Vous êtes sur la page 1sur 29

CHAPTER 3.

5
INTRODUCTION TO PROCESS CORROSION

THE ELECTROCHEMICAL MECHANISM OF CORROSION

Corrosive attack on the metals normally used in the low temperature (<300°F) sections of a refinery
processing system is an electrochemical reaction. This reaction is usually most severe in areas of the
water, or salt, dew points. Generally, this attack is in the form of acid attack on the active metal surface.

In water, the corrosion reaction occurs at the anode of a corrosion cell. This reaction is also called
oxidation or loss of electrons. An example of a corrosion reaction:

Fe --------> Fe++ + 2e-


iron atom iron ion electron

In order for corrosion to proceed, (1) the electrons must be used up, and (2) the iron ions must be able to
move away from the remaining iron so that new ions can form.

The electrons released at the anode are used up at the cathode. This using up of electrons is called
reduction. If dissolved oxygen is present:
O2 + 2H2O 4e -------> 4OH-
oxygen water electrons hydroxyl ions

If no oxygen is present, the reaction at the cathode requires acid (low pH)

H+ +e- --------> H -----------> H2


Hydrogen ion electron hydrogen atom hydrogen gas

The oxygen or hydrogen ions are dissolved in the water and they are the cathodic reactants or
corrodents.

The four requirements of corrosion:

1. Anodic reaction (corrosion reaction).

2. Cathodic reaction (uses up the electrons produced at the anode).

3. A metallic path for electrons to flow from the anode of the cathode. Electrons flow only
through a metallic path.

4. An electrolytic path for ions to flow between anode and cathode. Ions flow only through an
electrolytic path (water).

If any one of the four requirements are eliminated, the corrosion will stop. Note that the corrosion in
water is electrochemical in nature; this means that electricity (electrons) is involved as well as chemicals.

CORROSION TENDENCY

Corrosion tendency is a fundamental property of each metallic element and of each alloy. Consider bars
of metals immersed in water (no corrodent is present to provide a cathodic reaction):

Each metal has a definite tendency to go into solution as ions, leaving electrons behind on the metal. If
too many ions are found, the high negative charge so produced will bring some back until equilibrium is
restored.
Magnesium (Mg) has a strong tendency to go into solutions as ions; hence, the bar of Mg has many
excess electrons on it at equilibrium, i.e., it shows a high negative charge.

The tendency to go into solution as ions is the corrosion tendency. It can be measured by comparing the
electrode potentials of the metallic bars with each other. Electrode potentials (a measure of corrosion
tendency) are measurable only by comparison of two potentials. Potentials (synonymous with voltages)
are analogous to electron pressure.

If the electrode potentials of several metals and alloys are ranked in descending order, one gets a
"Galvanic Series”. (See Table 3.5-1) Any metal or alloy more active (more negative electrode potential or
having a greater corrosion tendency) than those metals or alloys below it in the series is able to supply
electrons to them when they are coupled together. (Whether this coupling will affect the corrosion rate
must be determined separately.)

TABLE 3.5-I

GALVANIC SERIES OF SOME COMMERCIAL


METALS AND ALLOYS IN SEA WATER

Active Magnesium
or Anodic Magnesium Alloys
Zinc
Galvanized Steel

Aluminum 1100

Aluminum 2024 (4.5 Cu, 1.5 Mg. 0.6 Mn)

Mild Steel
Wrought Iron

Cast Iron

13% Chromium Stainless Steel


Type 410 (Active)
18-8 Stainless Steel
Type 304 (Active)

Lead-Ti Solders
Lead
Tin

Muntz Metal
Manganese Bronze
Naval Brass

Nickel (Active)
76 Ni-16 Cr-7 Fe alloy (Active)

60 Ni - 30 Mo - 6 Fe - 1 Mn

Yellow Brass
Admiralty Brass

Red Brass
Copper
Silicon Bronze
70:30 Cupro Nickel
G-Bronze
Silver Solder
Nickel (Passive)
76 Ni - 16 Cr - 7 Fe
Alloy (Passive)

13% Chromium Stainless Steel Type


410 (Passive)
Titanium

18-8 Stainless Steel


Type 300 Series (Passive)

Silver
Noble or
Cathodic Graphite
Gold
Platinum

FORMS OF CORROSION

1. Uniform
2. Galvanic-Two Metal
3. Erosion
4. Fretting
5. Crevice
6. Pitting
7. Exfoliation
8. Selective leaching - Dezincification
9. Intergranular
10.Stress Corrosion Cracking
11. Embrittlement
12.Corrosion Fatigue

1. Uniform - Easiest to measure. Unexpected failures can be avoided by regular inspection.


Knowledgeable selection of materials and protection methods are used to control it.

2. Galvanic - Occurs when two dissimilar metals are in contact and exposed to a conductive
solution. Difference in electrical potential exists between different metals and is the
driving force to pass current through the corrodent. The larger the potential difference
the greater the corrosion.

Area effects: Large cathode/small anode = rapid corrosion.

CAN BE AVOIDED BY:

1. Insulation
2. Restricted use of metals together when they have a large potential difference.
(Refer to Table 3.5-I)
3. Favorable area effects.

Different microstructures can produce galvanic effects. Welding can produce galvanic
effects.
3. Erosion Corrosion - The movement of a corrodent over a metal surface increases the rate
of attack due to mechanical wear and corrosion. The role of erosion is usually
attributed to the removal of protective surface films, e.g. air-formed protective oxide
films or adherent corrosion products.

Appearance - smooth-bottomed shallow pits.

CAN BE AVOIDED BY:

1. Design changes
2. More corrosion resistant material
3. Cathodic protection can be helpful
4. Coatings
5. Changing the environment

Cavitation: Formation and collapse of vapor bubbles at the metal surface. High pressures
produced by this collapse can deform the metal, remove protective films, etc. Occurs on
pump impellers, agitators, piping.

4. Fretting - Occurs when metals slide over each other and cause mechanical damage to one
or both of them. Corrosion plays one of the following roles:

1. Heat of friction oxidizes the metal and this oxide, then wears away.

2. The mechanical removal of protective oxides or corrosion products results in the


continued exposure of fresh surface that continues to actively corrode.

FRETTING CORROSION IS MINIMIZED BY:

1. Using harder materials.


2. Designing to minimize friction (lubrication).
3. Increasing friction to a state that movement no longer occurs.

5. Crevice Corrosion - Environmental conditions in a crevice can become quite different to


those on a nearby clean, open surface.

Crevices commonly exist at gaskets, lap joints, bolts, rivets, etc. They are also created by
deposits, corrosion products, scratches in paint films, etc.

CREVICE CORROSION IS USUALLY ATTRIBUTED TO ONE OR MORE OF THE


FOLLOWING:

1. Changes in acidity in the crevice.


2. Lack of oxygen in the crevice.
3. Buildup of a detrimental ion in the crevice.
4. Depletion of an inhibitor in the crevice.

Some materials are more susceptible than others; namely, those that depend on an air-
formed oxide film to achieve their corrosion resistance, such as stainless steel and
titanium.

Many failures are the result of crevice corrosion. Pitting occurring under a deposit (salt).

WAYS TO COMBAT CREVICE CORROSION:

1. Alloy materials to improve their resistance (316SS).


2. Designing to minimize crevices.
3. Maintenance to keep surfaces clean (water washing).
4. Eliminate salt deposition.

6. Pitting Corrosion - Formation of holes in an otherwise relatively unattacked surface. A pit


can be considered a self-formed crevice. Pits continue to grow for the same reasons
mentioned in crevice corrosion.

To minimize pitting, a clean, homogeneous surface is desirable. For example, a pure,


homogeneous metal with a highly polished surface will generally be more resistant than
one with many inclusions, defects, and a rough surface.

