Sei sulla pagina 1di 24

100 Fundamentals of Corrosion Mechanisms

Author: J.W. Coombs

Abstract
This section describes some basic corrosion mechanisms of concern to the petroleum industry. It covers basic chemical corrosion theory and equations, and biological corrosion problems and control; more complete discussions are provided elsewhere in the manual. It also contains a glossary that is pertinent to the entire manual. Contents 110 120 121 122 123 124 130 131 132 140 141 142 143 150 160 Introduction Wet Corrosion Electrochemical Reactions Electrode Potentials Polarization and Corrosion Rates Forms of Wet Corrosion Dry Corrosion High-temperature Oxidation Hydrogen Sulfide Corrosion Biological Corrosion Influence of Microorganisms Control of Corrosion Caused by Microorganisms Influence of Macroorganisms Glossary References 100-20 100-23 100-16 100-15 Page 100-2 100-2

Chevron Corporation

100-1

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

110 Introduction
Corrosion is the deterioration or destruction of a material, or its properties, because of its reaction with an environment. Unfortunately, most environments are corrosive to some degree, so corrosion cannot be completely avoided. However, by addressing the fundamentals, corrosion can be controlled in an economical and safe manner. Corrosion control is essential for a facility to operate safely and productively. Approximately 25 billion dollars is believed to be spent annually on corrosion problems which could have been eliminated or reduced by addressing corrosion fundamentals. Ideally, corrosion fundamentals should be considered prior to plant construction to reduce apparent costs such as maintenance, shutdowns, and contamination/loss of valuable products, and also hidden costs such as safety and reliability. Proper inspection and maintenance is also essential to reduce the number of corrosion failures. The material in this section is presented under three general headings: 120 Wet Corrosion 130 Dry Corrosion 140 Biological Corrosion

Wet corrosion, as the name implies, occurs only in the presence of aqueous solutions or electrolytes. Dry corrosion occurs when no liquid is present, or the temperature is above the dew point of the environment. Biological corrosion is a specific type of wet corrosion. These classifications of corrosion phenomena are arbitrary. No single system of terminology has received general acceptance.

120 Wet Corrosion


Wet corrosion is an electrochemical process, and electrochemical reactions are used to explain it. Wet corrosion, as the name suggests, occurs only in the presence of aqueous solutions (which include condensation from moisture) or electrolytes and is the most commonly encountered type of corrosion. Wet corrosion can be further categorized by the appearance of the corroded metal. These categories, given with the relevant sections of this manual, include the following: Uniform Corrosion (Sections 100 and 210) Galvanic Corrosion (Sections 100 and 210) Crevice Corrosion (Section 210) Pitting (Section 210) Intergranular Corrosion (Sections 210 and 410) Selective Dealloying (Sections 100 and 210) Erosion corrosion (Section 100) Stress Corrosion Cracking (SCC) (Sections 210 and 420)

August 1999

100-2

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

121 Electrochemical Reactions


Metals corrode through the simultaneous occurrence of oxidation and reduction reactions. Oxidation reactions produce electrons and put ions into solution. They occur at the anode and are therefore also called anodic reactions. The anode of a corrosion cell corrodes. Reduction reactions consume the electrons produced by oxidation reactions. Reduction reactions are also called cathodic reactions; they occur at the cathode, which does not corrode. Examples of oxidation reactions are: Fe Fe2+ + 2e(Eq. 100-1)

Al Al3+ + 3e(Eq. 100-2)

The general form for the oxidation (or corrosion) of metal M is: M Mn+ + ne(Eq. 100-3)

Ions will also oxidize in the presence of oxidizing agents. For example, ferrous ions are oxidized to ferric ions: Fe2+ Fe3+ + e(Eq. 100-4)

Common reduction reactions are: (H2 evolution) 2H+ + 2e- H2


(Eq. 100-5)

(O2 reduction in acid) O2 + 4H+ + 4e- 2H2O


(Eq. 100-6)

(O2 reduction in neutral or basic solutions) O2 + 2H2O + 4e- 4OH(Eq. 100-7)

(Metal ion reduction) M3+ + e- M2+


(Eq. 100-8)

Chevron Corporation

100-3

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

(Metal plating) M + + e- M
(Eq. 100-9)

122 Electrode Potentials


When considered separately, anodic and cathodic reactions are referred to as halfcell reactions. By convention, the electrode potential for half-cell reactions is referenced against the standard hydrogen electrode (H2/H+) which is arbitrarily set to zero. The standard electrode potential of iron is minus 0.44 volts (see Figure 100-1); that of silver (Figure 100-2) is plus 0.80 volts. These figures illustrate standard electrode potentials (electromotive force or EMF). A potential tells you how fast and in which cell electrons are being produced. For example, because iron has a more active (negative) potential than silver, iron will corrode faster. Figure 100-3 lists the standard EMF for some other half-cell reactions. The standard EMFs are for the cases in which all reactants are at unit activity (1 gramatomic weight of metal ion per liter). The potentials of half-cells with reactants not at unit activity may be calculated using the Nernst equation.
Fig. 100-1 Potential Difference Between Iron and a Hydrogen Electrode Fig. 100-2 Potential Difference Between Silver and a Hydrogen Electrode

