUAL October 1999

At elevated temperatures, steam decomposes at metal surfaces into hydrogen and oxygen and
may cause oxidation of steel. It is more severe than air oxidation at the same temperature and, as
a result, the temperature limits given in Table 2-2 should be lowered by roughly 38oC (100oF) for
high temperature steam service. Fluctuating steam temperatures tend to increase the rate of
oxidation by causing scale to spall and, thereby, expose fresh metal to further attack.

The following table shows the approximate temperature to which materials can be exposed
without significant scaling in air, steam and flue gas. These temperatures should be considered
approximate and may change significantly depending on the specific environment.

TABLE 2-2 MAXIMUM SKIN TEMPERATURES

Material Maximum Skin Temperature


During Normal Operation During Thermal Decoking
C F C F

Carbon Steel 538 1000 677 1250


1 Cr Mo Steel 593 1100 732 1350
2 Cr 1Mo Steel 621 1150 732 1350
5Cr Mo Steel 649 1200 732 1350
7Cr Mo Steel 649 1200 732 1350
9Cr 1Mo Steel 677 1250 732 1350
18Cr 8Ni Steel 816 1500 816 1500
25Cr 20Ni Steel 1066 1950
Incoloy 801 1066 1950

5.1.9 Pitting
Pitting is highly localized corrosion in the form of small holes or pits having a diameter about the
same or less than the hole depth. Pitting can occur in isolated locations or be so concentrated
that it looks like uniform attack. Often, pits have a tendency to undercut the metal surface which
makes their detection difficult. Pitting often requires an incubation period, depending on the
metal and corrosive environment involved. Equipment failures are usually in the form of
perforation at one or more points, with less general loss of metal section. Since pits are usually
covered by corrosion product, they are hard to detect prior to failure.

From a practical standpoint, pitting often occurs under stagnant flow conditions in the presence
of chloride ions. Chloride ions are sufficiently small and mobile to penetrate protective films,
scale, or corrosion products.

It is important to note that pitting is an auto-catalytic process which stimulates the continuing
activity of the pit. Oxidation of the metal takes place within the pit, while the cathodic reaction
takes place on adjacent surfaces. This produces an excess of positive ions within the pit and
results in migration of chloride ions to maintain electrical neutrality. Subsequent hydrolysis lowers
the pH of the solution within the pit and this accelerates metal oxidation.

Page 2-21
CALTEX REFINERY MATERIALS MANUAL October 1999

Pitting is initiated at surface defects, emerging inclusions or grain boundaries of the metal and
breaks in mill scale or protective films. Excessive corrosion at these sites will promote chloride
migration and stimulate highly localized attack. In refineries, pitting has been a problem mostly
with stainless steels. Martensitic, ferritic, and austenitic stainless steels are very prone to chloride
pitting unless alloyed with molybdenum. As a general rule, metals and alloys that pit during
corrosion tests in a given environment should not be used to construct process equipment.

Page 2-22
CALTEX REFINERY MATERIALS MANUAL October 1999

TABLE 2-3 CORROSIVES FOUND IN MANY REFINING PROCESSES

Sulfur Present in raw crude. It causes high-temperature sulfidation of metals, and it combines
with other elements to form aggressive compounds, such as various sulfides and sulfates,
sulfurous, polythionic, and sulfuric acids.
Naphthenic A collective name for organic acids found primarily in crude oils from western United
Acid States, certain Texas, Gulf Coast, South American some Mid-East and Far East crudes.
Polythionic Acid Sulfurous acids formed by the interaction of sulfides, moisture, and oxygen, usually
occurring when equipment is shutdown.
Chlorides Present in the form of salts (such as magnesium chloride and calcium chloride)
originating from crude oil, catalysts, and cooling water.
Carbon Dioxide Occurs in steam reforming of hydrocarbon in hydrogen plants, and to some extent in
catalytic cracking. C02 combines with moisture to form carbonic acid.
Ammonia Nitrogen in feedstocks combines with hydrogen to form ammonia. Ammonia is used for
neutralization - which in turn may combine with other elements to form corrosive
compounds, such as ammonium chloride and ammonia bisulfide.
Cyanides Usually generated in the cracking of high-nitrogen feedstocks. When present, corrosion
rates are likely to increase.
Hydrogen Formed through hydrolysis of magnesium chloride and calcium chloride, it is found in
Chloride many overhead (vapor) streams. On condensation, it forms highly aggressive
hydrochloric acid.
Sulfuric Acid Used as a catalyst in alkylation plants and is formed in some process streams containing
sulfur trioxide, water and oxygen.
Hydrogen In itself not corrosive but can lead to fissuring, blistering and embrittlement of steel.
Also, it readily combines with other elements to produce corrosive compounds.
Phenols Found primarily in sour water strippers.
Caustic BFW treatment; concentration can cause cracking/corrosion. Used for carbon removal.
Used a corrision inhibitors and solvent in treating gases and light hydrogen for
removing H2S and/or CO2.
Hydrofluoric Used as alkylation catalyst. Handling problems because HF vaporizes at atmospheric
Acid pressure.
Phosphoric Acid Used in waste water treating; used as catalyst for light hydrocarbon polymerization.
Boiler Feed If not properly treated, can contain oxygen and acidic components
Water
Steam Can contain oxygen and carbon dioxide.
Condensate
Cooling Water Sea water is always corrosive but even fresh and recirculated water can be corrosive if
not treated.
Fuel Ash Occurs in furnaces and boiler firing high sulfur fuel.
Microbes Can occur in many environments.

Page 2-23
CALTEX REFINERY MATERIALS MANUAL October 1999

AVERAGE CORROSION RATES


High Temperature Sulphur Corrosion
Hydrogen Free Environment
McConomy Curves

450 500 550 600 650 700 750 800


200 200

l
tee
nS
100 rbo 100
Ca

80 80
70 70
Cr
60 3% 60
1-
50 50

40 40
Cr
%
30 4 -6 30

20 20
Cr
7%

Cr
9%
10 10
Corrosion Rate - Mils/ year

8 8
7 7
6 6
1 Mil = 1/1000 inch

5 5

4 4
Cr
%
12
3 3

2 2

1.0 1.0
Ni
Cr
/8
0.8 18 0.8
0.7 0.7
0.6 0.6
0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
450 500 550 600 650 700 750 800
o
Temperature F

FIGURE 2-2

Page 2-24
CALTEX REFINERY MATERIALS MANUAL October 1999

10

PRE
1.0

DIC IL S P
TED ER
'M
COR YEAR
RO
SIO
60

NR
'
50

ATE
40
30
Mole % H2S

25
20
15

0.1
10
5
3
2
1

0.01

NO CORROSION

0.001
400 600 800 1000
0
Temperature F
FIGURE A1- CARBON STEEL - NAPHTHA DILUENT

10 10
PRE 'MILS ORRO AR'

PRE

PRE 'MILS

PRE 'MILS
PRE

DICT

DICT ILS PE

DICT

DICT
DIC ILS P

'M
ED C ER YE ION RA

ED C R YE

ED C ER YEA

ED C ER YEA
1.0 1.0
TED ER
'M

PRE 'MILS
ORR AR'

ORR AR'
P

ORR

ORR
P

P
C

DICT
OSIO

OSIO
50

OSIO '

OSIO '
40

ED C ER YEA
YE
30

N RA

R
N

N RA
20

RAT

ORR
P
15

RAT 80
TE
10

TE
E

100

OSIO '
E
TE
5

60

R
N RA
2

40
Mole % H2S

Mole % H2S

30
1

TE
20

0.1 0.1
15
10
5

0.01 0.01

NO
NO CORROSION
CORROSION NO CORROSION NO CORROSION NO CORROSION
NO CORROSION

0.001 0.001
400 600 800 1000 400 600 800 1000
0
Temperature F 0
Temperature F
FIGURE A-3 5% CR STEEL -- NAPHTHA DILUENT FIGURE A-4 5% CR STEEL -- GAS OIL DILUENT

The above figures show isocorrosion curves based on correlations of American Oil data along with individual data points.

COOPER-GORMAN CURVES
Figure 2-3
Page 1

Page 2-25
CALTEX REFINERY MATERIALS MANUAL October 1999

10 10

PRE 'MILS

PRE 'MILS
PRE 'MILTE

PR E

DICT
PRE 'MILS

DICT
DIC

D
IC T E IL S P ER
1.0

DIC
TED PEROR

ED C PER YE

ED C ER YEA
'M
1.0

D CO YE A

PRE 'MILS ATE


SD C
COR YEAORS'ION

ORR AR'

ORR
P
DICT
RR O R '
ROS

OSIO
PER

OSIO
ED C
R

SIO
ION
YEA

R'
NR

N RA
50

N RA

ORR EAR'
40

PER
RAT
R'
30

TE
50
RAT
20

OSIO
TE

40

Y
E

30
15

20
E
10

10

N RA
Mole % H2S

Mole % H2S

5
0.1

TE
5

0.1
4
3
2
1

0.01 0.01

NONO
CORROSION
CORROSION NO CORROSION NO CORROSION NO CORROSION
NO CORROSION

0.001 0.001
400 600 800 1000 400 600 800 1000
0
Temperature F 0
Temperature F
FIGURE A-5 9% CR STEEL -- NAPHTHA DILUENT FIGURE A-6 9% CR STEEL -- GAS OIL DILUENT

10 10

10.0
8.0
6.0
4.0
2.0
1 .0
0.8
PREDIC S PER YEAR'

0 .6
0.4
0.2
'MIL
TED CO

1.0 1.0
PR E
PRE

DIC IL S P E
DIC

RROSION
PRTE ILS P'MEI

T ED
'M
'M
EDDICC R

CO R YEA R
TOERDR YPEEARR'

RATE

R
RO S '
LS
COOSIO YEAR
RRNO R '

ION
Mole % H2S

Mole % H2S
SIO

R
ATE

0.1
AT E

0.1
NR
ATE
25
20
15
10
5
2

0.01 0.01
1

NO CORROSION
NO CORROSION
NO CORROSION
NO CORROSION

0.001 0.001
400 600 800 1000 400 600 800 1000
0
Temperature F Temperature 0 F
FIGURE A-7 12% STAINLESS STEEL -- NAPHTHA DILUENT FIGURE A-8 18-8 STAINLESS STEEL -- NAPHTHA DILUENT
AND GAS OIL DILUENT AND GAS OIL DILUENT

The above figures show isocorrosion curves based on correlations of American Oil data along with individual data points.

COOPER GORMAN CURVES


Figure 2-3
Page 2

Page 2-26
CALTEX REFINERY MATERIALS MANUAL October 1999

5.2 SPECIFIC CAUSES OF GENERAL/LOCALIZED CORROSION

5.2.1 Aluminum Chloride


Aluminum chloride is used as a catalyst in refining processes such as butane isomerization,
ethylbenzene production and polybutene production. Aluminum chloride itself is not corrosive
provided it is kept absolutely dry. However, if water or water vapor is present in hydrocarbon
streams, aluminum chloride hydrolyzes to hydrochloric acid which is highly corrosive.

To control corrosion in the presence of aluminum chloride, the feed stock is dried in CaCl2 dryers.
During shutdowns, equipment should be opened for the shortest possible time and, on closing,
dried with hot air. Equipment which is exposed to hydrochloric acid will require extensive lining
with nickel alloys, such as Hastelloy B-2, C-276, or G-22.

5.2.2 Amine
Amines used in gas treating units have been a source of refinery corrosion problems. The
common amines used in refinery service include MEA (monoethanol amine), DEA (diethanol
amine), DIPA (diisopropenol amine) and DGA (diglycol amine) and MDEA (monodiethanol
amine). Corrosion is not caused by the amine itself but rather by dissolved H2S or CO2 and by
amine degradation products and by heat stable salts. Corrosion can usually be traced to faulty
plant design, poor operating practices and/or solution contamination. In general, corrosion is
most severe in systems removing only CO2 and least severe in systems removing only H2S.
Corrosion in amine plants using MEA is more severe than in those using DEA because MEA is
more prone to degradation and can be used at higher gas loadings. Locations most affected are
those where acid gases are desorbed or removed from rich amine solutions. Here, temperatures
and flow turbulence are the highest. This includes the regenerator reboiler and the regenerator
itself. Corrosion can also be a significant problem on the rich amine side of the lean/rich
exchangers, in amine solution pumps, and in reclaimers. Hydrogen blistering, hydrogen induced
cracking and stress corrosion cracking can be a problem in amine systems.

Except for the overhead system, the standard material of construction for amine gas treating
equipment is carbon steel. Welds should be postweld heat treated to resist stress corrosion
cracking. Pitting and groove type corrosion of carbon steel reboiler tubes may require a change
to Type 304 or 316 stainless steel. As a rule, copper alloys are not used in amine units due to the
risks of failure from stress corrosion cracking. See Section 2 5.3.8 for details of amine stress
corrosion cracking problems.

5.2.3 Ammonium Bisulfide (NH4HS)


Ammonium bisulfide (NH4HS) is a potent corrosive formed during the hydrotreating and
hydrocracking of hydrocarbons containing organic nitrogen and sulfur compounds. In
hydrodesulfurizers and hydroprocessing, high concentrations of ammonia and hydrogen sulfide
can heavily load wash water at the cold end of the process. The resulting NH4HS can cause
serious corrosion of carbon steel. Reaction section effluent streams, beyond the water wash
introduction point, are highly susceptible to NH4HS corrosion, especially piping components
seeing high velocity and turbulence. Monel, Incoloy 800, Incoloy 825, Alloy 20, and duplex
stainless steels have all been successfully used to combat NH4HS corrosion in hydroprocessing
cold end equipment.

In some sour water stripping units, exceptionally high concentrations of NH4HS build up in a thin
film of condensed water on overhead condenser tubes and can cause severe corrosion of carbon
steel. Titanium and other alloy tubes have been used to solve this problem.

Page 2-27
CALTEX REFINERY MATERIALS MANUAL October 1999

NH4HS will rapidly attack admiralty brass tubes, which, in some compressor aftercooler
applications, have been known to last for only 30 days. If pH values of the process water are
above 8, carbon steel tubes normally are not corroded because a protective iron sulfide film
forms on all metal surfaces. Unfortunately, in situations of high velocity and turbulence, the
protective film can be eroded and the steel corrodes rapidly.

Corrosion severity is based on a factor, Kp, which is the product of the mole percent of ammonia
and the mole percent of H2S.

A hydroprocessing licensor uses the following parameters for design to reduce ammonium
hydrosulfide corrosion:

Kp figured dry before water wash

Concentration wt. percent from the High Pressure Separator water

Piping balanced inlet and outlet, i.e., symmetrical

Water Wash 20 percent excess free water at injection point

Velocity Carbon Steel> 10 ft/sec to 20 ft/sec (good experience)

Alloy (825, 2205 etc.) > 30 ft/sec (very conservative)

Kp > 0.4 Alloy (i.e., 825)

Kp< .15 Carbon steel

Sour water 2 percent is the carbon steel limit usually gives too much water to handle; 8
percent maximum composition

Wash water 20 ppb 02 limit (usually no stripped sour water) Call for no cyanide (this
usually eliminates stripped FCC or Coker sour water) Some people feel this
is not realistic.

Most failures have been due to exceeding velocity limits on carbon steel outlet headers. There is a
corrosion survey by UOP that should be available to licensees.

5.2.4 Atmospheric Corrosion


Atmospheric corrosion can be a major problem at refineries, especially those located in coastal
zones. A certain amount of corrosion will be experienced by carbon steel and low alloy
equipment in the presence of air and moisture. Relative humidity would have to be below 60% in
order to have essentially no corrosion.

The normal rate of atmospheric corrosion ranges from 1 to 10 mils per year (O.025-0.25 mm/y),
but it may be as high as 50 mpy (1.2 mm/y), depending on location and time of year.
Equipment located near boiler or furnace stacks will corrode fairly rapidly because stack gas sulfur
dioxide and sulfur trioxide dissolve in moisture present on metal surfaces to form acids.
Chlorides, hydrogen sulfide, fly ash and chemical dusts in the atmosphere will also accelerate
corrosion. For example, brass valves in refinery atmospheres turn black within a matter of days
due to traces of sulfur compounds in the air.

Page 2-28
CALTEX REFINERY MATERIALS MANUAL October 1999

Atmospheric attack often is in the form of crevice corrosion. Any structural members which have
pockets from which water cannot drain will suffer severe corrosion. Obviously, structures exposed
to cooling tower spray are especially prone to this type of atmospheric attack.

Protective barriers, in the form of paints or other protective coatings, are the best way to stop
atmospheric corrosion. When personnel safety is involved, as with ladders, railings and flooring,
galvanized steel can be used for improved service life.

At coastal locations, special precautions need to be taken in view of the relatively high salt
content of airborne mist. On carbon steel and low alloy steels, zinc rich primer paints should be
used. These should be topcoated with maintenance type epoxy coatings. However, zinc rich
paints should not be used under insulation in hot or cold service or under fireproof coating
systems. Even stainless steel equipment should be considered for painting at coastal locations, to
prevent pitting or stress corrosion cracking. No paints containing metallic aluminum or zinc
powder should be used on austenitic stainless steels because of the danger of liquid metal
embrittlement in case welding has to be done at a later date or should fire exposure occur. For
the same reason, galvanized steel must be kept away from austenitic stainless steel.