Surface cleanliness and selection of materials known to be resistant to pitting in the given
environment are usually the safest ways of avoiding the problem.

7. Exfoliation - Exfoliation is subsurface corrosion that begins on a clean surface but spreads
below it. The attack is usually recognizable by a flaky and sometimes blistered surface. It
is well known in aluminum alloys and can be combated by alloying and heat treatment.

8. Selective Leaching - Selective leaching is the preferential removal of one element from an
alloy. The most common example is the removal of zinc in copper-zinc alloys
(dezincification). The corrosion is detrimental because it yields a porous metal with poor
mechanical properties. The remedy involves the use of nonsusceptible alloys.

9. Intergranular Corrosion - Localized attack at and adjacent to grain boundaries with


relatively little corrosion of the grains is intergranular corrosion. It can be caused by
impurities at the grain boundaries, enrichment of one of the alloying elements, or depletion
of one of these elements in the grain-boundary area. Austenitic stainless steels (300
series) are very susceptible to intergranular corrosion.

The almost universally accepted theory for intergranular corrosion is based on the
depletion of chromium in the grain-boundary areas. It takes about 11.5% chromium to
make a stainless steel. If the chromium is effectively lowered, the relatively poor corrosion
resistance of ordinary steel is approached. In the temperature range of 650 - 1550°F, the
chromium and carbon that are in solid solution combine and form Cr 23C6 precipitate out of
solution. The Cr is thereby removed from solid resolution and the result is metal with
lowered Cr content in the area adjacent to the grain boundaries.

THREE METHODS USED TO MINIMIZE INTERGRANULAR CORROSION OF


AUSTENITIC STAINLESS STEEL:

1. Employing high temperature solution heat treatment (commonly termed quench-


annealing or solution-quenching).
2. Adding elements that are strong carbide formers (called stabilizers -Cb -
columbium, Ti - titanium).
3. Lowering the carbon content to below 0.03% (304L).

Weld decay and knife-line attack are two other types of failures related to intergranular
corrosion.

10. Stress Corrosion Cracking - Stress-corrosion cracking refers to cracking caused by the
simultaneous presence of tensile stress and a specific corrosive medium.

SOME EXAMPLES OF STRESS-CORROSION CRACKING OF PARTICULAR METALS


IN CERTAIN ENVIRONMENTS ARE:

1. Chloride SCC of austenitic stainless steels.


2. Ammonia SCC of brass.
3. Caustic embrittlement of steel.

IN THE CASE OF CHLORIDE SCC OF STAINLESS STEELS, THERE ARE


SEVERAL ENVIRONMENTAL FACTORS THAT ARE NECESSARY FOR
CRACKING TO OCCUR:

1. Chloride concentration, not known exactly what the critical concentration is.
2. Elevated temperature, >150°F.
3. Stress, not known exactly what the minimum stress level is.
4. Must be oxygen present.
5. Must be an aqueous phase.

STRESS CORROSION CRACKING CAN BE REDUCED OR PREVENTED BY


APPLICATION OF ONE OR MORE OF THE FOLLOWING METHODS:
1. Lowering the stress, stress relief.
2. Eliminating the critical environmental species.
3. Changing the alloy.
4. Applying cathodic protection.
5. Adding inhibitors.

11. Hydrogen Damage - Hydrogen damage may be classified into four distinct types:
1. HYDROGEN BLISTERING.
2. HYDROGEN EMBRITTLEMENT.
3. DECARBURIZATION.
4. HYDROGEN ATTACK.

HYDROGEN BLISTERING: Hydrogen blistering results from the penetration of


hydrogen into a metal. Hydrogen can be produced at the metal surface by a corrosion
reaction or cathodic protection. Atomic hydrogen diffuses through the steel until it
encounters a defect (void) and then it combines to form molecular hydrogen. Since
molecular hydrogen cannot diffuse, the concentration and pressure of hydrogen gas
within the void increase. The equilibrium pressure of molecular hydrogen in contact
with atomic hydrogen is several hundred thousand atmospheres, which is sufficient to
rupture any engineering material.

Hydrogen Blistering Can Be Prevented By:

1. Using clean steel.


2. Using coatings impervious to hydrogen penetration.
3. Using inhibitors to reduce corrosion rate.
4. Removing poisons which increase hydrogen uptake (cyanide).
5. Substituting alloys.

HYDROGEN EMBRITTLEMENT: The exact mechanism of hydrogen embrittlement


is not as well know as that of hydrogen blistering. The initial cause is the same,
penetration of atomic hydrogen into the metal structure.

The necessary factors for hydrogen embrittlement are:

1. A critical concentration of hydrogen.


2. A tensile stress.
3. A susceptible steel, i.e., high enough strength level.
4. Temperature must be below 250°F.

Hydrogen embrittlement can be prevented by:


1. Reducing corrosion rate.
2. Altering plating conditions to reduce hydrogen pickup.
3. Baking the hydrogen out of the steel.
4. Substituting alloys.
5. Lowering the strength level of the steel.
6. Practicing proper welding, low H2 rods.

DECARBURIZATION: Decarburization may be defined as the loss of carbon from


the surface of a ferrous alloy as a result of heating in a medium that reacts with the
carbon. So far as refinery equipment is concerned, decarburization is usually the
result of restricted high-temperature oxidation (see next sub-topic for an exception).
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 appearance of the decarburized layer is usually not serious
unless fatigue is a problem. However, its occurrence in operating equipment is
evidence that the steel has been overheated, and other effects may be present.
Decarburization can be found only by metallurgical examination. For all practical
purposes, decarburization is limited in refinery service to the ferritic steels, and is
most often found in steam and hydrogen services.

Although decarburization of refinery process equipment is in itself not usually


harmful, it can be quite detrimental in some applications. Heat treating of steels
without the use of a proper atmosphere may result in a thin, decarburized layer. This
soft, weak layer has very poor fatigue properties and may result in failure of springs,
shafts, gears, and the like. Also, the presence of a decarburized layer on a cutting
tool will preclude proper surface hardening and result in poor performance.

HIGH-TEMPERATURE HYDROGEN ATTACK: Some refining operations involve the


use or production of hydrogen at high temperatures and pressures. Hydrogen at
elevated temperatures (above 450°F) can have a very destructive effect on steels.

At high pressures and temperatures, hydrogen dissociates into atomic hydrogen


which can penetrate a metal. Likewise, carbon in steel will migrate at high
temperatures. Thus, at low pressures but high temperatures, carbon can migrate
toward the surface and combine with hydrogen to form methane. Under these
conditions the hydrogen attack is confined to decarburization. However, if the stress
level is high, cracking may result from over stressing.

As pressure is increased and the temperature is still high, the dissociation of


hydrogen and the penetration of hydrogen into the metal is more rapid than the
migration of carbon to the surface. Under these conditions, carbon migrates to the
grain boundaries, dislocations, and the like, and combines with hydrogen to form
methane. This methane reaction causes internal pressure that can crack the
metal intergranularly. Under externally applied stress the cracks form and propagate
more readily.

12. Corrosion Fatigue - Fatigue failures occur in the absence of corrodents and are caused by
repeated cyclic stressing. Such failures are common in structures subjected to continued
vibration. Sucker rods almost always fail by fatigue or corrosion fatigue. A corrodent will
sometimes lower by half the stress normally required to cause fatigue failures in dry air.

Methods of avoiding corrosion fatigue usually attempt to prevent a fatigue crack from
starting -- e.g. design changes, protection, inhibition, etc. -since it is difficult to stop a crack
from propagating once initiated.