Direction of Electrochemical Reactions


In an electrochemical reaction the more negative or active half-cell tends to be oxidized, and the more positive or noble half-cell tends to be reduced. In Figure 100-4, for example, silver and iron are coupled. Iron has the more active half-cell potential, so it becomes the anode and corrodes. Silver is the more noble and becomes the cathode, although it does not enter directly into the reaction. The reactions at the two electrodes are as follows:

August 1999

100-4

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

Fig. 100-3

Standard EMF Series of Metals Metal-Metal Ion Equilibrium (Unit Activity) Au Au+3 Pt Pt+2 Pd Pd
+2

Electrode Potential vs. Normal Hydrogen Electrode at 25C, Volts +1.498 +1.20 +0.987 +0.799 +0.788 +0.337

Noble or Cathodic

Ag Ag+ Hg Hg
+2

Cu Cu+2 H2 H+ Pb Pb+2 Sn Sn Co Co Fe Fe
+2

0.000

-0.126 -0.136 -0.250 -0.277 -0.403 -0.440 -0.744 -0.763 -1.662

Ni Ni+2 Active or Anodic


+2

Cd Cd+2
+2

Cr Cr+3 Zn Zn
+2

Al Al+3 Mg Mg KK
+ +2

-2.363 -2.714 -2.925

Na Na+

Source: A. J. de Bethune and N. A. S. Loud, Standard Aqueous Electrode Potentials and Temperature Coefficients at 25C, Clifford A. Hampel, Skokie, Ill., 1964.

Anode. The loss of electrons from the iron causes the reaction in Equation 100-1 to proceed toward the right, with the result that the iron dissolves (i.e., corrodes). Cathode. Hydrogen ions derived from dissociation of water appear at the cathode, where they are removed from solution by electrons received from the external circuit. This is often called plating-out of hydrogen at the cathode. Figure 100-4 shows the reactions of electrochemical corrosion. These reactions normally proceed slowly because of the limited number of hydrogen ions available from the water dissociation reaction. If a greater number of hydrogen ions is made available (as by the addition of acid to the solution), the corrosion reaction will proceed much more rapidly, even to the

Chevron Corporation

100-5

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

point of evolving hydrogen gas. Iron corrodes more rapidly as the solution pH is lowered. Figure 100-4 shows a divided cell with a separated anode and cathode. Generally on a corroding metal, the oxidation and reduction reactions occur on the same metal surface. Corrosion usually involves more than just one oxidation and reduction reaction. There may be several anodes and cathodes on the same metal surface; these anodes and cathodes develop at different points due to local differences in material or environment. The same concepts of electrochemical potentials apply whether the anode and cathode are separated or on the same surface.
Fig. 100-4 Electrochemical Corrosion

ExampleRusting of Iron
The rusting of iron in oxygenated water is a common example of electrochemical corrosion. The reactions are: Oxidation: 2Fe 2Fe2+ + 4e(Eq. 100-10)

Reduction: O2 + 2H2O + 4e- 4OH(Eq. 100-11)

The sum of these partial reactions gives the overall reaction: 2Fe + O2 + H2O 2Fe2++ 4OH- 2Fe(OH)2
(Eq. 100-12)

August 1999

100-6

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

Ferrous hydroxide is further oxidized in the presence of oxygen to give ferric hydroxide or rust: 2Fe(OH)2 + H2O + O2 2Fe(OH)3
(Eq. 100-13)

The complete rusting reaction is summarized as: 4Fe + 3O2 + 6H2O 4Fe(OH)3
(Eq. 100-14)

An increase in oxygen content (aeration) normally increases the corrosion rate. The increased oxygen content forces Equation 100-11 to the right, produces more OH- ions, and removes more electrons, thereby accelerating corrosion at the anodes. Both these effects increase the corrosion rate of iron. Iron will not rust if it is submerged in oxygen-free water, nor will it continue to rust if the oxygen removed from water by the reaction is not replaced. For this reason firewater lines containing saltwater suffer little internal corrosion over many years if the lines are not used for purposes other than infrequent fire emergencies. Iron will also not rust in a dry atmosphere. Ships and other equipment can thus be mothballed and kept in good condition indefinitely if the atmosphere in contact with metal surfaces is maintained at a low relative humidity. The presence of oxygen and moisture in the environment will not alone cause rusting. Oxygen must have access to the surface of the metal. Without agitation (in a solution) or if the metal is protected by corrosion products, the reaction becomes diffusionlimited and proceeds slowly. Relative velocity between the metal and the solution tends to increase the corrosion rate of iron, especially in saltwater.

