You are on page 1of 79

200 Corrosive Environments

Author: R.I. (Bob) Piehl

Abstract
This section discusses corrosion caused by corrosive environments, and recom-
mended metallurgy. Section 210 covers general and localized corrosion of steel,
stainless steels, nickel alloys, copper alloys, titanium, and aluminum alloys in sea
water and brine. It also covers internal and external corrosion and specifically
discusses offshore platform corrosion. Section 220 discusses corrosion by cooling
water.

Contents Page

210 Corrosion by Sea Water and Brine 200-2


211 Introduction
212 Types of Corrosion
213 Corrosion of Equipment and Facilities
214 Corrosion of Carbon Steel
215 Corrosion of Other Alloys
220 Corrosion by Cooling Water 200-40
221 Typical and Ideal Cooling Water System
222 Make-up Water
223 Blowdown Control and Cycles of Concentration
224 Fouling, Scale and Corrosion
225 Biocide Monitoring and Control
226 Water Quality Monitoring: How to Read a Water Service Report
227 Chemical Treatment Program Control: Monitoring and Evaluation
228 Glossary
230 References 200-78

Chevron Corporation 200-1 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

210 Corrosion by Sea Water and Brine

211 Introduction
Sea water is among the worlds most common corrosives. For most kinds of struc-
tures and equipment, ordinary carbon steel is the economical construction material
despite a moderate corrosion rate in sea water. We accept this when we specify
carbon steel for salt water lines, wharf piling, or jacket legs on offshore structures.
Sea water is an aqueous solution containing many different dissolved salts in
varying concentrations. Typical sea water composition is given in Figure 200-1.

Fig. 200-1 Chemical Composition of Sea Water


Principal Constituents of Sea Water
Constituent Amount, ppm
Chloride 18,980
Sulfate 2,649
Bicarbonate 140
Bromide 65
Fluoride 1
Boric acid 26
Sodium 10,556
Magnesium 1,272
Calcium 400
Potassium 380
Strontium 13
Simplified Sea Water Composition (From British Standard 1170)
Sodium chloride 22,620 ppm
Magnesium chloride 3,300 ppm
Magnesium sulfate 1,960 ppm
Calcium sulfate 1,220 ppm
Calcium bicarbonate 180 ppm
Specific gravity 1.025
pH 7.6-7.8

Sea water corrosion rates are surprisingly uniform around the world, despite differ-
ences in temperature and salinity. It has been suggested that the controlling vari-
ables may change in compensating ways. For example, higher temperatures promote

May 2001 200-2 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

faster reaction rates, but they decrease oxygen solubility and increase calcareous
deposits which act as a partial barrier to the corrosion process.
Synthetic sea water, formulated from pure chemicals, is less corrosive than natural
sea water. This is attributable to the presence of marine organisms, including slime
and bacteria, in the natural water.
Brines may be more or less corrosive than sea water, depending on their chemical
composition. Produced waters are regarded as roughly similar to sea water.
However, the wide range of salinities and the presence of acid gases may greatly
alter corrosivity. Geothermal brines vary even more widely in corrosivity. Produced
waters and brines are usually oxygen-free when they come from the ground. This
may greatly reduce their corrosivity.
Corrosion in sea water or brine can take many different forms. Figure 200-2 shows
the types of corrosion that different alloys can suffer. For further information refer
to the paragraphs in Section 210 on the appropriate alloy or type of corrosion.

Fig. 200-2 Corrosion Mechanisms for Various Materials in Sea Water


Metal or Alloy
Type of Corrosion Carbon Stainless Nickel Copper
Steel Cast Iron Steels Alloys Alloys Titanium Aluminum
General Corrosion X X X
(1) (2)
Pitting & Crevice Cor. X X X X
(3)
Stress Cor. Cracking
Intergranular Cor. X
End Impingement X
Graphitization X
Dezincification X
Note See text on individual alloys for details.

(1) Some nickel alloys pit at low velocity, or suffer crevice corrosion.
(2) Titanium pits over 250F.
(3) Stainless steels may stress-corrosion crack over 150F.

212 Types of Corrosion


General Corrosion and Pitting
Corrosion in sea water and brines will occur through some combination of general
(uniform) corrosion and pitting. Copper alloys in low velocity sea water primarily
suffer general corrosion, but very little pitting. Carbon steel experiences both
general attack and pitting, with the pitting occurring at rates several times as high as
the general loss. Stainless steels corrode almost exclusively by pitting and suffer
little general corrosion. Aluminum, which suffers severe corrosion of both types,
should be avoided in sea water service.

Chevron Corporation 200-3 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

When corrosion rates are quoted in published sources, a clear distinction is not
always made between general corrosion and pitting. For example, 5 mils per year is
a commonly quoted corrosion rate for carbon steel in ambient temperature sea
water. However, this is the rate of general corrosion. Pitting is likely to be consider-
ably fasterby a factor as high as 5 or 6.
Whether general corrosion or pitting is more important depends on the application.
In a pipe that carries liquid, perforation is the obvious criterion and pitting rate is the
more important. In load-carrying structural members, such as wharf piling, the
average loss is important, not the maximum pit depth.
Figures 200-3 and 200-4 present some general information on relative pitting resis-
tance of various alloys, where the alloys can be used and the effectiveness of
cathodic protection. Figure 200-3 focuses on material selection as a function of
equipment design (i.e., whether or not crevices are built into the design).
Figure 200-4 addresses the effect of flow conditions on material selection.

Stress Corrosion Cracking


Sudden and severe failures have resulted from stress corrosion cracking, sometimes
in unexpected locations. The austenitic stainless steels, commonly called 300
Series or chrome-nickel stainless steels, have the potential for stress corrosion
cracking in sea water. If sensitized by welding or heat treatment, they can fail, albeit
in rare cases, at temperatures as low as U.S. coastal water temperature. Nonsensi-
tized stainless steels are generally considered resistant to chloride stress corrosion
cracking up to 140F. See Section 400 regarding sensitization and stress corrosion
cracking, respectively.
High-strength carbon and low alloy steels can fail by stress corrosion cracking in
sea water, but they are susceptible only when tensile strength exceeds about
180 ksiequivalent to a hardness of Rockwell C40 (also written HRC 40). For
critical structures in marine environments, a lower and more conservative threshold
value of 148 ksi (equivalent to HRC 34 or 320 Vickers) is commonly specified.
For welded structures, 320 Vickers is usually specified as a maximum allowable
heat-affected zone hardness for stress corrosion cracking resistance. For critical
structures such as offshore platforms and subsea pipelines, the question of strength
and hardness limits requires careful review. Consult the ETD Materials Division for
help in this area.

Impingement
Impingement is a common and important form of corrosion on copper alloys in
applications involving high velocity sea water, such as piping, valves, and heat
exchanger tubing. It causes preferential thinning, commonly indicated by small,
horseshoe-shaped marks, in turbulent areas such as the inlet end of a tube. These
horseshoe tracks always lead upstream.
Additional information on impingement is presented later, under Heat
Exchangers.

May 2001 200-4 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-3 Resistance to Pitting Under Fouling and to Crevice Corrosion in Sea Water (Courtesy of LaQue Center for
Corrosion Technology, Inc.)
Degree of Resistance Alloy Remarks
Crevices can normally Titanium These metals foul but rarely pit.
be tolerated in designs Alloy C Titanium will pit at temperatures above 250F.
using these materials
Alloy 625 Alloy 625 after 2 to 3 years shows signs of incipient
pitting in some tests in quiet sea water.
90/10 copper-nickel Shallow to no pitting.
Admiralty brass 90/10 copper-nickel is standard sea water piping alloy.
70/30 copper-nickel Good resistance to pitting.
Copper Useful in piping applications.
Tin and aluminum bronzes
Austenitic nickel cast iron
Cathodic protection Nickel-copper alloy 400 Pits tend to be self-limiting in depth at about 1/16 inch.
required on critical No protection required for heavy sections.
surfaces
Cathodic protection from steel or copper base alloys
will prevent pitting on O-ring, valve seat, and similar
critical surfaces.
CN7M (Alloy 20) Occasional deep pits will develop.
Alloy 825 Protection not normally required for all Alloy 20 pumps.
Cathodic protection from less noble alloys may be
necessary for O-ring and similar critical surfaces.
Type 316 stainless steel Cathodic protection from zinc, aluminum, or steel is
required except when part is frequently removed from
sea water and thoroughly cleaned.
Crevices cannot be Nickel Many deep pits develop.
tolerated in designs Cathodic protection from less noble alloys required.
(Excellent, however, in
above-the-waterline Type 304 stainless steel Many deep pits develop.
marine applications) Cathodic protection from steel may not be fully effec-
tive.
Precipitation hardened Many deep pits develop.
grades of stainless steel Cathodic protection with zinc or aluminum may induce
cracking from hydrogen.
Severe crevice corro- Type 303 stainless steel Severe pitting.
sion limits usefulness Cathodic protection may not be effective.
Series 400 stainless steel Severe pitting.
Cathodic protection with zinc or aluminum may induce
cracking from hydrogen.

Chevron Corporation 200-5 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-4 Comparative Localized Attack as a Function of Flow Conditions (Used with permission from MACHINE
DESIGN, 1-18-68. A Penton Publication.)

Graphitization
Graphitization is the usual form of corrosion of gray cast iron in salt water service.
The carbon in gray cast iron is in the form of small graphite flakes. As corrosion
occurs, the iron goes into solution, leaving the graphite behind. The result is a mass
of soft carbon in the corroded area that looks like solid metal at first glance but has
little strength. The carbon can readily be scraped away.
Graphitization may either be uniform or form deep pockets or pits. The graphite
layer acts as a cathode, setting up a galvanic cell with the uncorroded metal. This
tends to accelerate attack, but, on the other hand, the relative impermeability of the
graphite tends to slow down the attack. As a result, corrosion rates of cast iron in
sea water are not much different from carbon steel.
The graphite layer on cast iron can also affect galvanic attack. Apparently the iron-
graphite galvanic couple is strong enough that adding a metal of intermediate poten-
tial has little further effect.

May 2001 200-6 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Dezincification
Dezincification is a corrosion phenomenon associated only with brasses. It occurs
readily in salt water. The net effect of dezincification is selective dissolving of the
zinc, leaving in its place a porous, low-strength mass of copper. Actually, both
copper and zinc go into solution, and the copper reprecipitates. Dezincification may
occur uniformly over the entire surface (layer-type), or selectively in small spots
(plug-type).
Dezincification can be controlled by adding small quantities of arsenic, phosphorus,
or antimony to the alloy. Brasses containing these elements are said to be
inhibited.
Other copper alloys can suffer forms of corrosion similar to dezincification.
Aluminum bronzes may selectively lose aluminum, and copper-nickel alloys may
lose the nickel. In all cases, the rate of damage increases with temperature.

Crevice Corrosion
Crevices are tight, stagnant areas where accelerated corrosion can occur. Crevice
corrosion is similar to pitting, except that the pits initiate in crevices. They usually
form as a result of oxygen concentration cells (see Section 100). On susceptible
alloys, crevice corrosion can be reduced, though seldom eliminated, by careful
equipment design that avoids crevices. This is discussed in Section 215 under Stain-
less Steels. Other alloys, including carbon steel and Monel, are less susceptible to
crevice corrosion, as shown in Figures 200-3 and 200-4.

Galvanic Corrosion
Galvanic corrosion is a common problem in sea water, because of the high elec-
trical conductivity of the fluid. Figure 200-5 shows the corrosion potentials of
various alloys in sea water. Those with the highest (most positive) potentials act as
cathodes, and will be protected by contact with a material of lower (more negative)
potential, which will act as anodes and corrode sacrificially.
General rules for dealing with galvanic corrosion include:
Where possible make all parts of the same material.
When using different materials pick those closest in potential.
The more noble (higher potential) metal should be the one with the least surface
area. For example, use the more noble material for valve trim and the less noble
(lower potential) material for valve body, never the other way around. Simi-
larly, fasteners such as nuts, bolts and screws should be more noble than the
materials being joined.
In Section 100, Figure 100-7 illustrated the galvanic compatibilities of some
common alloy combinations. Though the chart was originally intended to apply to
fasteners, it equally applies to other uses, so long as base metal is understood to
mean the material which exists in large surface area, and fastener is the material
which exists only as a small surface area.

Chevron Corporation 200-7 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-5 Corrosion Potentials in Flowing Sea Water (Courtesy of LaQue Center for Corrosion Technology, Inc.)

May 2001 200-8 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Carbon steel can also suffer galvanic attack due to the presence of mill scale. The
mill scale is cathodic to the bare metal and accelerates attack at bare spots and
breaks in the scale layer. As the percentage of bare metal in a mill-scale-coated
surface increases, the rate of corrosion decreases [5] as shown in Figure 200-6.

Fig. 200-6 Corrosion Rate of Base Metal Increases as Percent of Bare Area Increases
Corrosion Rate in Bare Area, mpy (at
% Bare Metal Area ambient temperature)
5 45
10 35
25 15

The accelerating effect of mill scale is greatest during the first month or two of
exposure, and eventually decreases.
Several instances have been reported in which new storage tanks have suffered
1/32-inch to 1/16-inch pits in about 6 weeks, during initial water tests using sea
water. The worst pitting incident occurred in spots of bright metal left by local
grinding just prior to the water test.

Intergranular Corrosion
Austenitic stainless steels show some tendency toward intergranular corrosion in sea
water. For this reason, low-carbon or stabilized grades should be used for parts
subject to welding or heat treatment. Sensitization, which leads to intergranular
corrosion, can also make the metal less resistant to stress corrosion cracking, and in
some applications, such as pumps, it can accelerate corrosion fatigue.

Fouling
Fouling by marine organisms can be a serious problem in sea water. In piping,
marine growth can seriously impede water flow by increasing surface friction and
reducing effective line diameter. Fouling of ships bottoms reduces maximum speed
and increases fuel consumption. On the legs of offshore platforms, marine organ-
isms increase the effective area exposed to wave forces, thereby increasing stresses
on structural members. Common techniques to prevent or retard marine growth
include biocides, anti-fouling paints, or use of fouling-resistant materials.
Copper and copper-base alloys are inherently resistant to biofouling because slow
corrosion creates an environment high in copper ions, which are toxic to the organ-
isms. Pure copper has the best fouling resistance; adding alloying elements to the
copper substantially reduces resistance. Figure 200-7 shows relative fouling resis-
tance of some common alloys in quiet sea water. Above 3 feet per second contin-
uous velocityabout 1.8 knotsfouling organisms have increasing difficulty in
attaching themselves and clinging to the surface, unless they are already attached
securely.
Cathodically protected copper alloys have no fouling resistance because formation
of toxic copper ions is prevented. Attaching a copper alloy to steel without an

Chevron Corporation 200-9 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-7 Fouling Resistance in Quiet Sea Water (Courtesy of LaQue Center for Corrosion Technology, Inc.)
Arbitrary Rating Scale of Fouling Resistance Materials
90-100 Best Copper,
90/10 copper-nickel alloy
70-90 Good Brass and bronze
50 Fair 70/30 copper-nickel alloy
10 Very slight Nickel-copper alloy 400
0 Least Carbon and low alloy steels, stainless
steels, nickel-chromium-high molybdenum
alloys, Titanium
Note Above 3 ft. per sec. continuous velocityabout 1.8 knotsfouling organisms have increasing difficulty in attaching themselves and
clinging to the surface, unless already attached securely.

insulating layer between the two metals cathodically protects the copper alloy and
simultaneously destroys its fouling resistance. This is especially a problem on ships
and offshore platforms where cathodic protection is often used.

Corrosion Fatigue
Corrosion fatigue in sea water has been studied extensively in recent years, espe-
cially in connection with the design of offshore drilling structures in the North Sea.
Corrosion fatigue is also an important cause of pump impeller failures in saline
water service.
Normally carbon steel possesses an endurance limit. This is a stress level below
which fatigue failure will never occur, regardless of the number of cycles to which
the material is subjected.
The endurance limit depends on both material and environment. In sea water, the
endurance limit essentially disappears; the stress required to cause failure continu-
ously decreases in relation to the number of stress cycles. Deaeration of the water,
or application of cathodic protection, causes reappearance of the endurance limit.