5.2.5 Boiler Feedwater


Boiler feedwater for steam generation must be treated to protect boilers and auxiliary equipment
against corrosion during operation. Low temperature corrosion problems occur in the reheat
system, deaeration equipment, feed water piping and pumps, heaters and economizers. The
primary causes are dissolved oxygen and low pH conditions from the presence of acidic
constituents.

Even small concentrations of oxygen can cause serious pitting corrosion. Oxygen enters with
makeup water due to air leakage on the suction side of pumps, or as a result of the breathing of
supply water tanks. It can be removed by mechanical deaeration, followed by chemical treatment
with catalyzed sodium sulfite. For boilers operating above 1000 psi (6890 kPa), hydrazine is used
instead of sodium sulfite. Neutralization is usually done with soda ash or with organic neutralizers
such as morpholine or cyclohexylamine.

Deposition of various materials on boiler surfaces can not only cause failure by overheating but
also highly localized corrosion. As mentioned earlier, caustic concentrates under porous deposits,
resulting in caustic corrosion, gouging, and caustic embrittlement. Even when demineralized
makeup water is used, a coordinated pH/phosphate treatment may be required to control caustic
corrosion. In certain critical boiler applications, only volatile treatments can be used because no
boiler water solids whatsoever can be tolerated.

5.2.6 Cooling Water Corrosion


Cooling water corrosion can become quite costly if not properly controlled. Most refinery cooling
water systems are of the open recirculating type, with mechanical draft cooling towers. Cooling is
by evaporation of a portion of the water and this concentrates minerals in the circulating water.
Makeup water replaces water losses from evaporation, windage losses, and that blown down to
control dissolved solids concentration. Since makeup water often is scarce and expensive, many
cooling water systems operate at 2 to 4 cycles of concentration or higher.

Intimate contact of cooling water with air can create a multitude of corrosion problems. Airborne
contaminants such as hydrogen sulfide, ammonia, sulfur dioxide, fly ash or dirt are scrubbed
from the air in the cooling tower and can contribute to corrosion. The concentration of dissolved
minerals, such as chlorides and sulfates, increases the conductivity of cooling water as well as the

Page 2-29
CALTEX REFINERY MATERIALS MANUAL October 1999

tendency toward crevice corrosion beneath deposits. Relatively high temperatures also increase
the potential for corrosion.

Cooling water corrosion is not normally a problem with inhibited admiralty metal tubes or with
titanium tubes. These can foul, however, if scale formation is not controlled. In contrast, carbon
steel equipment, such as piping, heat exchanger tubes, channels, channel covers and tubesheets,
can become seriously damaged by cooling water corrosion.

Corrosion of carbon steel cooler and condenser tubes is especially troublesome for several
reasons:

Even relatively low corrosion rates of 1 to 2 mils per year can form enough corrosion
products in the form of tubercles on the tube wall to interfere with water flow.

Scale formation on tube walls is accelerated by the presence of corrosion products,


interfering further with water flow.

The resultant decrease in water flow can raise the temperature of the water to the point
where it boils in part of the bundle.

Under the above conditions, increased corrosion leads to premature tube failures, sometimes
within a matter of months.

Corrosion in open recirculating cooling water systems is controlled by maintaining small


concentrations of inorganic corrosion inhibitors in the water. These inhibitors retard corrosion by
formation of protective oxide films on carbon steel to passivate it. Common examples of
inhibitors include various combinations of chromate, polyphosphates and zinc compounds. More
recently, various organic inhibitors have been combined with certain inorganic materials to meet
regulations that limit air and water borne chromate discharges.

Refineries that rely on brackish or sea water for cooling should consider aluminum brass, copper-
nickel, or titanium tubes. These are normally rolled into carbon steel tubesheets which are solid
or clad with aluminum bronze, Monel, or titanium on the waterside. Clad carbon steel
tubesheets have had serious galvanic corrosion problems when tube roll leaks have occurred.
Monel 400 is an alternative tubesheet material and can be used to clad or weld overlay
components in salt water service.

5.2.7 Carbon Dioxide (CO2)


Carbon dioxide is a corrosive found in refinery steam condensate systems, hydrogen plants, and
in the vapor recovery sections of catalytic cracking units. In the case of boiler systems, carbonates
and bicarbonates remaining in boiler feed water decompose at the elevated temperatures to
form CO2, oxides, and hydroxides. The CO2 released in the decomposition, goes overhead with
the steam. In the vapor phase, no accelerated corrosion is observed. But when the steam
condenses, CO2 dissolves in the condensate and results in rapid acid corrosion of condensate
piping and equipment. The most effective corrosion mitigation procedure is to use an improved
boiler feed water treatment to prevent entrance of carbonates and bicarbonates into the boiler.

When removal of unwanted carbonate minerals is not practical, the use of neutralizing and
filming amines can be considered. Neutralizing amines are alkaline compounds having a suitable
boiling point and solubility such that they condense with the condensate. The neutralizing amine
then reacts with the CO2. Typically, enough neutralizing amine is added to raise the pH of the

Page 2-30
CALTEX REFINERY MATERIALS MANUAL October 1999

condensate to 8.5 to 9.0. If appreciable amounts of CO2 are present, the amount of neutralizing
amine can be prohibitively expensive.

Filming amines can be added to the feed water or directly into the steam. Typical products are
hexadecylamine or octadecylamine. These are volatile and, like neutralizing amine, condense
with the condensate. Unlike the neutralizing amines, however, the filming amines are not very
soluble in water and provide an inhibiting rather than neutralizing function. Chealants such as
EDTA can be very effective in minimizing boiler scaling but can be very corrosive if not properly
diluted before being injected into the system.

5.2.8 Corrosion Under Insulation


Corrosion under insulation (CUI) occurs when insulation or fire proofing is allowed to get wet.
Corrosion of underlying metal surfaces becomes a serious problem with piping and vessels
operating below about 149oC (300oF) because the metal does not get hot enough to keep
insulation dry during normal operation. Refrigeration systems and equipment in intermittent
service are especially affected by corrosion under insulation. CUI is more of a problem in humid
climates as is equipment downwind of cooling towers. Chloride can be introduced by breathing
in coastal areas. There is a limited life with insulation inhibitors. Water repelling materials are
available but the water repellent tends to be applied to the outside surfaces and the effectiveness
is drastically reduced if the insulation is cut or damaged. Corrosion is typically worse under
fibrous insulation like fiberglass or mineral wool. Closed cell, water resistant insulation like
cellular glass or perlite is usually the best for CUI prevention (polyurethane foam is an exception).
Calcium silicate is in between. CUI also occurs under fireproofing. CUI corrosion is generally
highly localized. Likely locations include:

flanges in vertical piping runs


low points in vertical piping
the 6 oclock position on horizontal vessels and exchanger shells
the area just above insulation support rings and vacuum stiffeners
at damaged weather jacketing
at insulation cutouts
under skirt fire proofing
around any structural attachments that penetrate the insulation to the shell or vessel heads

The best preventative approach is to keep insulation dry. This means proper wrapping and
caulking of joints. In general, it is advisable to paint new lines in CUI services. Metal surfaces
near flanged connections, valves and pumps should be painted since wetting of insulation due to
leakage is likely to occur at these locations. In the case of austenitic stainless equipment and
piping, only low chloride insulation should be used to minimize any potential chloride stress
corrosion cracking problem.

Typical inspection techniques include stripping the insulation, radiography, and real-time x-
ray/radiography. There are several new techniques being developed both in Europe and the
USA, but at the time of writing (1998) none of them are yet commercially available.

5.2.9 Hydrogen Chloride


Chloride salts are found in most production wells, either dissolved in water emulsified in the
crude oil, or as suspended solids. Salts also originate from salt water injected for secondary
recovery, or from sea water ballast in marine tankers. The amount of water varies rather widely,
but an approximate range might be 0.1 to 2.0% (by volume) of the crude oil. The amount of salt
contained in the emulsified aqueous phase may range from 10 to 300 pounds per thousand

Page 2-31
CALTEX REFINERY MATERIALS MANUAL October 1999

barrels of crude oil. Typically, the salt contains 75% sodium, 15% magnesium and 10% calcium,
mostly in the form of chlorides. On heating to above 149oC (300oF), hydrogen chloride is
evolved from magnesium and calcium chloride, but sodium chloride is essentially stable up to
roughly 427oC (800oF).

Hydrogen chloride evolution takes place primarily in the crude preheat exchanger train and
furnace. Dry hydrogen chloride, especially in the presence of large amounts of hydrocarbon
vapor or liquid, is not corrosive to carbon or low alloy steel. However, when steam is added to
the bottom of the crude tower to facilitate fractionation, dilute hydrochloric acid is produced in
the overhead condensing system and other locations where the water dew point is reached.

All equipment in this area is normally made from carbon steel and severe corrosion can occur at
temperatures below the initial water dew point. Corrosion rate increases with decrease in pH of
the overhead condensed water. To minimize chloride corrosion, the pH of the overhead
accumulator water should be maintained between 5 and 6 through the injection of a neutralizer.

In addition to neutralizers, filming amine type corrosion inhibitors are often used to combat
chloride corrosion in crude towers as well as other overhead systems in the refinery. Injection of
inhibitors into the top of the tower can lead to fouling problems in the tower and in downstream
heat exchangers. Ammonia used in overhead neutralization must be carefully controlled to avoid
the formation of solid ammonium chloride deposits and can cause corrosion problems with
copper based alloys. These deposits can be very corrosive if not removed by wash water.

If a change of materials appears necessary, replacement of carbon steel tubes with titanium tubes
will alleviate most crude unit overhead condenser corrosion problems. Monel has commonly
been used to line the top section of crude towers and reflux drums. This approach is dependent
on a low concentration of sulfides in the process. Admiralty and 70-30 CuNi tubes have also been
used with some success but may corrode rapidly when the pH is in excess of 7.5.

The most effective method for reducing chloride corrosion is the elimination of brine from crude
oil by proper tank settling and desalting. In desalting, water is injected into preheated crude to
dilute the brine and dissolve solid salt particles. The water is then separated from the crude oil in
a large separator or desalting drum. This drum contains electrodes which produce an
electrostatic field to help coalesce the brine droplets for withdrawal from the drum. Injection of
surface active chemicals can assist in the desalting operation.

At some refineries, dilute fresh caustic is injected into desalted crude oil to react with any
magnesium chloride and calcium chloride which easily hydrolyze to HCl. However, NaCl may
still form hydrogen chloride in the crude feed furnace but in reduced quantities. The resultant
sodium chloride leaves the crude tower with the reduced crude. This may not be acceptable for
downstream units or product specifications.

5.2.10 Hydrofluoric Acid


Hydrofluoric acid is used as a catalyst in some alkylation units instead of sulfuric acid. In general,
concentrated hydrofluoric acid is less corrosive than hydrochloric acid because it passivates most
metals by formation of protective fluoride scales. As long as feed stocks are kept dry, carbon steel,
with appropriate corrosion allowances, can be used for vessels, piping and valve bodies in
hydrofluoric alkylation units. Carbon steel is used up to about 71oC (160oF). Fluoriding gets
excessive above this temperature. Carbon steel welds should be postweld heat treated and
hardness limits similar to the limits used for preventing wet sulfide cracking must be imposed to
prevent stress corrosion cracking. Monel 400 is used selectively at locations where corrosion of
carbon steel is expected. Heat exchanger tubes, valve trim, valve stems and thermowells are

Page 2-32
CALTEX REFINERY MATERIALS MANUAL October 1999

often made from Monel 400. Teflon valve seats may be required where corrosion has caused
valves to freeze up. Pump internals are often made from Monel 400 (with Monel 500 shafts) or
Hastelloy B-2.

Experience has shown that most corrosion problems in HF alkylation units can occur after
shutdowns because pockets of water have been left in the equipment. This water originates with
the neutralization and washing operation which are required for personnel safety prior to
opening equipment for inspection and maintenance. It is very important that equipment be
thoroughly dried, by draining all low spots and by circulating hydrocarbon, prior to introduction
of the hydrofluoric acid catalyst. HF is very damaging to health and is a particular problem
because it vaporizes at atmospheric pressure. It is also imperative that good welding and
threading practices be followed because hydrofluoric acid has the capability to find the smallest
holes in welded or threaded connections. Similarly, flange gasket leakage can be a problem
unless flanged connections are carefully made up.

5.2.11 Naphthenic Acid


Naphthenic acid corrosion is an aggressive form of corrosion associated with some crude oils
from California, Trinidad, Venezuela, Mexico, Eastern Europe and Russia as well as some very low
sulfur Southeast Asian crudes. Naphthenic acid is a collective name for organic acids primarily
composed of saturated ring structures with a single carboxyl group. These, along with minor
amounts of other organic acids, are found in naphthenic based crudes. Their general formula
may be written as R(CH2)nCOOH where R is usually a cyclopentane ring. The higher molecular
weight acids can be bicyclic (12<n<20), tricyclic (n>20), and even polycyclic. Naphthenic acid
content generally is expressed in terms of neutralization or Total Acid Number (TAN), which is
determined by titration of the oil with potassium hydroxide (KOH), as described in ASTM D664.
The TAN is the milligrams of KOH required to neutralize one gram of stock.

Naphthenic acids are generally heavy organic carboxylic acids that occur in many crude oils.
They are corrosive in the 232oC-398oC (450o-750oF) temperature range. Typically, naphthenic
acid corrosion leaves a scale-free eroded looking surface with wastage rates similar to sulfide
corrosion on carbon steel. The corrosion is sensitive to velocity and, in higher velocity areas, rates
may approach an inch per year. Unlike sulfide corrosion, naphthenic acid attack does not
decrease with increasing chromium content . The mechanism appears to involve the continuous
removal of the normally protective sulfide scale. The standard 12Cr used for sulfide protection in
columns can suffer high corrosion rates. Type 316 austenitic stainless steel with its molybdenum
content generally offers good naphthenic acid corrosion resistance. However, higher
molybdenum alloys have been used to give reasonable corrosion resistance in some heavy high
naphthenic acid West Coast U.S. crudes. Problem areas typically occur in high velocity and
condensation areas, vacuum heater tubes and outlets, transfer lines and columns, although
atmospheric columns, heater tubes, transfer lines and side stream strippers are sometimes
involved. Pumps are often involved due to the high velocities around the impellers.

The old rule of thumb that was developed primarily from the heavy Venezuelan and West Coast
U.S. crudes was that neutralization numbers over 0.5mg KOH for the whole crude or 1 mg KOH
for any of the gas oil range cuts could be a problem. This rule was felt to be very conservative
since most problems occurred at neutralization numbers greater than one on the whole crude
and over two on the side cuts. However, as more Southeast Asian crudes have moved into the
market, naphthenic acid-like problems are cropping up with very low neutralization numbers [i.e.
0.1). Very high corrosion rates have been experienced and a switch of materials to Type 316 has
generally solved the problem. These Southeast Asian crudes generally have very low sulfur
contents. The theory is that the sulfide scale is somewhat protective and is needed to inhibit the
naphthenic acid corrosion.

Page 2-33
CALTEX REFINERY MATERIALS MANUAL October 1999

Type 304 austenitic stainless steel is marginally resistant to naphthenic acid corrosion under
conditions of low velocity, although it is subject to severe pitting. Type 316 and 317
molybdenum containing austenitic stainless steels have higher resistance to naphthenic acids and
provide adequate protection under most circumstances. Experience indicates a minimum
molybdenum amount of 2.2% is needed for effective resistance. Aluminum has excellent
resistance to naphthenic acid and can be used where its strength and erosion resistance
limitations are not a problem. Alloy 20 stainless steel is also highly resistant. Very high TAN
crudes and side cuts may require higher molybdenum alloys such as 625 and 904L.

Crude blending has been very successful in minimizing naphthenic acid corrosion problems.
Caustic neutralization generally has not worked because of the amount of caustic required to
neutralize causes corrosion and potential cracking problems. The molybdenum bearing austenitic
stainless steels and higher molybdenum nickel alloys are usually effective from a materials
standpoint. Onstream radiographic monitoring has been used to follow in-service corrosion.

While naphthenic acids are usually thought of as a crude unit problem, downstream units may be
affected. High neutralization number sidecuts can cause corrosion problems in the feed side of
FCC and hydroprocessing units prior to the reactors. The naphthenic acids break up over the
catalyst.

There have been curves developed for naphthenic acid corrosion but they usually reflect
experience with a specific crude and many times the velocity effect is not included. Except for
high velocity areas, it is reasonable to assume corrosion rates similar to sulfide corrosion on
carbon steel up through 12Cr alloys in naphthenic acid. Austenitic stainless steels generally are
somewhat better in low velocity areas but at least SS Type 316 with its molybdenum content is
needed in the higher velocity areas. Most Caltex refineries are not well protected for running
unblended crudes that are known for naphthenic acid corrosion problems. If a refinery is required
to run a known naphthenic acid corrosion causing crude or there is no experience with the
specific crude, it would be wise to set up an inspection program with radiography in known high
velocity areas to monitor possible high corrosion losses on line. Experience with some of the very
low sulfur Southeast Asian crudes with low neutralization numbers has resulted in rates on carbon
steel in the high velocity target areas on a atmospheric transfer line in excess of 1000 mpy.