123 Polarization and Corrosion Rates


The corrosion rate determines whether a material is usable in an environment. Corrosion rates are measurements of penetration or metal loss over a specific time period. Often, metals with a tendency to corrode do so slowly enough to be usable. Wet Corrosion. Wet corrosion involves reactions at the metal-electrolyte interface, movement of ions in the fluid to the interface, and movement of ions away from the interface. Any one of these steps may limit the rate of corrosion; thus, an electrochemical reaction is polarized, or limited, by these steps. Polarization is the limiting of the reaction by certain physical or chemical factors. Whichever step occurs slowest controls the corrosion rate. Polarization. Polarization is a change in potential as the result of current flow. There are two types of polarization: activation polarization and concentration polarization. In activation polarization, the electrochemical process is controlled by the reaction sequence at the metal-electrolyte interface. Either the anodic or cathodic reaction (or both) may become polarized, but cathodic polarization is typically more common in environments such as water. The reactions occurring at either the anode or the cathode are the slowest steps of the overall corrosion reaction and therefore

Chevron Corporation

100-7

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

control the corrosion rate. Concentration polarization is controlled by the diffusion of ions in the electrolyte; ions moving in solution to the anode and cathode limit the corrosion rate, and agitating the fluid will accelerate corrosion. Passivation. Passivation occurs when a protective film forms on the metal surface. In the passive state metals become relatively inert and corrosion rates are slow. If the protective film is damaged or removed, the metal may become active and corrode rapidly. However, under certain conditions, the protective film can reform and the metal is repassivated. Stainless steels and alloys of aluminum, chromium and titanium are examples of metals showing active-passive behavior. Passivity is usually achieved in moderately to strongly oxidizing solutions such as nitric acid or concentrated sulfuric acid. Figure 100-5 illustrates the active-passive behavior of a metal. In the active range, increased oxidizing power of a solution results in a higher corrosion rate, until the passive state is reached. In the passive range, increases in the oxidizing power of the solution produce little change in the corrosion rate. However, at very high concentrations of oxidizing agents, the metal enters the transpassive region and corrosion rate begins to increase with increasing concentrations.
Fig. 100-5 Polarization Diagram for an Active-passive Metal

Effect of Temperature on Corrosion Rates


Like chemical reaction rates, corrosion rates generally increase with temperature. As a rule of thumb, corrosion rates double for each 18F (10C) temperature increase. However, this is true only when corrosion is controlled by the rate of surface reaction at the anode or cathode. In the more common case of diffusionlimited corrosion the effect of temperature is not as great. A doubling of the corrosion rate for every 10C can be considered the upper limit for the temperature effect.

August 1999

100-8

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

124 Forms of Wet Corrosion


There are eight forms of wet corrosion, classified by the appearance of the corroded metal.

Uniform Corrosion
Uniform corrosion is a general thinning over the entire exposed surface of a metal object.

Galvanic Corrosion
Galvanic corrosion results when dissimilar metals form a corrosion cell or couple. This cell (Figure 100-4) results when metals with different electrochemical potentials are immersed in an electrolyte. If the two metals are connected electrically, electrons will flow in the metallic circuit to the cathode from the anode. The tendency of a metal to corrode in a galvanic cell is determined by its position in the galvanic series (or electrochemical series) of metals and alloys. Figure 100-6 gives the galvanic series for some of the common engineering metals and alloys in sea water. A metal tends to corrode when connected to a metal more cathodic than it (above it on the table). The alloys grouped in brackets are similar in composition and can be coupled together with little danger of galvanic corrosion. The farther apart the metals are in the series, the more tendency there is for corrosion of the more anodic metal. The galvanic series gives a more accurate indication of the galvanic corrosion of alloys than does the Standard EMF series. Because the rest potential of a metal depends on the solution it is exposed to, a metal's position in the galvanic series depends on what electrolyte is used. Note that many of the stainless steels appear toward the anodic (upper) end of the series when they are in the active condition and at the cathodic (lower) end when in the passive condition. The dual nature of the stainless steels is related to their ability to form protective films (the passive condition) in the presence of oxygen or other oxidizing agents such as nitric or carbonic acid. If the film is destroyed, these alloys are subject to rapid corrosion (the active condition) in the presence of hydrochloric, hydrofluoric, or other oxygen-free acids. To select the correct stainless steels the engineer must determine whether they will be in the passive or active state. The rate of corrosion resulting from galvanic action depends upon the relative exposed areas of the two metals in contact. For example, at normal atmos-pheric temperatures zinc connected to iron will corrode but tends to protect the iron by making it the cathode of the galvanic cell, as in the case of zinc-coated (galvanized) steel. However, a small piece of zinc will corrode very rapidly (a high corrosion rate per unit area) when coupled to a large area of iron, leaving the iron with little protection. Conversely, a large piece of zinc will not corrode significantly more rapidly as a result of being coupled to a small area of iron. Under these conditions the iron may receive excellent protection. However, it should be noted that in some hot water applications the polarities of zinc and iron are reversed. Figure 100-7 shows the compatibility of different alloys coupled together, based on the relative exposed areas.