CORROSIVE COMPONENTS OF CRUDE OIL


The crude oil to be processed in refining operations contains impurities that can produce corrosive
conditions within the refining equipment and can adversely affect the quality of the products. The
purpose of this section is to discuss the corrosive elements encountered in refining processes and how
these result in deterioration of equipment. The corrosion problems in refining operations that are of
concern to us can be divided into two major groups as follows:

1. Corrosion from components present in the crude oil.


2. Corrosion from chemicals used in refinery processes.

Corrosive Components of Crude Oil

A. HYDROGEN CHLORIDE AND ORGANIC AND INORGANIC CHLORIDES

Crude oil contains salts because brine is produced along with the crude oil and the separation
of crude from brine is not perfect. The brine is generally believed to be roughly the same
analysis as sea water. A typical analysis of sea water is shown in Table 3.5-2.

Table 3.5-2
The Major Constituents of Sea Water
Ion Parts per Million
Chlrodie, Cl- 18,980.0
Sulfate, SO 4-- 2,649.0
Bicarbonate, HCO3- 139.7
Bromine, BR- 64.6
Fluoride, F- 1.3
Boric acid, H3BO3 26.0
Total
Sodium Na+ 10,556.1
Magnesium, Mg++ 1,272.0
Calcium, Ca ++
400.1
Potassium, K+ 380.0
Strontium, Sr ++
13.3
Total

Magnesium and, to a lesser extent, calcium chlorides when dissolved in water and heated
form hydrochloric acid that is very corrosive:

MgCl2 + 2H2OD -----> Mg (OH)2 ß + 2 HCI Ý

This reaction, called hydrolysis, begins to take place at approximately 250°F. The extent of
hydrolysis increases as the salt concentration in the crude decreases. The relative
contributions to HCl evolution in order of decreasing hydrolysis is: MgCl 2 >> CaCl2) NaCl.
Sodium chloride does not hydrolyze and, consequently, does not cause serious corrosion in
crude distillation units.
Hydrogen chloride gas is normally not corrosive in process streams above the water dew
point. However, as water begins to condense, HCI is dissolved to form hydrochloric acid.
(Water is present in the system from the injection of stripping steam in addition to the small
amount that is entrained in the hydrocarbons being processed.) Hydrochloric acid produced
during crude oil distillation is the principle strong acid responsible for corrosion in crude unit
overheads. In the presence of hydrogen sulfide, the corrosion of iron by hydrochloric acid
becomes even more pronounced. The corrosion of iron by hydrochloric acid becomes a cyclic
reaction as follows:
Fe + 2 HCl <----> FeCl 2 + H2

FeCl2 + H2S <----> 2 HCI + FeS

The first equation demonstrates the hydrochloric acid (HCI) producing the primary corrosion of
iron (Fe). Hydrogen sulfide (H2S) then regenerates iron chloride (FeCl 2) to form additional HCl
and iron sulfide (FeS), as shown in the second equation. This, in turn, causes more HCl
corrosion and the cycle is repeated.

Corrosion by salts and hydrochloric acid can be reduced by decreasing the crude oil salt
content through the use of single or double stage desalting.

Keeping overhead chloride levels to a minimum is fundamental to many corrosion control


programs. Caustic addition to the crude charge is a common practice to control overhead
chloride levels. Where caustic use would be unsatisfactory or restricted, neutralizing amines
offer another means of control. A further discussion of caustic injection and its limitations may
be found on page 23 of this section.

The primary goal of both desalting and neutralizer addition to the crude charge is to reduce the
quantity of HC1 that passes overhead from the atmospheric tower. Condensing system
corrosivity varies directly with condensate HCl content and temperature. Chloride
concentrations are substantially higher in the area where water first begins to condense. Table
3.5-3 is a typical example of crude unit overhead condensate chloride levels as a function of
temperature. Clearly, dilution of the chlorides takes place as additional water is condensed.

Table 3.5-3
Neutralizer Titration Curves
HCl Alone HCl Plus
0 2.6 2.5
120 2.63 2.63
125 2.7
130 2.71
225 2.73
240 2.75
260 2.76
280 2.77
300 2.74
325 2.78
350 2.785
375 2.789
380 2.79
395 2.78
450 3
475 3.1 3
530 3.2
560 3.25 3.2
600 3.49
625 3.65
630 3.8 3.4
640 3.85
690 4.1
695 4.25
700 5
700 6.2
705 6.7
710 6.78
720 7.25
720 7.45
730 7.6
740 7.78
745 8.15
750 3.6
770 8.25
840 3.8
895 3.95
1000 4.15
1080 4.25
1125 4.35
1225 4.5
1325 4.6
1410 4.8
1530 4.95
1600 5.15
1625 5.22
1725 5.4
1800 5.65
1825 5.8
1835 6
1860 6.2
1875 6.4
1885 6.75
1890 6.85
1925 7.3
1950 7.5
1975 7.8
1995 8

Crude overhead accumulator condensate chloride levels ranging from under 5 ppm to over
400 ppm have been observed in the field. The wide variation in chloride levels is due to
different salt levels, desalting practices, dilution effects of stripping steam, water wash rate,
and crude neutralization techniques.

For example, a Midwestern refiner controls chloride levels in the 1 to 5 ppm range as a result
of desalting to 1 pound per thousand barrels (ptb) (2.85 mg/I), addition of 3 ptb (8.55 mg/I)
caustic, and 3 ppm oil soluble neutralizer to the crude charge. Another refiner on the West
Coast runs heavy California crude oils without desalting and, as a result, salts to the crude unit
average 25 ptb (71.28 mg/I). Here with 3 ptb (8.55 mg/l) caustic addition to the crude,
overhead chlorides average 150 ppm.

B. SULFUR COMPOUNDS
Usually, hydrogen sulfide is the major sulfur species that increases crude unit overhead
corrosion potential. Maintaining accumulator water pH in the proper range in sour
overhead systems can significantly reduce H2S related corrosion potential in the aqueous
phase, reduce fouling tendency and minimize neutralizer consumption.

Sulfoxy acids like H2SO3 and H2SO4 can also form near or above the initial
water condensation point either from direct condensation of acid gas (SO 2 and SO3), or
by oxidation of H2S. Higher concentrations of these sulfoxy species have been identified
in vacuum tower overheads where a higher oxygen concentration exists. Sulfur species
increase the overall system corrosion loading, increase salt formation, and increase
neutralizer demand.

Hydrogen Sulfide

Hydrogen sulfide is probably the most active of the sulfur compounds in causing
corrosion in refinery equipment. Hydrogen sulfide itself is present in crude oil and
additional hydrogen sulfide may be formed by the decomposition of organic sulfur
compounds at the process temperatures.

Temperature plays an important part in corrosion by hydrogen sulfide. In the absence. of


water, corrosion is not serious at low temperatures, and a metal temperature above
450°F is necessary for "dry" H2S corrosion to occur. This "dry corrosion is typically found
in flash zones and furnaces processing sour crude slates. Hydrogen sulfide gas is not
corrosive in the temperature ranges that are present in most crude unit overhead
systems. In high temperature areas, 5% and 9% chromium steels are used to provide
increased protection.

Hydrogen sulfide reactivity in the vapor phase has minimal effect on neutralizer demand
due to the instability of this salt at temperatures above the water dew point. This has
been confirmed through neutralizer bisulfide salt TGA data as well as field experience
comparing neutralizer demand between high and low hydrogen sulfide reactivity
neutralizing amines.

Hydrogen sulfide in the aqueous phase will accelerate corrosion in a crude overhead
system by the following mechanism:

Fe + 2HCl Û FeCl2 + H2+

H2S + FeCl2 Û FeS + 2HCl

Hydrogen sulfide corrosive attack is a function of the concentration of the acid in the
aqueous phase. Hydrogen sulfide solubility in water increases with increasing pH and
decreasing temperatures.