It should be recognized that many crudes are blends and that sulfur and neutralization numbers
can change depending on specific wells, field age and secondary or (tertiary) recovery
techniques.

Naphthenic acid corrosion is best controlled by blending crude oils having high neutralization
numbers with other crude oils. Blending is designed to reduce the naphthenic acid content of the
worst sidecut. In practice, blending means that the charge to the crude distillation unit should
have a TAN no higher than 1.0. When blending is insufficient to prevent attack, affected areas are
alloyed with Type 316 or 317 stainless steel or higher molybdenum materials. The very low sulfur
Southeast Asian crudes that have caused problems have to be blended with higher sulfur crudes.

5.2.12 Organic Chlorides


Organic chlorides that contaminate crude oil will produce various amounts of hydrogen chloride
at elevated temperatures, depending on the chlorides involved. While most major oil producers
are aware of the problem, other operators have been known to continue to use organic chloride
solvents to remove wax deposits. These solvents are also used extensively for metal degreasing
operations within and out of the refinery. Often, spent solvent is discarded with slop oil and later
mixed with crude oil charged to the crude unit. Some solvent dewaxing units use

Page 2-34
CALTEX REFINERY MATERIALS MANUAL October 1999

dichloroethyene as a solvent. Many of the organic chlorides breakup in the atmospheric column
and condense in the overhead system. Some of the chlorides deposit on the feedside of the
downstream hydrotreater feed/effluent exchangers causing severe pressure drop problems.

Contaminated crude oils have been found to contain as much as 7000 ppm chlorinated
hydrocarbons. Such contaminated crude oils not only cause severe corrosion in the overhead
system of distillation towers but also affect downstream reformer operations. Typical problems in
the latter case include cracking, rapid coke buildup on the catalyst and increased corrosion in
fractionator overhead systems. Obviously, every effort must be made to avoid charging
contaminated crude oil. If contaminated crude oil must be run off, the usual approach is to blend
it slowly into uncontaminated crude oil, so that the organic chloride content of the charge stays
below 1 to 2 ppm. However, many times you do not have advance warning. The Monel
cladding used in the top of the atmospheric column is rapidly corroded by any chloride
condensing in the tower due to a low overhead temperature operation. Monel is reasonably
resistant to dilute HCl but corrodes rapidly in the condensing more concentrated acid.

There are other corrosion problems caused indirectly by organic chlorides. For example, organic
chlorides are routinely used to regenerate reformer catalyst. Hydrogen chloride tends to be
stripped off the catalyst if excessive moisture is present in the naphtha feed. This causes increased
corrosion, not only in reformers, but also in desulfurizer sections which use hydrogen makeup gas
produced in reformers. This is particularly a problem on continuous reformers. There is chloride
being stripped off the catalyst during operation and chloride being added all the time. The net
result is that considerable chloride is available to cause corrosion on the reformer and on
downstream units that use the hydrogen generated in the reformer. There are chloride traps
available for both the hydrogen and liquid reformer streams in order to minimize downstream
chloride problems.

5.2.13 Phenol (Carbolic Acid)


Phenol or carbolic acid is used in refining operations to convert heavy, waxy distillates into
premium grade lubrication oils. The treating section where feed is contacted with phenol
operates at temperatures below 121oC (250oF) and carbon steel is satisfactory in this section. In
the raffinate recovery section, phenol is separated from the treated oil or raffinate. Here again,
there are few corrosion problems with carbon steel.

In the extract recovery section, spent phenol is separated from the extract by vaporization.
Experience indicates that every piece of equipment in this section may have a completely
different corrosion history during different periods of operation. Towers generally need to be
lined with Type 316 austenitic stainless steel and solid Type 316 tower internals are often used.

Above 232oC (450oF), both carbon steel and Type 304 austenitic stainless steel will corrode
rapidly in phenol service, especially when erosion corrosion is likely to occur. Tubes, return bends,
and headers in extract furnaces must be made of Type 316. In some cases, Hastelloy C-276 may
be needed to combat erosion corrosion.

In addition to phenols, furfural and NMP (N-methyl-2-pyrrolidons) are used in lube oil
production. Both have problems in the extract recovery section similar to phenol based units.
Many lube oil stocks have naphthenic or other similar heavy organic acids that can cause very
rapid corrosion of Type 316 stainless steel. Many of the phenol and furfural lube oil treating units
have been modified to run NMP.

Page 2-35
CALTEX REFINERY MATERIALS MANUAL October 1999

5.2.14 Phosphoric Acid


Phosphoric acid is sometimes used as a biological nutrient in refinery process water treatment
plants. Corrosion by phosphoric acid depends somewhat on the methods of manufacture and
the impurities present in the commercial finished product. Fluorides, chlorides and sulfuric acids
are the main impurities present in the manufacturing process and in some marketed acids. For
example, corrosion resistance of the high silicon irons, the austenitic stainless steels without
molybdenum, and tantalum are appreciably affected by the presence of small amounts of
hydrofluoric acid in phosphoric acid.

Two of the most widely used alloys for handling phosphoric acid are Type 316 stainless steel and
Alloy 20. These alloys show very little attack in acid concentrations up to 85% and temperatures
up to boiling. Lead and its alloys are also used at temperatures up to 93oC (200oF) and
concentrations up to 80% for pure acid. Lead forms an insoluble phosphate that provides
protection. High silicon irons, glass, and stoneware show good resistance to pure acid. Copper
and high copper alloys are not widely used in phosphoric acid services. High nickel- molybdenum
alloys exhibit good resistance to pure acids but are attacked when aeration and oxidizing
impurities are present. Aluminum, cast iron, steel, brass, and the ferritic and martensitic stainless
steels exhibit poor corrosion resistance.

Phosphoric acid is also used as a catalyst on clay pellets for polymerization of light ends for
octane blending stocks. Moisture strips the phosphoric acid off the pellets and may result in
severe pitting problems in the piping around the catalyst reactors.

5.2.15 Process Chemicals


Process chemicals can cause severe corrosion problems. Hydrogen chloride, stripped off reformer
catalyst by moisture in the feed, falls into this category. Corrosion can also be caused by caustic
and other neutralizers which, ironically, are added to control acid corrosion. Filming amine type
corrosion inhibitors are quite corrosive if injected undiluted into a hot vapor stream. Another
group of process chemicals which are corrosive, or become corrosive, are solvents in treating and
gas scrubbing operations.

5.2.16 Soil Corrosion


Soil corrosion is a major problem with underground piping and tank bottoms unless coatings and
cathodic protection are applied. Coatings must be resistant to various hydrocarbons since there is
always the likelihood of leaks or spillage. Cathodic protection is sometimes not practical because
of the multitude of buried piping involved and the concern for stray current corrosion.

Soil corrosion is caused by differential concentration cells involving oxygen, water and various
chemicals in the soil. Incomplete mill scale on piping and tank bottoms, bacterial action, pinholes
in protective coatings, and coupling of dissimilar metals all contribute to soil corrosion. This is
why we grit blast bottom sides of tanks. Corrosion rate is inversely proportional to the electrical
resistivity of the soil. Corrosion rate can be reduced by excavating and backfilling with clean,
nonconductive sand. Soil corrosion can also occur on the bottom of piping which is laid directly
on the ground. If grass or weeds are allowed to grow beneath and around piping, moisture will
remain for long periods of time and the piping or tanks are bound to corrode. The best practice
for preventing soil corrosion is to locate piping well above grade and to isolate tank bottoms
from the soil by using underside membranes, asphalt, or concrete pours. Current practice for
above ground storage tank leak prevention will dictate the tank bottom design and corrosion
mitigation features.

Page 2-36
CALTEX REFINERY MATERIALS MANUAL October 1999

5.2.17 Steam Condensate


Steam condensate corrosion is caused by dissolved oxygen and carbon dioxide. Oxygen
corrosion in condensate systems is in the form of pitting. In contrast, carbon dioxide corrosion
usually takes the form of uniform metal loss. Thinning and longitudinal grooving of the lower
portion of piping and heat exchanger tubes points to carbon dioxide corrosion as the most
probable cause. CO2 enters the steam condensate system either as dissolved gas or as
bicarbonate or carbonate alkalinity in boiler makeup water. Dissolved carbon dioxide normally
will be removed by properly operated deaeration equipment. However, external treatment
methods are required to reduce the alkalinity of the makeup water. Condensate corrosion can be
controlled by injection of filming amine corrosion inhibitors, usually in conjunction with
ammonia or organic neutralizers such as morpholine or cyclohexylamine.

5.2.18 Sulfuric Acid


Sulfuric acid is used as a catalyst in alkylation units and in the regeneration process for
demineralized water trains. Carbon steel is widely used for sulfuric acid concentrations above
85% and is usually not corroded provided temperatures are below 380oC (100oF) and velocities
are kept under 2 feet/second (0.6 m/s). However, attack in the form of erosion corrosion will
occur at locations of high turbulence. Examples of such locations are elbows and tees in
alkylation unit reactor effluent piping, especially just downstream of the points where dilute
caustic or wash water are added. In piping systems handling concentrated sulfuric acid, pipe
erosion is often seen around transfer pumps due to the increased turbulence.

Figure 2-1 shows corrosion of steel by strong sulfuric acid as a function of temperature and
concentration. The curves represent corrosion rates of 5, 20, 50, and 200 mpy. These are iso-
corrosion or constant corrosion rate lines. The corrosion of steel by strong sulfuric acid is
complicated because of the peculiar dip in the curves in the vicinity of 101% acid. The
narrowness of this range means that the acid must be carefully analyzed to reliably predict
corrosion. The dips or increased attack around 85% are more gradual and less difficult to
establish. Contaminated acid can behave very differently than pure acid.

Selective lining with Alloy 20, Hastelloy C-276 or B-2 may be required to handle low
concentration situations because of temperatures generated by the heat of dilution/mixing.
Carbon steel valves usually require Alloy 20 trim because even slight attack of the carbon steel
seating surfaces is sufficient to cause leakage. Line flange bolting, equipment and valves are also
often specified to be Alloy 20 for H2SO4 service.

Page 2-37
CALTEX REFINERY MATERIALS MANUAL October 1999

Steel

275 527

200 Over 200 mils

225 437

Temperature, oC
Temperature, oF

175 50 - 200 mils 347

50
125 257

20 20 - 50 mils

75 167

5
5 - 20 mils
0 - 5 mils
60 65 70 75 80 85 90 95 100 105 110
-3
mil = 10 in /year corrosion rate

CORROSION OF CARBON STEEL BY SULFURIC ACID


FIGURE 2-1

5.3 STRESS CORROSION CRACKING (SCC)


Stress Corrosion Cracking (SCC) is the spontaneous cracking of alloys by a combination of
corrosion and tensile stress. Failure is frequently caused by simultaneous exposure to a seemingly
mild chemical environment and to a tensile stress well below the yield strength of the material.
Fine cracks penetrate deeply into the part while the surface exhibits only faint signs of corrosion,
and often an apparent brittle fracture may occur in what would normally be a ductile material.

In SCC, the stresses involved may be residual stresses in the metal, such as from bending or
welding, or from uneven heating or cooling. Applied stresses, such as working stress from
internal pressure or structural loading, also can be involved. In general, however, residual stresses
are the major factor in SCC. Only tensile stress results in SCC and compressive surface stress, if
anything, has a beneficial effect. Peening to introduce compressive surface stress has been used
as a preventative measure under some circumstances.

Some aspects of SCC, pertinent to refineries are worth discussion:

Page 2-38
CALTEX REFINERY MATERIALS MANUAL October 1999

Specific combinations of corrosives and alloys result in cracking and practically all alloys will
crack under some conditions. Table 2-4 is a list of alloy systems and environments known to
cause SCC. Austenitic stainless steels crack in the presence of chlorides and strong caustic.
Carbon steel cracks in caustic, amines and carbonate solutions. Titanium alloys crack in
certain strong acids and copper alloys crack in ammonia. Usually, a certain alloy cracks in a
certain medium and, if the composition of the alloy or the medium is changed slightly,
cracking may become more or less severe. Trivial changes in residual or alloying elements
may have an effect on cracking way out of proportion to their concentration. For example,
pure copper is immune to SCC in ammonia, but if it is alloyed with as little as 0.1%
phosphorus, it becomes extremely susceptible.

Oxygen may play an important role in stress corrosion cracking of alloys used in refining
applications. For example, polythionic acids require oxygen to form and ammonia stress
corrosion cracking can occur when equipment is exposed to air.

There seems to be no consistent pattern as to whether the fracture path through an alloy is
along grain boundaries (intergranular) or through grains (transgranular). At times, both
modes of cracking occur simultaneously and the cracks can be heavily branched or
unbranched. This variation of crack appearance for a given alloy in a given environment can
cause confusion when investigating root cause failures in refinery process streams.

The cracking process has three distinct stages; initiation, propagation, and fast fracture. The
initiation stage can last a few minutes or several years. The propagation stage usually
proceeds at a relatively constant rate. As cracking progresses, the effective cross section or
wall thickness of the component is reduced, which leads to fast fracture. Mechanical rupture
then occurs. The minimum stress for SCC can be as low as 10% of the alloy's yield strength.
At stresses near the yield point, failure sometimes occurs almost immediately upon exposure
to the corrosive environment.

Page 2-39
CALTEX REFINERY MATERIALS MANUAL October 1999

TABLE 2-4 ALLOY SYSTEMS SUBJECT TO STRESS CORROSION CRACKING

ALLOY ENVIRONMENT

Aluminum Base Air


Sea Water
Salt & chemical combinations
Magnesium Base Nitric acid
Caustic
HF solutions
Salts
Coastal atmospheres
Copper Base Primarily ammonia & ammonia
hydroxide
Amines
Mercury
Carbon Steel Caustic
Anhydrous ammonia
Nitrate solution
H2S solutions
Martensitic and Sea water
Precipitation Hardening Chlorides
Stainless Steels H2S solutions
Austenitic Stainless Steels Chlorides - inorganic & organic
Caustic solutions
Sulfurous & polythionic acids
Nickel Base Caustic above 315oC (600oF )
Fused caustic
Hydrofluoric acid
Titanium Sea water
Salt atmospheres
Fused salt

5.3.1 Chloride Stress Corrosion Cracking


Aqueous environments containing chlorides at temperature above roughly 52oC (125oF) can
crack stressed austenitic stainless steels. All of the commonly used austenitic stainless steels such
as types 304, 304L, 316, 316L, 321 and 347 are susceptible. The resistance generally increases
with increasing nickel content. For example, alloy 825 is generally very resistant to chloride stress
corrosion cracking. The chromium stainless steels and the duplex ferritic/austenitic stainless steels
are generally quite resistant. There does not appear to be a practical lower chloride limit if
chlorides can concentrate in the specific environment. Chloride stress corrosion cracking must be
considered anytime austenitic stainless steels are considered for refinery service. In many
instances, the chloride stress corrosion cracking potential is highest during startup or shutdown,
downperiod and nitrogen/ammonia blanketing or soda ash washing for pH adjustment must be
considered. Chloride stress corrosion cracking in austenitic stainless steels generally shows up as
multiple branched transgranular cracking.

A similar transgranular cracking of austenitic stainless steels can also occur in caustic. From a
failure analysis standpoint, many steam service expansion joint bellows failures have been blamed
on chlorides which are usually present in the deposits found after the failure. However, cracking
due to caustic from the steam system is probably as likely a reason for the failures as the
chlorides. Either problem can probably be solved by using a higher nickel alloy such as Inconel

Page 2-40
CALTEX REFINERY MATERIALS MANUAL October 1999

625. The 625 alloy can be used up to a bellows metal temperature of about 538oC (1000oF).
Inconel 625 will embrittle due to a precipitation problem above this temperature.

While downperiod protection is usually aimed at minimizing polythionic acid cracking the higher
pH environment will also minimize chloride cracking problems.

CONDITIONS REQUIRING STRESS RELIEF


OF CARBON STEEL IN CAUSTIC SERVICE

260

Caution: General excessive corrosion of


240 carbon steel can occur at these 116
temperatures

220

93
200
Stress Relief Necessary
to Prevent Cracking
180

71
160

TEMPERATURE C
TEMPERATURE F

140
o

o
49
120

100

27
80

60 No Stress Relief Necessary

4
40

20

0 10 20 30 40 50

CONCENTRATION NAOH, % BY WEIGHT


Figure 2-4

Page 2-41
CALTEX REFINERY MATERIALS MANUAL October 1999

5.3.2 Caustic Stress Corrosion Cracking


Caustic SCC is typically seen in carbon steel boiler components. SCC of carbon steels under
tensile stress and exposed to caustic, amine, and carbonate solutions generally occurs at
temperatures above 66oC (150oF), 24oC (75oF), and 38oC (100oF), respectively. Cracks are
intergranular, oxide filled, and the fracture surface appears to have been embrittled.

Caustic SCC also occurs in ferritic steels and austenitic stainless steels. As with chloride SCC,
alternating wet and dry conditions can cause caustic to concentrate. Unlike chloride SCC, oxygen
need not be present for cracking to occur. Residual tensile stress has been clearly found to be a
major factor in caustic SCC and therefore, postweld heat treatment (stress relief) is often used to
provide resistance to cracking. Similarly, cold formed components are stress relieved for caustic
services. Caustic concentration of 50 to 100 ppm are believed sufficient to cause cracking.