Chevron Corporation

100-9

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

Fig. 100-6

Galvanic Series of Common Alloys in Sea Water (Data derived from Fontana and Green, Corrosion Engineering, McGraw-Hill Book Company, 1978. Used with permission)
Cathodic Platinum Gold Graphite Titanium Silver 12% Ni, 18% Cr, 3% Mo Steel 20% Ni, 25% Cr Steel 23 to 30% Cr Steel 14% Ni, 23% Cr Steel 8% Ni, 18% Cr Steel 16 to 18% Cr Steel 12 to 14% Cr Steel 80% Ni, 20% Cr Inconel 60% Ni, 15% Cr Nickel Monel Metal Copper Nickel Nickel-Silver Bronzes Copper Brasses 80% Ni, 20% Cr Inconel Nickel Tin Lead Lead-Tin-Solder 12% Ni, 18% Cr, 3% Mo Steel 20% Ni, 25% Cr Steel 14% Ni, 23% Cr Steel 8% Ni, 18% Cr Steel Ni-Resist 23 to 30% Cr Steel 16 to 18% Cr Steel 12 to 14% Cr Steel 4 to 6% Cr Steel Cast Iron Carbon Steel Aluminum Alloy 2024 (4.5 Cu, 1.5 Mg, 0.6 Mn) Cadmium Commercially Pure Aluminum Zinc Magnesium and Magnesium Alloys

Passive

Passive

Active

Active

Active

Anodic Notes:

1. The alloys grouped together in shaded areas are similar in composition and can be coupled with each other with little danger of galvanic corrosion.

August 1999

100-10

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

Fig. 100-7

Compatibility of Common Alloys in Sea Water (Used by special permission from CHEMICAL ENGINEERING, June 19, 1988 McGraw-Hill, Inc., New York)

Chevron Corporation

100-11

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

The use of corrosion at the anode to protect the cathode is the basis for cathodic protection by sacrificial galvanic anodes (see Sections 1100 to 1900). Galvanic corrosion is not limited to cells in which totally dissimilar metals are in contact while exposed to an electrolyte. Differences in the composition or surface condition of otherwise similar metals often result in galvanic corrosion cells. Examples of galvanic cells are: Corrosion of new steel electrically connected to old steel. (It is a common experience in pipeline operations to find that new pipe installed in a repair section or branch line corrodes considerably more rapidly than the old pipe to which it is connected) Corrosion of steel pipe (bare or poorly coated in soil or water) which is electrically connected to concrete-covered pipe Corrosion of steel pipe connected to copper pipe or tubing Corrosion of a steel propeller shaft operating in a bronze bearing

The conductivity of a solution (electrolyte) in which dissimilar metals are placed will have a significant effect on the corrosion rate. For example, more galvanic corrosion problems occur in saltwater, because of its higher conductivity, than in fresh water. In summary, galvanic attack is minimized or prevented by: Avoiding unfavorable metal combinations, that is, metals far apart in the galvanic series Avoiding combinations involving relatively small areas of the more active (anodic) metal Using insulation to prevent contact (and completion of the electrical circuit) between dissimilar metals Avoid coating the anode of the couple. This is important because gaps in the coating will expose small areas of anode creating unfavorable anode-to-cathode areas. Coat the cathode to reduce galvanic corrosion

Crevice Corrosion
Crevice corrosion is localized corrosion that occurs in crevices and on shielded areas of metal surfaces exposed to corrosives. This type of corrosion occurs under bolt and rivet heads, between lap joint surfaces, under gasket surfaces, and under surface deposits. Metals that depend on a passive film for corrosion protection, such as stainless steels, are most susceptible. Crevice corrosion is most severe in high chloride environments. Crevice corrosion has also been described as concentration cell corrosion because of differences between the concentration of some critical ion in the electrolyte versus the crevice (see Figure 100-8). In an oxygen concentration cell (Figure 100-9), the

August 1999

100-12

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

corrosion reaction is promoted whenever oxygen in the air comes in contact with a moist steel surface. The reaction at the cathode of the oxygen concentration cell is: 2H2O + O2 + 4e- 4(OH)(Eq. 100-15)

After a short time, corrosion reactions consume the oxygen in the crevice; because the solution is stagnant, the oxygen is not replenished. Therefore, the area in the crevice does not fully passivate and becomes the anode. Hence, the oxygen concentration cell accentuates corrosion where the oxygen concentration is lower. Corrosion is often accelerated in cracks, crevices, and other apparently inaccessible places because oxygen-deficient areas are the anodes for oxygen concentration cells. It has been theorized that a build-up of metal cations in the crevice attracts chloride anions to the crevice. As the chloride concentration builds, the local pH drops, and corrosion rates increase. As shown in Figure 100-9, an oxygen concentration cell can occur without a crevice.
Fig. 100-8 Concentration Cell

Pitting
Pitting is a localized corrosion that makes holes in the metal. Often, pitting is caused by chlorides, particularly on stainless steels. Pitting usually occurs in stagnant conditions, and the pitted metals will show little general or uniform corrosion. See Section 210.

Intergranular Corrosion
Intergranular corrosion is localized corrosion at grain boundaries while the grains remain relatively unattacked. The material basically disintegrates because grains fall out. Intergranular attack is a common problem with austenitic stainless steels. Section 410 discusses this topic in detail.

Chevron Corporation

100-13

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

Fig. 100-9

Oxygen Concentration Cells

Selective Dealloying
Selective dealloying is the selective attack on one element of an alloy. Dezincification, the removal of zinc from brass, is a common example. Often, inhibited admiralty brass will help prevent dezincification. Another example is graphitization of cast iron in sea water, in which the iron corrodes, leaving the carbon (graphite) in place. In severe environments, 90-10 Cu-Ni can suffer denickelfication.