The higher the concentration of hydrogen sulfide in the water, the more corrosive it
becomes. High concentrations of hydrogen sulfide will increase corrosion due to
increased formation of hydrochloric acid and iron sulfide.

Low levels of sulfides in an overhead system may be beneficial for corrosion control
purposes. When hydrogen sulfide reacts with the metal surface, it forms a brittle, but
somewhat protective, sulfide film on the metal surface. The least protective sulfide films
are formed in the pH range of 6 to 8; however, the sulfide film stability increases above a
pH of 8. The sulfide film is readily dissolved by stronger acids, such as hydrochloric
acid, when found in higher concentrations. If excessive amounts of hydrogen sulfide
dissolve in the water, the thickness of the brittle film will increase. When the film reaches
an inordinate thickness, the natural velocities in the system can strip it from the metal
wall surface. This exposes a fresh metal surface to the corrosive environment and results
in an acceleration of the corrosion rates in the overhead system.

Iron sulfide and other metal sulfides are insoluble in both hydrocarbon and water.
Therefore, they can deposit out and cause fouling on equipment surfaces, as well as
under-deposit corrosion.

In fact, iron sulfide is a major emulsion stabilizer in crude oil. The metal sulfide corrosion
product tends to float at the interface between hydrocarbon and water. Its presence will
increase water carryover in the hydrocarbon product streams coming from the
accumulator.

The hydrogen sulfide compound becomes dissociated in aqueous solution to the


bisulfide ion under mildly alkaline pH conditions by the following mechanism:

H2S (aq) Û H+ (aq) + HS- (aq)

The tendency of hydrogen sulfide to dissociate in the aqueous phase increases with
increasing pH. See Figure 3.5-1.
Figure 3.5-1

Hydrogen Sulfide Dissociation


100
H2S H+ + HS-
Percent Dissociation

80

60

40

20

0
4 5 6 7 8 9
pH of the Condensate

The bisulfide ion is considered mildly corrosive relative to hydrochloric acid with
increasing corrosivity at higher concentrations. Like hydrogen sulfide, the bisulfide
concentration increases at decreasing temperature. The highest bisulfide concentrations
are found at the lower temperatures present in the accumulator water while very low
levels of bisulfide are present at the water dew point.

Steel Corrosion Reactions

The overall corrosion reaction for steel is:

Reaction 1: Fe + 2HS- Û FeS + S-2 + 2Ho

The amount of bisulfide ion formed depends on pH, temperature and H 2S partial
pressure. The hydrogen sulfide and bisulfide concentrations are equal at approximately
6.8 pH, and the hydrogen sulfide concentration is nil above pH 9. The iron sulfide scale
that forms on the metal surface in the presence of FCC alkaline sour water at pH 8 to 9
is soft, porous and loosely adherent. If cyanides are not present, that scale will reduce
the overall corrosion rate.
The bisulfide ion will consume neutralizer. Therefore, in order to reduce the neutralizer
consumption from the bisulfide ion, the concentration of bisulfide in the accumulator
water must be minimized. This can be done by restraining the sulfides from dissociating
through accumulator pH control. In a sour system, the neutralizer demand can vary up to
50% by elevating accumulator water pH from 5.5 - 6.0 to 6.0 - 6.5. A lower accumulator
water pH control range in sour systems (5.5 to 6.0) will minimize hydrogen sulfide
dissociation and reduce neutralizer demand significantly without adversely lowering the
dew point pH. Because bisulfide tends to buffer the overhead water pH, accumulator pH
can be controlled in a tight range without much fluctuation with a nonbuffered neutralizer.

Bisulfide salt formation will typically occur below the water dew point and is considered to
be of minor importance in crude unit overhead corrosion due to the presence of free
water.

Copper Corrosion Reaction

The corrosion of copper by H2S generally results in the formation of copper sulfide, and
the reaction can be self-limiting. In the FCC where copper complexing agents such as
ammonia and ammonium salts are present in high concentrations, the copper sulfide film
cannot become permanently established. As the copper sulfide film is formed, it is
complexed and washed away, and further sulfidation of the base metal then takes place.

Cu+2 + 4NH4 + Û Cu (NH3)4++ + 4H+

(Copper is one metal that forms a very stable ammonium complex even in a weak
ammonium solution.)

C. OXIDIZED SULFUR SPECIES

Sulfoxy acid species are formed in the presence of oxygen in crude overhead systems.
Three of these species have been identified in significant quantities: sulfite, sulfate, and
thiosulfate.

Sulfite (SO 3-2) is the dissociated anion from sulfurous acid (H 2SO3). Its presence in
process condensate is the result of three probable reaction paths: 1) acid gas (SO 2)
condensation; 2) elemental sulfur oxidation; or 3) H 2S dissociation in the presence of
dissolved oxygen.

1. SO2 + H2O + H2SO3


2. H2S + SO2 Þ 3S° + 2 H2O (Claus Reaction)
S° + O2 + H2O Þ H+ + HSO3- Û H+ + SO3-2

3. 3HS- + 3/202 Û HS03-+ H+ Û SO3-2 + H+

Of these three reaction mechanisms, only the third may be controlled by proper pH
control in a crude overhead system. To explain, bisulfite (HSO 3-) may form from the
reaction of bisulfide with oxygen below the water dew point. The chances of bisulfite
formation increase with the rate of dissociation of hydrogen sulfide. Sulfite concentration,
like bisulfide concentration, increases at decreasing temperatures. The rate at which
bisulfide reacts with oxygen to form the sulfoxy acid is pH dependent. The rate of oxygen
uptake by bisulfide is highest between pH levels of 6 - 7. The rate of reaction decreases
between pH 8 - 9 and increases again over a pH of 9.

Controlling accumulator pH above 6.0 in systems with high HS concentration and


dissolved oxygen present will further increase neutralizer demand, not only by
increasing H2S dissociation, but also by increasing the rate of sulfoxy acid species
formation. These stronger acids, in turn increase the baseline corrosion potential. It has
been suggested that each form of reduced sulfur, H2S and HS-, is oxidized via a different
pathway which leads to a different distribution of reaction products.

Sulfite has been occasionally identified in accumulator waters in concentrations up to 50


ppm.

In the aqueous phase where excess oxygen is present, sulfite (SO 3-2) can be oxidized to
the sulfate ion (SO 4-2). Oxidation from SO3 = to SO4 is a very slow process relative to the
oxidation of H2S to SO3 =; therefore, low concentrations (around 5 ppm) have been found
in some overhead systems. High pH (above 8.5) conditions would increase the rate of
reaction.

Research indicates that the formation of sulfuric acid can occur approximately 80°F
above the water dew point temperature. Although H 2SO4 is a strong acid similar to HCl,
the concentrations found in crude unit overheads are too low to be considered a
significant factor in corrosivity and neutralizer demand. There are some exceptions;
however, in a West Coast refinery, sulfuric acid was added into the desalter wash water.
This resulted in a dew point pH of less than one on several occasions and high corrosion
rates in the crude unit overhead due to the presence of increased sulfuric acid.

Vacuum tower overheads are more likely to form sulfuric acid due to oxygen
contamination from equipment leakage. Sulfate concentrations of 20 - 50 ppm have
been identified in these systems. Corrosion from sulfuric acid is significant in combustion
processes with large volumes of excess oxygen, i.e., boilers and process furnaces
burning sulfur containing fuels. Unlike crude overhead systems, these systems have high
levels of oxygen which promote high concentrations of sulfuric acid.

Thiosulfate (S2O2-2) is the anion of thiosulfuric acid (H 2S203). Thiosulfuric acid can exist in
the acid form in low oxygen concentrations that are present in crude overheads.
Thiosulfate formation can result from the reaction between sulfur and sulfite in aqueous
solution.