213 Corrosion of Equipment and Facilities


Piping
Carbon Steel Piping. In sea water service Corporation Piping Standards call for
Schedule 80 carbon steel pipe in 2-inch to 4-inch pipe sizes, and extra-strong red
brass for 1-inch and smaller pipe. Valves in brass piping are typically specified as
ASTM B62, which is an 85-5-5-5 Cu-Sn-Pb-Zn bronze (Alloy C83600). Valves in
bare or cement-lined steel piping are typically cast iron with bronze trim.
In noncritical applications where a relatively short life is acceptable, carbon steel
can be used for ambient temperature sea water piping. Life predictions can be based
on an assumed corrosion and pitting rate of 35 mpy. Since penetration is mostly by
pitting, it is often reasonable to use the entire pipe wall as a pitting allowance. In
nominal pressure applications, it will leak before rupture. It is not unusual to use

May 2001 200-10 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

carbon steel for noncritical, small diameter lines. Schedule 80, -inch pipe should
last about five years before it leaks.
Carbon steel is often used successfully in oxygen-free sea water and brine. One
common application is fire water lines. If the lines are used only for fires, corrosion
is minimal because oxygen in the stagnant water is rapidly consumed in minor
rusting of the steel that stops when the oxygen is gone. However, if the fire lines are
used for purposes other than fires, such as wash-downs, more frequent replenish-
ment of oxygen and accelerated corrosion result.
Carbon steel lines are also used for sea water and for brine injection in oil fields.
Here, good deaeration and oxygen scavenging are essential if good service life is to
be obtained. In large, complex systems care and ingenuity are needed to prevent or
locate points of oxygen entry.
Carbon steel piping has suffered rapid pitting at flange faces having spiral-wound
gaskets. This may result from galvanic attack at a cell established between the steel
and the metal spirals of the gasket.
Cement-lined Steel Piping. This piping is widely used in sea water service. When
failures occur, they are generally the result of flaws in the cement lining, usually at
welded joints or branch connections. Reliability of cement-lined piping is much
greater when line diameter is large enough to permit internal access for joint
patching. The presence of frequent branch connections also reduces reliability.
Cement-lined piping is generally not used in piping smaller than 3 inches in
diameter.
Cast Iron Piping. In the past, cast iron has sometimes been used for large diameter
buried salt water lines. The corrosion rate of cast iron is not greatly different from
steel; however, service life may be quite long simply because the lines are thick. In
any event, leakage may not become apparent for years because the line is buried and
because pits may be partially or completely plugged with graphite residue from the
cast iron.
Cupro-nickel Piping. Where long life and high reliability are required, 90-10
cupro-nickel is a popular piping material. Its high resistance to corrosion allows use
of relatively thin walls. The resulting weight saving is an advantage in shipboard
use and on offshore platforms. The alloy is readily welded. Connections can either
be welded or use 90-10 stub ends and steel lap joint flanges (providing that the steel
flange is not water-immersed). Weld neck flanges of 90-10 are not readily avail-
able, and attempts to use red brass flanges have resulted in rapid galvanic attack of
the brass. Maximum recommended velocities for 90-10 are:
Diameter, in. Velocity, fps
to 1 5
1 to 3 6
4 to 10 8
12 to 36 10

Chevron Corporation 200-11 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Nickel aluminum bronze is becoming a popular valve material with 90-10 piping.
There is no significant potential difference between the two, minimizing the danger
of galvanic attack. In contrast, most brasses and bronzes are anodic to 90-10 and
valves of these materials would be preferentially corroded.
70-30 cupro-nickel is also a good piping material for sea water, though it is more
costly and has slightly less resistance to fouling. 70-30 can tolerate velocities about
2 fps higher than 90-10.
Austenitic Stainless Steels. These materials are seldom used for sea water piping
because of their tendency to pit. There are a few exceptions to this, as discussed
later under Desalination Plants.
Plastic Piping. The commonly available plastic piping materials are all immune to
corrosion by sea water, so selection of a particular plastic type is governed by such
factors as cost, temperature-pressure ratings, and mechanical integrity rather than by
chemical resistance. For larger line diameters, FRP is the most common. For small,
noncritical piping such as brine lines at zeolite water softeners, PVC is often used.
However, when handling oil-containing produced water or refinery waste water,
selection of a plastic with adequate oil resistance becomes important.
A recent application of plastic pipe in saline water service is a 24-inch diameter,
custom-fabricated, glass-reinforced vinyl ester line at El Segundo, installed as part
of the 1985-86 Effluent Diversion Project.

Heat Exchangers
The Richmond, El Segundo, and Perth Amboy Refineries have used sea water for
cooling in the older parts of the refineries. Until the 1970s, copper-base alloys were
used almost exclusively for heat exchanger tubing. Admiralty brass, aluminum
bronze, and 70-30 cupro-nickel were commonly used. In Richmonds experience, on
average, aluminum bronze lasted 1 times as long as admiralty brass, and
70-30 cupro-nickel lasted twice as long. The average life of 16 gage admiralty brass
tubes at Richmond Refinery was four years, according to the most recent survey.
Tubes lasted longer in coolers than they did in condensersa 4.5-year average
compared to 2.7 years.
In 1955, a detailed study was made of exchanger tube experience at Richmond to
identify conditions that led to premature tube failure. Based on this study, the limits
given in Figure 200-8 were placed on the common tube alloys when used in sea
water service.
The 1955 survey data still provide reasonable limits for design of exchangers in sea
water service. To avoid excessive tube wall temperatures, water velocity should not
be less than about 3 feet per second.
The most common mode of failure of the copper-base alloys is end impingement.
The usual cause of end impingement failure is excessive velocity. However, throt-
tling the salt water flow at the exchanger inlet rather than at the outlet is, in many
cases, a major contributing factor. The resulting pressure drop causes dissolved air
to come out of solution, and the cavitation effect at the tube inlet adds to the end
impingement problem.

May 2001 200-12 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-8 Common Tube AlloysSea Water Service Limitations


Shell Inlet Tube Outlet Avg. Heat Water
Temperature, F Temperature, F Density, Btu/hr.Ft.2 Velocity, fps
Admiralty brass:
coolers 300 120 3000 5
condensers 200 120 5000 5
Aluminum bronze:
coolers 400(1) 120 5000(1) 6
condensers 300(1) 120 7000(1) 6
70-30 Cupro-nickel:
coolers 300 120 3000 6
condensers 200 120 5000 6
(1) See text regarding validity of these data.

A common remedy for end impingement is installation of plastic or titanium


ferrules in the inlet ends of the tubes. While ferrules do not generally decrease the
rate of corrosion, when used in tubes which have already suffered attack the ferrules
cover the corroded area, and attack starts again on uncorroded metal downstream of
the ferrule. Thus, the life of the tube is extended.
It has been common practice to use naval-rolled brass (NRB) tube sheets with admi-
ralty or aluminum bronze tubes, and Monel or 70-30 cupro-nickel tube sheets with
cupro-nickel tubes. Using the wrong tube sheet material causes rapid corrosion. For
example, accelerated end impingement is likely to occur if admiralty or aluminum
bronze tubes are used in Monel tube sheets. Use of cupro-nickel tubes in naval-
rolled brass tube sheets results in severe dezincification of the tube sheet. End
impingement of cupro-nickel tubes is also accelerated by Monel tube sheets, but the
effect is not great enough to make this an undesirable materials combination. Past
attempts to use Monel-clad steel tube sheets were generally unsuccessful; sea water
seeping into the crevice between tube and tube sheet caused severe galvanic corro-
sion of the steel at the Monel-steel junction.
Heat exchanger channel sections for sea water service have usually been constructed
of Monel, using Monel-overlaid carbon steel flanges. Service life has been
outstanding, though there has been some concern over the years that the galvanic
effect might accelerate end impingement of copper alloy tubes.
In the early 1970s, the low cost and ready availability of welded titanium tubing
made it a natural choice for exchangers cooled with salt water. Early experience was
so successful that titanium became the preferred choice of tube material. Titanium
proved to be essentially immune to end impingement. Its one limitation was rapid
pitting where sea water temperatures were high. Most titanium tube failures were, in
fact, associated with water-side salt deposits due to water boil-out.
In most exchangers, titanium has been installed as a replacement for less resistant
alloys, reusing the original tube sheet material. This has usually been successful but
occasionally tube sheet corrosion or dezincification has been accelerated. Use of
solid titanium tube sheets will prevent this, but it is unlikely that the incidence of

Chevron Corporation 200-13 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

damage is high enough to justify discarding otherwise good tube sheets when
upgrading to titanium tubes.
A few other heat exchanger tube materials have been used in Company plants, with
poor results. A few aluminum brass bundles have been installed over the years.
Service life has been poor, because of a high incidence of stock-side environmental
cracking, rather than water-side attack. Since Aramco successfully used aluminum
brass tubes (from different suppliers) for many years, tube manufacturing problems
are the suspected cause. Their tubes were obtained from different suppliers.
Shortly after World War II, Type 316 stainless steel tubes were installed in some
salt-water-cooled heat exchangers in a chemical plant at Richmond to remedy
process-side corrosion. It was found that rapid pitting of Type 316 from the salt
water side could be prevented by opening the exchangers monthly and cleaning the
insides of the tubes with a power wire brush. Without this regular cleaning, failures
were rapid.
Monel tubes have also been tried in Company plants, with little success. General
corrosion was negligible, and Monel proved almost immune to end impingement.
However, isolated but rapid failures occurred from lodgement attack.

Pumps
The Pump Manual contains detailed recommendations for centrifugal pump mate-
rials in sea water service. Several different materials are specified, depending on
service conditions. For pumps in continuous service, Class D-2 (18-8 stainless steel)
is the preferred material. The austenitic stainless steels are essentially immune to sea
water corrosion under turbulent conditions, but they pit rapidly when the water is
stagnant. Hence, Class D-2 pumps should not be used for intermittent duty. If they
are to stand idle for more than a week or two, they should be flushed with fresh
water first.
Pump Class C-1 (bronze) is intended for intermittent duty pumps, such as fire
pumps, that may stand idle for long periods. Bronze will not pit under stagnant
conditions but has far less resistance than stainless steels to high turbulence. Most
grades of bronze will not last more than about a year in continuous service, though
service life will be long if use is occasional.
In the past, very large Class D-1 pumps have sometimes been used. Class D-1 is
similar to Class D-2, except that the case is carbon steel with a Lithgow lining. Cost
savings were significant; however, availability has been a problem in recent years,
and acceptable substitutes have not been found. Lithgow is a baked phenolic mate-
rial, like the Formica used for kitchen counters except thicker. It is both corrosion-
resistant and wear-resistant. Thin-film coatings such as epoxy have occasionally
been substituted for Lithgow, but they wear away rapidly.
Suppliers of submerged pumps often recommend epoxy-coated steel columns in an
effort to save money. These have typically shown poor life and required frequent
replacement. Fiberglass columns are a far better choice and are usually available in
diameters up to 12 inches.

May 2001 200-14 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

In some instances pumps are in intermittent service but have a high operating factor.
Such services require a material with better erosion-corrosion resistance than the
common bronzes, but better pitting resistance under stagnant conditions than stain-
less steels. Monel can be used, but C95800 or C95500 grade nickel aluminum
bronze is the preferred choice. C95500 is sometimes subject to surface dealloying.
Nickel aluminum bronze is a good choice for submerged pumps where the interior
of the bowl is subject to high velocity and the exterior sees stagnant conditions.
As a general rule, sea water pump impellers should be constructed either of a more
noble material than the case, or of the same material. Do not use materials that will
allow the case to cause galvanic corrosion of the impeller.
Laboratory tests and field experience show that alloys can be classified in four
groups, corresponding to their resistance to cavitation and erosion-corrosion in sea
water, as follows:

Group I (best)
1. Titanium
2. Austenitic stainless steels
3. Ni-Cr-Mo alloys (Hastelloy C, Inconel 625)

Group II
1. Monel
2. Nickel aluminum bronze

Group III
1. 70-30 cupro-nickel
2. G and M bronzes
3. Ni-Resist

Group IV (worst)
1. Carbon steel
2. Cast iron
3. Aluminum
Oil field water injection is a specialized pump application in sea water and brine
service. The water being injected may be more or less salty than sea water, is gener-
ally deaerated at least nominally, and may contain sand and scale. Type 316 stain-
less is the closest thing to an industry standard for this service, though other
materials have occasionally been used. Despite the supposed deaeration of the
waters, carbon steel cases have typically exhibited poor service life. CA-6NM
(13 Cr-4 Ni) impellers have also not been successful, with failures occurring from
corrosion fatigue.

Chevron Corporation 200-15 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Though Type 316 stainless usually performs well in this service, sand and scale in
the water may cause enough corrosion and abrasion to necessitate frequent mainte-
nance. In one instance at Inglewood, California, the Company achieved a marked
reduction in maintenance by using a cast titanium pump.

Desalination Plants
The Companys first desalination plants were the two 880 gpm MSF (multistage
flash) plants at Borco Refinery. The first of these was installed during initial
refinery construction (1970), and suffered severe corrosion until materials were
upgraded in 1975. A second desalination plant was built in 1974, incorporating
lessons learned from the first. It suffered far less corrosion.
MSF plants produce fresh water by heating sea water to about 250F and allowing it
to flash. Steam is condensed; this is the product water. Flashed brine is returned to
the ocean. Early desalination plants like Borco #1 utilized carbon steel for much of
the equipment, on the theory that careful deaeration of the sea water would render it
noncorrosive. This proved optimistic. Inability to maintain a very low oxygen level
(estimated to be 20 parts per billion or less) resulted in rapid internal corrosion of
the steel water boxes. The resulting water leakage caused severe external corrosion
of the evaporator boxes and of structural steel throughout the plant. This was even-
tually corrected by rebuilding the evaporator boxes, installing Type 316L stainless
liners on the floors, and lining the walls with Precrete, a proprietary non-portland
cement concrete.
Tubes in the first Borco plant were 90-10 cupro-nickel. These suffered slow internal
(feed water side) corrosion because of the presence of the iron-oxidizing bacteria
crenothrix polyspora in the feed water, which came from wells drilled into the coral.
These wells became badly contaminated with bacteria. In addition to corrosion, the
bacteria caused heavy ferric hydroxide slime deposits to form, necessitating
frequent shutdowns for cleaning. The tube corrosion problem was eventually solved
by retubing with titanium. The fouling problem was never effectively solved.
Worldwide MSF desalination plant experience shows that 90-10 cupro-nickel tubes
perform well in a majority of plants, but corrode in some plants where H2S is
present in polluted sea water feed. Carbon steel evaporator boxes have proven
almost invariably to be unsuitable because of the difficulty in maintaining a low
enough oxygen content in the feed water. Type 316L stainless, which readily under-
goes stress corrosion cracking in hot, aerated sea water, has been relatively free of
chloride cracking in MSF desalination plants. Evidently the deaeration achieved has
been sufficient to retard chloride cracking of stainless steel, though not to suppress
general corrosion of carbon steel.
A vapor compression type of desalination plant has been used on some of the
Company's offshore platforms to provide fresh water for both drilling and drinking.
In contrast to MSF plants, which provide for very large water demands, these small,
compact plants are intended for use where water demand is nominal. Vapor
compression plants typically have aluminum bronze tubes and 90-10 cupro-nickel
piping and water boxes. On Chevron Platform Grace a vapor compression unit
suffered severe corrosion because of a malfunction in the sulfuric acid feed system

May 2001 200-16 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

(acid addition to the feed water is required to control scaling). Although desalina-
tion plants on offshore platforms must be able to function unattended, it is question-
able whether corrosion can be adequately controlled in vapor compression units
operating in this service.
The use of reverse osmosis (R.O.) desalination plants for sea water desalination is a
relatively recent innovation, though they have been used for some time to desali-
nate brackish water. The biggest operating cost in R.O. plants is pumping energy,
and the pressure at which the water must be pumped is a direct function of its
salinity. To force sea water through an R.O. membrane, a pressure on the order of
1000 psi is needed to overcome osmotic pressure.
The reverse osmosis membranes, of which there may be hundreds in a large plant,
are housed in individual fiberglass canisters. The feed pumps that supply water to
the membranes are typically Type 316 stainless. The piping system is complicated,
providing many crevices for pit-sensitive materials. Product water, which is at low
pressure, typically has plastic piping. The high pressure feed water piping is gener-
ally metallic.
Since the reverse osmosis desalination process was first applied to brackish water
rather than sea water, Type 316L stainless became the standard material for feed
water piping and performed reasonably well. Type 316L continued to be used in the
early sea water plants, but less successfully. Feed water to an R.O. plant is fully
aerated, and 316L has been observed to pit in aerated sea water at rates of up to
1.5 mm (60 mils) per month. Pitting most often occurs at crevices, and plants are
designed to eliminate crevices as much as possible. However, this can only be
partially achieved because connections to the individual R.O. cells must be
demountable, and require O-ring compression fittings or threaded connections. To
eliminate or reduce the Type 316L pitting problem some newer plants use alloys
with higher chromium and molybdenum content in the feed water system. The
Companys R.O. plant at Gaviota uses 254 SMO (20 Cr-18 Ni-6 Mo) for feed water
piping.