Erosion Corrosion
Erosion corrosion is an increase in the corrosion rate caused by relative movement of the corrosive fluid with respect to the metal. Erosion corrosion is velocityassisted corrosion. Corrosion of the metal increases with velocity because protective films on the metal surface are stripped off. Often an abrupt critical velocity is associated with erosion corrosion. Above the critical velocity, corrosion will be severe; below the critical velocity, corrosion will be low. For example, carbon steel is generally acceptable for handling sulfuric acid at ambient temperature in low velocity conditions (less than 3 feet per second). However, as the velocity increases the protective sulfate film is stripped away, allowing corrosion. Similarly, the corrosion of brass in high-velocity sea water is accelerated because of the removal of the protective oxide film.

Stress Corrosion Cracking


Stress corrosion cracking (SCC) is cracking caused by tensile stress and a specific corrosive environment. Cracking occurs with little apparent ductility, and occurs below the metals yield strength. See Section 420 for details on SCC.

August 1999

100-14

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

130 Dry Corrosion


Dry corrosion, or direct chemical combination, occurs above the environment dew point. Dry corrosion is usually associated with high temperatures, and gases are the usual corrodents. Two examples follow.

131 High-temperature Oxidation


At high temperatures, nearly all metals will react with oxygen to form an oxide scale, as follows: Metal + O2 metal oxide
(Eq. 100-16)

In many industry applications, the general term oxidation will refer to this hightemperature reaction with air rather than to low-temperature electrochemical oxidation such as the rusting of steel. Carbon steel may be used in an oxidizing atmosphere without excessive scaling up to a temperature of about 1050F. Above this temperature various alloys must be added to increase oxidation resistance as well as to provide adequate mechanical properties and metallurgical stability. See Section 430.

132 Hydrogen Sulfide Corrosion


Many crude oils and refining unit charge stocks contain sulfur compounds. Most are organic compounds, but some crude oils contain significant amounts of dissolved hydrogen sulfide. Regardless of the form of sulfur compounds, most will decompose or combine with hydrogen in various process atmospheres to form hydrogen sulfide. Also, hydrogen sulfide dissolved in crude oil will be released when the crude is heated. At temperatures above 450F, hydrogen sulfide will react with iron: Fe + H2S FeS + H2
(Eq. 100-17)

The conversion of iron to iron sulfide (H2S corrosion) occurs more rapidly at higher hydrogen sulfide concentrations and at higher temperatures. Since hydrogen is involved in the corrosion reaction, hydrogen partial pressure also affects corrosion rate. Corrosion may be accelerated or retarded by hydrogen, depending on which of several iron sulfides is the stable scale species. If the partial pressure of hydrogen is high compared to H2S partial pressure, iron sulfide may decompose into metallic iron and hydrogen sulfide. Hydrogen sulfide corrosion, including control measures, is discussed further in Section 400. Section 3100, Crude Distillation Units, and Section 3420, Catalytic Reforming Process, discuss specific areas where this corrosion occurs in process

Chevron Corporation

100-15

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

plants. (These sections are in Volume 2 of the Corrosion Prevention and Metallurgy Manual.)

140 Biological Corrosion


Biological corrosion is not given detailed coverage in a separate section of this manual. The following material will describe the principal organisms and conditions associated with this type of corrosion. Some comments on control and prevention are included. Certain types of living organisms in contact with metals influence corrosion rates. The chemical reactions by which living organisms ingest food and eliminate waste affect corrosion conditions by: Influencing anodic and cathodic reactions Changing protective surface films Creating corrosion conditions or producing deposits

Biological corrosion most often results in localized (crevice or pitting) corrosion. Many types of organisms are closely identified with corrosion, and primarily with the corrosion of iron and steel. These organisms may be either microscopic or macroscopic in size: Microscopic Organisms Bacteria Algae Fungi Macroscopic Organisms Marine algae Calcareous and Siliceous Sponges Barnacles Mollusks Bryozoa Coelenterates

141 Influence of Microorganisms


Bacteria
Bacteria are simple one-celled organisms. They reproduce rapidly by fission and generally grow best at 70F to 120F, although some grow at subfreezing temperatures and others near boiling. They are found in soil and water environments. Bacteria are very hardy, surviving for long periods under unfavorable conditions. They are classified either as aerobic (requiring oxygen to survive) or anaerobic. Bacteria cause corrosion more often than fungi and algae do. Some types of bacteria known to cause corrosion follow. Sulfate-reducing Bacteria. Sulfate-reducing bacteria are anaerobic microorganisms. They use sulfate ions (instead of oxygen) as oxidizing agents to assimilate organic nutrients. The best known and probably most widespread is Desulfovibrio desulfuricans. These extremely common bacteria can reduce sulfates to sulfides in

August 1999

100-16

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

the absence of oxygen. This action is commonly found in association with a source of ferrous ionsfrequently a buried pipe line, well casing, or a water-bottom in an oil storage tank. The sulfide combines with dissolved iron to form the black layers of ferrous sulfide characteristic of this process. Symptoms of corrosion caused by sulfate-reducing bacteria include: Presence of hydrogen sulfide gas Bright corroded areas under the layers of corrosion product (ferrous sulfide and iron oxide)