SO3-2 (aq) + S(s) Þ S2O3-2 (aq)

This reaction favors thiosulfate formation as the pH increases above 8.5. Thiosulfate
cannot condense from the vapor phase as could sulfite (SO3-2) and sulfate (SO 4-2). The
concentration of thiosulfate increases at decreasing temperatures; therefore, the highest
concentration of this anion is present in the accumulator water in the crude unit
overhead. Thiosulfate has been identified in several overhead systems in concentrations
up to 400 ppm.

D. OXYGEN

Oxygen is present in refining operations as a result of increased handling of the crude.


Oxygen can enter the system due to breathing of the equipment during temperature
changes and pumpings, open air water sources and steam leaks.

The corrosion mechanism of iron is aggravated by the presence of oxygen. Iron will react
with oxygen in the presence of water to form ferric hydroxide by the following equation:

4 Fe + 6 H20 + 3 02 Þ 4 Fe (OH)3
Iron Water Oxygen Ferric Hydroxide

Ferric hydroxide is insoluble in water and precipitates out of solution. The precipitate
formed by ferric hydroxide is not protective. The effects of temperature and time produce
ferric oxide, a red rust, as the final end product of the corrosion mechanism as shown
below:

2 Fe (OH)3 + Temp/Time Þ Fe2O3 + 3 H2O


Ferric Hydroxide Ferric Oxide Water

Laboratory studies of this type of corrosion indicate that globules of water form on the
metal, and corrosion occurs under each droplet in the form of a pit. Sometimes, the pits
are so numerous that corrosion appears to be general.

Oxygen contamination of up to 6 ppm has been seen in vacuum tower overhead systems
as a result of vacuum ejector or pump leakage. Oxygen initiated corrosion generally
occurs at low operating temperatures in these units. This is due to the increase in
solubility of the primary corrodents present in the aqueous phase at decreasing
temperatures.

E. CARBON DIOXIDE

Carbon dioxide combined with water causes corrosion in refining processes. There are
two major sources of carbon dioxide, being: 1) From the decomposition of bicarbonates
contained in, or added to, the crude oil; and 2) from steam used to aid distillation. As
carbon dioxide dissolves in water, it causes the pH to be depressed by increasing the
hydrogen ion concentration as shown in the reaction sequence below:

CO2 + H2O Þ H2CO3


Carbon dioxide + Water Þ Carbonic Acid

H2CO3 Þ H+ + HCO3-
Carbonic Acid Þ Hydrogen + Bicarbonate ion

HCO3- Þ H+ + CO3-2
Bicarbonate ion Þ hydrogen ion + Carbonate ion

Carbonic acid promotes the iron corrosion reaction by supplying a reactant, H+ The
overall reaction is:

2 H2CO3 + Fe Þ Fe (HCO3)2 + H2

Carbonic acid Iron Ferrous Hydrogen Bicarbonate

In most crude unit systems, the concentration of carbon dioxide in water is low, and
corrosion rates are low relative to those associated with other corrodents. The corrosion
rate from carbon dioxide depends on the partial pressure, pH, and temperature of the
unit. The pH of streams does not give a good indication of the anticipated corrosion when
a weak acid such as carbonic acid is the corrodent. It has been found that appreciable
corrosion can occur in carbonic acid solutions at a pH of 6. Total acidity, rather than pH,
determines the extent to which an acid corrodes iron.

Carbon dioxide corrosion of condensate piping is a generalized loss of metal. Thinning


of the pipe wall can occur, particularly in the lower portion of the pipe circumference.

Carbon dioxide corrosion is of major importance in reforming processes for generating


hydrogen that employs catalyst over which hydrocarbon and steam are passed.
Corrosion by carbon dioxide in these plants cause metal dissolution, intergranular attack,
and stress corrosion cracking.
F. ORGANIC ACIDS

The presence of low molecular weight organic acids in oil and gas production has been
documented for some time. In this environment these acids, often referred to as lower
fatty acids, contribute to the acidity of the produced water which is responsible for
corrosion. In all of the crude unit overheads studied, some level of low molecular weight
fatty acids in the water condensate were found. Acetic, propionic, and butyric are the
most prevalent compounds with acetic usually present in the highest concentration.
Acids as high as C7 have been reported by our laboratories. Water solubility of the acids
drops off significantly above C5.

Total fatty acid concentration varied between 10 ppm and 1000 ppm in accumulator
water. Average concentrations were between 100 to 150 ppm which is equivalent to
about 50 to 70 ppm as chloride ion with respect to neutralizer demand. The highest
readings to date have come from California refineries which process heavy naphthenic
California crude oils.

Figure 3.5-3A shows the acid composition of a Midwestern refinery overhead


condensate. In this study, Horican crude had the highest level of organic acids with acetic
acid present in greatest amount.
Figure 3.5-3A
Crude Unit Overhead Water Analysis
% Acids expressed as HCl Equivalents

40% Cabinda,
20% Domestic,
Louisiana Light 40% Foreign Mix

Legend
Hydrochloric Acid

Hdrogen Sulfide

Sulfoxy Acids
Bow River, Wyoming Horican
Carboxcylic Acids

Low molecular weight carboxylic acids can contribute to crude unit overhead corrosion
potential, especially in the area of the water dew point. Also, as with other acids, salts
can form above the dew point to promote corrosion and fouling conditions. Their
presence can increase corrosion rates and have a significant impact on neutralizer
demand. Frequently, we find that high or fluctuating neutralizer requirements can be
explained by the presence of these weak acids. Organic acid corrosion has been studied
in the laboratory using acetic acid.

It should be noted that the hydrogen ion activity of all the fatty acids is not too variant
from that of acetic acid. The dissociation constants or pKa values indicate these acids
are partially ionized and, as a result, are classified as weak acids. The pKa values and
other physical properties for several organic acids are shown in Table 3.5-4.

Table 3.5-4
Physical Properties of Organic Acids
Solub.
Organic Acids MW B.P. °F pKa g/100 g H2O
Formic 45 213 3.75 ¥
Acetic 60 245 4.75 ¥
Propionic 74 287 4.87 ¥
Butyric 88 326 4.83 ¥
Pentanoic 102 401 4.88 3.7

Carbon steel corrosion rate curves have been developed at 20°C for dilute solutions of
formic and acetic acid. This data is presented in Figure 3.5-2. Corrosion is of a general
nature with no pitting observed. Increased temperature would be expected to shift the
line upward, about doubling for each 20°C rise. On this basis, 1000 ppm of acetic acid
would generate corrosion rates of about 80 mpy (2.03 MM/y) at 100°C. While
concentrations of this magnitude are not found in accumulator water, such levels have
been observed at the initial condensate point as would be expected since acetic acid
would condense prior to water.
Figure 3.5-2

The Effect of Organic Acids


on the Corrosion Rate
Carbon Steel at 20°C
8
Corrosion Rate (mpy)

ci d
6
m icA
For tic A
ci d
A ce
4

0
0 200 400 600 800 1000
Acid Concentration (ppm)

From a corrosion control viewpoint, the most significant impact will be an increased
demand for neutralizers. In many systems (see Figures 3.5-3A and 3.5-3B), the
neutralizer demand, due to organic acids, can be equal to or greater than that of HCl.
The impact on the neutralizer demand decreases as the molecular weight of the acid
increases. Data provided in Figure 3.5-3B gives an example of a West Coast refinery
overhead water analysis of organic acids (as HCl equivalent). Note that organic acidity is
much greater than HCl acidity.
Figure 3.5-3A
Crude Unit Overhead Water Analysis
% Acids expressed as HCl Equivalents