NaOH is widely used in refinery operations to neutralize acid constituents. At ambient


temperature, caustic can be handled in carbon steel equipment. Carbon steel is also satisfactory
for aqueous caustic solutions up to roughly 66oC (150oF). For caustic service at temperatures
above roughly 49oC (120oF), carbon steel must be postweld heat treated to avoid SCC at welds
and heat affected zones. Austenitic stainless steels, such as Type 304, may be used up to 93oC
(200oF), while nickel alloys such as Monel 400 or Nickel 200 are required at higher temperatures.
If sulfur compounds are present in caustic at elevated temperatures, Nickel 201 should be used.

Dilute caustic (3-6% aqueous solution) is normally injected into hot, desalted crude oil to
neutralize any remaining hydrogen chloride. When concentrated caustic is used, severe corrosion
of the crude piping just downstream of the caustic injection point can occur. Appropriate
dispersion of the dilute solution in the hot crude oil prevents puddles of molten caustic from
collecting along the bottom of the pipe where caustic droplets can cause severe attack.

There are some unusual situations in refineries where caustic corrosion is encountered. For
example, traces of caustic can become concentrated in boiler feed water and cause SCC,
sometimes referred to as caustic embrittlement. This occurs in boiler tubes which alternate
between wet and dry conditions (steam blanketing) due to overfiring or poor design. Locations
such as cracked pressure welds or leaky tube rolls can concentrate caustic and lead to caustic
stress corrosion cracking.

Caustic corrosion or gouging is found under deposits in heat exchangers which generate steam.
Vertical heat exchangers are especially vulnerable if deposits are allowed to build up on the
bottom tubesheet. Boiler feed water permeates these deposits and evaporates, concentrating
caustic in the liquid that is left behind. The caustic content of the trapped liquid can reach several
percent, more than enough to destroy the normally protective iron oxide film on the boiler steel
and thus cause severe corrosion. Severe caustic corrosion has been found on the bottom
tubesheet of some of the large true countercurrent flow combined feed exchangers on catalytic
reforming units. The caustic collects on the tubesheet during regeneration.

Steam tracing may raise the temperature enough to cause cracking problems. Sometimes simply
the heat of dilution of the caustic is enough to cause problems. Proper water chemistry control
minimizes boiler problems although slight leaks can concentrate caustic and cause a cracking
problem (i.e., tube roll leaks at the drums). Carryover from a caustic scrubbing operation to a
non-stress relieved column has caused severe cracking in sulfuric acid alkylation units. Over-
injection of caustic in crude unit overhead chloride control has caused cracking in fuel oil tanks.

Page 2-42
CALTEX REFINERY MATERIALS MANUAL October 1999

5.3.3 Carbonic Acid


In steam reformer hydrogen plants, the hydrogen is produced by the water-gas reaction of
methane and steam at high temperature in conjunction with a catalyst. The steam and methane
reform into hydrogen, carbon monoxide and carbon dioxide. In a subsequent stage, the carbon
monoxide reacts with additional water to form carbon dioxide and hydrogen. The effluent gas is
then contacted with an alkaline solution, such as potassium carbonate, to remove the carbon
dioxide.

Whenever the system operates below the dew point of water, corrosion of steel caused by CO2 is
likely and can be severe. Where condensation occurs, corrosion rates exceeding 1 inch/year (2.5
cm/y) have been experienced. Corrosion resistant alloys such as Monel, aluminum, stainless steels
and copper-nickel must be used in these cases. The addition of at least 12% chromium as an
alloying element confers corrosion resistance to steel from attack by CO2.

In vapor recovery sections of catalytic cracking units, CO2 may react with ammonia, forming
ammonium carbonates. When the pH is over about 8.5, the ammonium carbonate may cause
SCC of steel piping and equipment.

5.3.4 Polythionic Acid Stress Corrosion Cracking (PASCC)


The sulfur acids that cause the severe intergranular corrosion are a downperiod problem on most
hydrodesulfurizers and hydrocrackers. However, on FCCs dew point conditions can occur during
operation that expose some equipment like slide valve bonnets and catalyst withdrawal lines to
cracking environments.

The stress required for PASCC of a sensitized 3xx Series 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 PASCC.

There are three basic means to prevent PASCC:

use alloys which resist sensitization,


isolate sensitized stainless steels from the sulfur-derived acids,
prevent polythionic acid formation.

In terms of alloy, lowering the carbon content of standard austenitic stainless steels (using 'L'
grades such as Type 304L) or specifying stabilized 3xx Series stainless steels (Type 321 or 347)
improves sensitization resistance. Both modifications reduce the tendency of 3xx Series stainless
steels to form low corrosion resistant chromium-depleted zones. Specifying low carbon material
minimizes the amount of chromium carbide that can form. However, the low carbon grades will
sensitize in a few hours with exposure above about 538oC (1000oF). Specifying, stabilized
stainless steels introduce active carbide formers which, compared to chrome, react preferential
with carbon.

Polythionic acids (PTA) are formed during shutdowns and turnarounds when equipment that has
been exposed to H2S/H2, at elevated temperatures is opened to the atmosphere. Sulfide scale
with steamout during the shutdown and moisture and oxygen in the air while down form acids
of the type H2SxOy. (Refer to Section 3- 5.5.4 and 8.5.7)

Page 2-43
CALTEX REFINERY MATERIALS MANUAL October 1999

Some austenitic stainless steels such as Types 304 and 316 and some high nickel alloys are very
susceptible to cracking when exposed to these environments in a stressed and sensitized
condition.

Sensitization is the precipitation of a continuous grain boundary carbide network. In susceptible


alloys the precipitation is time, temperature and chemistry dependent in the 399o-871oC (750o-
1600oF) temperature range. The continuous network of grain boundary chromium carbides
depletes the adjacent area of chromium allowing a combination of stress and certain corrosives
to rapidly attack or crack the equipment. Polythionic acid cracking will show up as intergranular
cracking in sensitized materials.

Considerable care must be taken in selecting alloys for various services since exposure to
sensitizing temperatures during fabrication or operation can cause subsequent rapid cracking and
failure. Type 304 can sensitize in 3 minutes at 649oC (1200oF) whereas it may take 100,000 hours
at 427oC (800oF). The stabilized grades of austenitic stainless steel Type 321 with titanium and
Type 347 with niobium (columbium) form stable carbides preventing networks of chromium
carbides in the grain boundaries. It is possible to destabilize Types 321 and 347 during mill
solution anneals, and "GPS-A9" specifies that a mill stabilizing heat treatment be run after
solution anneal. This 885 14oC (1625 25oF) heat treatment positively stabilizes the materials
so that-subsequent exposure will not sensitize them. Some refiners call for a stabilizing heat
treatment after welding to restabilize the heat affected zone. Caltex believes that this is necessary
only if one pass welds are used. Multipass welds will tend to restabilize the pass heat affected
zones beneath them.

If the austenitic stainless steels will be exposed to sensitizing conditions only during welding, use
L or low carbon grades. The low carbon grades will not form carbide networks during the
welding temperature cycle but will sensitize with longer exposure times.

Because of the potential chloride stress corrosion and polythionic acid cracking problems, many
refiners specify either a nitrogen/ammonia blanket during shutdown or wash with a 1.5% soda
ash solution if the equipment is opened to the atmosphere. The ammonia and soda ash tend to
raise the pH making chloride stress corrosion cracking less likely and will tend to neutralize the
polythionic acids that might be formed during exposure to air.

Proper shutdown procedures to exclude oxygen and moisture are also effective deterrents to PTA
cracking. NACE RP0170-93, "Recommended Practice for Protection of Austenitic Stainless Steels
in Refineries Against Stress Corrosion Cracking by Use of Neutralizing Solutions During
Shutdown" discusses prevention measures.

Page 2-44
CALTEX REFINERY MATERIALS MANUAL October 1999

1500

1400 760

1300

1200 649

Heating Temperature, C
Heating Temperature, F

o
1100

1000 538

900

800 427
1 10 100 1000 10,000
Time at Heating Temperature, Hours
347

304L

304

321
FIGURE 2-5
TIME-TEMPERATURE-SENSITIZATION (TTS) CURVES FOR THE COMMONLY-USED
300-SERIES STAINLESS STEEL
Curves Correspond ro 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 guidance purposes only.

5.3.5 Ammonia Stress Corrosion Cracking


Ammonia Stress Corrosion Cracking in refineries is typically seen in copper-zinc alloy heat
exchanger tubing. Fractionation tower overhead systems often have admiralty brass condenser
bundles installed for cooling water and process side corrosion resistance. If the tubes contain
residual stress from tubing fabrication or tube expanding, ammonia stress corrosion cracking can
occur. Ammonia bearing deposits can be especially aggressive when oxygen is introduced when
equipment is opened. Copper alloys are used in a variety of refinery water applications. Some of
these contain organic matter, which on decay, produces ammonia in systems where it would not
normally be expected.

Page 2-45
CALTEX REFINERY MATERIALS MANUAL October 1999

Cu-Zn alloys like admiralty and aluminum brass are very susceptible to stress corrosion cracking
in aqueous environments containing air and ammonia or ammonium compounds. The cracking
is generally transgranular. A common remedial action is to change materials to 70-30 Cu-Ni that
is not susceptible to cracking or to water wash the deposits before exposing the equipment to
air.

Carbon steel is susceptible to cracking in anhydrous ammonia. Ammonia cargos are shipped with
a small amount of water which inhibits cracking.

Ammonium chloride will also stress corrosion crack Monel. Several of the early catalytic reformer
recycle gas compressors had Monel 500 impellers that stress corrosion cracked when exposed to
wet ammonium chloride deposits during shutdowns. Thorough water washing the compressors
before opening them minimized the cracking problems. In fact., the low alloy Cr-Mo-V impellers
used in modern compressors are still washed if they tend to build up deposits. When wet, these
deposits will corrode the low alloy impellers rapidly when exposed to air.

Ammonia is not widely used in boilers as a neutralizer for CO2, mainly because it is corrosive to
copper alloys which are commonly found in steam and condensate exchange equipment. For an
all-steel system, however, ammonia would be a suitable neutralizer.

5.3.6 Low Temperature Hydrogen Related Cracking


At temperatures up to about 232oC (450oF), hydrogen sulfide under dew point conditions is a
relatively mild acting corrosive to carbon steel and general corrosion rates tend to be low.
However, during the corrosion process, considerable amounts of hydrogen can be liberated and
it can have several significant, detrimental effects on welded pressure containing components.

Equipment in light ends service on Fluid Catalytic Cracking, Crude, Visbreaker and
Hydroprocessing units have experienced sulfide stress corrosion cracking, hydrogen blistering,
hydrogen induced cracking, stress oriented hydrogen induced cracking, and carbonate stress
corrosion cracking as well as fabrication related cracking and quality problems. Recent industry
surveys of approximately 6000 pieces of equipment have shown that cracking of some sort has
been found (primarily by wet fluorescent magnetic particle inspection] in about 30-40% of the
equipment. Service related cracking has accounted for 10-20%, fabrication problems the
remainder. In some reports, service related cracking was attributed to causing a much higher
percentage of problems.

Hydrogen embrittlement affects many materials but, in refinery service, normally causes problems
with carbon steel, low alloys and the chromium stainless steels. The cracking is a result of high
stresses and hydrogen pickup due to HCl or wet sulfide corrosion and is often promoted by
cyanides in the FCCU gas recovery streams. Hydrogen embrittlement or sulfide stress corrosion
cracking usually affects hardened or heat treated parts that exceed 240 Brinell hardness. Care
must be taken in selecting materials for heat treated parts to insure they are in a metallurgical
condition that will minimize cracking problems. Problems also occur in relief valve springs,
compressor valve springs, compressor impellers and hardware. Remedial action includes lowering
stress, lowering hardness and/or by heat treating restrictions. Sometimes modifying the
environment by steam jacketing, adding water or by specifying plastic coatings is used. Very
specific heat treating restrictions have been placed on compressor manufacturers to limit the
yield strength of impellers and other highly stressed parts to 90,000 psi. The added requirement
of utilizing a heat treatment such that no untempered martensite is formed greatly minimizes
potential sulfide stress corrosion problems.

Page 2-46
CALTEX REFINERY MATERIALS MANUAL October 1999

In the late 1960's, sulfide cracking of carbon steel vessels occurred as a result of fabricators using
certain welding wire-flux combinations during fabrication. Some of the welding procedures
resulted in abnormally high weld hardness and caused unexpected sulfide cracking failures. As a
result, a hardness limit of 200 Brinell and certain welding restrictions have been placed on
fabricators of carbon steel equipment. Since the early 1970's, LPG spheres and bullets and certain
light ends processing equipment have experienced sulfide stress corrosion and other hydrogen
related cracking problems.

Sulfide stress corrosion cracking and fabrication related hydrogen cracking tend to show up as
heat affected zone or weld cap toe cracking in hard and untempered areas. Proper preheat
during fabrication and suitable postweld heat treatment minimize these problems. The preheat
and post heat treatment requirements of the construction codes may not be adequate to prevent
cracking. Steel chemistry restrictions to reduce elements in the steel composition that contribute
to hardenability and tempering resistance is recommended. Restrictions of C, Mn, Cr, Mo, Nb, V
and Ti are required. Minimizing these elements and at the same time retaining the generally
good notch toughness of today's modern steels can be a problem to steelmakers.

Hydrogen Blistering, Hydrogen Induced Cracking (HIC) and Stress Oriented Hydrogen Induced
Cracking (SOHIC) are a result of corrosion taking place in the system. Corrosion generates atomic
hydrogen which permeates through the steel and recombines at laminations and sulfide
inclusions. The recombined hydrogen molecule is too large to move through the steel's
microstructural lattice. Pressure is built up internally and blistering or cracking results. The
blistering and cracking are usually parallel to the surface but may be linked through wall by local
stress fields. Wet hydrogen sulfide and cyanides are the chief corrodents. Inhibitors have generally
not been effective, however, water washing and polysulfide injection have worked in some
instances. Stainless steel overlays and cladding and some coatings such as SS and nickel alloy
spray and guniting have been reasonably effective. Recent work by Chevron suggests that
aluminum spray should be avoided. Postweld heat treating has not been effective in preventing
HIC problems but does help with SOHIC. The specification of very low sulfur steels to minimize
inclusions plus inclusion shape control has shown some promise in minimizing the problems but
these steels have been generally expensive and sometimes hard to obtain.

Carbonate cracking has been experienced in some equipment in FCC fractionator overhead and
light ends service where draw water pH's have exceeded 8.5. Many refineries have found major
cracking problems. The cracking in carbon steel equipment is intergranular and located in weld
heat affected zones. This is similar to amine and caustic cracking and can be prevented by
postweld heat treatment.

From a mechanism standpoint the wet sulfide corrosion problems start as follows:

Atomic hydrogen (H) and molecular hydrogen (H2), are produced in the corrosion reaction of
steel with aqueous H2S as follows:

Fe + H2S FeS + 2H followed by 2H H2

Under ordinary acidic conditions, molecular H, produced by the above corrosion reaction, forms
at the surface of steel and, if produced slowly at low corrosion rates, it harmlessly evolves away.
But, when sulfide scale is present, the sulfide acts as a negative catalyst and discourages the
reaction 2H -+ H. This permits the atomic hydrogen to penetrate the steel where its accumulation
in the crystal structure affects mechanical properties. Other poisons that promote the entrance of
atomic hydrogen into steel are cyanide, phosphorous, antimony, selenium and arsenic ions.

Page 2-47
CALTEX REFINERY MATERIALS MANUAL October 1999

Atomic hydrogen is relatively small compared to the atomic space occupied by molecular
hydrogen and, as a result, the atomic form will diffuse through the steel's microstructure. Upon
diffusing into lattice defects, non-metallic inclusions, laminations, and other voids, the hydrogen
can form molecular hydrogen which is too large to diffuse out of the defect. This causes a
pressure build up sufficient to produce localized ruptures and fissuring.

Aside from corrosion, sources of atomic hydrogen capable of charging a steel with hydrogen are
acid pickling and cleaning operations, plating, steel mill furnace atmospheres, welding, cathodic
protection and hydrogen storage systems.

5.3.6.1 Wet H2S Cracking


Wet H2S cracking of the steel occurs during the advanced stage of hydrogen saturation. The
structure becomes brittle as a result of the many strains imposed on the metal lattice by the
presence of microbubbles of hydrogen gas. In such cases, the structure will fracture instead of
deforming when subject to stress. Micro-cracks exist in most structures, introduced by
fabrication, heat treatment or welding. In the absence of hydrogen, these are unlikely to
propagate. In the presence of hydrogen., sudden brittle failure at low stress levels can result. In
general, harder and higher strength steels are more susceptible to hydrogen embrittlement than
lower strength steels.

Embrittlement of the charged steel can be removed by low temperature heat treatment once
the component is removed from the hydrogen producing source. Molecular hydrogen
trapped in the steel may be very difficult to remove. For example, 200 oC for 2-4 hours is a
standard hydrogen out-gassing treatment.