The sulfate-reducing action requires that organic materials be present. The precise metabolic functions of these nutrients are not known. Little is known about the specific reactions that cause the removal (depolarization) of cathodic hydrogen from the metal surface. Investigators have suggested direct bacterial action as a causepossibly including metabolic use of hydrogenbut this has not been proven. Some authorities believe that the sulfate-reducing bacteria do no more than develop strongly reducing (oxygen-free) areas that become anodic and corrode by oxygen concentration cell effects. Under conditions favorable to their development, anaerobic sulfate-reducing bacteria can form large colonies which may become dispersed in water. Either alone or with other organic growths, these colonies and their associated corrosion products can contribute to the fouling of water systems and filters, and to plugging of producing formations into which the water is injected. A favorable environment for growth of sulfate-reducing bacteria is one with: A pH range of 5.5 to 8.5 A supply of soluble iron Absence of oxygen An adequate supply of sulfate ions An adequate supply of organic nutrient

These conditions are encountered in producing formations, in poorly drained soils, and in water used for secondary recovery operations. Anaerobic sulfate-reducing bacteria are sometimes found in aerated waters and soils in coexistence with oxygen scavengers that create locally anaerobic conditions. In these cases, they may cause corrosion by developing oxygen-concentration cells. Water from source wells commonly contains sulfate-reducing bacteria. Such water is normally kept in a closed system and injected into the formation (often after filtering) without being intentionally exposed to air. Nevertheless, extensive growths of sulfate-reducing bacteria may occur even when the obvious sources of air have been eliminated. Corrosion with which sulfate-reducing bacteria is associated can be controlled by conventional methods. Chemical treatments, coatings, and cathodic protection can all be effective.

Chevron Corporation

100-17

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

Fouling, including plugging by biological corrosion products, may be controlled by various types of biocides. Although specific bactericides can be used, general biocides such as chlorine are frequently used for shotgun control of all organic growths. Some studies have shown that sulfate-reducing bacteria do not survive well in brines containing over 100,000 ppm chlorides. However, biocides may still be necessary to control other microorganisms (such as halophilic microorganisms) which thrive in this type of environment. Iron-depositing Bacteria. Iron-depositing bacteria are the best known of the aerobic microorganisms associated with corrosion and fouling. They thrive in water that is slightly acid and contains dissolved iron. Materials in contact with water carrying these bacteria become coated with nodules of ferric hydroxide or with various scums impregnated with ferric hydroxide. The ferric hydroxide frequently gives the water a reddish tinge. These bacteria do not appear to directly attack steel. Instead, they assimilate soluble iron which is oxidized to ferric hydroxide by oxygen contained in the water. The ferric hydroxide may form tubercles which act as sites for concentration cell corrosion attack. By causing depletion of oxygen, the tubercles may also provide an environment for the growth of anaerobic bacteria. Some corrosion resulting from tuberculation probably is caused by the joint action of iron bacteria and sulfatereducing bacteria. Examples of iron depositing bacteria are Gallionella, Sphaerotilus, and Crenothrix. These bacteria most commonly cause a pitting-type attack on steel and stainless steel. Pits will usually be found beneath brown to reddish ferric hydroxide deposits or tubercles. Gallionella, in particular, damages stainless steel because it tends to concentrate chlorides. Crenothrix has been responsible for both ferric hydroxide fouling and corrosion of 90-10 cupronickel tubes in the sea water desalination plants at Borco Refinery. Feedwater for these plants comes from salt water wells in the coral upon which the refinery is built. The coral formation is badly contaminated with Crenothrix, which enters the desalination plant with the feedwater. Sulfur-oxidizing Bacteria. Sulfur-oxidizing bacteria, such as Thiobacillus ferroxidans, are aerobic. They oxidize sulfides to sulfates and can drive the pH to 1 or less. The formation of acid promotes corrosion. Thiobacillus can coexist with sulfatereducing bacteria. Another form, Beggiatoa, may exist where there is an abundance of H2S. It oxidizes sulfides to elemental sulfur. It forms bulky slime masses that are well suited for harboring sulfate reducers. This seems to be its primary role in causing corrosion. Slime-forming Bacteria. Pseudomonas is the type of slime-forming bacteria most prevalent in industrial waters. Slime-forming bacteria are aerobic oxygen scavengers that can harbor sulfate reducers. The category slime-forming is misleading because many types of bacteria, algae, and fungi can form bulky slime. Control of

August 1999

100-18

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

bacteria-induced corrosion is frequently not possible without first controlling the slime formers with chemical treatments.

Algae and Fungi


Algae are plant-like organisms that get their energy from light (photosynthesis) and produce oxygen, which can accelerate corrosion. Fungi are plant-like organisms that get their energy through either fermentation (anaerobic) or oxidation (aerobic). Organic acids are often by-products. Algae and fungi cause corrosion by the formation of differential concentration cells at the metal surface. Both are slime-formers. In sunlight, algae growing in contact with metal give off oxygen, which contributes to corrosion as a depolarizer of cathodic areas. Fungi develop primarily in acid environments and promote corrosion by creating differential concentration cells. Even where the slime-formers do not contribute to corrosion, they have other undesirable effects. Algae and fungi growing in cooling water systems may greatly reduce circulation and heat transfer capacity. In secondary recovery operations, they tend to plug formations, and in wastewater discharges, they contribute to unsightly conditions. Water treatments, therefore, will usually include specific algaecides and fungicides to control organic growths.