Geothermal Sites
Geothermal brines vary widely in composition and corrosivity. Some, such as those
at the Companys Beowawe (Nevada) geothermal site, are low in salinity (1100 ppm
TDS) and high in pH (9.5), and therefore are relatively noncorrosive. Brines at
Heber (California) are more saline (15,000 ppm TDS) and approximately neutral in
pH. The Heber brines are totally oxygen-free as they come from the wells, and are
noncorrosive to carbon steel as long as oxygen entry is prevented.
In general, the geothermal sites that are being utilized commercially are those where
the produced fluids are noncorrosive to carbon steel, either inherently or by the
careful exclusion of oxygen. There are some geothermal reservoirs where, owing to
low pH, the brine is corrosive. Others, such as those in Californias Salton Sea area,
are corrosive because of their high salinity (up to 20% dissolved salts).
At most geothermal sites, especially those in the United States, spent brine from
power plants must be reinjected to prevent subsidence, avoid contamination of
surface waters, or maintain the reservoir. Return brine from the geothermal power

Chevron Corporation 200-17 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

plant must be monitored for corrosivity and oxygen intrusion, to avoid corrosion of
reinjection lines, pumps, and well casings.
Each geothermal site is unique and requires an individual corrosion evaluation.
Even though carbon steel may be suitable as the basic construction material, many
other potential corrosion problems have been shown to exist to varying degrees:
Stainless steels may stress-corrosion crack, especially in the event of oxygen
intrusion.
High-strength, low alloy steels may sulfide crack where the produced fluids
contain H2S.
H2S may also cause corrosion of Monel and copper-base alloys.

Concrete Structures
Concrete itself is not affected by sea water, but deterioration of concrete structures
may occur in sea water environments owing to corrosion of the reinforcing steel.
The currently popular theory is that alkalinity in the concrete passivates the carbon
steel rebar, but chlorides destroy this passivity. When concrete is placed in service
in a salt-containing environment, these salts immediately begin to permeate the
concrete. A salt concentration gradient develops from the surface inward, with the
salt content at any point within the concrete rising with time. Eventually the salt
content at the rebar may reach a critical level at which corrosion initiates.
It has long been recognized that chlorides in concrete can cause rebar corrosion.
Efforts to control it have taken two directions. In one, chloride content of the
concrete mix is minimized, especially in concrete intended for marine environ-
ments where further salt pickup is probable. This involves control of mixing water
and avoidance of chloride-containing admixtures. Alternatively, dense, imperme-
able concrete is used with adequate cover over the rebar (at least 3 inches for marine
service) to minimize the quantity of chloride that reaches the rebar.
Experts disagree on the chloride concentration necessary to initiate rebar corrosion.
Values accepted by the Portland Cement Association and some other experts are
0.20% total chloride based on weight of cement, or 1 pound of chloride per cubic
yard of concrete. Prestressed concrete requires even more rigid control of chloride
owing to the greater sensitivity of the prestressing tendons to corrosion and stress
corrosion cracking.
For concrete near or submerged in sea water, prevention of rebar corrosion requires
consideration not only of chloride concentration in the mix water but also the
increase in chloride content that will occur during prolonged exposure. A thick
(3-inch or greater) concrete cover over the rebar, as well as a dense, impermeable
concrete, are required to prevent rebar corrosion for the life required of most struc-
tures. Impermeable concrete not only slows the rate at which chlorides enter, but it
also limits corrosion by acting as a diffusion barrier for oxygen.
Epoxy-coated rebar is available and is another effective answer to the rebar corro-
sion problem.

May 2001 200-18 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Offshore Platforms
Offshore structures operate in an especially hostile environment. Parts of the struc-
ture are continually immersed in sea water while others are subject to splash, spray,
and wave action. Corrosion protection requirements are commonly defined by
dividing the structure into zones, as shown in Figure 200-9. In the absence of any
corrosion prevention measures, corrosion rates of carbon steel will usually vary
from point to point, roughly as depicted in the figure.

Fig. 200-9 Corrosion Rate of Offshore Structures, by Zone (Courtesy of LaQue Center for
Corrosion Technology, Inc.)

Splash Zone. The worst corrosion typically occurs in the splash zone, which is
alternately exposed to sea water and the marine atmosphere. A value commonly
used for estimating is 17 mpy. This is equivalent to a -inch corrosion allowance
for 30 years' life. However, the severity of splash zone corrosion varies around the
world.
Water and air temperature do not seem to have much effect on corrosion rate, but
there is reason to believe that rougher water leads to higher rates. Splash zone rates
as high as 38 mpy (maximum penetration) have been measured on Company plat-
forms in the Gulf of Mexico. On the other hand, Aramco platforms in the Arabian
Gulf are reported to corrode at 17 mpy maximum.
Extremely high splash zone corrosion rates have been reported in Alaska. At Cook
Inlet corrosion is increased by the abrasive effect of 3- to 4-foot thick ice moving
past the jacket legs at a speed of 6 knots. In the -5 feet to +25 feet tide zone, 15 mpy
general attack and 60 mpy pitting have been reported by Union Oil. A Shell Oil
platform in the same area showed 35 mpy general corrosion and 80100 mpy
pitting.

Chevron Corporation 200-19 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Several other highly effective techniques have evolved over the years for control of
splash zone corrosion. For maximum reliability, Monel jacketing is preferred. The
usual thickness of the Monel is 18 gage. The original Monel jacketing installed on
some Company platforms on the California coast as early as 1958 is still in service.
One advantage of Monel is that it is not much affected by temperature and can
therefore be used to protect hot risers. A disadvantage is its lack of resistance to
fouling by marine organisms.
Vulcanized rubber (Splashtron, for example) has also been used extensively. Most
Company experience with this has been in the Gulf of Mexico. It has proven to be
an effective method of corrosion protection. To assure good adhesion the rubber
should be shop-applied rather than field-applied, since the degree of control
required for optimum properties can seldom be assured in the field.
Filled epoxies, such as Ameron's Tideguard 171 have achieved some success for
splash zone protection. These are sprayable materials, typically applied to a thick-
ness of about 3/16 inch. Their record within the industry has been good. Some uses
within the Company include Platforms Edith and Grace on the California coast and
several platforms in West Africa. Because they are brittle there has been some
concern about resistance to mechanical damage.
It is frequently necessary to fill structural steel voids and inside vessel skirts with
resin to prevent accumulation of saltwater in the void or skirt. Filled epoxies may
not work best because they are often too stiff to pour into a cavity and there may be
insufficient room to trowel them in properly. A solution is to use a catalyzed poly-
ester resin. This material pours readily yet hardens within hours and is also less
expensive than epoxy. A major disadvantage with polyester resin is that it shrinks
during curing, requiring follow-up applications around the perimeter of the void or
skirt.
Heavy marine growth can add significantly to the effective surface area of jacket
legs and greatly add to the force exerted by waves. Elimination of fouling can result
in important savings in structural steel or eliminate costly cleaning. One of the
newer techniques for splash zone protection employs a composite material
consisting of a sheet of 90-10 cupro-nickel bonded to the steel jacket with a layer of
vulcanized rubber. The rubber provides the basic corrosion protection, while the
cupro-nickel protects against fouling by marine organisms. To provide protection
against fouling the cupro-nickel must be electrically insulated from the steel by the
rubber layer, not welded directly to the steel. If in direct contact, the steel will
cathodically protect the cupro-nickel. This prevents the slow corrosion that gener-
ates the copper ions that are toxic to marine growth.
Underwater Parts. The underwater parts of offshore platforms are almost always
cathodically protected and suffer almost no corrosion. Also, they may occasionally
be coated. Coating improves distribution of cathodic protection current, allows rapid
polarization of any bare spots in the coating, and protects new structures until the
cathodic protection system goes into service. Corrosion fatigue is an important
consideration on the submerged part of the structure, since the effect of corrosion is
to eliminate the fatigue endurance limit. Coatings and cathodic protection provide

May 2001 200-20 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

protection against corrosion fatigue, but excessive cathodic protection potentials can
have a detrimental effect.
Marine Atmosphere. Parts of the structure exposed only to the marine atmosphere
are protected by organic or inorganic coatings. This subject is beyond the scope of
the Corrosion Prevention and Metallurgy Manual; see the Coatings Manual.

Wharves and Docks


Wharf and dock piling are subject to the same corrosive effects as jacket legs on
offshore structures. However, corrosion rates tend to be lower, probably because the
water is, on average, much calmer. Because corrosion rates are lower, and because
wharf piles are structurally less critical than platform jacket legs, unprotected or
poorly protected steel is often used. As a result, corrosion of wharf piling may
develop into a difficult maintenance problem over time.
The U.S. Navy has run extensive corrosion tests on sheet piling at various locations
around the United States and the Canal Zone. Figure 200-10, based on 13 to 27
years of exposure, illustrates the variation in corrosion rate that can be expected,
rates in all cases being measured at the worst point along the length of the pile. See
Figures 200-11 and 200-12 for additional data.

Fig. 200-10 Corrosion Rates of Sheet Piling in Various Locations, mpy


Location General Corrosion Pitting and General
Boston, MA 5 8
Norfolk, VA 8 17
Key West, FL 14 19
Coco Solo, C.Z. 10 18
Puget Sound, WA 5 10
Alameda, CA 5 8
San Diego, CA 7 13
Pearl Harbor, HI 5 10

Time is a factor in the corrosion rate. United States Steel tests at Wrightsville
Beach, North Carolina, showed 25 mpy maximum penetration rate after 5 years, and
18 mpy after 9 years. A 6-year test in the same location, on a steel pile in a shore-
line bulkhead, showed a 20-mpy rate.
Corrosion rates in San Francisco Bay are not greatly different. In some tests -inch
steel plate has been perforated in 18 years (14 mpy) and 1-inch rods have become
severed in 26 years (19 mpy).
Underwater corrosion is commonly prevented by cathodic protection. However, this
cannot prevent splash zone corrosion. In sheltered waters the splash zone may be
almost nonexistent, making it possible to coat virtually to the water level. Rough-
ness of the water has a dominant effect on whether coatings can be maintained on
this part of the piling. At Pascagoula, the piles on the original wharf were provided

Chevron Corporation 200-21 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-11 Corrosion Rate Profile on Piling (Courtesy of the Naval Civil Engineering Laboratory)

May 2001 200-22 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-12 Pitting Profile on Piling (Courtesy of the Naval Civil Engineering Laboratory)

with 1/16-inch corrosion allowance, and zinc/vinyl coating. This approach was
reasonably successful, giving about 15 years service before major maintenance was
required. The Kenai (Alaska) Wharf was similarly protected, except that the coating
was a coal tar epoxy. This was ineffective in the tidal zone because of the scouring
effect of moving ice. Major maintenance problems were also encountered on the
Perth Amboy Wharf after many years of operation, owing to wharf design and use
of H-piles, both of which made coating maintenance difficult.

Chevron Corporation 200-23 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

214 Corrosion of Carbon Steel


General Behavior
In sea water, carbon steel corrodes through a combination of general attack and
pitting. However, there is no widely accepted value for the corrosion rate of steel in
ambient temperature sea water. For example, sea water piping does not corrode at
the same rate as wharf piling, and the corrosion criteria are different. In piping,
pitting and perforation are the primary concerns. On piling, general corrosion is
more important because it best correlates with strength.
The best way to predict the rate of sea water corrosion is to compare similar
services. Extensive corrosion data are available in published sources and in
Company records on corrosion of wharf piling and offshore platform jackets. Such
data provide more accurate information than generalized data on the corrosion of
steel in sea water. Data on some special applications are given later in this chapter.
Where specific data are lacking, Company experience indicates that the following
corrosion rates be used for piping and equipment handling aerated sea water at
ambient or ocean temperatures:
5 mils per year general corrosion, plus
30 mils per year pitting
These rates provide a reasonable basis for estimating the life of piping and process
equipment in sea water service. However, they should not be used for corrosion rate
predictions on structures in ocean environments, such as wharf piles or platform
legs. Here, test panels and measurements have shown maximum penetration rates
(pitting plus general corrosion) from a low of 6 mpy to a high of 40 mpy. The
reasons for these wide variations were discussed in Section 213.
While corrosion rates in sea water are often treated as constant over time, experts
agree that, in fact, corrosion rates of steel in salt water decrease with time.
Longterm losses can be predicted by the equation:

Corrosion (mils) = K1 + K2t


(Eq. 200-1)
where:
K1 , K2 = constants which depend on geographic location and exposure
condition
t = time in years

May 2001 200-24 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

U.S. Navy tests in Panama lasting 16 years showed general corrosion and pitting
rates of immersed panels at 80F and low flow. The losses could be predicted by the
following equations:

General corrosion loss (mils) = 5 + 2.7t

Pitting (mils) = 64 + 3t

Similar tests by U.S. Steel at Wrightsville Beach, North Carolina showed pitting
rates of 25 mpy during a 5-year exposure and 18 mpy during a 9-year test. These
rates correspond to the following equation:

Pitting (mils) = 80 + 9t

Effect of Temperature
Reliable data on the effect of temperature on sea water corrosion of carbon steel are
hard to find and often conflicting. For predictive purposes Figure 200-13 should be
used. The recommended design curve on this graph is based on judgment and expe-
rience, for predicting corrosion rates of both carbon steel and cast iron. The indi-
cated corrosion rates represent the sum of both general corrosion and pitting
(i.e., maximum penetration).

Fig. 200-13 Effect of Temperature on Corrosion in Aerated Sea Water (Courtesy of LaQue
Center for Corrosion Technology, Inc.)

Effect of Salinity
Though sea water is generally treated as a fluid of fixed composition, there are
substantial variations in salinity from place to place. Where the climate is hot and
waters are shallow, as in the Arabian Gulf, the high evaporation rates increase

Chevron Corporation 200-25 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

salinity. In other locations, such as Richmond Refinery on San Francisco Bay, there
is some dilution by fresh water runoff. However, in neither case is the salinity varia-
tion enough to cause the water to be significantly different in corrosivity from sea
water of normal salinity.
Dilute brines are commonly encountered in such services as waste water effluent in
coastal refineries, brackish water wells, and some produced waters. Experience
clearly shows that corrosivity of diluted sea water and brines is not directly propor-
tional to salinity. In the absence of good published data, a common practice has
been to assume that a logarithmic relationship exists between corrosion rate and
salinity such that each order of magnitude reduction in salinity cuts the corrosion
rate in half.
For produced or brackish well waters, special allowances are necessary for factors
other than salinity that represent important differences from sea water. These factors
are:
Oxygen content
Dissolved acid gases (H2S, CO2)
Differences in specific salts

Effect of Oxygen
In sea water or neutral brine, corrosion will not occur unless oxygen is present. It is
common to take advantage of this phenomenon and deaerate the water to permit use
of carbon steel piping and equipment. However, only very small quantities of
oxygen can be tolerated without corrosion occurring, and oxygen tolerance
decreases as the temperature rises.
Sea water injection is a common oil field practice. Injection water must be carefully
deaerated to prevent corrosion of piping and well casing and to avoid side-effects
such as bacteria growth underground. Corrosion in water injection systems is
usually controlled by deaeration and oxygen scavenging. This is no easy task,
however, and extreme diligence is required to keep oxygen levels at a minimum.
Even at best, it is generally not possible to use carbon steel for water injection
pumps.
At elevated temperatures, carbon steel's tolerance for oxygen is even lower. Early
attempts to use carbon steel equipment in flash evaporation desalination plants were
unsuccessful because the very low oxygen levels needed to stifle corrosion could
not be consistently maintained. The desalination plants at Borco Refinery are prime
examples. Figure 200-14 was plotted from Dow Chemical Company data in an
operating desalination plant [6], and shows the extreme sensitivity of carbon steel
corrosion to temperature and oxygen content. Data from other sources are given in
Figures 200-15 through 200-17.
Geothermal brines vary widely in composition and corrosivity. Fields of direct
interest to the Company include Beowawe (Nevada) at one extreme with only
1100 parts per million total dissolved solids (60 ppm chloride) and Brawley (Cali-
fornia) at the other with over 200,000 ppm solids. Geothermal brines being
commercially utilized generally have lower salinity than sea water, are neutral in

May 2001 200-26 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-14 Corrosion of Carbon Steel in Deaerated Sea Water; Effect of Oxygen Content
and Temperature [6] (Copyright NACE International. All rights reserved.)

pH, and are noncorrosive to carbon steel providing oxygen is kept out. At the Heber
Geothermal Field, Company specifications limit the oxygen content to 10 parts per
billion (ppb) to prevent corrosion of brine reinjection equipment. Most geothermal
brines are totally oxygen-free when they are produced. A crucial aspect of corro-
sion control in geothermal systems is the prevention of oxygen intrusion.