5.3.6.2 Hydrogen Blistering


Hydrogen blistering is another effect of the entrance of atomic hydrogen into steel. Relatively soft
steels, containing inclusions from cold rolling or other fabrication steps, are very susceptible to
hydrogen blistering. Castings containing inclusions or blow holes are also subject to blistering.
Blisters begin as atomic hydrogen enters inclusions and forms molecular hydrogen. The
expansion pressure of the accumulating hydrogen gas produces a separation in the components
throughwall and becomes apparent as a blister on the metal surface. Hydrogen blisters usually
propagate parallel to the surface and do not pose a serious threat to pressure containment.
However, if 'stepwise cracking' at the edges of a blister occurs, especially if blisters exist at
different depths in the plate, the through-wall cracking must be considered. (Refer Section 2-
5.2.6.4)

Hydrogen blistering is controlled or eliminated by the reduction or elimination of hydrogen


activity. This can be done by using alloy or alloy clad materials resistant to hydrogen producing
corrosion or by inhibition of the corrosion process. The use of steels processed to minimize
inclusions has also been somewhat successful in combating hydrogen blistering.

Sulfide Stress Cracking (SSC), Hydrogen Induced Cracking (HIC), and Stress Oriented Hydrogen
Induced Cracking (SOHIC) are other low temperature forms of damage caused by hydrogen in
conjunction with high material hardness, dirty steel, and/or stress. In refineries, the source of
hydrogen is usually the exposure to wet hydrogen sulfide environments.

Page 2-48
CALTEX REFINERY MATERIALS MANUAL October 1999

5.3.6.3 Sulfide Stress Cracking (SSC)


Sulfide Stress Cracking (SSC) is cracking attributed to hydrogen in high strength, low ductility
microstructures that can be identified by high hardness. It is highly dependent on a steel's
composition, microstructure, strength, residual stress and applied stress levels. Small, localized
hard zones in welds and weld heat affected zones can initiate SSC even if the bulk material
hardness is quite low. In refineries, sulfide stress cracking is seen in hardened components such as
valve trim and compressor springs that are exposed to low temperature wet sulfide
environments. Cracking generally seems to be restricted to temperatures below 82oC (180oF). In
welded components, resistance to SSC can be provided by tempering or postweld heat
treatment which reduce hardnesses to 200 Brinell Hardness Number (BHN) or lower. A Rockwell
hardness limit of HRC 22 is often imposed as a control measure. This and other methods of
control to prevent SSC in carbon steel welds are discussed in NACE Standard RP0472. Postweld
heat treatment both reduces residual stresses and tempers the microstructure. Postweld heat
treating and the control of welding parameters are the best approach to providing SSC
resistance. Where cracking of standard 12% chromium steel valve trim is a problem, a change to
type 316 austenitic stainless steel can be considered or double heat treatment of the 12 Cr trim
used. High strength bolting can be given a modified temper to reduce hardness and thereby the
tendency to crack, although setting the gaskets in joints with lower strength bolting has been a
problem.

5.3.6.4 Hydrogen Induced Cracking (HIC)


Hydrogen Induced Cracking (HIC) is cracking that results from parallel hydrogen laminations
that link up to produce a throughwall crack with no apparent interaction with applied or residual
stress. It is a function of steel cleanliness and relates back to the method of steel manufacture,
impurities present, and their form. Typically associated with HIC are non-homogeneous,
elongated sulfide or oxide inclusions occurring parallel to the plate rolling direction. These serve
as sites for formation of microscopic hydrogen blisters that grow and eventually connect via
stepwise cracks. Since HIC is not stress dependent nor associated with hardened microstructures,
postweld heat treatment is of little value. HIC resistance is provided by restricting trace elements
and controlling manufacturing variables for the steel.

5.3.6.5 Stress Oriented Hydrogen Induced Cracking (SOHIC)


Stress Oriented Hydrogen Induced Cracking (SOHIC) is similar to HIC except that cracking is
stress driven and has a crack direction perpendicular to the primary stress direction. Since stress is
involved in crack development and propagation, postweld heat treatment is considered to be of
some benefit for resistance to SOHIC along with the control of manufacturing variables and trace
elements.

5.3.7 Hydrogen Cyanide


Hydrogen Cyanide (HCN), like ammonia, is formed from nitrogen bearing feed stocks in fluid
catalytic cracking and delayed coking units. HCN is a significant factor in hydrogen blistering and
cracking of pressure containing equipment, especially in the vapor recovery sections of these
units. Affected equipment includes fractionator overhead drums, compressor interstage and high
pressure stage separator drums, absorber and stripper towers, and light ends debutanizer and
depropanizer towers. HCN destroys the protective iron sulfide film normally present on carbon
steel and converts it into soluble ferrocyanide complexes. The exposed steel corrodes rapidly and

Page 2-49
CALTEX REFINERY MATERIALS MANUAL October 1999

the resulting atomic hydrogen can readily penetrate the metal and can cause blistering and/or
cracking.

The effect of HCN on corrosion and blistering can be minimized by reducing its concentration
through water washing. Conversion to harmless thiocyanates by the injection of dilute solutions
of ammonium polysuifide has also been found effective. Filming amine type corrosion inhibitors
have also been used with some success.

5.3.8 Amine
Experience has shown that carbon steel can stress corrosion crack in both lean and rich amine
solutions. Because of the uncertainty of cracking susceptibility in the various amines, Caltex as
well as other refiners have been stress relieving all new equipment. The cracking is usually
intergranular and typically shows up in heat affected zones including arc strikes. Most refiners
now call for stress relief in MEA, DEA and ADIP solutions, although some do not relieve stress
below certain temperatures in specific amines. MEA and ADIP solutions generally seem to crack
unstress relieved carbon steels more than DEA, DGA or MDEA solutions. A NACE survey of amine
cracking problems and API Publication 945 cover the subject quite well. While proper stress relief
minimizes the cracking problems, torch stress relief of small bore connections has generally not
been successful. Water washing unstress relieved equipment and piping before steaming out is
generally considered good practice to minimize cracking. Refer to Section 3 10.0 for a
complete description.

Page 2-50
CALTEX REFINERY MATERIALS MANUAL October 1999

TABLE 2-5 OPERATING CONDITIONS FOR C-O.5 MO STEELS


(That Experienced Hydrogen Attack Below The 1977 Edition 0.5 Mo Steel Curve)

Temperature Hydrogen Partial Service Hours Degrees Below


Pressure (Approx) 0.5Mo Curve
(Absolute) (Approx)
o o o o
Point F C psi Mpa F C

36 A1 790 421 350 2.41 80,000 20 11


1
37 B 800 427 285 1.97 57,000 30 17
38 C1 640 338 270 1.86 83,000 180 100
39 D1 700 371 300 2.07 96,000 125 69
41 F1 760 404 375 2.59 85,000 40 22
42 G1 750 399 350 2.41 150,000 60 33
1
43 H 625 329 350 2.41 150,000 185 103
44 I1 7302 3882 313 2.16 116,000 90 50
45 J3 6202/6404 327/338 457 3.15 70,000 147/167 82/93
46 K1 6262/6804 330/360 350 2.41 131,000 130/184 72/102
3
47 L 6842 362 738 5.09 61,000 54 30
5
48 M 550/570 288/299 1060/1100 7.31/7.59 79,000 105/125 58/69
* 655/670 346/354 --- --- 17,500 5/20 3/11
49 S3 750/770 399/410 390 2.69 67,000 30/50 17/28
* 650 343 --- --- 163,000 150 83
51 U3 690 366 397 2.74 --- 100 56
53 W5 545 285 2190 15.1 140,000 45 25
54AA1 7252/7604 3852/4044 3002/3804 2.072/2.62 105,000 40/100 22/56
4

55BB1 8002/8504 4272/4544 1752/1904 1.212/1.31 124,000 80/30 44/17


4

56CC1 8102/8254 4322/4414 2752/3004 1.902/2.07 223,000 15/0 8/0


4

57DD1 8504 4544 2254 1554 158,000 10 6


1 2 4 2 4
58EE 810 /855 432 /457 170 1.17 138,000 70/25 39/14
59FF3 5502/6004 2882/3164 2000 13.79 210,000 50/0 28/0
3 2 4 2 4
60GG 550 /600 288 /316 2000 13.79 210,000 50/0 28/0
61HH3 5302/6004 2772/3164 2200 15.17 210,000 60/0 33/0
62II3 6702/7004 3542/3714 190 1.31 192,000 180/150 100/83
3 2 4
63JJ 600 /750 3162/3994 500 3.45 235,000 180/30 100/17
64KK3 6002/7704 3162/4104 525 3.62 283,000 170/0 94/0
65LL3 775 413 550 3.79 --- 0 0

1
Catalytic reformer service
2
Average
3
Hydrodesulfurizer service
4
Maximum
5
Ammonia plant

Note: Numbers and letters in the first column (labeled Point) refer to the references and comments for Figure 2-8.

Page 2-51
CALTEX REFINERY MATERIALS MANUAL October 1999

Page 2-52
CALTEX REFINERY MATERIALS MANUAL October 1999

5.4 HIGH TEMPERATURE HYDROGEN DAMAGE

There are two types of hydrogen damage in refinery type environments. One type is the result of
hydrogen combining at elevated temperatures and pressures with carbon from carbides in steels
to form methane (CH4) which cannot diffuse through the steel lattice. The methane pressure
builds up forming cavities, microfissures, fissures and finally cracking that affects the load carrying
capacity of the pressure containing part. The other type of hydrogen damage is due to absorbing
hydrogen during welding or as a result of an aqueous corrosion process and was previously
discussed in the wet sulfide section above. (refer Section 2 5.3.6)

In high temperature hydrogen attack, the process environment hydrogen reacts with carbon
from carbides in the metal microstructure to form methane. This reaction can cause surface
decarburization . If the diffusion of carbon to the surface is limiting, this reaction can result in
internal decarburization, methane formation and cracking. Chromium and/or molybdenum
additions to carbon steel increase carbide stability and resistance to hydrogen attack. For a
specific material, hydrogen attack was shown to be temperature/hydrogen partial pressure/time
dependent. The data on hydrogen attack was plotted on temperature-pressure-materials specific
curves by George Nelson of Shell Oil in the 1940's. The Nelson curves plotted hydrogen attack
experience by industry and have been widely accepted as design criteria in hydrogen
environments. The American Petroleum Institute took over maintaining the curves in the early
1970's. The current fifth Edition of API 941 "Steels for Hydrogen Service at Elevated Temperatures
and Pressures in Petroleum Refineries and Petrochemical Plants" was published in January 1997
(See Figures 2-6 and 2-7). The document is currently on a five-year revision schedule. For
background on the development of API 941, see the 1993 NACE paper "50 years of Hydrogen
Attack - An Overview." API 941 shows a temperature versus hydrogen partial pressure safe
operation envelope for carbon and low alloy steels. It is generally assumed that austenitic
stainless steels are not hydrogen attacked at conditions seen in refinery units. Normal practice is
to use a 19-28oC (35-50oF) approach safety factor for each material. The hydrogen attack curves
form the basis for material selection for hydrogen processing units. Materials may need to be
upgraded, clad or overlaid for corrosion protection, but the upgraded material or backing steel
must be resistant to hydrogen attack at the design, startup, shutdown or emergency operating
conditions.

As a result of several failure incidents with carbon -0.5Mo material, the carbon -0.5Mo line has
been removed from the main 941 curves. It is generally recommended that carbon -0.5Mo not
be used on new construction for hydrogen processing equipment above the carbon steel curve.
In fact, "GPS-A9", along with specifications from several refineries advise not to use carbon -0.5
alloys in hydrogen services. The known carbon -0.5Mo incidents are shown in Figure 2-8 and
listed in tabular form in Table 2-5. Inspection of existing carbon -0.5Mo equipment has proved
difficult. Some failures have been in the weld heat affected zones but others have been away
from known welds. The attack and resulting damage have been very localized in some cases.
Inspection technology is constantly improving. Currently, ultrasonic techniques using a
combination of velocity ratio and backscatter have been most successful in finding
fissuring/serious cracking. The variable attack problems with carbon-0.5Mo steels seem to
involve the type of carbide formed during mill or fabrication exposure.

Page 2-53
CALTEX REFINERY MATERIALS MANUAL October 1999

5.4.1 Outgassing Hydroprocessing Equipment


There are two concerns with brittle failure potential with heavy wall hydroprocessing equipment.

The first is the temper embrittlement problem with 2Cr-1Mo reactors and other heavy wall
equipment. With a material susceptible to temper embrittlement, the impact strength can shift
to a higher temperature by as much as 149oC (300F). This raises concern over potential defect
propagation during startups and cool downs. The other concern is the potential hydrogen
embrittlement problem associated with equipment cool down rates. Heavy wall equipment may
be cooled down rapidly enough to seriously affect the notch toughness due to supersaturation
with hydrogen.

While onstream, low alloy equipment absorbs relatively large amounts of hydrogen; the exact
amount is related directly to the pressure shell temperature and hydrogen partial pressure. The
hydrogen concentration decreases linearly through the wall and is zero at the outside diameter.
At typical hydrocracker pressure and temperatures [2000 psi and 441oC (825F)], the absorbed
hydrogen is soluble in the steel and appears to be innocuous. However, on cooling, the
hydrogen solubility of the steel is reduced drastically (it is essentially zero at ambient
temperature) and, unless the cooling rate is sufficiently slow to permit diffusion of the hydrogen
out of the shell through both the inside and outside surfaces (called outgassing), a
supersaturation condition will result. While outgassing to the surfaces continues below 149oC
(300F), much of the hydrogen tends to diffuse to areas of lattice strain such as defects and crack
tips. As a result, slow subcritical crack growth may occur.

The maximum amount of supersaturated hydrogen which can be tolerated by a reactor wall at
temperatures below 149oC (300F) without resulting in crack growth is defined as the safe level
of hydrogen, CS. The safe hydrogen concentration is known to decrease with increasing strength
level, tensile stress across the crack and flaw size. If the end of run hydrogen concentration of a
reactor head or shell or other equipment pressure boundary, Cr, exceeds the safe hydrogen level,
CS, special shutdown procedures should be designed to ensure sufficient outgassing takes place
so that the hydrogen concentration decreases to Cs before the metal temperature is below about
149oC (300F). To cater to potential temper embrittlement problems, cool down rates should not
exceed 10oC (50F)/hr with depressurization to less than 1/3 design at temperatures below about
149oC (300F). Going through an outgassing calculation may further restrict the cool down rate.
The suggested cool down rate applies to old non-pedigreed material that would have a high
potential for temper embrittlement. Modern (post 1975) reactor steels may not require such
care in shutdown. Experienced materials engineering help may be required to assist in
evaluating each individual piece of equipments situation and how it relates to the minimum
pressurization temperature. In general, special outgassing is usually not required with wall
thicknesses less than 5 inches (150 mm).

There are different calculation ground rules for clad or weld overlaid and unclad equipment. The
Chevron-developed outgassing calculation procedure can be obtained through CSC. Other
companies have developed similar procedures which use a fracture mechanics and hydrogen
diffusion approach to estimate the maximum amount of hydrogen that can be tolerated below
149oC (300F) without crack growth.