40% Cabinda,
20% Domestic,
Louisiana Light 40% Foreign Mix

Legend
Hydrochloric Acid

Hdrogen Sulfide

Sulfoxy Acids
Bow River, Wyoming Horican
Carboxcylic Acids

Figure 3.5-3B
Crude Unit Overhead Water Analysis
% Acids expressed as HCl Equivalents

80% Maya,
San Joaquin Valley 20% Empire

Legend
Hydrochloric Acid

Hdrogen Sulfide

Sulfoxy Acids
West Texas, Sour California Naphthenic
Carboxcylic Acids

As can be seen by the titration curves in Figure 3.5-4, organic acids can buffer pH in the
range of 3.5 to 5.5. These titrations were carried out with 100 ppm HCI and various
levels of acetic, propionic, and butanoic acids. The organic acids increase the neutralizer
demand by about 2.5 times compared to HCI alone. This buffering effect can result in
little pH change with increasing neutralizer addition rates. This condition was observed at
a West Coast refinery where high levels of acetic acid caused the pH to buffer in the 4.5
to 5.0 range.
Figure 3.5-4

Neutralizer Titration Curves


9
8
HCl Plus
7 HCl Alone Organic Acids
6

pH
5
4
3
2
0 500 1000 1500 2000
Neutralizer Concentration (ppm)

G. NAPHTHENIC ACIDS

Naphthenic acids is a collective name for organic acids present in various crude oils.
Naphthenic acids constitute about 50% by weight of the total acidic components in crude
oil. Naphthenic acids are straight chain alkyl cycloparaffin carboxylic acids as shown
below:

CR2
R2C CR (CH2)n-COOH R = H or an alkyl radical
R2C CR2

The average molecular weights of the naphthenic acids range from 200 - 300. The flash
points may vary from 280 - 300°F. Naphthenic acids are barely soluble in water but are
readily soluble in most organic solvents, particularly hydrocarbons.

The bulk of the naphthenic acids in crude oils generally is found in the gas oil and light
lubricating oil fractions.

Naphthenic acid content of a crude generally is determined by titration with potassium


hydroxide (KOH). This is called the neutralization number. The neutralization number
for dark products is determined using the standard ASTM-D664 which requires a
potentiometric titration with KOH. For light products, the color titration procedure ASTM
D-974 can be used. Neutralization numbers are expressed in milligrams of KOH to
neutralize the acid constituents in a gram sample.

The neutralization number will include phenols and other carboxylic acids in addition to
naphthenic acids. Either method will give:

1) Total Acid Number: KOH required to neutralize all acidic constituents present.

2) Strong Acid Number: KOH required to neutralize the strong acid constituents
present. Carboxylic acids are not included in this value.

The appearances of naphthenic acid corrosion are generally highly characteristic,


varying with increase in stream velocity from sharp-edged, crater-like holes to sharp-
edged streamlined grooves. The corrosion products from iron and copper and their
common alloys are easily dissolved in moving hot oil and washed off, leaving no scale
behind. These facts explain the characteristic appearance of the attacked metal. One is
inclined to think of erosion, whereas actual velocity corrosion has taken place. Evidence
for this velocity corrosion is found in the fact that acid-free, cracked or neutralized oil
does not give any such wear and, also, that a soft material-like aluminum, which does
not react with naphthenic acids, remains practically unaffected. There is some difficulty
of identifying specific corrosion in refinery equipment as to whether this corrosion is due
to high temperature sulfur or to naphthenic acid.

One of the most reliable methods for confirming that corrosion in an existing installation
is due to naphthenic acids would be the use of Type 304 stainless steel corrosion
coupons. High corrosion rates on Type 304 coupons would indicate naphthenic acids; low
corrosion rates would indicate high temperature sulfur corrosion.

Naphthenic acid corrosion is encountered mainly in crude and vacuum distillation units
processing certain sour crudes or derivatives thereof. Type of equipment chiefly subject
to this corrosion includes pumps, return bends and tube inlets to furnaces, transfer lines,
feed and reflux sections of columns, heat exchangers, condensers, and objects which
hinder a smooth stream. Naphthenic acid is generally not a problem in cracking plants.
The main factors that influence the corrosion by naphthenic acids are the following:

Neutralization Number
Temperature of Operation
Velocity and Turbulence
Physical State, Vapor or Liquid Phase

Neutralization Number

Naphthenic acid corrosion only becomes a problem when the neutralization number
exceeds 0.5 mg KOH per gm. Acid values as high as 10 mg KOH per gm have been
reported At a given set of conditions, corrosion increases with increasing naphthenic
acid content (or neutralization number) of a hydrocarbon stream.

Temperature of Operation

The corrosion caused by naphthenic acids is found in the range of 450°F to 750°F. At
higher temperatures the naphthenic acids decompose and at lower temperatures the
corrosion rate is not high enough to cause problems.

Velocity and Turbulence

Fluid velocity has a major impact on the corrosion rate. Often, very severe corrosion is
found at places where a high, turbulent stream - possibly combined with impact - exits,
whereas at neighboring places free from intensive turbulence only little corrosion occurs.
This effect is magnified if the crude contains light products and/or water because
vaporization of the lighter products or steam causes increased velocity. Increased
corrosion rates are generally found at elbows, tees, return bends, and in transfer lines.

Physical State, Vapor or Liquid Phase

The corrosion activity of the naphthenic acid will increase with decreasing molecular
weight. These smaller molecules are more active in the liquid than in the vapor phase
because corrosion only takes place in the liquid phase. Naphthenic acids will become
partially vaporized at temperatures above 500°F, depending on temperature, pressure,
and crude type.
The severity of corrosion appears to be higher when the physical state of the acids is
changing, for instance, in a vaporization situation such as a transfer line where the linear
velocities are higher.

The corrosion activity also increases in a condensing situation because the neutralization
number of the hydrocarbon liquids is increasing because the naphthenic acid vapors are
condensing out. This should occur on a limited number of trays in atmospheric and
vacuum towers of crude distillation units, usually from 550°F to 650°F.

Basically, four methods exist for reducing corrosivity of naphthenic crudes or their
fractions:
Blending
Topping
Neutralization
Appropriate Selection of Metallurgy’s

Blending

Crudes with a high content of naphthenic acids can be blended with a low neutralization
number crude to produce a less corrosive blend.

Topping

Topping removes the naphthenic acids at lower temperatures with lower corrosion rates.

Neutralization

Caustic soda can be added to the crude to neutralize the naphthenic acids before
distillation. This method can, however, result in emulsifier formation and caustic
cracking of carbon steel equipment.

Appropriate Selection of Metallurgy

Austentic type 316 SS is normally used for resistance to naphthenic acid corrosion.

H. NITROGEN COMPOUNDS

The nitrogen compounds that are found in crude oils are generally divided into two types:
basic and non-basic. The non-basic are of the carbazole, indole, and pyrrole type. These
compounds are heterocyclic amines; however, they are non-basic because the nitrogen
is not available for bonding. See Figure 3.5-5. The basic compounds are primarily of
quinoline structure and, as can be judged from their dissociation constants, the strengths
of these bases are such that they can be virtually removed from the oil by acid extraction
Figure 3.5-5
Prototypes of Nitrogen Compounds in Petroleum Distillates

Prototype pKb Structure


N
N
Non-Basic: Pyrrole 13.6
N
N

Indole 14

N
N
Carbazole 15

N
Basic: i-Quinoline 8.9
N

Pyridine 8.8

Note: The actual compounds occurring in petroleum often have


one or more hydrocarbon groups as side chains.

The nitrogen content in crude oil varies greatly with the source of the oil and the greatest
concentrations are found in California petroleums.

Typical nitrogen compound concentrations in crude oil are in the range of 1% for the
lowest and 13% for the highest oils from California.