142 Control of Corrosion Caused by Microorganisms


In addition to good housekeeping, approaches that are effective in controlling or preventing microbiological corrosion are: Change to materials resistant to the corrosion. This is not always effective, because some bacterial corrosion mechanisms attack stainless steels and copper alloys more rapidly than they do steel and cast iron Use coatings to insulate the metal surface from the bacteria. Thin-film linings are often successfully used; however, pinholes in the lining are likely to occur, inviting the development of bacteria colonies. If corrosion of bare metal is high, a thick coating may be used to avoid rapid attack of the metal. However, if corrosion of the metal at coating pinholes can be tolerated, a thin-film lining may be more economical. Coatings should be tested to determine that they will not be attacked by the bacteria Use cathodic protection. Like coatings, cathodic protection is very successful, especially in preventing corrosion caused by sulfate-reducing bacteria. It may not be effective in preventing corrosion of metals other than steel and iron Use chemical treatments. Chemical inhibitors sometimes work. Biocides control and kill microorganisms if the right ones are used and if injection is properly monitored Avoid using source waters known to be contaminated

A key step in solving microbiological corrosion is identifying the problem and the organism(s) responsible. It is fairly inexpensive to analyze a representative

Chevron Corporation

100-19

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

water sample to determine if harmful microorganisms are present. You may need the help of a consulting laboratory to identify organisms and effective biocides.

143 Influence of Macroorganisms


The primary sources of corrosion by macroorganisms are the larger growths that become attached to a metal surface early in life and must maintain this attachment in order to survive. Three general growth patterns distinguish the types of corrosion caused by macroorganisms. The organism grows evenly over a smooth surface, resulting in increased protection The attachment is irregular, or the growth is uneven, causing the development of differential oxygen-concentration cells The growth establishes anaerobic conditions, which can be associated with sulfate-reducing bacteria (see Section 141)

150 Glossary
These definitions are largely based on the extensive Definitions Relating to Metals and Metalworking in the American Society of Metals Metals Handbook, Volume 1, and are supplemented by those in the Corrosion Handbook, edited by H. H. Uhlig. Note that a cathodic protection glossary is located at the end of Section 1100. aerobic: containing free air or uncombined oxygen anaerobic: free of air or uncombined oxygen anhydrous: devoid or purged of water anion: a negatively charged ion or radical (e.g., Cl- or SO4=) that migrates toward the anode under the influence of a potential gradient anode: the electrode that supplies electrons to an external circuit and at which oxidation (and normally corrosion) occurs anodic polarization: that portion of the polarization of a cell that occurs at the anode biological corrosion: deterioration of a metal as a result of the activity of living microorganisms such as bacteria or macroforms such as algae cation: a positively charged ion or radical (for example Fe++ or NH4+) that migrates toward the cathode under the influence of a potential gradient cathode: the electrode that receives electrons from an external circuit and at which reduction occurs cathodic polarization: that portion of the polarization of a cell that occurs at the cathode

August 1999

100-20

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

cathodic protection: protection of a metal from corrosion by making it a cathode, using either a galvanic source or impressed current (cathodic protection is covered in Sections 1000 through 1900) caustic embrittlement: stress corrosion cracking in an alkaline solution cavitation corrosion: damage of metal associated with the formation and collapse of bubbles in the liquid at a solid-liquid interface concentration cell: an electrochemical corrosion cell established by differences in electrolyte concentration at anode and cathode areas corrosion: the deterioration of a metal by a chemical or electrochemical reaction with its environment corrosion rate: the rate of penetration or thinning due to corrosive attack, usually expressed as mils per year (mpy) (one mil equals 0.001 in.) corrosion-erosion: corrosion occurring in conjunction with and accentuated by erosion and the removal of corrosion products with resulting exposure of fresh surfaces to corrosion corrosion-fatigue: reduction in fatigue life in a corrosive environment couple: two dissimilar conductors in electrical contact crevice corrosion: localized corrosion as a result of the presence of a crevice current: rate of flow of electrical charge, measured in amperes (coulombs/sec). Charge is measured in coulombs (1 coulomb = 6.3 1018 electrons). Current in this manual generally refers to conventional current which is the reverse of electron flow current density: current flow per unit area depolarization: reduction or elimination of polarization by physical means or by change in environment dezincification: corrosion of a zinc-containing alloy (usually brass), in which zinc preferentially goes into solution, leaving other elements behind differential aeration cell: oxygen concentration cell electrode potential: voltage developed at an electrode as measured against a standard reference electrode electrolyte: a liquid, most often a solution, that will conduct an electric current erosion: mechanical abrasion by solids suspended in a fluid (no chemical reaction is involved) fatigue: the tendency of a metal to fracture in a brittle manner under conditions of repeated cyclic stressing at stress levels below its tensile strength

