Você está na página 1de 30

CORROSION OF METALS

222R-1

ACI 222R-96

Corrosion of Metals in Concrete


Reported by ACI Committee 222

ACI COMMITTEE 222 Corrosion of Metals in Concrete Brian B. Hope Keith A. Pashina Chairman Secretary John P. Broomeld Kenneth C. Clear James R. Clifton Israel Cornet Marwan Daye Bernard Erlin John K. Grant Kenneth C. Hover Stephen L. Amey Steven F. Dailey Stephen D. Disch Hamad Farzam Per Fidjestol Rodney R. Gerard Michael P. Gillen Bret James Thomas D. Joseph David G. Manning Walter J. McCoy Theodore L. Neff Charles K. Nmai William F. Perenchio Randall W. Poston Associate Members Odd E. Gjorv Clayford T. Grimm Alan K. C. Ip Andrew Kaminker Mohammad S. Khan Philip J. Leclaire Joseph A. Lehmann Robert E. Price D. V. Reddy William T. Scannell David C. Stark Wayne J. Swiat Thomas G. Weil Richard E. Weyers David A. Whiting Mohamad A. Nagi Morris Schupack Ephraim Senbetta Robert E. Shewmaker Bruce A. Suprenant William F. Van Siseren Michael C. Wallrap

This committee report has been prepared to reect the state of the art of the corrosion of metals, and especially steel, in concrete. Separate chapters are devoted to the mechanisms of the corrosion of metals in concrete, protective measures for new concrete construction, procedures for identifying corrosive environments and active corrosion in concrete, and remedial measures. A selected list of references is included with each chapter.
Keywords: admixtures; aggregates; blended cements; bridge decks; calcium chlorides; carbonation; cathodic protection; cement pastes; chemical analysis; chlorides; coatings; concrete durability; corrosion; corrosion resistance; cracking (fracturing); deicers; deterioration; durability; marine atmospheres; parking structures; plastics, polymers and resins; portland cements; prestressing steels; protective coatings; reinforced concrete; reinforcing steels; repairs; resurfacing; spalling; waterproof coatings.

CONTENTS Chapter 1Introduction, p. 222R-1 1.1Background 1.2Scope 1.3References Chapter 2Mechanism of corrosion of steel in concrete, p. 222R-3 2.1Introduction 2.2Principles of corrosion 2.3Effects of the concrete environment on corrosion 2.4References Chapter 3Protection against corrosion in new construction, p. 222R-11 3.1Introduction 3.2Design and construction practices 3.3Methods of excluding external sources of chloride ion from concrete 3.4Methods of protecting reinforcing steel from chloride ion
ACI 222R-96 replaces ACI 222R-89 and became effective May 23, 1996. Copyright (c) 1997, American Concrete Institute. All rights reserved including rights of reproduction and use in any form or by any means, including the making of copies by any photo process, or by any electronic or mechanical device, printed or written or oral, or recording for sound or visual reproduction or for use in any knowledge of retrieval system or device, unless permission in writing is obtained from the copyright proprietors.

ACI Committee Reports, Guides, Standard Practices, and Commentaries are intended for guidance in designing, planning, executing, or inspecting construction and in preparing specifications. Reference to these documents shall not be made in the Project Documents. If items found in these documents are desired to be part of the Project Documents, they should be phrased in mandatory language and incorporated in the Project Documents.