Effect of Velocity
Figures 200-17 and 200-18 show the effect of velocity on sea water corrosion.
Figure 200-18 indicates a much more pronounced velocity effect than
Figure 200-17. This conflict in the data is probably the effect of marine organisms
and slimes. Generally, organisms cannot attach to a metal surface when the velocity
is over 3 feet per second (fps). However, they may attach during periods of lower
velocity, after which they can maintain adhesion at velocities over 3 fps. Slimes can
attach themselves at higher velocities, and the velocity effect on corrosion is shown
in Figure 200-17. Some Company experience in the Santa Barbara (California) area
suggests that slimes wash off the surface at velocities exceeding about 12 feet per
second.

Effect of pH
Within the pH range of 5 to 10, the corrosion rate of carbon steel is not affected
much by pH when the acidity is due to the presence of strong mineral acids. This is
illustrated in Figure 200-19. Below a pH of 5, corrosion occurs through a hydrogen
evolution mechanism, and oxygen is no longer required for a corrosion reaction to
occur.
When acidity is from a weak acid such as carbon dioxide, acid corrosion can initiate
at a pH value somewhat higher (closer to neutral) than 5. Carbon dioxide is often
present in produced water and geothermal brine. Solubility of CO2 is a function of
pressure, and much of the CO2 may flash out of solution when system pressure is

Chevron Corporation 200-27 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-15 Effect of Oxygen Concentration on Corrosion of Mild Steel in Pacific Ocean
Water at Ambient Temperature [6] (Copyright NACE International. All rights
reserved.)

reduced. For this reason, corrosion and pH measurements in produced fluids taken
after the system is depressured may not represent the higher-pressure downhole
environments.

May 2001 200-28 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-16 Corrosion Behavior of AISI 1010 Mild Steel in 250F Sea Water at Several
Dissolved Oxygen Levels Under Recirculating Flow Conditions (Courtesy of
NACE International)

Chevron Corporation 200-29 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-17 Effect of Sea Water Velocity on Corrosion of Steel at Ambient Temperature,
Exposed 38 Days (Courtesy of LaQue Center for Corrosion Technology, Inc.)

215 Corrosion of Other Alloys


Cast Iron
Cast iron corrosion in sea water takes the form of graphitization, as discussed in
Section 212. The rate of corrosion is generally assumed to be the same as for carbon
steel. Some comparative data on carbon steel and cast iron are shown in
Figure 200-13.

Low and Intermediate Alloy Steels


High strength, low alloy steels such as AISI 4140 are sometimes used in sea water
service where a high strength material is required. They corrode at about the same
rate as carbon steel. Some specialty steels like Cor-Ten are said to corrode at lower
rates than carbon steel (see Figure 200-18). As discussed in Section 212, high-
strength low alloys may also fail by stress corrosion cracking.
Intermediate alloy steels are rarely used in sea water service. They cost more than
carbon steel and offer no significant corrosion benefit, although published data are
sometimes conflicting because long test exposures give different results than short
exposures. Sixteen-year immersion tests conducted by the U.S. Navy showed 2%
and 5% nickel steels to have about the same general corrosion rate as carbon steel,
but higher pitting rates. Two percent and 5% chromium steels showed higher
general corrosion rates than carbon steel, but the same pitting rate.

Conventional Stainless Steels


Stainless steels contain chromium, which contributes to the formation of a passive
surface film. As a result, general corrosion of stainless steel in sea water is negli-

May 2001 200-30 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-18 Weight Loss Corrosion Rates of Mild Steel Fig. 200-19 Effect of Sea Water pH on the Corrosion
and Cor-Ten A in High Velocity High Rate of AISI 1010 Mild Steel at a Typical
Temperature Sea Water (Courtesy of NACE Deaerator Temperature (Courtesy of NACE
International) International)

gible. However, chlorides in sea water can lead to a local breakdown of passivity,
resulting in pitting and crevice corrosion.
The 400 Series stainless steels are rarely used in sea water or brine service. They are
more difficult to weld and fabricate than the austenitic (300 Series) grades, and
pitting rates are far higher.
Austenitic stainless steels are also subject to pitting and crevice corrosion in sea
water but are sometimes used under special conditions. Type 316 is generally
thought to be superior to Type 304, but some authorities have pointed out that pits in
316, while fewer, are almost as deep. Pitting rates of 50 to 100 mpy are common,
and rates as high as 700 mpy have been reported occasionally.
While the rate of crevice corrosion is affected little by fluid velocity, pitting is
reduced or eliminated if the velocity is high enough (5 to 6 fps). However, Company
experience shows that it is not practical to rely on high velocity to prevent pitting of

Chevron Corporation 200-31 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

stainless steel piping or heat exchanger tubes. Marine growth, lodgement of a sea
shell, or local obstruction by a foreign object can result in localized low velocity and
rapid pit formation.
Pumps are among the few successful applications of stainless steel in sea water. The
extreme turbulence within the pump keeps surfaces scoured clean and prevents the
formation of pits. Type 316 stainless is the usual pump material for sea water
service. However, stainless steel should be limited to pumps that operate continu-
ously or are flushed and drained when not in use.
Austenitic stainless steels are subject to chloride stress corrosion cracking in hot sea
water or brine. The threshold temperature for chloride cracking is generally 140F.
However, there have been isolated cases of cracking at ambient temperature when
the metal was sensitized by welding or heat treatment. A high degree of cold work
may also lower threshold temperature.
The probability of chloride cracking can be reduced by thermal stress relief of welds
and cold bends. However, under very severe conditions, the threshold stress for
chloride cracking may be so low that even stress relief does not assure that cracks
will not form. A hot surface on which sea water is boiling and evaporating is an
example of a service condition for which stress relief does not guarantee immunity
from chloride cracking.
Deaeration retards pitting, crevice corrosion, and stress corrosion cracking. None of
these phenomena can occur in the absence of oxygen. Experience in the Borco
Desalination Plants shows that at temperatures up to 250F neither general corro-
sion, crevice corrosion, nor pitting occurred on Type 316L liner plates in the multi-
stage flash evaporator boxes. On one occasion, a few stress corrosion cracks were
found in the first of the 34 consecutive stages. These were attributed to traces of
oxygen in the feed water that had not been removed by deaeration and oxygen scav-
enging. This small amount of oxygen flashed off in the first evaporation stage, and
there were no cracks downstream.
One source reports a pitting rate of 3 mpy on Type 316 stainless in 220F sea water
deaerated to an oxygen content of 25 ppb [8].

Specialty Stainless Steels


A number of specialty stainless steels have been introduced which contain higher
percentages of chromium and molybdenum for increased resistance to pitting and
crevice corrosion. These alloys are generally sold under trade names. Some are
austenitic and some are ferritic or duplex. Examples are:
Trade Name Composition
Sanicro 28 (27 Cr, 31 Ni, 3 Mo)
254 SMO (20 Cr, 18 Ni, 6 Mo)
6X (20 Cr, 24 Ni, 6 Mo)
904L (20 Cr, 25 Ni, 4 Mo)
SAF 2205 (26 Cr, 4 Ni, 1 Mo)

May 2001 200-32 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Ferralium 255 (26 Cr, 5 Ni, 3 Mo)


29-4-2 (29 Cr, 2 Ni, 4 Mo)
29-4C (29 Cr, 4 Mo)
Sea Cure (27 Cr, 1 Ni, 3 Mo)
Monit (24 Cr, 4 Ni, 4 Mo)
To rate the relative pitting and crevice corrosion resistance of the stainless family of
alloys, a pitting index number was developed in Germany called the wirksumme.
The equation for calculating the pitting index number of a metal is:

PI = % Cr + (3.3 % Mo)
(Eq. 200-2)
Thus, the pitting index number for Sanicro 28 would be 27 + (3.3 3.5) = 38.5.
Figure 200-20 gives pitting index numbers for both specialty stainless steels and a
variety of conventional stainless steels and high alloys. The higher the pitting index
number the more resistant a material is to chloride pitting and crevice corrosion.
The pitting index that corresponds to complete immunity lies between 36 and 44,
the higher value being correct where crevice corrosion is a factor, and the lower
value correct for systems essentially free of crevices.
Relative performance of the specialty stainless steels can be evaluated on the basis
of a critical pitting temperature. This is determined by a laboratory test in ferric
chloride solution. Critical pitting temperatures for a variety of alloys are tabulated in
Figure 200-20. Those alloys with a proven history of performance in sea-water-
cooled surface condensers have a pitting index number of at least 38 and a critical
pitting temperature of 122F or better.
Welds in these specialty stainless steels often have poorer corrosion resistance than
the parent metal, if made with electrodes of matching chemical composition. For
this reason, care is required in selection of welding electrodes. It is common to use
electrodes that are more highly alloyed than the base metal.

Fig. 200-20 Relative Pitting and Crevice Corrosion Resistance (1 of 2)


Composition(2) Pitting Index Critical Pitting
UNS Micro- Number Temperature
Alloy Number structure(1) Cr Ni Mo Cr+3.3 Mo F(C)(3)
Type 304L S30303 A 18 8 0 18 41 (5)
Type 316L S31603 A 17 12 2 25 73 (23)
Alloy 20 Cb-3 N08020 A 20 34 2 28 91 (33)
Type 317L S31703 A 18 14 3 30 86 (30)
SAF 2205 S31803 D 26 4 1 31
Alloy 825 N08825 A 21 42 3 31 120 (49)
CD4MCu J93370 D 26 5 2 33
904L N08904 A 20 25 4 35 108 (42)

Chevron Corporation 200-33 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-20 Relative Pitting and Crevice Corrosion Resistance (2 of 2)


Composition(2) Pitting Index Critical Pitting
UNS Micro- Number Temperature
Alloy Number structure(1) Cr Ni Mo Cr+3.3 Mo F(C)(3)
Ferralium 255 S32550 D 26 5 3 36
MONIT S44635 F 25 4 4 38 131 (55)
Sanicro 28 N08028 A 27 31 3 38 127 (53)
Sea-Cure S44660 F 27 1 3 39
254 SMO S31254 A 20 18 6 40
29-4-2 S44800 F 29 2 4 42 122 (50)
29-4C S44735 F 29 0 4 42 131 (55)
Hastelloy G N06007 A 22 45 6 43 171 (77)
Alloy 625 N06625 A 21 62 9 51 207 (97)
Hastelloy C N10002 A 15 54 16 68 307 (153)
(1) The letters indicate microstructure. (A - Austenitic, F = Ferritic, D = Duplex).
(2) Balance is largely iron. Many of these alloys also contain small amounts of Cu, Cb, or W.
(3) Temperature above which pitting occurs in a standard laboratory corrosion test using 10% ferric chloride.

At elevated temperatures some specialty stainless steels are susceptible to chloride


stress corrosion cracking in sea water and brines. In general, the ferritic grades are
immune to chloride cracking except when sensitized by welding. Austenitic grades,
although very susceptible when nickel contents are low (8-12%), become more
resistant at higher nickel contents. Alloy 20 Cb-3 (34% Ni) is virtually immune.
Duplex grades of stainless are intermediate in chloride cracking resistance between
austenitic and ferritic grades.

Nickel Alloys
Pure nickel is subject to pitting and crevice corrosion in sea water. It is seldom used
because it is not a cost-effective material in sea water or brine services.
The chromium-containing nickel alloys vary in sea water resistance according to
their chromium and molybdenum contents. Resistance rises as the parameter
(Cr + 3.3 Mo) increases, as previously discussed under Specialty Stainless Steels.
Of the alloys in this grouping, Inconel 600 and Incoloy 800 suffer pitting and
crevice corrosion at rates similar to the 300 Series stainless steels. Incoloy 825
achieves somewhat better results but is rarely used. Hastelloy G, Hastelloy C, and
Inconel 625 have excellent resistance to sea water. Test panels of Hastelloy C,
polished to a mirror finish, have retained their appearance without change after
several decades exposure to sea water.
Monel, a 70% nickel, 30% copper alloy, has long been an industry standard for sea
water service. However, it is not a perfect material, having some tendency to pit in
stagnant sea water. It is commonly used for channel sections of heat exchangers,
where life is so long that replacements are rare. Monel also has long been used as
sheathing on the jacket legs of offshore platforms. On Chevron Platforms Hazel and

May 2001 200-34 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Hilda off the Southern California coast, the original 18 gage Monel jacketing
installed in 1958-59 is still in good condition, though heavily fouled with marine
growth.
The addition of nickel to copper results in passivity if nickel content is over 40%.
This greatly increases resistance to erosion and impingement but introduces the
possibility of pitting. Company experience has been that Monel heat exchanger
tubes perform poorly in sea water service. They are free of end impingement but
suffer isolated pitting failures, most often where a sea shell or other foreign object
becomes trapped in a tube. There have also been isolated cases at Richmond
Refinery where Monel piping has pitted badly, though such cases are rare.
Nickel-molybdenum alloys such as Hastelloy B do not have outstanding resistance
to sea water and are seldom used.
Nickel alloys possess good resistance to corrosion at high velocities. Monel is
essentially immune to impingement and similar types of high velocity attack that
can occur in piping and heat exchangers. However, Monel will suffer some loss
under conditions of extreme turbulence, as in pumps. Nickel alloys containing chro-
mium, such as Inconel 625, Hastelloy G, and Hastelloy C will resist high velocities
under almost any condition likely to be encountered in sea water service. One recent
report mentioned a 3-inch Hastelloy C gate valve which had been examined after
19 years service in sea water. The only damage was severe rusting of the cast iron
handwheel.
Most nickel alloys are immune to chloride stress corrosion cracking. Incoloy 800 is
an exception. Although this alloy has much better chloride cracking resistance than
the 300 Series stainless steels and is often used to replace them after a cracking
failure, it will chloride-crack under severe conditions.
K-Monel is susceptible to stress corrosion cracking in sea water, but only at very
high strength and hardness levels. Failures have occurred at hardnesses on the order
of HRC 36, but should not occur if hardness is limited to a maximum of HRC 32.
For marine bolting, INCO recommends that K-Monel be ordered in the annealed
and aged condition rather than cold-drawn and aged. Some care is necessary to
avoid abuse, since some failures have been reported in badly overtorqued bolts even
though they conformed to the HRC 32 hardness limit.