Nitrogen compounds in crude oil alone do not contribute to a corrosion problem.


However, after decomposition in catalytic cracking, the ammonia and cyanides formed
contribute to problems such as high pH hydrogen blistering in which the cyanides
remove the protective polysulfide scale. This shall be discussed in detail in the “Fluid
Catalytic Cracking Unit Chapter.

In hydrocracking, the ammonia and hydrogen sulfide form ammonium hydrosulfide


which causes serious corrosion problems at temperatures below the water dew point.

The presence of ammonium hydrosulfide in sour water stripping systems can also be a
problem.

Ammonia formed in any cracking or hydrocracking unit limits the use of any copper
alloys in product coolers because of their attack in areas where protective scale may
have spalled. Both ammonia and cyanides attack copper and brasses (and other copper
alloys) in a pitting or worm-holing type attack. With both compounds present, there have
been reported incidents of reversal of anode and cathode where brass is in contact with
steel, with the brass being corroded.

Ammonium chloride and certain organic amine salts have been identified in several
crude unit overhead deposit samples collected above the dew point. These neutralization
salts will cause localized attack beneath the deposit when formed above the dew point.

Nitrogen compounds can raise the pH of the process stream in systems such as crude
units, vacuum units, Fluid Catalytic Cracking Units, hydrocracking units, and sour water
strippers, thereby eliminating the need for a neutralizing amine. Conversely, certain
nitrogen compounds can consume neutralizer such as isocyanate compounds which
have a structure

O=C=N-R

Nitrogen compounds can poison the catalysts used in many cracking operations that are
downstream of the crude unit.

I. AMMONIA

The most common sources of ammonia are breakdown of nitrogenous organic


contaminants. In addition, ammonia is used in the refining industry as a refrigerant and
for the neutralization of acidic components in such locations as the overhead streams
from crude units and cracking units.

Severe corrosion in the form of general metal loss and stress corrosion cracking can
occur when ammonia is permitted to contact copper base alloys in pH ranges of 8.0 and
above. The general attack may be identified by the appearance of a blue salt. Ammonia
attack of copper alloy equipment may result in severe fouling.

Ammonia has been the most commonly used neutralizer because of its strong
neutralizing ability, its availability, and low cost.

A disadvantage of ammonia, however, is that the product formed by its neutralization of


HCI is the crystalline salt, ammonium chloride (NH 4Cl), which is insoluble in hydrocarbon
streams. Solid ammonium chloride deposits can form at temperatures above the water
dew point of overhead streams, thus contributing to fouling deposits.

These salts are highly hygroscopic. The water that is absorbed by these salts contains
acids such as hydrochloric acid and certain organic acids. The acid concentrations
underneath the salt deposit are high, being close to the saturation concentration for each
compound involved. For this reason, localized acid attack can occur beneath the deposit
when formed above the water dew point by the following equations:

NH4Cl + H2O Û NH3 + HCl + H2O


2HCI + Fe  FeCl2 + H2

FeCl2 + H2S  + FeS ß + HCl

This under deposit attack is usually in the form of pitting and eventually leads to
equipment failure.

J. CAUSTIC

Caustic refers to soda ash or sodium hydroxide. Sodium hydroxide is often fed to the
crude unit charge to reduce the neutralization costs at the crude distillation unit.
Hydrochloric acid is produced by the hydrolysis of inorganic salts, primarily magnesium
chloride (MgCl 2) as follows:

MgCl2 + 2H2O  Mg(OH)2 ß + 2HCl Ý

Sodium hydroxide is used as a charge neutralizer to reduce hydrochloric acid evolution


by the following mechanism:
2NaOH + MgCl2  Mg(OH)2 ß + 2 NaCl
The sodium chloride salt that is formed is hydrolytically stable at crude tower flash zone
temperatures and will exit with the bottoms streams from the crude distillation unit.

The usage rate of sodium hydroxide generally exceeds the theoretical requirements.
The exact causes of the inefficiency are not completely known; however, two possible
reasons are: 1) acidic materials in the crude which partially neutralize the caustic, and 2)
the poor dispersibility of caustic in the crude.

Caustic treatment, however, has several disadvantages, including stress corrosion


cracking that occurs in carbon steels which is commonly called "caustic embrittlement".
The term "caustic embrittlement" derives from the fact that failure occurs in the presence
of a highly concentrated caustic solution, and the metal itself fails in a sudden or brittle
manner with no apparent deformation of the metal prior to actual failure. Carbon steel
that is not stress relieved and stainless steel type 304 in, or adjacent to, welds or at other
points of high stress concentration at temperatures over 400°F are subject to this type of
failure.

Other adverse effects of caustic treatment are increased preheat exchanger fouling due
to sodium breaking down and causing polymerization, resulting in reduced heat transfer
and increased fuel usage.

Caustic increases furnace coking and causes tube embrittlement and rupture. Also, in
coker furnaces, sodium lowers the quality of the coke product.

Field studies have shown that roughly 85% of the sodium that is fed to the crude
distillation tower exits with the bottoms or resid stream while approximately 1% to 4% of
the sodium leaves with each side cut. The increased amount of sodium in these streams
due to caustic injection can cause the product fuels to be out of spec.

Sodium poisons the Fluid Catalytic Cracking (FCC) catalyst as well as the
Hydrodesulfurization (HDS) catalyst. The FCC catalyst will lose approximately 0.5%
conversion per each additional 0.5 ppm sodium in the FCC charge or for each additional
500 ppm of sodium on the FCC catalyst.

The HDS catalyst will lose approximately 50% of its activity when the sodium content on
the catalyst exceeds 1 wt % and when the catalyst loses over 70% of its activity, it must
be replaced.

CORROSION THEORY

General Corrosion

The attack by HCl-H2S is pH dependent. In general, four pH ranges may be distinguished. Refer to
Figure 3.5-6.

1. Below 4.5 pH there is a region of severe acid corrosion due to an excess of hydrogen ions.

2. In the pH range of 4.5 to 6.0, because of low concentration of both hydrogen and bisulfide ions,
corrosion rates reach a minimum.

3. The least protective sulfide films are formed at pH 6.0 to 8.0.

4. At a pH level above 8.0, very low corrosion rates occur because of the protective sulfide film.
Figure 3.5-6

Corrosion Rate vs. pH in H2S,


HCl, Water Environment
1000
900
800

Corrosion Rate (MPY)


700
600
500
400
300
200
100
0
0 1 2 3 4 5 6 7 8 9 10
pH

It must be remembered that the pH in a crude overhead system, for example, is measured at the
reflux drum water draw. pH in the bundles themselves may be different. This will be explained later
in this section when “the initial condensation point theory” is discussed.

The importance of closely controlling pH can be better understood by examining Figure 3.5-6. As
pH falls below 4.5, corrosion rates increase dramatically due to the increase in hydrogen ion
concentration. Above a pH of 6, the concentration of bisulfide ions increases, again, increasing the
corrosion rate.

The need for a good filming inhibitor can also be seen in Figure 3.5-6. Although the curve reaches a
minimum between 5 to 6 pH, the system corrosion rate may be unacceptable. A good inhibitor will
reduce this corrosion by 90 to 95 percent.

Mechanism of Acid Dew Point Corrosion

DEW POINT TEMPERATURE

The dew point temperature of a vapor species is the temperature at which the rate of evaporation of
the condensate from a clean metal surface is equal to the rate of condensation of the vapor. The
vapor pressure at that temperature is known as the saturated vapor pressure (SVP) and from a
knowledge of the SVP, the dew point temperature can be deduced from the partial pressure of
various vapor species present in the gas. In nature, the effect is most frequently seen when water
vapor condenses in the air, becomes saturated and is deposited on the leaves of plants. Typical
atmospheric water vapor levels are 0.5 to 1% giving dew point temperatures of 32°F to 50°F. Crude
unit overhead vapors have water vapor levels of 2% to 5% and observed water dew points typically
fall in 170°F to 240°F range.