222R-1

222R-2

ACI MANUAL OF CONCRETE PRACTICE

3.5Corrosion control methods 3.6References Chapter 4Procedures for identifying corrosion environments and active corrosion in concrete, p. 222R-21 4.1Introduction 4.2Methods of evaluation 4.3References Chapter 5Remedial measures, p. 222R-23 5.1Introduction 5.2General 5.3Applicability 5.4The remedies and their limitations 5.5Summary 5.6References Chapter 6References to documents of standardproducing organizations, p. 222R-28 Appendix AStandard Test Method for WaterSoluble Chloride Available for Corrosion of Embedded Steel in Mortar and Concrete Using the Soxhlet Extractor, p. 222R-28 CHAPTER 1INTRODUCTION 1.1Background Concrete normally provides reinforcing steel with excellent corrosion protection. The high alkaline environment in concrete results in the formation of a tightly adhering film which passivates the steel and protects it from corrosion. In addition, concrete can be proportioned to have a low permeability which minimizes the penetration of corrosion-inducing substances. Low permeability also increases the electrical resistivity of concrete which impedes the flow of electrochemical corrosion currents. Because of these inherent protective attributes, corrosion of steel does not occur in the majority of concrete elements or structures. Corrosion of steel, however, can occur if the concrete is not of adequate quality, the structure was not properly designed for the service environment, or the environment was not as anticipated or changes during the service life of the concrete. The corrosion of metals, especially steel, in concrete has received increasing attention in recent years because of its widespread occurrence in certain types of structures and the high cost of repairs. The corrosion of steel reinforcement was first observed in marine structures and chemical manufacturing plants.1.1,1.2,1.3 More recently, numerous reports of its occurrence in bridge decks, parking structures, and other structures exposed to chlorides has made the problem particularly prominent. The consequent extensive research on factors contributing to steel corrosion has increased our understanding of corrosion, especially concerning the role of chloride ions. It is anticipated that the application of the findings of this research will result in fewer instances of corrosion in new concrete structures and improved methods of repairing corrosion-induced damage in existing structures. For these improvements to occur, the information must be disseminated to individuals responsible for the design,

construction, and maintenance of concrete structures. The main purpose of this report is to present the state of the art. While several metals may corrode under certain conditions when embedded in concrete, the corrosion of steel reinforcement is of the greatest concern, and, therefore, is the primary subject of the report. Chloride ions are considered to be the major cause of premature corrosion of steel reinforcement. Corrosion can occur in some circumstances in the absences of chloride ions, however. For example, carbonation of concrete results in reduction of its alkalinity, thereby permitting corrosion of embedded steel. Carbonation, however, is usually a slow process in concrete which has a low water-cement ratio and carbonation-induced corrosion is not as common as corrosion induced by chloride ions. Chloride ions are common in nature and small amounts are usually unintentionally contained in the mix ingredients of concrete. Chloride ions also may be intentionally added, most often as a constituent of accelerating admixtures. Dissolved chloride ions also may penetrate unprotected hardened concrete in structures exposed to marine environments or to deicing salts. The rate of corrosion of steel reinforcement embedded in concrete is strongly influenced by environmental factors. Both oxygen and moisture must be present if electrochemical corrosion is to occur. Reinforced concrete with significant gradients in chloride ion content is vulnerable to macrocell corrosion, especially if subjected to cycles of wetting and drying. Other factors that affect the rate and level of corrosion are heterogeneities in the concrete and the steel, pH of the concrete pore water, carbonation of the portland cement paste, cracks in the concrete, stray currents, and galvanic effects due to contact between dissimilar metals. Design features also play an important role in the corrosion of embedded steel. Mix proportions, depth of cover over the steel, crack control measures, and implementation of measures designed specifically for corrosion protection are some of the factors that control the onset and rate of corrosion. Deterioration of concrete due to corrosion results because the products of corrosion (rust) occupy a greater volume than the steel and exert substantial stresses on the surrounding concrete. The outward manifestations of the rusting include staining, cracking, and spalling of the concrete. Concurrently, the cross section of the steel is reduced. With time, structural distress may occur either by loss of bond between the steel and concrete due to cracking and spalling or as a result of the reduced steel cross-sectional area. This latter effect can be of special concern in structures containing high strength prestressing steel in which a small amount of metal loss could possibly induce tendon failure. The research on corrosion to date has not produced a steel or other type of reinforcement that will not corrode when used in concrete and that is both economical and technically feasible. However, research has pointed to the need for quality concrete, careful design, good construction practices, and reasonable limits on the amount of chloride in the concrete mix ingredients. Other measures that are being investigated include the use of corrosion inhibitors, protective coatings on the steel, and cathodic protection. There has been some

CORROSION OF METALS

222R-3

success in each of these areas though problems resulting from corrosion of embedded metals are far from eliminated. 1.2Scope This report discusses the factors that cause and control corrosion of steel in concrete, measures for protecting embedded metals in new construction, techniques for detecting corrosion in structures in service, and remedial procedures. Consideration of these factors, and application of the discussed measures, techniques, and procedures should assist in reducing the occurrence of corrosion and result, in most instances, in the satisfactory performance of reinforced and prestressed concrete elements. 1.3References
1.1. Tremper, Bailey; Beaton, John L; and Stratfull, R. F. Causes and Repair of Deterioration to a California Bridge Due to Corrosion of Reinforcing Steel in a Marine Environment II: Fundamental Factors Causing Corrosion, Bulletin No. 182, Highway Research Board, Washington, D.C., 1958, pp. 18-41. 1.2. Evans. U. R., The Corrosion and Oxidation of Metals: Scientic Principles and Practical Applications, Edward Arnold, London, 1960, 303 pp. 1.3. Biczk, Imre, Concrete Corrosion and Concrete Protection, 3rd Edition, Acadmiai Kiad, Budapest, 1964.