Copper Alloys
Copper alloys as a class have good resistance to sea water and are commonly used
in sea water service. Common applications are piping, valves, pumps and heat
exchanger tubes.
The corrosion resistance of copper can be greatly improved by the addition of
certain alloying elements, of which zinc, tin, aluminum, and nickel are the most
common. These alloys are referred to as follows:
Name Alloy
Brass copper + zinc
Tin bronze copper + tin

Chevron Corporation 200-35 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Aluminum bronze copper + aluminum


Cupro-nickel copper + nickel
Sometimes these alloying elements are used in combination as in nickel aluminum
bronze. Other alloying elements (silicon, manganese, lead) may be added to achieve
certain mechanical properties such as strength and ease of machining, but they do
not add to corrosion resistance.
Nickel and aluminum additions improve resistance to corrosion by high velocity salt
water (cavitation and impingement attack). Zinc and aluminum reduce H2S attack in
polluted sea water. Copper alloys can be roughly ranked in order of decreasing
resistance as follows:

High velocity corrosion:


1. Nickel aluminum bronze
2. Cupro-nickel
3. Aluminum bronze
4. Brass
5. Copper

Polluted sea water:


1. Brass
2. Aluminum bronze
3. Cupro-nickel
4. Copper
Because copper ions are toxic to marine organisms, copper is nonfouling and marine
organisms will not adhere to it. To resist fouling the copper or copper alloy must
corrode at a finite rate (about 1 mpy is optimum). More corrosion-resistant copper
alloys tend to be less fouling-resistant because they do not put enough copper ions
into solution. Relative fouling resistance of some alloys is as follows (a rating of
100 is the best obtainable; a zero rating means no fouling resistance):

May 2001 200-36 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Alloy Relative Fouling Resistance


Copper 100
90-10 Cupro-nickel 190
Brass 70-90
Tin bronze 70-90
70-30 Cupro-nickel 150
Aluminum bronze 50
Monel 10
Carbon steel 10
To control fouling, a 90-10 cupro-nickel was selected for a sea water intake line at
the Companys Gaviota (California) Gas Plant.
One of the newer techniques for providing both corrosion resistance and fouling
resistance in the splash zone area on jacket legs of offshore structures uses a vulca-
nized rubber lining with attached facing sheet of 90-10 cupro-nickel. The vulca-
nized rubber lining between 90-10 and steel is vital if fouling resistance is to be
maintained. Without the rubber, the 90-10 would be in direct electrical contact with
the steel. This would cathodically protect the 90-10, prevent copper ions from going
into solution, and eliminate resistance to fouling.
Monel jacketing fouls badly in the same service.
Brasses may be subject to dezincification in sea water. Other copper alloys may
suffer selective loss of one element or one metallurgical phase. This attack is best
described by the general term dealloying.
The severity of dezincification of brasses is a function of zinc content, those with
the highest zinc content being the worst. Muntz Metal, a common heat exchanger
tube sheet material containing 40% zinc, is readily dezincified. Yellow brass
(35% zinc) is nearly as bad. Red brass (15% zinc) is quite resistant, though at some
sacrifice in general corrosion and pitting resistance. Resistance to dezincification
can be improved by adding about 1% tin to the alloy. Naval brass, which is basi-
cally Muntz Metal with 1% tin, is substantially more resistant and is therefore
commonly used by the Company for heat exchanger tube sheets in preference to
Muntz Metal. Similarly, admiralty brass, a 30% zinc brass with 1% tin, is much
more resistant than a similar alloy without the tin.
Admiralty brass can be made more resistant to dezincification by addition of small
quantities of antimony, arsenic, or phosphorus. The resulting alloys are called inhib-
ited admiralty. There is no significant difference between the three types of inhib-
ited admiralty; they are used interchangeably by the Company for heat exchanger
tubes whereas the uninhibited admiralty is not used.
Other copper alloys may also be subject to dealloying, though generally not to such
a degree. Wrought aluminum bronze and cupro-nickel occasionally dealloy, usually
in services with elevated temperature salt water exposure such as heat exchanger
tubes. Correction is best accomplished by reducing metal wall temperatures, since
rates of dealloying increase rapidly with temperature.

Chevron Corporation 200-37 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Aluminum bronzes can be subject to dealloying if aluminum content exceeds about


8%. Aluminum bronzes are single-phase alloys if aluminum content is low. High
aluminum content promotes the formation of a second phase that is subject to selec-
tive attack. A tempering heat treatment can be used to alter the nature of this second
phase and improve resistance to dealloying. Such a heat treatment is often appro-
priate for castings in sea water service.
Nickel aluminum bronze is becoming a popular material for cast propellers, pumps,
and valves in sea water service. It has only a slight tendency toward dealloying,
which can be improved by careful control of chemistry. Resistance to dealloying can
be further improved by a tempering heat treatment, but this is seldom necessary and
the small incremental benefit is justified only for the most critical services.
See Section 213 for additional information on copper alloys under both Piping and
Heat Exchangers.

Titanium
Because of its reasonable cost and excellent corrosion resistance, titanium is widely
used in sea water service. However, problems in manufacture or fabrication confine
the use of titanium almost exclusively to heat exchanger tubes. Welded tubes are
used almost exclusively because they are far less costly than seamless and of excel-
lent quality. Tubes are purchased to ASTM B338. The grades given in
Figure 200-21 are of interest for heat exchanger applications.

Fig. 200-21 Titanium for Heat Exchanger Tubes


Threshold Pitting/Crevice
Grade Composition Relative 1989 Cost Corr. Temp., F
2 Commercially pure Lowest 220
titanium
12 Ti Alloy: 20-25% higher 350
0.8% Ni
0.3% Mo
7 Ti Alloy: Twice cost of Grade 350-450
0.2% Pd 2

Grade 2 is used most often because of its reasonable cost and ready availability.
Grade 12 can be obtained for a nominal cost premium and offers better corrosion
resistance than Grade 2 in hot sea water. Grade 7 is much more costly than Grade 12
but has even better corrosion resistance.
In ambient temperature sea water, these grades of titanium show essentially zero
corrosion. They will not suffer pitting, crevice corrosion, or stress corrosion
cracking. At higher temperatures pitting or crevice corrosion may occur. The
threshold temperature is a function of chloride content, alloy composition, and
sharpness of the crevice. Threshold pitting/crevice corrosion temperatures in sea
water are given in Figure 200-21.

May 2001 200-38 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Brines of higher salinity have a lower threshold pitting/crevice corrosion tempera-


ture. Salt concentration can increase locally underneath a salt plug in sea water
service. The primary effect is from increased MgCl2 concentration, which lowers
the pH. The effect of pH on crevice corrosion behavior is shown in Figure 200-22.
Increased NaCl concentrations also affect the threshold temperature, though not as
much.

Fig. 200-22 Immunity of Titanium (Grade 2) From Pitting (Courtesy of Titanium Metals Corp.)

Chevron Corporation 200-39 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Titanium is likely either to pit rapidly or not at all. Cases of minor pitting are rare in
sea water service. At temperatures over the pitting threshold, pitting rates are likely
to be very high and titanium could easily have a shorter life than copper-base alloys.
Below the threshold pitting temperature, corrosion resistance is so good that essen-
tially no corrosion allowance is needed. Unless high operating pressures dictate a
need for greater metal thickness 20-gage tubes are commonly used. Bundles using
such thin tubes require careful handling.
Titanium is also good in most types of HC service, making it an ideal choice for salt
water coolers.

Aluminum Alloys
Aluminum is seldom used in sea water applications. It is not an acceptable material
for vessels, piping, or heat exchangers in sea water service because of its tendency
to pit. Successful use of aluminum for such things as boats depends on special tech-
niques such as coatings, cathodic protection and avoidance of crevices. Otherwise,
regular removal from the water and subsequent cleaning or fresh water wash is
necessary.
Aluminum is subject to severe galvanic corrosion in contact with heavy metals, one
of the worst of which is copper. In sea water aluminum is likely to lose its protec-
tive oxide film, achieve an electrochemically active potential, and sacrificially
corrode in contact with another metal.
Usually zinc chromate primers are used when painting aluminum; primers
containing lead can cause corrosion of aluminum, especially in marine environ-
ments. Antifouling paints containing copper can also corrode aluminum unless
isolated from the metal by a special primer. Aluminum weatherjacketing should
only be used if it is coated.

220 Corrosion by Cooling Water


Without proper treatment to remove or neutralize contaminants, cooling waters can
adversely impact the performance of refinery heat transfer surfaces and create human
health risks. This section discusses cooling water in general and ways to reduce or
eliminate the instances of heat exchanger plugging, pitting and, fouling and subse-
quent plant shut down.

221 Typical and Ideal Cooling Water System


Figure 200-23 is a trouble shooting guide that lists some of the most common
problems caused by cooling water. Suggested remedial actions are also provided.
Examples of when to use the guide include: high measured corrosion rates,
detecting a hydrogen leak, a high microorganism count, or when there is pH excur-
sion. The reader is encouraged to review the guide and then read the appropriate
part of this section for more detailed information.

May 2001 200-40 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-23 Trouble Shooting Guide (1 of 3)


PROBLEM ACTION
Low pH (<6.5) Check acid addition, pH meters - cross check biocide slug.
Add caustic incrementally.
Shut off acid; allow tower pH to cycle up.
High pH(>7.5) Check acid addition pump, pH meters.
Add acid incrementally.
High iron counts Check pH.
On-line cleaning occurring?
Shut in recovery?
Phosphate residual levels?
Adjust dispersant level?
Contact water vendor to determine remedial action.
Fouling Factor increase High iron counts?
Increase in calcium or magnesium?
Higher cycles?
When was last dip slide run?
In or out of control range?
If due to microorganisms, shock treatment of biocide needed.
Hx leaks See oil, scum floating on basin water?
High bug counts?
Use an activated charcoal bomb attached to side-stream water.
Send out for GC-MS fingerprinting to identify where leak is occurring.
Fix leak.[21]
Scale Loss of acid?
Changes in Ryznar or Puckorius indices?
Silica levels?
Sulfate levels?
On-line clean out?
Check with water vendor to modify treatment chemicals.
High microbiologic counts Loss of biocide?
(greater than 105) Pump?
Hx leak?
Shut-in?
Other upset?
React quickly with shock treatment to prevent further fouling. If time
permits, run serial dilutions to determine if acid or sulfate reducers
present.

Chevron Corporation 200-41 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-23 Trouble Shooting Guide (2 of 3)


PROBLEM ACTION
Pitting factor increase Increase in chlorides? High fouling factor? On-line cleaning may be
required.
Low velocities High fouling factor? May need on-line cleaning.
Additional pump may be needed to bring velocities up to >3 ft/sec.
Hard to treat exchangers Consider a satellite chemical addition pump upstream of those
exchangers.
It will be less expensive than adding more inhibitors to the whole system.
Flow constriction Blockage?
Scaling index?
If really bad, may need to shut in and inspect.
Sediment May need more dispersant.
Contact water vendor to alter treatment chemical levels.
Backwash Exchangers.
Murky basin water Indicates flocculation/sedimentation agent is working. If oily sheen
appears, need to check for HX leak.
Fouled tower fill Must be cleaned during next turn around (or sooner).
Make sure tower is still delivering necessary delta t drop (20 degrees or
more).
High ammonia (>1PPM) Check for nitrifiers.
Check TTA delivery (for increased demand).
Blowdown to reduce effect on yellow metals.
High TDS What caused change?
Change in makeup water?
Higher cycles?
Loss of filtration?
Blowdown to prevent fouling while finding source of increase.

May 2001 200-42 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-23 Trouble Shooting Guide (3 of 3)


PROBLEM ACTION
High chlorides (>850PPM) New water source?
Higher cycles?
Is pitting index rising?
Slug of chlorine used for biocide excursion? If no increase in corrosion or
pitting is occurring, note increase and watch other corrosion indicators
closely.
High Corrosion Rates On either coupons or instantaneous.
pH excursion?
Loss of inhibitor?
Cycles too high?
Loss of biocide?
On-line cleaning occurring?
May need to increase blowdown.
Consult water vendor to decide.

Typical Cooling Water System


Most Chevron cooling water systems are open recirculating systems that consist of
the following components: a make-up water source; evaporative cooling tower with
drift eliminators and fill; cold water basin; blowdown valve to sewer; acid dilution
system and trough; inhibitor and biocide addition systems; and recirculating pumps
to supply the water to the heat exchangers.
Cooling towers can be forced draft counter flow, induced draft cross flow, or
induced draft counter flow. Figure 200-24 is a rough schematic of a cooling tower
system.
Cooling tower fill can either be splash fill or film fill. If the cooling tower is splash
fill, it is important to know the type of wood used as splash fill columns because
pressure-treated lumber can have adverse reactions with treatment chemicals. All
tower fills are subject to biofouling and should be routinely inspected to ensure
proper water cooling, distribution, and tower integrity.
The amount of cooling water used by a tower often provides clues to the state of the
system. Rapid increases in the amount of cooling water usage indicates a problem
with the system, often due to fouling, scale, or corrosion. The following commonly
used equations help determine how much cooling water is used by the tower.

Chevron Corporation 200-43 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-24 Cooling Tower System

Evaporation = Recirculation rate (GPM) F/1000 or


Recirculation rate (GPM) (C/556)
(Eq. 200-3)

Make-up = Evaporation + Loss


(Eq. 200-4)

Losses = Windage + blowdown + leaks


(Eq. 200-5)

Concentration Cycles = Total tower water/Make-up water, or


%Evap. + %BD
Cycles = -----------------------------------------
%BP
(Eq. 200-6)

Blowdown = Evaporation/cycles 1
(Eq. 200-7)

Times per cycle = Capacity/Recirculation rate


(Eq. 200-8)

Half life = 0.6933 (Capacity/blowdown)


(Eq. 200-9)

Make-up = Evaporation cycles/cycles 1


(Eq. 200-10)[22]

May 2001 200-44 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Ideal Cooling Water System


An ideal cooling water system would contain:
Automated blowdown based on conductivity (TDS) or other ions
Automated pH control
Sidestream filtration
Chlorine injection pump tied to an Oxygen Reduction Potential (ORP)
On-line inhibitor residual monitors tied to automated injection pumps
Sidestream manifold containing a fouling monitor, delta-p filter, Bio-Monitor,
and Corrator for corrosion rate measurements; inputs to go directly into control
room
No exchangers with tube-side velocities less than 3 ft/sec
No exchangers with tube skin temperatures greater than 160F or bulk water
outlet temperatures greater than 120F
No shell-side cooled exchangers
Temperature delta of greater than 20F between the top and bottom of the tower

222 Make-up Water


Cooling water can come from a variety of sources, including: municipal water,
reclaimed or reuse water, river water, or well water. Water quality and dissolved and
suspended solid components vary from region to region and must be thoroughly
documented through water testing. Impurities which cause problems can then be
identified to determine if make-up water pretreatment is necessary. Water impuri-
ties that can cause cooling water corrosion problems include:
Algae
Alkalinity
Aluminum
Ammonia
Bacteria
Barium
Calcium
Chlorides
Copper
Fluorides
Iron
Macrobiofouling
Manganese
Nitrates
Organic Hydrocarbons

Chevron Corporation 200-45 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Phosphates
Silica
Suspended and Dissolved solids
Sulfate
TSS
TDS
Zinc
A complete water analysis should be run monthly on all make-up water sources as
well as weekly on recirculated tower waters.

Make-up water pretreatment


Some make-up water may require pretreatment due to inherent impurities. The
amount and type of pretreatment needed, if any, must be determined by analyzing
the constituents present in the water sources used in the cooling water loop. The
decision to pretreat the make-up water is made by balancing the cost of the pretreat-
ment against the savings in water and chemical costs for the overall cooling water
treatment program. Often it is better to expend resources pretreating the make-up
water to achieve a desired composition as opposed to trying to resolve heat
exchanger corrosion problems caused by untreated water after it is already in the
system.
A description of various pretreatment methods is located in the Glossary
(Section 228).

Reclaimed Water
There is a growing need to secure reliable water supplies for the refining industry.
External water sources such as treated sewage effluent, called reclaimed or reuse
water, is becoming a popular adjunct to fresh water sources, especially in the
drought prone western United States. Other internal reuse sources can be reverse
osmosis reject, softener backwash and rinse, and boiler blowdown. These sources
should be considered in the same way that any water source is.
You must establish how much pretreatment of the proposed water is necessary to
meet the following criteria:
Ammonia <1ppm
0-3 colony forming units (cfu) or <3 Most Probable number (Mpn)/100ml of
coliform bacteria
Phosphate levels between 010 ppm
Total iron < .1ppm
pH dependent on treatment program (ex: phosphate between 6.57.5)
Chlorine residual of .3.5 ppm
COD 50 ppm
Fish survival rates of > 90%
Consistency of composition
Acceptable TDS

May 2001 200-46 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

High levels will dictate whether the water will require further treatment before it can
be used.
To be a cost effective water source, reclaimed water must meet a minimum of cycles
of concentration. Control limits must be specified based on maximum cycles of
concentration of the following:
TSS
TDS
Total Ca and Mg
Total alkalinity and hardness
Chlorides
Silica
Sulfate
Monitoring becomes even more critical with reclaimed water because it will have
higher TDS and potential foulants. Baseline data must be gathered on the system
without reclaimed water for a minimum of 3 months prior to introduction.