Page 2-54
Hydrogen partial pressure, megapascals absolute

Temperature, degrees

Temperature, degrees
Hydrogen partial pressure, pounds per square inch absolute

1. The limits described by these curves are based on service experience


originally collected by G.A. Nelson and on additional information
gathered by or made available to API. LEGEND
Surface Decarburization - - - - - - - - - - - - - - - - - -
2. Austenitic stainless steels are generally not decarburized in hydrogen at Internal Decarburization ____________________
any temperature or hydrogen pressure. OPERATING LIMITS FOR STEEL IN Carbon 1.0Cr 2.25Cr 6.0Cr
Steel 0.5Mo 1.00Mo 0.5Mo

HYDROGEN SERVICE TO AVOID Satisfactory


3. The limits described by these curves are bast on experience with cast
DECARBURIZATION AND FISSURING Internal decarburization and fissuring
steel as well as annealed and normalized steels at stress levels Surface decarburization
described by Section VIII, Division I, of the same ASME Code. See comments
Figure 2-6
Temperature, degrees Fahrenheit Hydrogen partial pressure, megapascals absolute

Temperature, degrees Celcius


Hydrogen partial pressure, pounds per square inch absolute

I. T.C. Evans, Hydrogen Attack on Carbon Steels, Mechanical Engineering, 1948, Vol. 70, pp. 414-416.
A. Amoco Oil Company, private communication to API Subcommittee on Corrosion, 1960.
J. Air Products, Inc., Private communication to API Subcommittee on Corrosion, March 1960.
B. A.R. Ciuffreda and W.D. Rowland, Hydrogen Attack of Steel in Reformer Service, Proceedings, 1957,
K. API Refinery Corrosion Committee Survey, 1957.
Vol. 37, American Petroleum Institute, New York, pp. 116-128.
L. I. Class, Present State of Knowledge in Respect to the Properties of Steels Resistant to Hydrogen
C. C.A. Zapffe, Boiler Embrittlement, Transactions of the ASME, 1944, Vol. 66, pp. 81-126.
Under Pressure, Stahl and Eisen, August 18, 1960, Vol. 80, pp. 1117-1135.
D. R.E. Allen, R.J. Jansen, P.C. Rosenthal, and F.H. Vitovec, The Rate of Irreversible Hydrogen Attack on
Steel at Elevated Temperature Proceedings, 1961, Vol. 41, American Petroleum Institute, New York, pp. 74-
84.
E. L.C. Weiner, Kinetics and Mechanism of Hydrogen Attack of Steel, Corrosion, 1961, Vol. 17, pp. 109-
115.
F. J.J. Hur, J.K. Deichler and D.R. Worrell, Building a Catalytic Reformer?, Oil and Gas Journal, October 29,
1956, Vol. 54, No. 78, pp. 103-107. TIME FOR INCIPIENT ATTACK OF CARBON
G. F.K. Naumann, Influence of Alloy Additions to Steel upon Reisistance to Hydrogen Under High Pressure, STEEL IN HYDROGEN SERVICE
Technische Mitteilungen Krupp, 1938, Vol. 1, No. 12, pp. 223-234.
H. M. Hasegawa and S. Fujinaga, Attack of Hydrogen o OIl Refinery Steels, Tetsu To Hagane, 1960, Vol. 46,
No. 10, pp. 1349-1352. Figure 2-7
Notes: EXPERIENCE WITH C-0.5Mo AND Mn-0.5Mo
1. Reference and comments are shown on Table A-1. STEEL in HIGH TEMPERATURE
2. Curves for carbon steel,1.0Cr-0.5Mo steel, and 1.25Cr-0.5Mo HYDROGEN SERVICE
steel are included for reference.
3. The symbol is retained as a reference against previous Figure 2-8
revisions of this publication.
4. Reference numbers are the same as in previous editions of this
publication.
5. The 0.5Mo steel curve is the same as the one shown in the Reprinted courtesy of the American Petroleum Institute from API Publication 941, Steels for Hydrogen
fourth edition of this publication. (1990) Service at Elevated Temperature and Pressures in Petroleum Refineries and Petrochemical Plants.
Fourth Edition, April 1990.
Hydrogen partial pressure, megapascals absolute

Temperature, degrees Celcius


Temperature, degrees Fahrenheit

Hydrogen partial pressure, pounds per square inch absolute

Notes:
1. This figure was adapted from Figure 3, Fourth Edition (1990) of
Steels for Hydrogen Service at Elevated Temperatures and TIME FOR INCIPIENT ATTACK OF 0.5Mo
Pressures in Petroleum Refineries and Petrochemical Plants. STEELS IN HIGH TEMPERATURE
HYDROGEN SERVICE
2. Numbered and lettered references for points in this figure refer to
data listed in Table A-1 and Figure A-1 of the above publication.
Figure 2-9
CALTEX REFINERY MATERIALS MANUAL October 1999

5.5 METALLURGICAL AND


AND ENVIRONMENTAL FAILURES

The metallurgical properties, such as strength, ductility/strain capability, toughness and corrosion
resistance can change in service due to microstructural changes as a result of thermal aging at
elevated temperatures. For example - carbon steels can strain age embrittle or graphitize; ferritic
and austenitic stainless steels can form sigma phase; ferritic stainless steels can suffer 885oF
embrittlement; 1-1/4 Cr -1/2 Mo can creep embrittle; and 2-1/4Cr -1Mo steel can temper
embrittle. Properties can also change as a result of hydrogen charging.

Elevated temperature service exposure can cause metallurgical changes to occur in many of the
metals and alloys used in refinery service. Many of these microstructural changes are time and
temperature dependent. Some of the changes are reversible but may recur faster during more
service exposure. The possibility that these microstructural changes may result in decreased
strength or notch toughness needs to be considered when selecting materials.

These changes in properties are difficult to identify, since detectable physical damage may not
have occurred. Sometimes inferences can be made from examining samples or surface replicas.
Steel composition and microstructure, operating temperature, process environment and
accumulated strain/stress are the most important factors that determine susceptibility to
metallurgical change. Often an equilibrium state of change is reached and further changes will
not occur. Hydrogen charging, even without material damage, will in many cases lower the
ductility and possibly even the toughness of the material. Hydrogen charging is a reversible
reaction and if it does not cause damage, has no permanent effect.

Once the metallurgical properties are changed in service, they usually are not recovered. Heat
treatment can be effective, although is often only temporary. To prevent further degradation,
operating conditions can be adjusted to a lower severity. If the degradation in properties is
known, the operating precautions such as start-up and shut-down procedures can be altered to
prevent failure or further damage from occurring despite the degraded physical properties.

5.5.1 Microstructural Changes and Embrittlement


In addition to low creep rates and high stress rupture strengths, metals and alloys used in high
temperature refinery services must have good structural stability. The most serious structural
changes that may occur as a result of exposure to high temperatures include the following:

Grain Growth
Graphitization
Hardening
Sensitization
Sigma Phase Formation
885oF Embrittlement
Temper Embrittlement
Liquid Metal Embrittlement
Carburization
Metal Dusting
Decarburization
Selective Leaching

Page 2-59
CALTEX REFINERY MATERIALS MANUAL October 1999

(i) Grain Growth


Grain growth occurs when steels are heated above a certain temperature, beginning about
593oC (1100oF) for carbon steel and most pronounced at 732oC (1350oF). The amount of
growth depends on the maximum temperature reached and the length of time at temperature.
Austenitic stainless steels and high nickel-chromium alloys do not become subject to grain
growth until heated to above 899oC (1650oF).

Grain growth lowers the tensile strength but increases both creep strength and rupture strength.
In practice, grain growth usually has not been a significant factor in refinery failures. It is,
however, very useful for pinpointing furnace operational problems that have led to localized
overheating failures of furnace tubes. Metallographic examination of the microstructure of failed
components can reveal, through grain growth, the temperature to which the component was
exposed. This technique is also applied to refinery fire damage evaluations.

(ii) Graphitization
Graphitization problems can occur with certain chemistry (usually aluminum killed) carbon and
carbon -0.5Mo steels operating above about 474oC (800oF). Graphitization occurs when iron
carbides in the normally pearlite/ferrite microstructure separate into iron and graphite. The
problems tend to occur on certain temperature isotherms in weld heat affected zones. If the
graphite particles line up through thickness, they can form a plane of weakness affecting the
ductility of the equipment, significantly reducing its load carrying ability. Typical graphitization
problems have occurred in power station steam piping and in Fluid Catalytic Cracking Unit
reactors, regenerators and catalyst piping. Chromium increases the carbide stability and therefore
minimizes graphitization. 1.25Cr-0.5Mo alloys are very resistant to graphitization problems.
While considerable graphite particles can be seen in a particular microstructure, it usually takes
almost a solid line of graphite to seriously affect mechanical properties. A Canadian refiner had a
massive failure of a catalytic reformer heater to reactor line due to severe graphitization. The C-
1/2Mo line failed for several feet in the heat affected zone of a long seam weld. Most of the
failures ascribed to graphitization have usually been due to a combination of causes such as
graphitization and creep rupture. Several FCC reactors had severe cracking problems in the late
1940s and early 1950s. The refining industry ran a detailed survey of graphitization problems
and concluded that although considerable graphitization was occurring in many of the high
temperature carbon steel vessels and large lines on the FCCs, very little in-line or eyebrow
graphitization was occurring. Many refiners still take samples or do surface metallography to
check for significant graphitization. Some samples may show that the pearlite is completely gone
with a microstructure of ferrite, some large carbides and considerable graphite. The typical room
temperature tensile strength of a microstructure like this will be in the 40 ksi range while it was in
the 60 ksi range when it went into service. Significant spheroidization is time dependent and
depending on the specific chemistry of the steel can occur in roughly ten hours at 649oC
(1200oF) and will take up to 100,000 hours at 474oC (885oF). It is not unusual to see significant
graphite in the matrix of a sample of carbon steel that has operated for long times at
temperatures above about 474oC (885oF). It is fairly unusual to see significant property
degradation other than the lowering of the tensile strength caused by the breaking up of the
pearlite colonies. Since graphitization does not occur uniformly, it is difficult to get representative
samples. That is, one end of a plate formed into a vessel or line may have significant graphite in a
heat affected zone while the other end may show little or no graphite in the microstructure.

(iii) Hardening.
Hardening of steels is the result of martensite formation after heating carbon and low alloy steels
to above the upper critical temperature followed by rapid cooling. A brittle martensitic/carbide

Page 2-60
CALTEX REFINERY MATERIALS MANUAL October 1999

structure is formed which is undesirable for refinery piping, furnace tubes, or pressure vessels.
Hardening can occur in the course of welding fabrication or when steels are exposed to severe
overheating, such as in a fire. Hot bending can also be a source of hardening.

Welding of carbon steels having less than 0.25% carbon generally presents few general
hardening problems because the usual cooling rates are not fast enough to permit martensite
formation. However, carbon steel with more than 0.35% carbon, low alloy steels and the
martensitic straight chromium stainless steels will harden simply by air cooling after welding.
Similarly, during fire exposure, these hardenable materials can become extremely hard and brittle
to the extent they are not serviceable.

To prevent cracking of hardened metal after welding, preheat and postweld heat treatments are
used. In the case of fire damaged material, hardness surveys using portable testers can be used to
identify equipment and piping hardened by overheating and quenching.

Conversely, softening can also be a problem with refinery equipment. Some pressure vessels are
made of low alloy steels that are quenched and tempered or normalized and tempered to
optimize design strength. Subsequent welding, heating for bending, or exposure to fire can
degrade strength properties such that replacement or re-heat treatment will be required.
Commonly used bolting ASTM A-193 Grade B7 is an example of an intentionally hardened
component. Hydroprocessing reactors made of 2 1/4 Cr - 1 Mo material are another example.

(iv) Sensitization
Sensitization of austenitic stainless steels has been mentioned earlier in the section on
intergranular cracking and polythionic acid SCC. (Refer Section 2 - 5.3.4) Sensitization occurs
when austenitic stainless steels are heated in the temperature range of 371 to 816oC (700 to
1500oF). For optimum corrosion resistance, these steels are normally supplied in the solution heat
treated condition with carbides fully dissolved in the austenitic matrix. During elevated
temperature exposure, either in service or at the time of welding, chromium carbides precipitate
at the grain boundaries. As a result, the grain boundaries are depleted of chromium and become
more susceptible to corrosion. Sensitizing does not appreciably affect the mechanical and heat
resisting properties of stainless steels.

Sensitization can be avoided by using low carbon and stabilized grades of austenitic stainless
steels when welding and elevated service temperatures are involved. Sensitizing can be reversed
by solution heat treatment after welding, but this is usually impractical because the component
needs to be heated to above 1093oC (2000oF) and water quenched.

(v) Sigma Phase


Sigma is a Cr-Fe intermetallic compound formed in higher chromium alloys between the
temperatures of 593-927oC (1100-1700oF). Formation is time dependent. Alloys with
considerable sigma will be very brittle at room temperature and, depending on the amount of
sigma, may be brittle at high temperatures. Sigma can be a problem with alloys containing more
than 17% chromium. Serious ductility loss may occur in austenitic stainless steels due to sigma
and carbide formation. Austenitic stainless steel welds and corrosion resistant weld overlays can
suffer extreme in-service ductility loss if proper deposit chemistries are not achieved during
fabrication.

FCC regenerator cyclones are usually 304H. Carbide precipitation and sigma formed during
service can decrease the toughness and ductility of the 304H low enough that weld repairs have

Page 2-61
CALTEX REFINERY MATERIALS MANUAL October 1999

to be made with more care. Maximum sigma in 304H is somewhere around 10%. This level of
sigma and precipitated carbides will cut the ductility from 60-70% as fabricated to around 10%
or less. This is still enough ductility to weld repair successfully but may require improved
technique.

Cast austenitic stainless steels can form considerable sigma. Typical furnace hanger alloys, HH
(25Cr-12Ni) and HK (25Cr-20Ni) both form considerable sigma along with carbides. The general
ductility of these alloys is very close to zero after service exposure in the 593-927oC (1100o-
1700oF) temperature range. Cast austenitic stainless steel slide valve bodies were used for several
years by the slide valve vendors. Even when the ferrite was limited to 10% by specification,
foundries had considerable difficulty meeting the limit in the multi-thickness castings. The thinner
sections typically met the specification while the thicker sections might have up to 35% ferrite.
After service exposure to flue gas or regenerated catalyst, the slide valve became very brittle and
a few actually cracked and leaked onstream. While sigma is usually considered to affect only the
room temperature ductility, the very high ferrite contents in the thicker slide valve sections
definitely cause a low ductility situation at operating temperatures of 593 -732oC (1100 -1350oF).

SS 347 weld overlay is used for sulfide corrosion protection of the low Cr-Mo equipment required
by hydrogen attack considerations in hydroprocessing units. If the overlay is deposited without
the proper base metal dilution, a Cr rich overlay can be deposited that will form sigma during
intermediate or final equipment heat treatment. The net result is a very brittle overlay that will
easily crack and will be next to impossible to weld repair. Typical dilution with submerged arc
overlay is about 30%. With hot wire TIG, the typical dilution is about 10%. Any SS 347 overlay
should have inspection provisions to make sure it is laid down essentially crack-free and has had
the proper amount of dilution. Because of the dilution required, submerged arc overlay can be a
real sigma problem if not properly done. The Cr rich alloy wire or strip will be very sigma prone if
not properly diluted.

(vi) 885oF Embrittlement


885oF embrittlement is the general loss of ductility that occurs in the ferrite phase of stainless
steels. It is caused by a microstructural change that occurs between roughly 343 - 566oC (650o-
1050oF). It is time and chemical composition dependent. The problems with Types 405 and 409
may result in poor ductility which shows up as an inability to weld without cracking and
breakage problems when trying to straighten upset trays. A standard de-embrittling heat
treatment at about 621oC (1150oF) for two hours usually restores sufficient ductility to weld or
straighten. Typical problems have occurred with trays and tray support beams in atmospheric,
vacuum and FCC main fractionator columns as well as the corrosion resistant cladding such as SS
405. After several years of service, the lining can become very crack sensitive making a retraying

job very tedious or impossible. It is possible to heat treat the column area with the embrittled
cladding, restoring its ductility so it can be successfully welded.

Austenitic stainless steel castings can have as much as 30-35% ferrite in them as cast. Even with
specifications restricting ferrite to 10% some of the thicker sections of the casting may have
considerably more ferrite. The ferrite phase can embrittle due to 427oC (885oF) problems up to
roughly 566oC (1050oF) and due to sigma formation up to 927oC (1700oF). The 885oF
embrittlement can be reversed by a heat treatment of roughly 566oC (1150oF) for about two
hours. However, it has been shown that the embrittlement can return with considerably less
exposure time. Another place that 885oF embrittlement has been a problem has been when 12Cr
has been used for small bore nozzles in clad areas. That is why 'GPS-A9' does not allow 12Cr for

Page 2-62
CALTEX REFINERY MATERIALS MANUAL October 1999

pressure parts at temperatures above 343oC (650oF). A Caltex refinery had a brittle failure of a SS
405 heater transfer line when it was hit during a shutdown.

(vii) Temper Embrittlement


Temper embrittlement is a microstructural segregation problem with low alloy steels operating
between about 343-566oC (650o-1050oF). It is time/temperature/composition dependent. The
temper embrittlement in the low Cr-Mo alloys became a very serious problem in the refineries
during the late 1960's when many heavy wall hydrocracking/hydroprocessing reactors were
installed. API sponsored research has since shown that the serious loss of notch toughness
suffered by the 2.25Cr-1Mo steels was primarily due to grain boundary segregation of P, As, Sb
and Sn. The segregation was time, temperature and concentration dependent. Newer steels with
lower tramp element levels perform in a satisfactory manner. The older, higher tramp element
reactors are handled with special startup and shutdown procedures which minimize stress on the
reactors until notch toughness is high enough [usually at temperatures greater than 149oC
(300oF) on startup]. Reactors, heavy wall vessels and equipment built up to 1975 are generally
considered old steel technology and require special startup conditions before full pressurizing.
Steels produced after 1975 should have lower tramp elements and not be susceptible to temper
embrittlement problems. However, it is advisable to investigate the pedigree of specific
equipment when setting up or reviewing equipment minimum pressurization temperatures.

One measure of the concentration of tramp elements is the J factor. J factors above about 125 are
considered high with current steel making practice. One of the reactors in a Caltex refinery was
purchased in the 1960s and has a J Factor of about 350. With the time in service, this reactor
would be expected to have a shift in the 40 ft lb notch toughness of up to 66-93oC (150o-200oF).
Special startup procedures call for warming the reactor system up to about 149oC (300oF) before
fully pressurizing the reactor. This procedure effectively caters to the expected decreased notch
toughness. This reactor has been a concern because of its previous cracking problems internally
at the catalyst bed beam supports. The concern is that a crack could reach critical size and
propagate during a low temperature full pressurization causing a brittle failure. Crack
propagation at operating temperature is very unlikely.

At the time of writing this manual (1998), J factors of 80 and below are easily achievable. "GPS-
A9" currently specifies 100 max. When specifying these 2 1/4 CR-Mo steels for heavy wall
reactors, it is recommended to require step cooling tests to confirm resistance to temper
embrittlement.

(viii) Creep Embrittlement


The creep embrittlement cracking problems with 1.25Cr-0.5Mo reactors, piping and exchangers
have been studied. Tentative conclusions are that some heats of steel have very low heat affected
zone rupture ductility because of tramp element grain boundary segregation. This appears to be
a somewhat different mechanism than the 2.25Cr-1Mo temper embrittlement problem.