one method of expressing electromotive forces. The Fe++ in Eq. (2-1) is subsequently changed to oxides of iron by a number of complex reactions. The volume of the reaction products is several times the volume of the iron. At the cathode, reduction takes place. In an acid medium, the reaction taking place at the cathode is the reduction of hydrogen ions to hydrogen. However, as will be shown later, concrete is highly basic and usually has an adequate supply of oxygen, so the cathodic reaction is
1/ 2

H2O + 1/4 O2 + e-

OH (2-2)

E = 0.401 Standard Redox Potential

CHAPTER 2MECHANISM OF CORROSION OF STEEL IN CONCRETE 2.1Introduction This chapter is divided into two sections. In the first, emphasis is placed on the processes responsible for corrosion of steel reinforcement in concrete. The corrosion mechanism is generally accepted to be electrochemical in nature. Some of the major causes of corrosion of steel in concrete are examined and related phenomena are discussed. In the second part, a discussion is given on the concrete variables that influence corrosion, including concrete mix proportions, quality, and cover, and the effects of corrosion inhibiting admixtures and carbonation. 2.2Principles of corrosion 2.2.1 CorrosionAn electrochemical processAlthough iron can corrode by chemical attack, the most common form of corrosion in an aqueous medium is electrochemical.2.1 The corrosion process is similar to the action which takes place in a flashlight battery. An anode, where electrochemical oxidation takes place, a cathode, where electrochemical reduction occurs, an electrical conductor, and an aqueous medium must be present. Any metal surface on which corrosion is taking place is a composite of anodes and cathodes electrically connected through the body of the metal itself. Reactions at the anodes and cathodes are broadly referred to as half-cell reactions. At the anode, which is the negative pole, iron is oxidized to ferrous ions. Fe Fe++ + 2e(2-1)

E = - 0.440 Standard Redox Potential

The Standard Redox Potential is the potential generated when the metal is connected to a hydrogen electrode and is

The corroding iron piece has an open circuit potential, also called a rest potential, related to the Standard Redox Potentials of the reactions in Eq. (2-1) and (2-2), to the composition of the aqueous medium, the temperature, and to the polarizations of these half-cells. The rate of corrosion is related to the polarizations as will be discussed later. 2.2.2 Availability of oxygen in concreteAlthough the availability of oxygen is one of the main controlling factors for corrosion of steel, there appears to be little quantitative information on its effect in concrete. Some information is shown, however, in Fig. 2.1, where the rate of oxygen diffusion through water-saturated concrete of varying quality and thickness is illustrated.2.2 The rate of oxygen diffusion through concrete is also significantly affected by the extent to which the concrete is saturated with water. A number of investigations indicate that if the steel passivity is destroyed, conditions will be conducive to the corrosion of steel reinforcement in those parts of a concrete structure that are exposed to periods of intermittent wetting and drying. Investigations also indicate that, although chlorides are present, the rate of steel corrosion will be very slow if the concrete is continuously water-saturated.2.3 In wet concrete, dissolved oxygen will primarily be diffusing in solution, while in a partly dry concrete the diffusion of gaseous oxygen is much faster. For oxygen to be consumed in a cathodic reaction, however, the oxygen has to be in a dissolved state. Since it is the concentration of dissolved oxygen that is important, all factors affecting the solubility of oxygen should also affect its availability. The effect of salt on the corrosion rate has been demonstrated by Griffin and Henry2.4 (see Fig. 2.2). Corrosion increased as the sodium chloride concentration increased until a maximum was reached. Beyond this, the rate of corrosion decreased despite the increased chloride ion concentration. This change in relationship between corrosion and sodium chloride concentration is attributed to the reduced solubility and diffusivity of oxygen, and, therefore, the availability of oxygen to sustain the corrosion process. This represents corrosion in a salt solution; however, the availability of oxygen in wet concrete may be different. Problems due to corrosion of embedded steel have seldom been observed in concrete structures that are continuously submerged, even as in seawater. 2.2.3 The importance of chloride ionsAs will be discussed later, concrete can form an efficient corrosion-preventive