223 Blowdown Control and Cycles of Concentration


Blowdown is the intentional wasting, either continuous or intermittent, of a portion
of the tower water specifically to control the concentration of dissolved solids in the
water. Blowdown is expressed as a percentage of the make-up water and can be
calculated by comparing the concentration of a specific component ion in the make-
up water to the circulation water, such as chloride, silica, sulfate or calcium, or by
taking readings of conductivity expressed in micromhos/cm.

%Blowdown = ppm ion (e.g., silica) make-up/ppm ion circulating 100


(Eq. 200-11)
Cycles of concentration is the inverse of blowdown and is the degree of concentra-
tion of the circulating water as compared to the make-up water. It is calculated by:

Cycles = ppm ion circulating/ppm ion make-up, or


micromhos circulating/micromhos make-up
(Eq. 200-12)
Blowdown is critical in water and chemical use conservation. High blowdown rates
equate to high make-up water and chemical use, whereas low blowdown rates use
less make-up water and treatment chemicals in the system.
Blowdown based on conductivity is easy to automate using an on-line conductivity
monitor with the output tied to the blowdown control valve. Conductivity limits
should be established for both make-up and tower control waters. These limits can
range from 130 micromhos ms/cm in a fresh water make-up to 8,000 micromhos/cm
found in five cycles of reclaimed municipal treatment water.
Blowdown rates should also be established on return water silica (SiO2). For
example, ranges of 100150 ppm prevent silicate scales from forming. Sulfate

Chevron Corporation 200-47 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

(SO4) levels must be controlled to prevent degradation of cement basins and kept to
levels of 1,500 ppm for type V concrete. Chloride ppm must be regulated because of
its effect on the pitting of stainless steels, and blowdown should be set to between
500850 ppm of chloride.
Operators must check conductivity of return headers once each shift. Vendors
should follow up once a week with portable conductivity probes that are calibrated
monthly. The operator should report all readings-on-line and hand-held to verify
accuracy. On-line conductivity probes should also be calibrated monthly.
Additionally, TDS, calcium, and chloride grab samples should be run once each
week by either shift personnel or the water vendor and must appear on the vendor
service report. These reports should be issued weekly.
Critical parts of blowdown control include: assessing the total system requirements,
determining heat load velocities, and assessing metallurgy and at-risk exchangers.
Knowing this information will assist in setting proper limits on whether the blow-
down should be tied to the TDS or other water components.

224 Fouling, Scale and Corrosion


This section discusses the main reasons for heat exchanger problems caused by
cooling water; fouling, scale, and corrosion. Control and monitoring of these
phenomena are also covered.

Fouling
Fouling is defined as inorganic and organic material causing deposits that restrict
water flow and reduce heat transfer. Fouling is caused primarily by heat, pH, reten-
tion time, organisms, flow rate and concentration of ions and impurities that can
come from the water, air and the system.
Fouling can consist of:
Silt and mud
Organic hydrocarbons
Minerals-scale
Micro/Macro organisms
Clarifier addition use
Phosphates
Detergents
Sewage
Corrosion products
Corrosion inhibitors
Fouling Control. Fouling can be controlled by either chemical or mechanical
means. Chemical control attempts to suspend the particulate material until it can be
removed from the system via blowdown, sidestream filtration, or allowed to settle
out in the cooling tower basin. Control chemicals fall into three categories: dispers-
ants, sludge fluidizers, and surfactants/wetting agents. Most chemicals are types of

May 2001 200-48 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

polymers with a treatment dosage of typically 510 ppm, but the dose can exceed
this in upset conditions. Consult the water vendor as to the type and amount of anti-
foulant needed for the particular system after the foulant is identified. If the system
has potential foulants and the chemical is working, the basin water will appear
murky or muddy.
Mechanical control consists of either filtration or mechanical scrubbing of the
exchanger tubes using brushes, sponge ball systems, or acidizing. When the
cleaning is completed, the bundle must be passivated with inhibitor prior to restart.
Fouling Monitoring. Commercially available fouling (or scale) monitors connected
to sidestream piping off of the return headers are one of the best ways to determine
fouling and deposition rates on any type of tube metal at the desired temperature
and flow rate. A sidestream can be fitted with clear plastic filters, flow meters,
tubing, and site glasses to provide visual indication of deposition problems. If
tubing shows fouling forming at return water temperatures, exacerbated problems
are likely occurring at higher temperatures. Most of the on-line fouling units (avail-
able from Bridger Scientific, Drew Associates, CML Inc., among others), have
cooling water passing through a heated element or tube (usually " diameter). The
flow rate is controlled to existing operating conditions. Three common types of
monitors are pictured in Figures 200-25, 200-26, and 200-27.
Note The Hawaii Refinery uses the monitor pictured in Figure 200-25 and this
monitor is the preferred type because it can more closely simulate conditions in
actual exchangers. The device pictured in Figure 200-26 is essentially an Annular
Flow Fouling Monitor and is used in El Segundos LSFO Towers.

Fig. 200-25 Typical Fouling Monitor

Chevron Corporation 200-49 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-26 BETZ Monitall Apparatus for Monitoring Fouling (Courtesy of BetzDearborn, a division of Hercules, Inc.)

May 2001 200-50 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-27 CorrDATS 3 Fouling/Corrosion Monitor (Courtesy of Bridger Scientific, Inc. &
Rohrback Cosasco Systems, Inc.)

The fouling monitor results are calculated as Heat Transfer resistance (fouling
tendency) from:

Q = Ka T/L
(Eq. 200-13)
where:
Q = the heat transferred per unit time
K = the conductivity constant of the type of material and its
temperature
a = the area perpendicular to the heat flow
T = the temperature difference between the hot and cold sides of the
substance through which the heat is being transferred (water, in
this case)
L = the pipe length
Other Heat Exchanger calculations are:

Mass flow rate (lb/s) = Density (lb/ft3) Velocity (ft/s) Area (ft2)

or

A = A A AA
(Eq. 200-14)

Chevron Corporation 200-51 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

B = B B AB
(Eq. 200-15)

Q A = Q B = A CP ( T A T A ) = BCP ( TB TB )
A Inlet Outlet B Outlet Inlet
(Eq. 200-16)

Q = U Design A Design log ( T )


(Eq. 200-17)

( TA TB ) ( TA TB )
Inlet Outlet Outlet Inlet
Q = UA. ---------------------------------------------------------------------------------------------------
-
log ( ( T A T B )/ ( T A TB ) )
Inlet Outlet Outlet Inlet
(Eq. 200-18)
Equation 200-18 is the basic equation for a typical heat exchanger. The CFM 500
(Corrosion Fouling Monitor 500 by Capcis March Ltd., England) reports the change
in heat transfer resistance where:

1
HTR = --------
U
(Eq. 200-19)
The CML corrosion/fouling monitor and the Bridger Scientific fouling monitor
report heat transfer resistance in hour ft2 F/BTU. A typical output screen is
shown in Figure 200-28.
When using a fouling monitor:
1. Establish a base operating condition, commonly the average temperature, most
common metallurgy, and an average flow rate of the system.
2. Look at the highest temperature and lowest flow rates of that system. On-line
deposition rates may point to changes that can be made in operating conditions
or to the chemical treatment level in order to minimize fouling.
3. Consult the water vendor to determine the type of fouling monitor and level of
assistance they can provide to optimize the treatment program.

Scale
Scale is defined as primarily inorganic material. All scale is a foulant but all
foulants are not necessarily scales. Scale can provide a point of attachment for
microorganism growth and restrict and reduce the effectiveness of corrosion control
added to the cooling water.
Scale results from super-saturated cooling water minerals that exceed their solu-
bility limits and form inorganic, strongly adherent deposits in heat exchangers.
Scale is caused by a build up of minerals, cations, and anions due to evaporation,
cycles of concentration, heat load, velocity and water pH. An example of a plugged
tube in an exchanger due to calcium carbonate scale is found in Figure 200-29. The

May 2001 200-52 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments
Fig. 200-28 Typical Output Screen from CML Corrosion/Fouling Monitor
Chevron Corporation 200-53 May 2001
200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

scaling depicted in the Figure 200-29 resulted from low water velocities and high
water temperatures.

Fig. 200-29 Calcium Scaling From El Paso Rheniformer Reactor Effluent Cooler

Note Phosphate based corrosion inhibitors function by forming a calcium-iron-


phosphate inhibitor film on steel surfaces. Calcium phosphate scale can occur
in phosphate-treated systems, but this is a visible layer of material on the tube
surface.
Scale composition can be determined by sending a small portion of the scale layer
for x-ray diffraction and fluorescence spectroscopy. (Figure 200-30 depicts a fluo-
rescence spectroscopy report.) This type of report assists in determining what type
of scale is forming deposits and point to mitigation methods, discussed later in this
section. The most common forms of scale are:
Calcium carbonate (CaCO3): Limestone, forms hard oyster shell deposit
Calcium phosphate (CaPO4): Phosphate rock deposit
Calcium sulfate (CaSO4): Cement, cementous deposit in low pH
Calcium fluoride (CaFl2)
Silica (SiO2): Glass, deposits commonly found on tower fill
Iron oxide (Fe2O3)
Iron phosphate (FePO4)
Manganese oxide (MnO)
Magnesium silicate (MgSiO4)
Magnesium carbonate (MgCO3)

May 2001 200-54 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments
Fig. 200-30 Sample Fluorescence Spectroscopy Report
Chevron Corporation 200-55 May 2001
200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

There are three ways to determine the scaling tendency of cooling water as calcu-
lated by formation of calcium carbonate scales:
Ryznar Stability index: RSI=2pHs-pHact - where pHs is the saturation pH
determined by water temperature and analysis and pHact is the measured pH.
Values obtained range from 3 (severe scaling) to 10 (severely scale dissolving).
Puckorius Stability index: PSI=2pHs-pHe - where pHe is the equilibrium pH
based on the total alkalinity value, a modification that can improve the accu-
racy of assessing scaling tendencies.[23]
Langelier Saturation Index (LSI): The actual pH of a water minus its pH of
saturation for calcium carbonate. A positive quantity indicates a scale-forming
tendency; a negative quantity indicates a scale-dissolving tendency.
The following information is required to calculate the scaling tendency of cooling
water:
TDS (in ppm)
Calcium hardness (as CaCO3)
Total Alkalinity as (CaCO3)
Temperature
Actual pH
Another way to determine scaling tendencies is to calculate the pH saturation of
calcium phosphate, then determine the solubility constants of calcium sulfate and
silica. However, pH saturation is not as accurate as either the Ryznar or Puckorius
indices.
Scale Control. Scale can be controlled by either keeping scale-forming ions below
their saturation limits or by using chemical compounds called crystal modifiers to
change the nature of the crystal structure. Crystal modifiers incorporate into the
scale crystals and change their shape so they will not adhere to the metal surface.
Common Crystal modifiers are:
Polymaleic anhydride
Sulfonated polystyrene
Maleate/acrylate copolymers
These chemicals act on calcium phosphate scales to change the crystal shape from
cubic to spherical. These chemicals are used in dosage ranges of 15 ppm. Most
crystal modifiers work in systems above pH 7.5, calcium hardness of >350 ppm,
and orthophosphate levels of >15 ppm.
Another method of scale control is solubilization, particularly of calcium. Solubili-
zation can be effected by automated or manual pH control.
Scale Monitoring. Scale can be monitored by using standard weight loss coupons
(see Figure 500-25 in the Corrosion Prevention and Metallurgy Manual) attached to
the outlet of the hottest exchangers. If two coupons are put together, the resulting

May 2001 200-56 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

crevice allows scale to attach. The scaling indices should be calculated weekly to
monitor scale formation.
The calculation of scaling indices, as well as addition of phosphonates or polymers
to the chemical treatment program, are other methods for determining if scale
control is required. Phosphonates are particularly good at controlling and bonding
calcium carbonate scales and calcium phosphate and calcium sulfate scales. Typical
dosages range form 15 ppm. Polymers are also very good at controlling calcium
scales. Typical dosages here also range from 15 ppm. Consult the water vendor to
determine the proper type of chemical and dosage for the cooling water system in
question.

Corrosion
Corrosion is defined as the deterioration or dissolution of materials by their environ-
ment. Aqueous corrosion of metals is due to two electrochemical reactions: one at
the anode that oxidizes metallic iron to ferrous ion; the other at the cathode which
reduces hydrogen ion to molecular hydrogen.
Corrosion is a function of system pH, temperature, time, alkalinity, velocity,
dissolved solids, and electronic coupling of dissimilar metals. Although corrosion
and scaling can be independent phenomena, they can also occur together, and do so
frequently in cooling water systems. Corrosion can result in general or uniform wall
loss, localized pitting or cracking events.
Corrosion Control. Corrosion inhibitors attempt to short circuit the electrochem-
ical reactions by preventing either the dissolution of the metal into the water, or
prevent the reduction of hydrogen at the cathode by laying down a protective film.
Phosphate is the most common corrosion inhibitor in open recirculating cooling
water systems and primarily protects carbon steel. Calcium phosphate and iron
phosphate form a film on the exchanger tube surface to prevent further metal loss to
the environment. Typical dosage levels of ortho and polyphosphate range from 10 to
20 ppm. This level is measured as residual in the system in order to achieve mild
steel corrosion coupon rates of 02.5 mpy in the return water. Phosphates are a
nutrient source for microorganisms so it is essential to optimize the phosphate treat-
ment program along with the biocide control program.
Phosphate-based corrosion inhibitors require a pH range of 6.57.5 pH units. This
range provides just enough calcium to bind to phosphate for corrosion protection.
This is done by addition of sulfuric acid adjusted to the pH set-point. Figure 200-31
details an acid addition system.
When adding inhibitor, you should strike a balance between injecting a sufficient
amount to form a film on a new or recently cleaned surface, and maintaining that
film without adding so much inhibitor that unwanted phosphate rock deposits form.
A partially filmed surface has an even greater tendency to corrode because it sets up
a potential difference from the filmed to unfilmed portion of the metal surface.
Tolyltriazole (TTA) is the corrosion inhibitor used for yellow metal protection. It is
normally measured as a residual of 11.5 ppm, or enough residual to produce
corrosion rates on admiralty brass of less than .2 mpy. TTA can be scavenged by

Chevron Corporation 200-57 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-31 Typical Acid Feed System

chlorine, however, so that during chlorine shock treatments at high chlorine resid-
uals, more TTA may be required to maintain a measurable residual in the system. It
is essential to establish the proper dosage level of TTA because of its expense.
Corrosion Monitoring. Corrosion weight loss coupons are a good way to monitor
corrosion in the system. Coupons can be located in three places:
Inserted into a 1" or " PVC sidestream assembly in the tower return water
riser.
Inserted into the piping system downstream of a critical heat exchanger.
It is generally preferred to place coupons on return risers because corrosion is
affected by temperature. Return water will be warmer. Also, chlorine injected
into the tower basin will affect corrosion rates and biological fouling.
Corrosion measurements that are made close to true tube wall temperatures will be a
better diagnostic tool than a coupon in the cooling tower basin. At minimum, install
coupons in return headers first, then consider installation at heat exchanger outlets.
A typical design is shown in Figure 200-32. Standard coupon assembly is pictured
in Figures 200-33 (below) and 500-22 of the Corrosion Prevention and Metallurgy
Manual.