Research work has shown that rather than a time dependent embrittlement problem, the very
poor rupture ductility seems to be present in the steel due to the chemistry and fabrication
procedures. Limiting the P, As, Sn and Sb similar to the pedigreed 2-1/4Cr-1Mo, seems to
minimize the heat affected zone rupture ductility problems. The basic problem is that the coarse
grain heat affected zone tends to crack instead of creeping as would be expected at temperatures
above about 441oC (825oF). The usual problem is cracking in the heat affected zones around the
higher stressed welds. It usually takes wet fluorescent magnetic particle inspection to find the

Page 2-63
CALTEX REFINERY MATERIALS MANUAL October 1999

tighter cracks although, many cracks have been found visually. Typical problem areas are as
follows:

Reactors Inlet and outlet nozzles, skirt attachments, other attachments and main seam
welds
Exchangers Nozzles
Piping Longitudinal seams and circumferential welds near geometry and thermal
stress risers

The solution to the problems is to specify the pedigreed 1-1/4Cr-1/2Mo material with annealed
properties (not to Class 2) and postweld heat treat at about 704oC (1300oF). Class 2 materials
would probably not meet tensile requirements with the high heat treating temperature. Many of
the incidents reported seem to have been triggered by a fabrication problem like reheat cracking
or hydrogen cracking that was not detected during fabrication inspection. Another solution has
been to use pedigreed 2-1/4Cr-1Mo which has much better and more consistent rupture
ductility.

(ix) Liquid Metal Attack


The most common liquid metal attack problem in refinery and/or petrochemical service is
exposure of austenitic stainless steel to zinc during welding or under fire conditions. During
welding, the zinc forms a brittle intergranular compound in the stainless steel that basically falls
apart when stressed. Galvanizing is the usual source of zinc because it is used to protect
structurals from coastal atmospheric corrosion. Welding a galvanized piece to an austenitic
stainless steel structural (for example, a platform brace for a cold column) will result in an almost
instantaneous embrittlement. Most refiners and shop specifications do not allow zinc anywhere
around austenitic stainless steel during welding, but it does happen. During a fire situation, the
zinc can melt off the structurals and fall on hot exposed austenitic stainless steel causing instant
embrittlement and a very difficult inspection problem.

Mercury is another embrittling element that might affect some refinery alloys. Many crudes are
appearing that contain trace amounts of mercury in various forms. When these crudes or
condensates are run, the mercury tends to collect in the atmospheric overhead accumulator
drum. This presents a major health problem for the inspectors when the equipment is opened up
during shutdowns and is a very potent stress embrittling agent for copper zinc and nickel-copper
alloys. The mercury is used in an ASTM test for proper heat treatment/stress relief of these alloys.

Mercury is also a very aggressive corrodent of aluminum. Since mercury tends to follow the light
ends, it tends to show up in gas condensate. LNG plants processing the condensate typically use
very large aluminum heat exchanger in the process. Any mercury severely attacks and basically
wastes the aluminum. This has caused gas and petrochemical plants to install mercury
(sulfur/activated carbon) traps in the feed sections of units that contain cold sections with
aluminum equipment. The large LNG aluminum exchangers have suffered severe amalgam
corrosion. Operators in Algeria and Indonesia have had serious problems. Arun condensate from
Indonesia is known to have mercury problems. Problems have occurred in the exchanger tubing
and cold box piping. The corrosion tends to occur at -50oC (-58oF) and above. This probably
corresponds to the solidification of mercury at -39oC (-38oF).

Page 2-64
CALTEX REFINERY MATERIALS MANUAL October 1999

(x) Carburization
At high temperatures above 540oC (1000oF), metals can absorb carbon from the surrounding
atmosphere to form metal carbides, a process called 'carburization'. Bulging and flaking of
carburized layers can lead to gradual metal loss.

Carburization is caused by carbon diffusion into the steel at elevated temperatures. Coke deposits
on furnace tubes are an example of a source of carbon for carburization. Carburization depends
on the rate of diffusion of elemental carbon into the metal and increases rapidly with increasing
temperature. This increase in carbon content results in an increase in the hardening tendency of
ferritic steels and, when carburized steel is cooled, a brittle structure can result. The presence of
such a hard, brittle structure may result in spalling or cracking.

All steels are susceptible to carburization under the proper conditions. However, susceptibility
decreases with increasing chromium in the steel. The austenitic stainless steels appear to be more
resistant to carburization than the straight chromium steels, partly because of their higher
chromium content and partly because of the nickel content.

CO and methane/hydrogen are environments that can cause carburizing. The CO can carburize
the low Cr-Mo alloys and will carburize austenitic stainless steels in FCC regenerator service. The
methane/hydrogen in catalytic reformer furnaces will carburize the 9Cr-1Mo tubes under some
operating conditions.

(xi) Metal Dusting


Metal dusting is catastrophic, highly localized carburization of steels exposed to mixtures of
hydrogen, methane, carbon monoxide, carbon dioxide and other light hydrocarbons in the
temperature range of 482 to 816oC (900 to 1500oF). Attack is in the form of small pits filled with
carbon or general, uniform wastage that yields a crumbly residue composed of graphite, metal,
carbides, and oxides. Trace amounts of sulfur seem to inhibit metal dusting. Failures resulting
from metal dusting have been known to occur in dehydrogenation units, fired heaters, coker
heaters, cracking units and gas turbines.

Metal dusting reactions result from a complex series of steps in which a reducing gas, rather than
an oxidizing agent, is usually the reactant. Copper, for example, has poor oxidation resistance
but is not affected by metal dusting, while ordinary stainless steels are known for their oxidation
resistance but are susceptible to metal dusting. For the most part, the rate of attack of metal
dusting increases linearly with temperature.

(xii) Decarburization.
Decarburization is the loss of carbon from the surface of a ferrous alloy as a result of heating in a
medium that reacts with carbon. As far as refinery equipment is concerned, decarburization is
usually the result of high temperature oxidation or hydrogen attack.

Decarburization can be found only by microscopic examination. When carbon is removed from
the surface of a steel, the surface layer is converted to almost pure iron, which results in
considerably lower tensile strength, hardness and fatigue strength. The presence of a
decarburized layer is usually not serious unless creep and fatigue are a problem. However, its
occurrence in operating equipment is evidence that the steel has been overheated and suggests
other effects may be present. Some decarburization will be found on any equipment that was hot
formed or postweld heat treated during fabrication.

Page 2-65
CALTEX REFINERY MATERIALS MANUAL October 1999

(xiii) Selective Leaching.


Selective leaching is the preferential loss of one alloy phase in a multiphase alloy. In the brasses,
such as admiralty tubes used in refinery cooling water systems, it is called 'dezincification'. In
copper-nickel alloys it is called 'dealloying', and in cast iron, the selective loss of iron is termed
(incorrectly) 'graphitization'.

There are two common types of dezincification seen on brass: uniform layer-type and localized
plug-type. In each case, the brass first dissolves by corrosion and copper, being more noble than
zinc, and subsequently plates out. As a result, the dezincified areas contain as much as 95%
copper but have become brittle and possess essentially no strength. In the case of plug-type
dezincification, exchanger tubes are suddenly discovered perforated when dezincified small areas,
the 'plugs', are blown out by water pressure or during bundle hydroblast cleaning.

Selective leaching is favored by stagnant flow conditions that allow deposits to settle out on
tubing surfaces. This can produce leaching as a result of crevice corrosion mechanisms. Oxygen
need not be present for selective leaching, but deaeration reduces the likelihood of attack in most
waters. In brass, the addition of small amounts of phosphorus, arsenic or antimony as an inhibitor
greatly reduces dezincification except in highly aggressive waters.

5.6 MECHANICAL FAILURES


FAILURES

In the absence of internal corrosion, most equipment will eventually deteriorate due to external
conditions. This deterioration normally occurs very slowly, unless incorrect or defective materials
were initially installed or process conditions exist that exceed the material mechanical properties.
Major pieces of equipment should be inspected and tested before being placed in operation.
However, mixing of materials can often occur with smaller items such as valves and fittings.
Mechanical damage, overloading of structural members, and over-tightening of bolts represent a
large portion of mechanical failures. Accidental overpressuring or brittle fracture of equipment
occasionally occur in fixed equipment. In contrast, fatigue failures are fairly common with
machinery having highly stressed, rotating or reciprocating parts.

Operational changes in process temperature or pressure, upsets, overfiring of furnaces to increase


throughput, control instrument failures, or exposure to fire often occur and can result in a
mechanical failure. For example, furnace tubes start to sag or bulge, vessel walls become
distorted and develop cracks or blisters, and piping becomes embrittled. These types of failures
are often accelerated by cyclic changes, including periodic shutdowns.

5.6.1 Incorrect or Defective Materials


Even in the absence of corrosion of metallurgical degradation, equipment will still eventually
deteriorate due to other factors. The major cause of this type of problem is due to incorrect
material is mix-ups by suppliers (positive material identification is needed). For example, when
one refining unit was built prior to positive materials evaluation programs currently being
implemented in the industry, roughly 30% of piping and fittings failed to meet specifications in
one way or another. While this figure included minor deviations as well as faulty weld metal,
approximately 5% of the total had to be replaced.

Often, vendors may substitute what they consider to be an equivalent or better material than
specified. They do not realize that a stainless steel fitting is not necessarily an improvement over a
carbon steel fitting, especially as far as pitting corrosion or SCC is concerned.

Page 2-66
CALTEX REFINERY MATERIALS MANUAL October 1999

Substitution of castings for wrought or forged shapes may lead to problems. Casting defects,
such as shrinkage, sand holes or blowholes, can create unforeseen cracking and corrosion
problems. Shrinkage cracks are often found in the thinner sections where the cast metal cools
faster. Sharp corners and abrupt changes in cross sectional area are stress raisers, and shrinkage
cracks can occur at such points. Sand holes are caused by molding sand trapped within the
casting. Blowholes are caused by gas trapped within the casting during solidification. The sand
and gas create crevices or holes, within the metal, that may not be visible from the exterior of the
casting.

Discontinuities in wrought material are excellent crack initiators. The discontinuities may be
laminations and crevices which may cause hydrogen blistering in certain applications.

In order to expedite repairs during a shutdown, material substitutions may have to be made.
Often, the correct material simply cannot be obtained because of long lead times and
unreasonably high minimum quantity purchase requirements. Intentional upgrading can also
lead to problems. For example, substitution of titanium tubes for admiralty metal tubes may
resolve corrosion problems at the expense of vibration problems if care is not exercised. Because
of the lower wall thickness of titanium tubes, baffle spacing and tube hole clearances should be
checked to prevent titanium tube fatigue failures.

As a direct result of an incident, a major refiner Positive Material Identified (PMI) all alloy circuits
in their piping systems. They found about 1% errors in metallurgy on existing units, some could
have caused major failures. Usually, the problem was carbon steel in place of a Cr-Mo alloy
installed for sulfide corrosion or hydrogen attack resistance. The incident that triggered the work
was a carbon steel elbow in a 5Cr-Mo line on a Delayed Coker. These failures point toward the
need to do a 100% PMI inspection on all new construction alloy equipment (in place in the case
of piping) and follow up with a 100% alloy PMI on existing units that were not construction
PMId where improper materials create a significant risk.

5.6.2 Mechanical Fatigue


Fatigue is the failure of a component by cracking after the continued application of cyclic stress.
Below a definite stress limit, cyclic stressing of a metal does not affect the material and no
cracking occurs regardless of the number of cycles. This stress limit is called the endurance limit
or fatigue limit. At stresses higher than the endurance limit, a crack initiates and is propagated by
continued application of stress cycles. Eventually the component fails, usually from a single
crack. Little deformation of the metal takes place and the failure appears to be brittle.

Generally, the endurance limit of steels is roughly 50% of the tensile strength, while the
endurance limit for non-ferrous alloys ranges from 35-50%. Fatigue properties are related to
notch toughness, deep scratches, sharp corners and weld intersections which will lower fatigue
strength by locally concentrating stresses. Brittle steels are more likely to fail by fatigue than
ductile steels. High ductility permits relief of concentrated stresses through plastic flow. In
refineries, many rotating equipment failures have been attributed to fatigue, or as discussed
earlier, corrosion fatigue. Prime examples are reciprocating parts in pumps and compressors and
the shafts of rotating machinery.

A significant improvement in minimizing fatigue failures can be achieved through the elimination
of stress raisers. A radius should be used instead of sharp corners on rotating or cyclically stressed

Page 2-67
CALTEX REFINERY MATERIALS MANUAL October 1999

parts. Stampings and other sharp edged marks should be avoided, as well as cold straightening
of bent parts that will later be subjected to in-service cyclic stresses.

5.6.3 Corrosion Fatigue


Corrosion Fatigue is a form of fatigue in which a corrosion process, typically pitting corrosion,
adds to or promotes the mechanical fatigue process. Pure mechanical or dry fatigue is a failure
mechanism that results from cyclic stress applied to a structural component. Corrosion fatigue
results in a shorter life than would occur with either dry fatigue or in the corrosive environment
alone.

Whereas dry fatigue takes the form of a single stepped crack, corrosion fatigue usually takes the
form of several or many cracks emanating from the base of pits. It is thought to be a two-stage
process in which the first stage is the formation of corrosion pits and the second stage is the
development of cracks. Failures are associated with environments that favor pitting, probably
because pits act as stress raisers. Cracking by corrosion fatigue is transgranular without
branching. Final fracture is strictly mechanical.

Since the development of corrosion pits is the first step in corrosion fatigue, mitigation of
corrosion is the best approach to prevention. Elimination of the cyclic stresses causing fatigue is
the next best approach. Shot peening, a process involving the cold working of a metallic surface
with a high velocity stream of steel shot, introduces residual compressive stresses a few mils deep
in the surface. Although the surface finish produced by shot peening is rougher than that of
machining or grinding, the resulting compressive surface layer improves fatigue and corrosion
fatigue resistance. Stress relieving, corrosion inhibitors and protective coatings have also been
successfully used to combat corrosion fatigue.

5.6.4 Deaerator Cracking


NACE has issued an updated RP on the deaerator cracking problems. It is RP0590-96 titled
"Recommended Practice for Prevention, Detection and Correction of Deaerator Cracking."
Information made available to NACE through industry surveys, testing laboratories and insurance
companies indicate a cracking rate of 41%. Cracking that extended into the minimum wall
thickness included 269 out of 650 deaerators. A sample survey of 174 deaerators found 89 or
51% had cracking of some kind. The cracking was predominantly corrosion fatigue and was
generally best found with proper wet fluorescent magnetic particle inspection. Residual welding
stresses are a prime factor in promoting crack growth. Therefore, post weld stress relief is
beneficial. As in all cracking situations, the cost of replacement must be evaluated against
detailed inspection and repair costs.

5.6.5 Cavitation Damage


Cavitation damage is caused by the rapid formation and collapse of vapor bubbles in liquid at a
metal surface as a result of pressure variations. Calculations have shown that bubble collapse can
produce shock waves with impact pressures sufficiently high to produce plastic deformation in
most metals. In brittle metals, cracking and metal loss occurs as grains are torn out of the surface.
Corrosive conditions accelerate cavitation damage. In refineries, cavitation occurs mostly on the
backside of pump impellers. Certain areas of piping components, such as elbows, can also
become subject to cavitation damage. In addition, vibration can lead to cavitation. Damage is
usually in the form of closely spaced pitting. Cavitation work hardens the surface layer of most
metals and can be detected by metallurgical examination of the damaged part. Cavitation

Page 2-68
CALTEX REFINERY MATERIALS MANUAL October 1999

damage can be reduced by techniques similar to those listed for erosion corrosion. To eliminate
cavitation damage, the conditions causing the cavitation (i.e., insufficient NPSH for pumps, valve
throttling, etc.) must be eliminated.

5.6.6 Mechanical Damage


Mechanical damage to refinery equipment is a reasonably common cause of failure. Typical
examples are the misuse of tools and other equipment, wind damage and careless handling
when equipment is moved or erected. Structural columns are normally designed for compressive
loading and other types of loading may lead to bending. Supports may be damaged when used
as anchors for winches. During earth-moving work, underground pipelines and electrical
conduits may be damaged if they are not carefully located and properly identified.

Flange faces and other machined seating surfaces may be damaged when not protected with
covers or when not handled with care. Material may be thrown from truck beds in such a
manner that it is bent, crushed or cracked. Tubes of heat exchanger tube bundles may be
crushed if the bundles are not lifted with proper slings. Foundations, piping or heat exchanger
shells may be damaged when an attempt is made to pull bundles without adequately anchoring
the shells.

Equipment and structures are normally designed to withstand anticipated wind loads. During
construction or repairs, however, wind damage may result if components are not properly guyed
or reinforced. Loose sheets of metal, boards and the like may be blown about by high winds if
they are not properly secured.

5.6.7 Abrasion
Wear or mechanical abrasion is a significant problem in refineries and accounts for numerous
failures. Catalyst movement in FCC units and coke handling in coking units are examples of wear
situations associated with refinery processes. Wear in pumps, compressors and other rotating
machinery is commonly seen in the refining industry. Abrasive wear can be classified into three
types.