222R-4

ACI MANUAL OF CONCRETE PRACTICE

Fig. 2.1Effect of water-cement ratio and thickness on diffusion of oxygen through mortar and concrete2.2 added. Chlorides may be contained in other admixtures such as some water-reducing admixtures where small amounts of chloride are sometimes added to offset the set-retarding effect of the water reducer. In some cases, where potable water is not available, seawater or water with a high chloride content is used as the mixing water. In some areas of the world, aggregates exposed to seawater (or that were soaked in seawater at one time) can contain a considerable quantity of chloride salts. Aggregates that are porous can contain larger amounts of chloride. 2.2.3.2 Diffusion into mature concrete. Chlorides can permeate through sound concrete, i.e., cracks are not necessary for chlorides to enter the concrete.2.8 Approximate determinations of the diffusion coefficients for chloride in concrete have been published,2.9 but largely the literature neglects the interaction between concrete and chloride. 2.2.3.3 Electrochemical role of free chloride ions. There are three modern theories to explain the effects of chloride ions on steel corrosion: (a) The Oxide Film TheorySome investigators believe that an oxide film on a metal surface is responsible for passivity and thus protection against corrosion. This theory postulates that chloride ions penetrate the oxide film on steel through pores or defects in the film easier than do other ions (e.g., SO4-). Alternatively, the chloride ions may colloidally disperse the oxide film, thereby making it easier to penetrate. (b) The Adsorption TheoryChloride ions are adsorbed on the metal surface in competition with dissolved O2 or hydroxyl ions. The chloride ion promotes the hydration of the metal ions and thus facilitates the dissolution of the metal ions. (c) The Transitory Complex TheoryAccording to this theory, chloride ions compete with hydroxyl ions for the

Fig. 2.2Effect of concentration of sodium chloride on corrosion rate2.4 environment for embedded steel. However, it is well documented2.5-2.7 that the intrusion of chloride ions in reinforced concrete can cause steel corrosion if oxygen and moisture are also available to sustain the reaction. No other common contaminant is documented as extensively in the literature as a cause of corrosion of metals in concrete. Chloride ions may be introduced into concrete in a variety of ways. Some are intentional inclusion as an accelerating admixture; accidental inclusion as contaminants on aggregates; or penetration by deicing salts, industrial brines, marine spray, fog, or mist. 2.2.3.1 Incorporation in concrete during mixing. One of the best known accelerators of the hydration of portland cement is calcium chloride. Generally, up to 2 percent solid calcium chloride dihydrate based on the weight of cement is

CORROSION OF METALS

222R-5

ferrous ions produced by corrosion. A soluble complex of iron chloride forms.2.9 This complex can diffuse away from the anode destroying the protective layer of Fe(OH)2, permitting corrosion to continue. Some distance from the electrode, the complex breaks down, iron hydroxide precipitates and the chloride ion is free to transport more ferrous ions from the anode. Evidence for this process can be observed when concrete with active corrosion is broken open. A light green semisolid reaction product is often found near the steel which, on exposure to air, turns black and subsequently rust red in color. Since corrosion is not stifled, more iron ions continue to migrate into the concrete away from the corrosion site and to react with oxygen to form higher oxides that result in a fourfold volume increase. It is the expansion of iron oxides as they are transformed to higher oxidation states that produce internal stress, which eventually cracks the concrete. Formation of iron chloride complexes may also lead to disruptive forces. 2.2.4 Corrosion rate and pHAs illustrated in Fig. 2.3, the corrosion rate of iron is reduced as the pH increases. Since concrete has a pH higher than 12.5, it is usually an excellent medium for protecting steel from corrosion. Only under conditions where salts are present or the concrete cover has carbonated does the steel become vulnerable to corrosion. 2.2.5 Pourbaix diagramsPourbaix2.10 devised a compact summary of thermodynamic data in the form of electrical potential versus pH diagrams, which includes the electrochemical and corrosion behavior of iron in an aqueous medium. Fig. 2.4 indicates that iron is in a passive state at a pH in the range of 8 to 13. This accounts for the protective properties of concrete and the absence of steel corrosion in concrete when no chloride ions, or only small amounts, are present. A careful inspection of Fig. 2.4 indicates, however, that corrosion may begin if the pH of the system is raised to above about 13. In this case, a soluble ferrite, HFeO2-, forms. Any mechanism where iron dissolves at an appreciable rate in concrete can be considered serious corrosion. Thus, the addition of materials that further increase the pH of concrete may be detrimental; however, the occurrence of this phenomenon in concrete has not been confirmed. 2.2.6 High-strength steels and other metalsThere is little information in the literature suggesting that the newer high-strength steels are either more or less subject to corrosion than the previous standard and lower strength steels. Although information is lacking, it is expected that whatever differences might exist would be secondary to the major factors influencing corrodibility of the steel. Studies have been made on the corrosion of prestressed steel in concrete2.11 that demonstrate the hazards of adding chlorides to concrete containing prestressed steel. Aluminum corrodes in concrete and the rate of corrosion is higher if the aluminum is in contact with steel and chlorides or if alkalies are present.2.12 Lead and tin (e.g., tin solder) can corrode similarly. At pH about 12.5, zinc reacts rapidly to form soluble zincates and hydrogen gas is liberated