Chevron Corporation

100-21

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

fretting: deterioration resulting from repetitive slip at the interface between two surfaces (when deterioration is further increased by corrosion, the term fretting corrosion is used) galvanic cell: a cell containing two dissimilar metals in an electrolyte galvanic (electrochemical) series: a series of metals and alloys arranged according to their relative electrode potentials in a specified environment graphitization 1: a corrosion phenomenon occurring on cast iron that dissolves the iron and leaves the graphite 2: metallurgically, the term describes the formation of graphite in iron or steel, from the decomposition of iron carbide (cementite) at high temperatures hydrogen overvoltage: overvoltage associated with the liberation of hydrogen gas inhibitor: a substance that retards corrosion when added to an environment in small concentrations, usually an additive promoting passivation in an acid solution or other electrolyte intergranular corrosion: corrosion occurring preferentially at grain boundaries (see Section 410) ion: an atom or group of atoms that carries an electric charge because of the addition or removal of electrons (see anion and cation) local action: corrosion caused by local cells on a metal surface, i.e., galvanic cells resulting from inhomogeneities between adjacent areas on a metal surface exposed to an electrolyte microbiological corrosion: biological corrosion caused by aerobic or anaerobic microorganisms such as bacteria Nernst equation: An equation that expresses the exact electromotive force of a cell in terms of the activities of products and reactants of the cell oxidation 1: The reaction in which there is an increase in valence resulting from a loss of electron(s) 2: oxide scale formation on metals at elevated temperatures: a form of dry corrosion oxygen concentration cell: a galvanic cell resulting from differences in oxygen concentration passivation: the changing of the chemically active surface of a metal to a much less reactive state pH: a measure of hydrogen ion concentration (pH 7 is neutral, below 7 is acidic, and above 7 is alkaline) pitting: corrosion tending to occur in localized areas, usually as small, sharp cavities formed on a metal surface polarization: the production of counter-emf by-products formed by concentration changes resulting from passage of current through an electrolytic cell

August 1999

100-22

Chevron Corporation

Corrosion Prevention and Metallurgy Manual

100 Fundamentals of Corrosion Mechanisms

potentiostat: an electronic device that maintains a metal at a given potential with respect to a reference electrode, often used in corrosion studies pourbaix diagrams: potential-pH diagrams of thermodynamic data, showing conditions of potential and pH under which a metal does not react or can react to form specific oxides or ions quaternize: reaction of an amine with organic radicals that form a salt that has four radicals attached to the nitrogen atom. These compounds are often surface-active reduction: a reaction in which there is a decrease in valence resulting from a gain in electron(s) rusting: corrosion of iron resulting in the formation of products on the surface, consisting largely of hydrous ferric oxide solute: the component of either a liquid or solid solution that is present to a lesser or minor extent; it is the component that is dissolved in the solvent solvent: the component of either a liquid or solid solution that is present to a greater or major extent; it is the component that dissolves the solute standard hydrogen electrode: standard reference electrode consisting of platinum electrode immersed in solution with unit H+ activity and under one atmosphere H2 pressure. By convention, the potential of the reaction H2S 2H+ + 2e- at this electrode is assigned a value of zero stray current corrosion: corrosion caused by current through paths other than the intended circuit or by an extraneous current in the earth stress corrosion: corrosion of a metal accelerated by stress stress corrosion cracking: failure by cracking under combined action of corrosion and stress

160 References
1. 2. 3. 4. 5. 6. 7. Speller, Frank N. Corrosion: Causes and Prevention. McGraw-Hill, 1951. Starkey, Robert L. The General Physiology of Sulfate-Reducing Bacteria in Relation to Corrosion, Producers Monthly, June 1958, pp. 1230. Uhlig, Herbert H. The Corrosion Handbook. John Wiley & Sons Inc., 1948. Uhlig, Herbert H. Corrosion and Corrosion Control. John Wiley & Sons Inc., 1963. Patton, Charles C. Oilfield Water Systems. Campbell Petroleum Series, 1977. Iverson, W.P. Biological Corrosion, Advances in Corrosion Science and Technology, Vol. 2. Fontana & Staehle, 1973. Tatnall, Robert E. Fundamentals of Bacteria Induced Corrosion, Materials Performance, September 1981.

Chevron Corporation

100-23

August 1999

100 Fundamentals of Corrosion Mechanisms

Corrosion Prevention and Metallurgy Manual

8. 9.

Fontana, Mars G. Corrosion Engineering, 2nd edition. McGraw-Hill, 1978. Pope, Daniel H., David Duquette, Peter C. Wayner, Jr., and Arlana H. Johannes. Microbiologically Influenced Corrosion: A State-of-the-Art Review. MTI Publication No. 13, MTI of the Chemical Process Industries, Inc., 1984.

10. Greene, Richard W. The Chemical Engineering Guide to Corrosion, McGraw-Hill, 1982. 11. Metals Handbook. American Society of Metals. 12. Lee, T. S., Preventing Galvanic Corrosion in Marine Environments. Chemical Engineering, April 1, 1985, pp 89-92. 13. Shreir, L.L., Corrosion, Volumes 1 and 2, Newnes-Butterworths, 1979. 14. Dillon, C.P., Corrosion Control in the Chemical Process Industries, McGraw-Hill, 1986.

August 1999

100-24

Chevron Corporation

Potrebbero piacerti anche