ACID DEW POINT TEMPERATURE

In the case of the crude overhead vapor stream, the situation is more complicated owing to the
presence of many other condensable constituents in addition to water vapor. Low concentration of
hydrogen chloride cause acid containing condensate (azeotropes) to form at higher temperatures
than would occur with the water vapor alone. Figure 3.5-7 shows the boiling point curve for water
and HCI acid mist. The presence of the azeotrope, which is 20.2% HCl at 1 atmosphere pressure,
helps explain the very high corrosion potential that exists in the “dew point zone”. Corrosion rates of
200 mpy have been observed on some systems.

Figure 3.5-7

Boiling Point of Acid Mist (HCl/H2O)


and Water vs. Pressure
Water Acid Mist (HCl/H2O)
320

300
Temperature (°F)

280

260

240

220

200
0 500 1000 1500 2000 2500 3000 3500
Pressure (mm Hg)

The Importance of pH Profiles

The pH profile can be defined as the pH variation observed as a function of water condensate
temperature. The pH profile in a crude unit overhead condensing system varies greatly depending
upon basic species naturally present, the type of neutralizer used, level and nature of acid species
present, and system operating conditions. It is not uncommon to observe initial water condensate
pH significantly higher or lower than the accumulator water. Sour systems (>100 ppm sulfides) with
low chloride concentrations (>20 ppm) which are using proprietary neutralizers generally have Initial
Water Condensate pH higher than the accumulator water. The difference in pH depends on the
composition of the neutralizer product used and becomes greater with higher neutralizer feedrate
and accumulator pH greater than 6.0. At these conditions, excess neutralizing amine is available to
react with HCI and other strong acids at the Initial Condensate Point (ICP). This requires tailwater
pH control to be maintained below 6.0 on a low chloride high H 2S system.

Maintaining high condensate pH increases corrosion and fouling potential, reduces the
effectiveness of certain filming inhibitors, and increases neutralizer cost.

Sweet Systems (relatively low H2S loading) can have initial condensate pH lower than accumulator
pH when ammonia is used for neutralization. Data from a West Coast refinery revealed initial
condensate pH of 3.0 during a survey when accumulator pH was maintained at 8.0. Good pH
control and knowing the pH profile across the entire condensing system is important in a good
corrosion control program. As mentioned, proper pH levels must be maintained in order to minimize
corrodent levels, reduce fouling tendency, and to minimize neutralizer consumption.

Aggressive Corrosion Above the Initial Condensation Point

Several studies have been conducted which confirm that the initial water condensate point is the
most aggressive corrosion location within a crude unit overhead system. In order to better define
this problem location, surveys were conducted in many crude unit overhead Systems processing a
variety of crude types using standard corrosion control programs

Very high corrosion rates in several systems were observed between the "free" water dew point
temperature up to 70°F above this location. This extended temperature range is now referred to as
the "water dew point region".
Corrosion rates experienced within the water dew point region have been recorded from 200 - 500
mpy (5.08- 12.7 MM/y) without treatment and in poorly treated systems. Laboratory testing of
several untreated field naphtha and overhead accumulator water samples (using naphtha/water
ratios commonly observed in the field) confirms the laboratory baseline corrosion potential is within
this same range. Corrosion potential experienced above the Initial Water Condensation Point (ICP)
can be associated with two possible mechanisms:

1. Salt deposition.
2. Aqueous acid corrosion on cooler metal surfaces.

Ammonium chloride and certain organic amine salts have been identified in several crude unit
overhead deposit samples collected above the ICP. These neutralization salts will cause localized
acid attack beneath the deposit when formed above the ICP. The sublimation temperature of these
salts primarily depend on the partial pressure of ammonia or neutralizer and amount of acids
present in the overhead system.

Computer programs have been developed to theoretically calculate the ammonium chloride
sublimation temperature. This computer calculation in combination with other methods can predict
the impact of ammonium chloride salt deposition above the ICP. Computer water wash
optimization calculations can predict the proper amount of water needed to reduce salt fouling and
corrosion in this "dry" location.

Hot crude unit overhead vapors may be "shock" condensed on the metal surface of the overhead
exchanger equipment. Low skin temperatures will condense small water droplets on the metal
surface along with the low Kva acid groups.

It has been observed that crude unit overhead systems with high acid loading experience the higher
corrosion rates within this region. It has also been determined that high concentrations of HCl,
organic acids, and sulfoxy acid species contribute to the increased corrosion potential found within
this region.

Hydrogen Blistering

Hydrogen blistering is a phenomenon resulting from hydrogen atoms becoming trapped within the
boundaries of mild steel vessel. The hydrogen builds up to pressures as high as several thousand
atmospheres which ultimately blister or laminate the vessel. The blisters can be very tiny and
invisible to the naked eye, or as large as several inches in diameter. Usually, hydrogen blistering is
evidenced with mild steel. High strength steels are subject to hydrogen attack but because of their
low ductility, rupture rather than blister.

The hydrogen molecule (H2) will not penetrate steel. However, atomic hydrogen can pass through
the grain boundaries of steel. It can pass completely through a vessel unless it is stopped by a
discontinuity or inclusion in the metal. At such irregularities in the wall of the vessel, the atomic
hydrogen recombines to form molecular hydrogen and hence is trapped. As more and more
molecular hydrogen is trapped, the pressure increases until a blister is formed. Hydrogen blistering
can occur in both acid and alkali conditions. However, this discussion will be confined to alkaline
conditions such as occur in catalytic cracker vapor recovery units.

THERE ARE SIX CONDITIONS NEEDED FOR HYDROGEN BLISTERING:

1. A steel surface subject to corrosion.

2. Conditions which permit the existence of a liquid water phase.


3. A source of reactive hydrogen. Hydrogen sulfide in the presence of water will corrode steel
as depicted by reaction A shown below. Atomic hydrogen is one of the products of this
corrosion reaction.

A) Fe° + H2 Û FeS + 2H°


4. A poisoning agent. Normally, the atomic hydrogen that is formed from corrosion is readily
converted to molecular hydrogen which leaves the reaction site as a gas. However, in the
presence of sufficient quantities of H2S cyanides and other sulfur compounds, reaction B
shifts to the left. The hydrogen remains as atomic hydrogen and is free to enter the vessel
wall.

B) 2H° Û H2

5. A promoter to maintain an active surface. High pH and the presence of cyanides promote
hydrogen blistering.

a. Cyanide ions dissolve the iron sulfide scale exposing bare metal to corrosion. This
is shown in reaction C.

(C) FeS + 6CN Û Fe(CN)6-4 + S-2

Moreover, cyanide ions can act directly with iron to form the soluble iron cyanide
complex and produce still more atomic hydrogen. Thus, a snowballing effect
occurs.

(D) Fe + 2H+1 + 6CN (presence of S-2) Û Fe(CN)6-4 + 2H°.

b. The higher the pH, the greater the potential for hydrogen blistering as shown in
Figure 3.5-8. As pH increases, the solubility of H2S in the water phase increases
due to the solubilizing effect of the alkali (such as ammonia). Therefore, H 2S is
readily transferred from the voluminous hydrocarbon phase into the water phase.
This makes the water more aggressive.
Figure 3.5-8

Hydrogen Blister Potential vs. pH


Hydrogen Penetration

7 7.5 8
pH
6. Imperfections in the steel. Any inclusions, voids, or discontinuities in the steel surface
causes the atomic hydrogen, which has penetrated the metal, to be converted to
molecular hydrogen.

Vous aimerez peut-être aussi