Gouging abrasion - a high stress phenomenon that is likely to be found under conditions
of high compressive stress coupled with impact loads.
Grinding abrasion - high stress abrasion that pulverizes fragments of the abrasive
substance that then becomes sandwiched between mating metal faces.
Erosion - a low stress, scratching abrasive action.

Most parts designed for gouging abrasion service are made of some grade of austenitic-
manganese steel because of its outstanding toughness coupled with good wear resistance. It,
along with hardenable carbon and medium alloy steels and abrasion resistant cast irons, are also
used to resist grinding abrasion. Gouging and grinding abrasion are not generally seen in
refineries.

Erosion is commonly observed in refineries and the abrasive is likely to be gas borne (as in
catalytic cracking units), carried by liquid (as in slurries), or gravity pulled (as in catalyst transfer
lines or coke handling equipment). Because velocity and kinetic energy of abrasive particles are
associated, the severity of erosion typically increases as a function of the velocity. The angle of
impingement also influences severity of attack. There is some evidence that a metal's abrasion
resistance is influenced by whether it is ductile or brittle. Most abrasion involved with
hydrocarbon processing is of the erosive type.

Page 2-69
CALTEX REFINERY MATERIALS MANUAL October 1999

There is a large assortment of alloys available for abrasive service in the form of wrought alloys,
sintered metal compacts, castings and hardsurfacing materials. These materials can be roughly
classified, in descending order of abrasion resistance and ascending order of toughness, as
follows:

Tungsten carbide and sintered carbide compacts


High chromium cast irons and hardfacing alloys
Martensitic irons and hardfacing alloys
Austenitic cast irons and hardfacing alloys
Pearlitic steels
Ferritic steels
Austenitic steels, especially 13% manganese type

Since toughness and abrasion resistance are likely to be opposing properties, considerable
judgement is required in deciding where, in the above series, a given materials candidate lies.
Hardness is often thought to be a property indicative of good wear resistance. It must, however,
be considered with discretion when evaluating an alloys' suitability in abrasive situations.
Hardness should only be considered after its relation to a given service has been proven. Simple
and widely used hardness tests, such as Brinell or Rockwell, tell almost nothing about the
hardness of microscopic constituents which are very important to good wear resistance.

5.6.8 Overloading
Overloading occurs when loads in excess of the maximum permitted by design are applied to
equipment. Hydrostatic testing of vessels can overload supporting structures due to the excess
weight applied. Excessive bending stresses may be induced in vessel shells when pipe support
brackets are attached. Addition of piping to existing pipe supports, or piping that is left
overhanging on supports, may present overloading problems. Overloads can also occur where
metal members have been weakened as a result of corrosion, wear, fire or change in shape or
position. Supports are sometimes bent or shifted in position by accidents or through use as
hitches.

Thermal expansion and contraction cause many overloading problems unless flexible connections
are properly provided. Piping subject to thermal expansion may force a centrifugal pump or
steam turbine out of alignment and warp the shaft, unless the pipe is anchored near the
equipment. Failures result from fatigue stresses that build up at supports, piping and equipment
in which sharp corners exist and in which anchoring attachments are undersized for vibration
loading.

There are other areas where overloading occurs. Gaskets between flanges may be crushed by
uneven or over tightened bolting. Furnace stacks, flare stacks, or similar structures are subject to
overstressing by unevenly tightened guy lines. Failure of equipment may result where wooden
supports decay or burn. Severe impact loads occur in machinery, such as compressors, when
bolts become loose or defective parts fail. Excessive loading is usually apparent because of visible
distortion, change of shape, or change of position. Typical evidence of overloading includes the
following:

Sagged or bent support beams


Cracked welds
Slipped bolts on bolted surfaces
Excessive springing of piping as it is being disconnected

Page 2-70
CALTEX REFINERY MATERIALS MANUAL October 1999

Repeated failures of bolting


Loose guy lines
Distorted expansion joints

5.6.9 Overpressuring
Overpressuring may be defined as the application of pressure in excess of the maximum
allowable working pressure of the equipment under consideration. With low excess pressure,
there is little chance of damage occurring. When excess pressures are high, failures causing loss
of life and property can occur. Overpressuring causes buckling, bulging, ruptures and splits. It
develops in a number of ways, including the following:

Excess heat develops as a result of abnormal operating conditions or upsets. Excess heat
can also be caused by failure of controls or loss of flow as has happened in furnaces.
Blocking off equipment which is not designed to handle full process pressure
Hydraulic hammer or resonant vibration
Inadequate or defective vents and pressure relief valves
Thermal expansion of trapped liquid
Expansion of freezing ice plugs

5.7 TEMPERATURE-RELATED
TEMPERATURE-RELATED FAILURE MECHANISMS

5.7.1 Notch Toughness


Notch toughness is a measure of the ability of a material to absorb rapid or impact loading.
Significant factors are the material used, the operating and ambient temperatures and the
required pressure containing thickness. Body Center Cubic materials (including carbon steel and
most low alloy steels) go through a ductile/brittle transition in a general temperature range of
-46 to 10oC (-50 to +50oF). Specific chemistry, processing, product form, heat treatment and
service conditions produce microstructures that can do a good job of absorbing impact loading
or a poor job. These factors combine and basically determine whether a given defect will
propagate resulting in a brittle failure or not propagate and withstand the stress conditions
imposed on the structure. Notch toughness is typically measured by notched bars broken on
impact at a specific temperature. The energy absorbed in breaking the bar is recorded. Although
considerations can become very complex, energy absorbed greater than 15-20 ft/lbs at a specific
temperature is usually considered adequate to minimize notch toughness problems which could
lead to brittle failure. Other lattice configurations such as Face Center Cubic (austenitic stainless
steels) can have very good notch toughness down to -157oC (-250oF) or below. When exposure
to conditions below -46oC (-50oF) has to be considered, austenitic stainless steels, low nickel alloy
steels and aluminum are used for pressure containing parts.

Twenty five years ago, notch tough carbon steels were easier to obtain from Japan and European
mills than from U.S. mills. Now, U.S. mills can routinely furnish the fine grained steels that give
good notch toughness properties. Notch toughness considerations are discussed in detail in "API
RP 920" and "ASME Section VIII Division 1". "API RP 920" will be dropped when a new fitness for
service document has been approved. The section on minimum pressurization temperatures goes
into more detail on notch toughness and how to figure minimum pressurization temperatures.

Page 2-71
CALTEX REFINERY MATERIALS MANUAL October 1999

Since the notch tough carbon steels have become more readily available, we have had more wet
sulfide cracking and hydrogen cracking problems. Many of the notch tough steels have alloying
elements added to them to enhance their notch toughness. Many of these elements enhance the
notch toughness by increasing hardenability. This increase in hardenability greatly increases
susceptibility to hydrogen cracking during fabrication and tends to yield microstructures
particularly in the heat affected zones that are susceptible to wet sulfide cracking or hydrogen
effects. In many instances, the heat affected zones will not temper at normal carbon steel
postweld heat treating temperatures and therefore, remain vulnerable to cracking problems even
after heat treating. We have had to change our approach to welding ordinary carbon steel
pressure vessel materials. As a result of the sensitivity to hydrogen pickup and cracking during or
right after fabrication, we have had to call for preheat and, in some cases, a hydrogen bake-out
treatment after welding to minimize potential cracking problems. Some fabrication cracking has
gotten into service because we never used to inspect new vessels/equipment with dry magnetic
particle, much less the much more sensitive wet fluorescent magnetic particle inspection. A
cracking survey in Caltex refineries and marketing terminals run in the mid-80's showed that up
to 30% of the spheres and bullets had cracking problems. (Refer to Caltex E&C Bulletin No. 384)
A much higher percentage of the refinery spheres and bullets were cracked than were the
marketing vessels. We have assumed this was due to the increased likelihood that H2S
breakthrough occurred in the refinery and caused the cracking. In analyzing the information, we
estimated that about 75% of the problems were fabrication oriented.

5.7.2 Elevated Temperature Strength, Creep and Rupture


When designing fixed equipment for operation within the creep range, greater than 425oC
(800oF) for carbon steel and 590oC (1100oF) for 3xx Series stainless steels, allowable stresses are
normally chosen to prevent material failure in a given time period (100,000 hours in most cases).
Much of the stress on FCC vessel internals, however, is due to forces generated by thermal
expansion. Although past failure prediction was largely a matter of trial and error, modern stress-
analysis techniques can remedy this problem. Adhering to design conditions is important, as
small increases in temperature can affect creep life significantly. For example, a regenerator
component designed to have a 10 year life at 700oC (1300oF) would be susceptible to creep
failure in little more than 1 year at 760oC (1400oF).

Components which operate at temperatures below the creep range are generally designed on
the basis of yield strength, tensile strength and fatigue strength by applying suitable safety
factors to these values. Because deformation and fracture are not time dependent under these
circumstances, there is no specific value of design life associated with them. In principle, as long
as the applied stresses do not exceed the design stresses, these components should last
indefinitely, although in practice various factors cause reductions in life. In the case of high
temperature components operating in the creep range, both deformation and fracture are time
dependent. They are therefore designed with respect to a target life usually based on a specific
amount of allowable strain or rupture in 100,000 hours. A further factor of safety is applied in
selecting the stress. Many metallurgical and operational factors can extend the actual life beyond
the design life and in like manner can reduce the actual life. Designs are usually based on
minimum or mean values of mechanical properties after other safety factors have been applied.
This can result in actual life exceeding design life. On the other hand, unanticipated stresses and
metallurgical degradation can result in much shorter actual than design life.

(i) Creep
Creep is a high temperature mechanism wherein continuous plastic deformation of a metal takes
place while under applied stresses below the normal yield strength. Creep strengths are usually

Page 2-72
CALTEX REFINERY MATERIALS MANUAL October 1999

expressed as the stress which produces a strain rate of 1% in either 10,000 hours or 100,000
hours at a given metal temperature. Creep strength data are the controlling mechanical property
when metals are exposed for continuous service at high temperatures, such as with furnace tubes
and supports. Creep failures are often found in the form of badly sagged furnace tubes. For
steels, creep becomes evident at temperatures above about 343oC (700oF). In general, the higher
the creep strength, the lower the ductility. A creep brittle material will be characterized by a
higher second stage creep rate, a shorter time to rupture and a lower rupture ductility. Creep
starts as microvoids which can grow under stress to form fissures and cracks. Most refinery
materials tend to be ductile and visibly deform before they fail.

TABLE 2-6 MINIMUM TEMPERATURE FOR CREEP

Material Minimum Temperature for


Creep

Carbon Steel 800o F 427oC


2 Cr - 1 Mo / 5 Cr - Mo 900o F 482oC
9 Cr - 1 Mo 950o F 510oC
T347 SS 1100o F 593oC

MICROSTRUCTURE REPLICAS for


CREEP-LIFE ESTIMATION

Figure 2-10 Fracture

Microstructure Corresponding Action Plan


A Observe
B Observe, Establish Inspection Intervals D
C Limited Services Until Repair
3

D Immediate Repair
ge
St a
Creep Strain

A
2
Stage
STAGE 1, 2, 3 CREEP STAGES
1
Stage

Exposure Time

Page 2-73
TEMPERATURE oC
482 593 704 816
100,000

10,000
STRESS PSI

304

347

9 Cr 1 Mo
Carbon Steel 321
1000 5 Cr 1/2 Mo

2 1/4 Cr 1 Mo

1 1/4 Cr 1/2 Mo

800 900 1000 1100 1200 1300 1400 1500


o
TEMPERATURE F

RELATIVE 100,000 HOUR


RUPTURE STRENGTH

Figure 2-11
100

Wrought
40 Forging (Code No. 2-9)
Casting

Mean

10

Minimum

5
STRESS -

42
-3
P = T (20 - log r) x 10

VARIATION of LARSON-MILLER CREEP RATE


PARAMETER with STRESS for 5 Cr - 1/2Mo STEEL

Figure 2-12
-3
(Td + 460) (20 + log Ld) 10

5 Cr-1/2 Mo
Specified minimum
ASTM A 213 T5

metal temperature
tensile strength Tensile strength A 335 P5

Limiting design
A 200 T5

Specified minimum
yield strength Yield strength

Elastic allowable stress Se

Elastic design governs


above this stress

Elastic allowable stress Se


(for A 200 only)

Average rupture strength

Minimum rupture strength

Rupture allowable stress Sr

Stress (kips per square inch)


Stress (kips per square inch)

-3
Design life, Ld (hours x 10 )
Rupture exponent, n

Design metal temperature, Td (degrees Fahrenheit)

5Cr - 1/2 Mo Steel


ASTM A213 T5, A 335 P5, A200 T5

Figure 2-13
CALTEX REFINERY MATERIALS MANUAL October 1999

Page 2-77
CALTEX REFINERY MATERIALS MANUAL October 1999

(ii) Stress Rupture


Stress rupture is defined by the time it takes for a metal at elevated temperature to fail under
applied stresses below its normal yield strength. Stress rupture data are usually expressed as the
stress which causes rupture in either 100, 1,000, 10,000, or 100,000 hours at a given metal
temperature. Actually, stress rupture tests are accelerated creep tests which have been carried to
failure. Stress rupture data are used extensively in the design of furnace tubes. For carbon steel,
the long term stress to cause rupture at 482oC (900oF) is 1,500 psi (79.5 MPa). This can be
compared to the short term tensile strength of 54,000 psi (373 MPa) for steel at 482oC (900oF).

Grain size and alloy composition are two factors that influence creep and stress rupture. Coarse
grain size materials tend to possess the greatest creep strength at elevated temperatures. Slight
changes in composition often alter the creep strength appreciably, with carbide-forming
elements being the most effective in improving the rupture strength. The relative magnitude of
the effects of small changes in stress and temperature are important to understand. For materials
operating in the creep range, small changes in temperature above design can drastically reduce
service life. Small pressure changes are less significant.

Stress rupture failures in the refinery are usually associated with fired heater tubes and fired
boilers. Most of these are a result of overheating and local hot spots in the furnace as a result of
faulty burners, inadequate control of furnace temperature and coke or scale deposits within or on
the outside of the tubes. Bulging or hot spots are generally signs of impending failure. In the case
of hydrogen producing steam methane reforming furnaces, improper catalyst loading can result
in tube hot spots and ruptures.

For comparison, the 100,000 hour stress to rupture versus temperature curves for several
commonly used alloys are shown in Figure 2-11. Elevated temperature strength as stress to
rupture and percent creep in 100,000 hours set the allowable stress for recognized codes such as
ASME and API RP 530 when operating in the creep range. For practical purposes, the creep range
is material specific and can be a factor above 399o-427oC (750o-800oF) for carbon steel to above
593oC (1100oF) for the austenitic stainless steels. Most refinery equipment is designed to operate
below the creep range. Fired heater tubes are an exception and are typically designed to the
Larsen Miller parameter curves in API RP530. Other creep range designed equipment are the hot
wall Catalytic Reformer and Fluid Catalytic Cracking Reactors and any associated feed effluent
exchangers and piping.

API RP 530 is usually used for heater tube design. The data is presented on curves that plot stress
versus a Larsen-Miller parameter that includes time and temperature (refer to Figure 2-12 and
Figure 2-13) This approach allows a calculation of remaining life for the heater tube or
component. Although the calculations are straightforward, the assumptions that need to be
made are very complex. Remaining life estimates must be made with care. When removing
equipment from service basis a calculated remaining life, there needs to be thorough backup
evidence such as scale thickness, microstructural creep void formation and localized bulging.

5.7.3 Thermal Shock


Thermal shock occurs when large and non-uniform thermal stresses develop over a relatively
short time in a piece of equipment due to differential expansion or contraction. If movement of
the equipment is restrained, this can produce stresses above the yield strength of the metal. In
refineries, thermal shock can be caused by occasional, brief flow interruptions or during a fire.

Page 2-78
CALTEX REFINERY MATERIALS MANUAL October 1999

5.7.4 Thermal Fatigue


Thermal fatigue differs from thermal shock in that the rate of temperature changes experienced is
much greater and the magnitude of the temperature gradient is much less. Every time a
processing unit is started up or shut down, thermal stresses will be set up in equipment.
Repeated application of thermal stresses can lead to progressive cracking, not unlike that of pure
mechanical fatigue. Coke drums are an example of refinery pressure vessels subject to thermal
cycling and associated thermal fatigue cracking. Bypass valves and piping with heavy weld
reinforcement on reactors in cyclic temperature service are also prone to thermal fatigue.

'Fatigue' refers to the initiation and growth of cracks in a component at regions of stress
concentration (notches, defects) under the action of a fluctuating (cyclic) stress. Fatigue cracking
results from the repeated action of very localized microstructural movements (deformations) at
the tip of the notch or defect with each stress cycle. If the fluctuating stresses are the result of
transient thermal gradients within the component, the damage mechanism is termed 'thermal
fatigue'. Unlike creep voids or fissures, which often initiate internally, fatigue cracks initiate along
the component surface and progress inward, perpendicular to it.

Preventing thermal fatigue cracking is best done through design rather than by upgrading
metallurgy. The best means of preventing thermal fatigue are avoiding stress concentrators
(selecting radiused bends in lieu of miters) and limiting temperature cycling.

Page 2-79
CALTEX REFINERY MATERIALS MANUAL October 1999

Page 2-80

Вам также может понравиться