May 2001 200-58 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments
Fig. 200-32 Examples of Corrosion Test Rack Installations
Chevron Corporation 200-59 May 2001
200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-33 Standard Coupon Assembly

Coupons should not be pretreated with inhibitor prior to insertion in the holder or
basin. Pretreated coupons should be used for comparison with untreated coupons
only. Phosphate levels should be adjusted to attain mild steel corrosion rates in
tower basins less than 2 mpy. Admiralty rates of .1.2 mpy and 304SS rates of less
than .5 ppm are considered acceptable. Visual inspection of two crevice coupons on
the outlet of hot exchangers is an excellent tool to determine under deposit attack
due to scaling and/or fouling.
Stressed U-bend coupons inserted in piping on the outlet of a hot exchanger can
yield valuable information on the cracking tendencies of yellow metals.
On-line instantaneous corrosion measurements using either a Corrator or other elec-
trochemical devices can assist in evaluating process and inhibitor effects. (Figures
500-16 and 500-17 of the Corrosion Prevention and Metallurgy Manual provide
information on polarization meters and corrator probes.) A small (1020 mv) poten-
tial is applied to one electrode (anode) and the current is measured at the cathode. If
you know the current and potential, you can solve for the polarization resistance
value, or Rp, which is inversely proportional to corrosion rate. This probe assembly
is pictured in Figure 200-34. Measurements made at ambient temperatures will not
be an accurate measure of the corrosion in exchangers under heat load. The closer
any measurement is made to operating temperature, the more indicative it will be of
heat exchanger performance.
On-line pH monitors must be routinely calibrated once a month with a hand held pH
meter so that no over- or under-deliver acid can occur. Significant corrosion prob-
lems arise when acid over-addition occurs. Likewise, significant scaling problems
can arise when recovering from an acid over-addition by addition of caustic to
modulate the pH.
Two new on-line monitors are available that can allow for measurement of corro-
sion rates at the temperature and flow rates of the hottest exchangers. These moni-
tors are available from Bridger-Rorhbach and CML.

May 2001 200-60 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-34 Detail View of Corrosion Coupon/LPR Probe Holder

High corrosion rates in the tower basin indicate acid addition problems. Acid addi-
tion pumps, on-line pH meters and inhibitor addition pumps should be checked as
soon as a problem is detected. High corrosion rates can also occur with a shock
addition of biocide after dip slides or plate counts of microorganisms are out of the
specified control range.

225 Biocide Monitoring and Control


There are three main classes of organisms that can cause corrosion or fouling, or
pose a serious human health threat in the cooling system:
Fungi
Algae
Bacteria (sulfate reducers, acid producers, nitrifying bacteria, iron bacteria, and
legionella)
Sessile microorganisms produce a biofilm layer made up of complex sugars. These
microorganisms secrete these substances to attach themselves to each other and to
tubes, piping and other equipment, such as cooling tower fill. This is seen as slime
formation which may contain bacteria, algae, and fungus. Slime is the principal
biologic problem in cooling water systems.
Planktonic bacteria (free floating) like legionella, can pose a health risk to refinery
workers through tower drift and evaporation. Nitrifying bacteria oxidize ammonia in
the water to nitric acid, thereby reducing the pH of the water. Rapid general thin-
ning of the carbon steel and copper-based alloys can occur.

Chevron Corporation 200-61 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fungi have the most adverse effects on cooling tower lumber. To mitigate fungal
growth, treat the lumber with copper-arsenite or creosote, usually during the design
and construction phase of the tower.
Algae require sunlight, carbon dioxide, and trace minerals as a food source, all of
which are present in cooling waters. Algal growth can be seen on decks, down-
comers, and piping systems which can change or restrict water distribution. Algae
can live in pH ranges from 4 to 9, and in temperatures of up to and sometimes
exceeding 120F. Dead algae absorb biocides, increase the thickness of fouling
layers already formed and plug screens filters and exchangers.
Biocide monitoring. A delta p-monitor is one tool for monitoring bacteria (see
Figure 200-35). This type of monitor examines the change in pressure across a side-
stream filter. Delta p-monitors typically measure pressure drop over a given length
of pipe. Higher delta p indicates biological growth. The monitor takes water
samples from the basin, return header, sessile monitor, or scrapes a fouling layer off
of a piece of equipment or filter. The delta p-monitor then runs either dip slides or
plate counts in order to get a general accounting of the amount of both sessile and
planktonic organisms in the system.

Fig. 200-35 Biofouling Monitor

Control ranges should be in the order of 102 104 microorganisms per milliliter.
Above 105 the system is considered out of control and biocide addition must be
increased until the count returns to normal levels. These tests should be run on a

May 2001 200-62 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

weekly basis and must be run after a hydrocarbon leak is detected (hydrocarbon is a
food source for microorganisms). High ammonia levels in a tower basin can point to
a nitrifying bacteria. In this case, experts should be consulted and treatment initi-
ated. Examples of microorganism attack are seen in Figure 200-36.

Fig. 200-36 Examples of Microorganism Attack. From Chemical Processing, Bob Strack, ed.,
Putnam Publishing.

Serial dilutions, another bacteria monitoring tool, distinguishes between different


classes and amounts of bacteria and should be run when tubercles or pitting is found
in an exchanger during shutdown. This test will indicate the type and level of
bacteria present and will assist both the water vendor and the environmental expert
in mitigating the problem.
Biocide control. Oxidizing chemicals are the most common and least costly type of
biocide control in use in open recirculating systems because they effect a wide range
of slime producing organisms. These agents destroy both the micro and macro
organisms and their food source, and are compatible with other treatment chemi-
cals. Chlorine gas, chlorine dioxide, sodium hypochlorite and chlorine-activated
bromine are the most common oxidizers. Oxidizing chemicals are added to the back
end of the tower by slug (shock) or by continuous chemical injection. Residuals
should be measured at the return header where free Cl2/Br2 range between .1.5
ppm by grab sample; total Cl2/Br2 is commonly between 1.04.0 ppm, except under
upset conditions where it can go as high as 10 ppm. Chlorine residual grab samples
should be taken at least once a shift.

Chevron Corporation 200-63 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

A better approach to biocide addition is to control the biocide injection pump with
readings from an Oxygen Reduction Potential (ORP) electrode placed in the PVC
sidestream. This device, pictured in Figure 200-37, measures the potential differ-
ence between gold and platinum electrodes in conductive solutions. Standard test
solutions of between 300-600 millivolts calibrate the ORP output to correspond to
the needed chlorine residual of .5 ppm. The reading output is tied to the chlorine
metering pump. This control loop ensures that the minimum biocide needed is
always present in the system.

Fig. 200-37 Aromax Cooling Tower ORP Controller for Cl2

226 Water Quality Monitoring: How to Read a Water Service Report


The water service report is one of the most important tools in addressing cooling
water problems. Water reports assess the components present in both the make-up
water and recirculated return water. Analyses of make-up water should be run by the
water vendor monthly. Because of the possibility of biocontaminents and ammonia,
water analyses should be run weekly for recirculated or reclaimed water. Inter-
preting the water reports will highlight potential corrosion, fouling, hydrocarbon,
and microorganism problems.

May 2001 200-64 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

A water report should contain most of the following information:


Free halogen (chlorine or bromine) ppm
pH
Turbidity - Nephelometer (Normal) Turbidity Unit (NTU)
Phosphate
Ammonia
Microbiologic counts (can include total coliform bacteria and sulfate reducing
bacteria levels)
Total suspended solids (TSS)
Total dissolved solids (TDS)
Carbon oxygen demand (COD)
Total Organic Carbon (TOC)
Anions
Cations
Conductivity (on-line and portable)
Alkalinity
Calcium
Magnesium
Total hardness
Total phosphate
Azole residual
Weight loss coupon corrosion rates; Corrator or instantaneous rates
Iron counts
Copper counts
Zinc
Chlorides
Sulfate
Silica
Oxygen Reduction Potential (ORP); on-line and portable
Cycles of concentration
Mechanical equipment operating checks
Ryznar or Puckorius Stability index
Water analyses can be reported and calculated as shown in the three examples
depicted in Figure 200-38. Examples of two vendor water reports are contained in
Figure 200-39 and Figure 200-40.

Chevron Corporation 200-65 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-38 Sample Water Analysis

Example 1 Calcium 50 PPM


Magnesium 15 PPM
Ionic-anions and cations Chloride 50 PPM
Reported as PPM (Mg/L)
Sulfate 20 PPM
Bicarbonate200 PPM

Example 2 Calcium 50/.4=125 PPM


Magnesium 15/.24=62.5 PPM
Equivalents to Calcium Chloride 50/.71=70.4 PPM
Carbonate
(temporary hardness) Sulfate 20/.96=20.8 PPM
Bicarbonate200/1.22=164 PPM
Conversion factors of ions expressed
as CaCO 3
for Example 2
CaCO 3=Ca/.4
CaCO 3=Mg/.24
CaCO 3=OH/.34
CaCO 3=CO 3/.6
CaCO 3=HCO 3/1.22
CaCO 3=Cl/.71
CaCO 3=SO 4/.96
CaCO 3=Na/.46

Example 3 Calcium= 50PPM/20.05= 2.5 epm

Equivalents per Million Magnesium=15 PPM/12.15=1.23 epm


Based on the Equivalent Chloride=50/35.5=1.41 epm
Weight of the Ion Sulfate=20/48=.42 epm
Bicarbonate=200/61=3.3 epm

Equivalent weight calculations for Example 3


ion Molecular weight valence electrons equivalent weight
Ca 40.1 2 40/2=20.05
Mg 24.3 2 12.5
Na 23 1 23
OH 17 1 17
CO3 60 2 30
HCO 3 61 1 61
Cl 35.5 1 35.5
SO4 96.06 2 48.03

May 2001 200-66 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-39 Ashland Chemical Cooling Water Service Report

Chevron Corporation 200-67 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Fig. 200-40 Betz Industrial Cooling Water Service Report

May 2001 200-68 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

227 Chemical Treatment Program Control: Monitoring and Evaluation


Cooling water chemical treatment programs do not treat the water; rather, the water
carries the treatment chemicals to the metal surface of the heat exchanger. The
chemical treatment program determines how chemicals will mitigate fouling, scale,
and corrosion problems in the cooling water system. Typical chemical treatment
programs include:
Corrosion inhibitors - phosphates for carbon steel, tolyltriazole for yellow metal
Scale inhibitor - phosphonates
Dispersant/flocculation agent - surfactants
sulfuric acid, caustic
oxidizing biocide - chlorine or bromine
For purposes of cost containment it is important ensure that the correct amount of
chemical is delivered to the system. Each water vendor has their own method of
monitoring chemical addition. NALCOs Tracer technology adds a dye to the
dispersant which is proportionally mixed and injected along with the phosphate
inhibitor. Residual dispersant and phosphate are measured by a fluorescence spec-
trometer on the recirculating water to maintain a control range of inhibitor residual
that produces mild steel corrosion rates of less than 2.5 mpy. Other companies auto-
mate feed based on make-up and blowdown rates. Even with automation, it is neces-
sary to sample the recirculating water for phosphate residuals at least once a shift as
a control check.
TTA can be measured on-line with a UV spectrometer to a specified residual. It
should be noted that chlorine will react with TTA causing it to absorb at a different,
unmeasured wave length. If low TTA levels are found, chlorine residual should be
checked before increasing the TTA to a readable residual.
Biocide addition should be checked on a per shift basis by a hand held ORP meter
or an automated on-line ORP meter tied to a logic pump for continuous injection.
The ORP meter should be calibrated monthly and the reading verified by a hand
held meter and standard millivolt solutions to register .5ppm residual chlorine.
Run weekly dip slides or bug bottles and monthly serial dilutions to determine any
type of microbiologic activity. Sulfate Reducing Bacteria (SRB) or coliform, if
suspected, should also be run weekly.
Hydrocarbon leaks can promote bug growth. Total organic carbon (TOC) samples
should be taken once a day and the source of the leak detected as soon as possible.
See A New Approach to Detect and Identify Hydrocarbon Leaks in Refinery
Cooling Systems, by Chris Spurrell for assistance with hydrocarbon leak source
detection. [21]
pH addition should be automated with instructions available to control-room
personnel. Excursions should be indicated by an alarm. pH should be checked
weekly with a hand-held, calibrated pH meter. All on-line pH meters should be cali-
brated once a month. pH should not vary by more than .2 pH units.

Chevron Corporation 200-69 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Corrosion and scale inhibitors should be maintained at adequate levels to minimize


corrosion and prevent the formation of scale. pH excursions below the control range
should be viewed as promoting corrosion, while pH excursion above the control
range promote scale formation. The vendor should be consulted if upset conditions
persist. Additional acid, dispersant, and flocculating agents may be needed to
control the excursion.
Changes in the magnitude of the heat transfer resistance as measured by a fouling
monitor, delta p filters, or by visual inspection of weight loss coupons on the outlet
of the hottest exchangers may point to scaling problems. The water vendor should
be consulted if sufficient deposition occurs.
Treatment program baseline data gathering and subsequent optimization is neces-
sary for both cost containment (costs are delineated in Figure 200-41) and program
effectiveness. Baseline data gathering should be conducted for a minimum of three
months and should include:
Measurements of corrosion rates with weight loss coupons, LPR instantaneous
corrosion rates, pitting rates by inspection, Rorhbach pitting imbalance read-
ings, or electrochemical noise localization indices calculated from the CML
corrosion-fouling monitor
ppm control limits for current chemical inhibitors
Fouling factors calculated through a deposit monitor and/or C factor
calculations
Flow and skin temperature conditions of average exchanger of most predomi-
nant metallurgy
Microbiologic control ranges, composed of serial dilutions and dip slides
Tower fill and basin condition
Number and type of at-risk exchangers - metallurgy, flow rates, skin
temperatures
Passivation procedures
Start-up, lay-up requirements
Upset conditions and their resolutions
This information can then be used to optimize the treatment program or to evaluate
a new program. Most vendors have research facilities that can simulate cooling
towers at true operating conditions in order to prevent adverse effects on heat
transfer surfaces when changing treatment programs or adjusting injection levels.

May 2001 200-70 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Fig. 200-41 Cooling Water/Chemical Savings by Increasing Cycles

228 Glossary
Acidity: The acid content of water in excess of that needed to neutralize the alka-
linity expressed as ppm or mg/l calcium carbonate.
Adsorption: Physical adhesion of molecules to surfaces of solids without chemical
reaction.
Aeration: Blowing air through water to sweep out other dissolved gases and to
equilibrate it with oxygen and carbon dioxide.
Aeration: .Aeration causes precipitation of iron and manganese oxides which can
then be removed by either in-line or sidestream filtration.
Aerobic: Bacteriological life which requires the presence of air for growth and life
cycle.
Algae: Simple forms of aquatic life which use sunlight for growth, known as photo-
synthesis; always colored green or blue-green and form "stringy" deposits on
cooling tower beams or sides and also suspended solids in piping systems. They are
the cause of most tastes and odors in water.
Alkalinity: Also known as basicity. The ability of water to react with hydrogen
ions. In water treatment it means amounts of carbonate, bicarbonate and hydroxide
that react with acid to a titration endpoint of 4.5.
Anaerobic: Bacteriological life which does not require air for growth. Oxygen is a
poison to these bacteria.
Anionic: An ion, colloid particle or metallic surface containing a negative charge.
Anionic/Cationic: Anionic/cationic ion exchange resins can be used to remove
excess components and alkalinity in much the same way as the lime-soda softening
process.

Chevron Corporation 200-71 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Anode: In the corrosion process, the area where metal is removed. The pH at the
anode is generally lower than the pH of the water.
Backwash: The process of reversing the normal direction of flow through filters,
heat exchangers, and ion exchange equipment at a rate sufficient to expand the bed,
scour the medium, and remove any material deposited upon the bed or in the
exchanger.
Biochemical Oxygen Demand: The concentration of oxygen required for the
biological and chemical oxidation of organic matter and reducing agents in the
effluent.
Biological Fouling: Deposits or growths of biological organisms such as algae,
bacteria and fungi and the products of their life cycles, often called "slime".
Blowdown: Bleed-off; the releasing of water from a system, such as a boiler or
cooling tower to decrease the concentration of suspended and/or dissolved solids.
Cathode: In the corrosion process, the area of metal which is not removed but
controls the rate of corrosion. The pH at the cathode is generally higher than the
water.
Cationic: An ion, colloid particle or metallic surface containing a positive charge.
Chlorination: The treatment of water with chlorine or chlorine releasing
compounds.
Clarification: Clarification is the process of adding coagulating and flocculating
agents to speed up settling. Alum is an example of a flocculation agent. Jar testing is
the most common way to evaluate clarification agents. Add surfactants to the jars of
make-up water, agitate and observe the settling time.
Coagulation: The neutralization of a single colloidal particle which allows it to
agglomerate with other like neutralized particles and/or settle under the influence of
gravity.
COD: Chemical oxygen demand; the quantity of oxygen, furnished by a chemical
oxidant, needed to oxidize the carbon compounds in solution.
Colloidal Matter: Matter of very fine particle size, usually in the range of 10-5 to
10-7 cm in diameter. It forms dispersions intermediate in character between a true
solution and a suspension.
Conductivity: The ability of water to conduct electricity. When measured with a
standard apparatus, it is called specific conductivity and is a function of the total
dissolved solids. TDS=2/3 specific conductance in micromhos/cm.
Controlled Scale: The adjustment of the chemical properties of a water to permit a
thin protective scale to be deposited on a metal surface.
Corrosion: Destruction of a metal by a chemical or electrochemical reaction with
its environment.

May 2001 200-72 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Cycles of Concentration (COC): The ratio of any one or all of the minerals in the
source water to that in the system water. This is obtained by dividing the system
water minerals by the source water. Two cycles mean that the system water contains
twice the mineral content as that of the source water.
Dealkalization: Exchange of bicarbonate for chlorides in an ion exchange process.
Deionization: Same as demineralizing; the removal of all of the ions in the water.
Demineralization: Removal of all ions from water; reduction of the mineral content
of water by a physical or chemical process; removal of salts.
Deposits: Any materials which form accumulations such as scale, sludge; may be
mineral, microbiological or oils.
Dispersant: Chemicals added to water systems to keep insoluble solids suspended
or dispersed. Dispersants are used to prevent accumulation of deposits or sludge.
Dissolved Solids: That matter, exclusive of gases, which is dispersed in water to
give a single phase of homogeneous liquid. (May be determined by calculation
(addition of individual constituents) or by evaporation to dryness and weighing the
residue). (Note: these may not agree; certain adjustments are required to offset
losses or gains of CO2 in drying, etc.) See also, TDS.
Electrolyte: Water containing charged ions.
Equivalent Weight: Molecular or atomic weight divided by valence. Calcium has
an atomic weight of 40 and a valence of 2. Hence, its equivalent weight is 20.
Equivalents per Million: EPM; a unit chemical equivalent weight of solute (the
substance dissolved) per million unit weights of solution. Concentration in EPM is
calculated by dividing concentration in parts per million by the equivalent weight
(combining weight) of the ion or substance.
Filter: A device of structure for removing solid or colloidal matter (which usually
cannot be removed by sedimentation) from water. This uses a straining process
where the solids are held on a medium of some kind (granular, diatomaceous earth,
woven, porous) while the liquid passes through.
Flocculation: The neutralization of more than one colloidal particle by the same
agent. The resulting, relatively large neutral particle will fall out of suspension as a
result of gravitational attraction.
Foulant: Usually any suspended material that deposits on a heat exchanger. This
causes loss of efficiency in either heat transfer or flow.
Fouling: The deposition of heat transfer surfaces of material normally in suspen-
sion.
Free CO2: A term used to designate the carbon dioxide gas dissolved in water and
not combined in the form of bicarbonates or carbonate ions.

Chevron Corporation 200-73 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Galvanic Corrosion: The corrosion occurring when two dissimilar metals are in
contact and corrode. An example of galvanic corrosion is when copper and mild
steel are in direct contact resulting in rapid mild steel corrosion.
Hardness: A characteristic of water generally accepted to represent the total
concentration of calcium and magnesium ions. It usually is expressed in terms of the
equivalent amount of CaCO3, e.g., 50 ppm as CaCO3, originally understood to be
the capacity of water for precipitating soap. It was measured by the amount of a
standard soap solution required to form a stable lather on a measured amount of
water, with alkalinity adjusted to eliminate effect H+ ions. (Alkalinity adjusted to
about pH 8.3.) Iron, manganese, aluminum, and hydrogen ions also precipitate soap
solution, so this is not considered a precision method of hardness determination.
Hardness, Calcium: The calcium compounds dissolved in water, expressed as
calcium carbonate.
Hardness, Carbonate: The calcium and magnesium carbonate and bicarbonate
dissolved in water, expressed as calcium carbonate.
Hardness, Magnesium: Magnesium compounds dissolved in water, expressed as
calcium carbonate.
Hardness, Non-Carbonate: The difference between the total hardness and the total
alkalinity of a water.
Hardness, Permanent: The hardness that cannot be removed from water by
boiling. Essentially, the same thing as the non carbonate hardness.
Hardness, Temporary: The hardness that can be removed from water by boiling.
Essentially the same as the as the carbonate hardness.
Hardness, Total: The sum of the calcium and the magnesium hardness. Also the
sum of the permanent and temporary hardness.
Hydrogen Ion Concentration: Commonly expressed as the pH value which repre-
sents the logarithm of the reciprocal of the hydrogen ion concentration.
Inhibitor: A material that reduces a normal tendency to cause scale or corrosion.
Usually used to describe chemicals that minimize corrosion through formation of
protective films on a base metal.
Ion: An atom or radical in solution carrying an integral electrical charge, either
positive (cation) or negative (anion).
Ion Exchange: A process where water is passed through a granular material and
ions on the granular material are replaced by ions contained in the water. For
example, in the zeolite softening process, the sodium ions (Na+) of the granular
zeolite are replaced by the calcium ions (Ca++) in the water, to leave the water free
of calcium (the cause of hardness, but with an increased amount of sodium).
Ion Exchange Material: An insoluble solid which has the ability to reversibly
exchange certain ions in its structure (or attached to its surface as functional groups)
with ions in a surrounding medium.

May 2001 200-74 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Lime Soda Softening: Lime-soda softening pretreatment reduces hardness, alka-


linity and silica. The water is slurried with the lime-soda mixture so that precipita-
tion can occur. Silica is removed by co-precipitation with magnesium in the water.
Some magnesium may be added in the form of dolomite or magnesium oxide if not
enough magnesium is naturally present.
Lime-Soda Ash Softening: A process of softening water by adding lime to precipi-
tate carbonate hardness and soda ash to precipitate noncarbonate hardness, with
subsequent removal of precipitate by sedimentation and filtration.
Makeup: A term used to distinguish that portion of water which is added to a boiler
or cooling system to compensate for water losses by blowdown (bleed-off), evapo-
ration, condensate losses, and stream consumption in process. Makeup can be raw
or pretreated water.
Micromho: An electrical unit of conductance (one millionth of an mho) which is
the reciprocal of electrical resistance as measured by a megohm-meter. It represents
an indirect measure of the dissolved solids in water.
Open Recirculating Water System: A system in which water continually flows
through piping and heat exchangers and is open to the atmosphere where evapora-
tion takes place such as in cooling towers, air washers, spray ponds, etc.
ORP: Oxidization-reduction potential. It is a characteristic of water useful in deter-
mining its corrosive property or lack of it.
Oxidizing Biocide: A biocide whose effectiveness depends upon its ability to
oxidize and thus, destroy organic material. Examples: chlorine, bromine, ozone.
P Alkalinity: Where the basicity of water consumes acid to a pH of 8.5. It can be
expressed as the carbonate alkalinity and the total OH alkalinity.
pH: The logarithm of the reciprocal of the hydrogen ion concentration of water. It is
a measure of the acidity or alkalinity. Water with a pH of 7.0 is neutral. A pH
greater than 7.0 indicates an alkaline water, a pH less than 7.0 indicates an acid
water.
(pH)e: Equilibrium pH; the pH a circulating water assumes when in equilibrium
with the contacting air. It is a function of pH alkalinity and temperature.
Phosphates: Chemicals used for corrosion control in cooling towers, and for
deposit control in boilers. Commonly these occur as orthophosphates or polyphos-
phates. The level of the active phosphate chemical is reported either as percent P2O5
(phosphorus pentoxide) or as PO4 (phosphate) with these related by a factor as
follows: PO4 = 1.34 x P2 O5.
Polymer: A naturally occurring or synthetic substance consisting of giant mole-
cules formed from smaller molecules of the same substance and often having a defi-
nite arrangement of the components of the giant molecules.
PPM: Parts per million; a part per million is one pound (or any other weight unit)
per million pounds (or any other weight unit) of water.

Chevron Corporation 200-75 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

Practical Index: The same as the Puckorius Index.


Puckorius Scaling Index (PSI): A modification of the Ryznar Stability Index
(RSI), indicating the tendency of water to form calcium carbonate scale. This is
determined in the same way as the RSI, except the pH of the water is not used. An
estimated pH (pHe) is used based on the total alkalinity of the water. The resulting
values obtained give the same indications of scaling tendencies.
Rough Screening: This method removes large pieces of debris (such as logs) and
organisms (such as fresh water clams) from the cooling water. The screens are most
commonly located in the cooling tower basin, upstream of the pump suction.
Ryznar Stability Index (RSI): Two times the (pH)s of a water minus the actual pH.
If this value is less than 6.0, the water has a scale dissolving tendency.
Saturation: The condition of a liquid when it has taken into solution the maximum
possible quantity of a given substance at a given temperature and pressure.
Scale: The deposition on heat transfer surfaces of material normally in solution.
Sedimentation: Sedimentation can occur in a holding reservoir, such as a tower
basin, and removes silts and sands or any particles of .011mm in diameter. Smaller
particles must be removed by clarification.
Sidestream Filtration: Filtration is effected by passing a slipstream or sidestream
from the circulation loop over a bed of porous material and returning it back to the
loop after filtering. As a cake of material forms on the sand or coal, it becomes the
primary filter medium. When the flow through that filter decreases, the water flow
must be reversed (backwashed). See Figure 200-42 for a schematic of a filtration
system.

Fig. 200-42 Self-Cleaning Automatic Cooling Water Filter

May 2001 200-76 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

Slime: Gelatinous deposits that may be chemical, but usually are microbiological
growths which entrap other insoluble materials from the water and cause loss of
heat transfer.
Sludge: Suspended or precipitated solids formed in the main body of boiler water
by reactions with corrosion products or by water softening reactions. If these drop
out onto heat exchanger surfaces, they may "bake on" and resemble "in situ" (crys-
talline) scale formation.
Softening, Water: The process of removing hardness minerals from water. There
are two softening processes in general use: chemical precipitation (lime-soda soft-
ening) and the zeolite ion exchange process.
Solubilization: Anything that increases the solubility of scale forming materials,
including pH, lowering, and sequestering.
Supersaturated Solution: A solution which contains more of a substance than can
normally be held in solution at a given temperature.
Suspended Solids: That matter, exclusive of gases, existing in the non-liquid state
which is dispersed in water to give a heterogeneous mixture. (Can be determined by
the difference between weighted evaporation residues from filtered and unfiltered
samples.
TDS: Total dissolved solids; the sum of the organic and inorganic materials
dissolved in water.
TOC: Total organic carbon; the carbon contained in organic materials dissolved in
water.
Total Alkalinity or M alkalinity: Where the basicity of water consumes acid to a
pH of 4.5. It is the sum of the carbonate, bicarbonate, and hydroxide alkalinities and
can be expressed as either PPM or equivalents of CaCO3.
Turbidity: The reduction of transparency of a liquid due to the scattering of light by
suspended particles. (The terms turbidity and suspended solids are closely
allied, but not identical. Suspended Solids is the quantity of material in water that
can be removed by filtration. Turbidity is a measurement of the optical obstruction
of light passed through the water sample. The primary standard for measuring
turbidity or for calibration of turbidimeters is the Jackson candle turbidimeter. The
former insoluble silica suspension method has been superseded as the measurement
standard.)
Valence: The number of charges, either positive or negative, associated with an ion.
Water Quality: The chemical, physical, and biological characteristics of water with
respect to its suitability for a particular purpose. The same water may be of good
quality for one purpose or use, and bad for another, depending on its characteristics
and the requirements for the particular use.
Zeolite Process: The same as ion exchange. The process of softening water by
passing it through a substance known in general as a zeolite, which exchanges
sodium ions for hardness constituents in the water.

Chevron Corporation 200-77 May 2001


200 Corrosive Environments Corrosion Prevention and Metallurgy Manual

230 References
1. Machine Design, January 18, 1968.
2. Guidelines for Selection of Marine Materials, International Nickel Company,
March 1975.
3. Materials for Sea Water and Brine Recycle Pumps, INCO, January 1972.
4. Uhlig, H. H., ed., The Corrosion Handbook. Sponsored by the Electrochem-
ical Society, Inc., New York, John Wiley, 1948.
5. Evans, U. R., Metallic Corrosion, Passivity and Protection. New York, Long-
mans, Green & Company, 1948.
6. Manning, J. A. & Carleton, S. V., A Study of Corrosion Rates in an Operating
Desalination Plant, Materials Performance, August 1975, pp. 9-14.
7. Schrieber, C. F. & Coley, F. H., Behavior of Materials in Desalting Environ-
ments - Seventh Progress Report (Summary), Materials Performance,
Volume 15, No. 7, 1976.
8. Todd, B. & Oldfield, J., The Use of Stainless Steels and Related Alloys in
Reverse Osmosis Desalination Plants, International Maritime Conference,
Bahrain, October 1964.
9. Oldfield, J. W. & Sutton, W. H., A New Technique for Predicting the Perfor-
mance of Stainless Steels in Sea Water and Other Chloride-Containing Environ-
ments, Br. Corros. J., 1980, Volume 15, No. 1.
10. INCO Nickel Topics, Volume 16, No. 7, 1963.
11. Bates, J. F. & Poplewell, J. M., Corrosion of Condenser Tube Alloys in
Sulfide Contaminated Brine, Corrosion, Volume 31, No. 8, August 1975.
12. Schmitt, R. J. & Phelps, E. H., Corrosion Performance of Constructional
Steels in Marine Applications, Journal of Metals, March 1970.
13. Corrosion Resistance of Wrought 90/10 Copper-Nickel-Iron Alloy in Marine
Environments, INCO, February, 1975.
14. May, T. P. & Weldon, B. A., Copper Nickel Alloys for Service in Sea Water.
Paper given before the International Congress on Fouling and Marine Corro-
sion, Cannes, Frances, June 8-13, 1964.
15. Copson, H. R., Effects of Velocity on Corrosion. Corrosion, Volume 16,
No. 2, February 1960, pp. 86t-92t.
16. Field, P., Recommendation for Using Steel Piping in Salt Water Systems.
Journal of the American Society of Naval Engineers, Volume 57, No. 2,
February 1945, pp. 1-20.
17. LaQue, F. L. & Tuthill, A. H., Economic Considerations in the Selection of
Material for Marine Applications. Transactions of the Society of Naval Archi-
tects and Marine Engineers, Volume 69, 1961, pp. 619-639.

May 2001 200-78 Chevron Corporation


Corrosion Prevention and Metallurgy Manual 200 Corrosive Environments

18. Rheingans, W. M., Resistance of Various Materials to Cavitation Damage.


Report of the 1956 Cavitation Symposium, American Society of Mechanical
Engineers, 1957.
19. Brouillete, C. V. & Hanna, A. E., Second Corrosion Survey of Sheet Steel
Piling. U. S. Naval Civil Engineering Laboratory, Port Hueneme, CA 1965.
20. Morton, B. B., Splash Zone Protection for Offshore Structures. World Oil,
Volume 144, No. 1, January 1957, pp. 125-128.
21. Spurrell, Chris, A New Approach to Detect and Identify Hydrocarbon Leaks
in Refinery Cooling Systems, NACE Annual Conference, Paper 610, 1996.
22. The source for Equation 200-3 through Equation 200-10 is Puckorius and Asso-
ciates Cooling Water Seminar, Evergreen Colorado, 80439, 1994.
23. Puckorius, P.R. and Brooke, J.M., A New Practical Index for Calcium
Carbonate Scale Prediction in Cooling Tower Systems, Corrosion, Volume 47,
No. 4, April 1991.

Chevron Corporation 200-79 May 2001

You might also like