Fig. 2.3Effect of pH on corrosion of iron in aerated soft water at room temperature2.1

Fig. 2.4Pourbaix diagram for iron2.10 Zn + OH- + H2O -> HZnO2- + H2

(2-3)

If galvanized steel is to be used in concrete, appropriate measures should be taken to prevent hydrogen evolution in the fresh concrete. A small amount of a chromate salt is generally used.2.13 In submerged concrete structures having freely exposed steel components in metallic connection with the reinforcing steel, galvanic cells may develop with the freely exposed steel forming the anode and the embedded steel the cathode. This may cause an increased corrosion rate on the freely exposed steel.

222R-6

ACI MANUAL OF CONCRETE PRACTICE

Fig. 2.5Iron electrode in concrete2.16 Copper, chromium, nickel, and silver generally do not corrode in concrete. However, they are somewhat susceptible to corrosion when the concrete is located in a marine environment.2.14 2.2.7 Polarization of half-cellsThe equilibrium of a half-cell is characterized by its reversible potential which depends on the Standard Redox Potential, on the concentrations (activities) of the species participating in the half-cell reaction (e.g., OH-, Fe++, etc.), and on the temperature. When a current flows through the half-cell, there is a shift of its measured potential away from the reversible potential. This shift is called polarization. For a given current, the polarization is large for half-cell reactions that are irreversible or nearly so. For instance, the oxygen half-cell reaction [Eq. (2-2)] is almost irreversible and can have a relatively high degree of polarization. Some corrosion processes, although thermodynamically more favored than others as indicated by the reversible potentials of the half-cells, are actually slower in practice because of polarization effects. There are three general kinds of polarization that apply to the anode as well as to the cathode. These are concentration polarization, ohmic polarization, and activation polarization. These three kinds of polarization can be present simultaneously. (a) Concentration polarization occurs when the concentration of the electrolyte changes in the vicinity of the electrode. An example of this would be depletion of oxygen concentration at the cathode. (b) Ohmic polarization occurs because of the ohmic resistance of the electrolyte (e.g., moist concrete) and of any films on the electrode surface. This produces an ohmic potential drop in accordance with Ohms law (IR drop). (c) Activation polarization occurs due to kinetic hindrance of the rate controlling step of the electrode reaction. Tafel2.1 has shown experimentally that for large currents in the absence of concentration and ohmic polarization, the measured polarization , which is the activation polarization, is directly proportional to the logarithm of the current density i = a + blog i where a and b are the so-called Tafel intercept and Tafel slope parameters, respectively. These parameters can be obtained by plotting versus i on semilogarithmic paper. The

Tafel intercept parameter a is related to the exchange current density io, which is the equilibrium current flowing back and forth through the electrode electrolyte interface at equilibrium and a measure of the reversibility of the reaction. The Tafel slope parameter b, on the other hand, gives an insight according to modern electrode reaction theory2.15 into the mechanism of the electrode reaction. 2.2.8 Passivity and transpassivitySpecific to the anodic half-cell in the corrosion process are the concepts of passivity and transpassivity. Passivity of a metal is generally characterized by a thin and tightly adherent oxide film on the metal surface, which tends to protect the metal against further corrosion. The exact composition of the thin and normally invisible film has been difficult to determine. It seems clear, however, that it is made up of chemical combinations of oxygen and is simply called an oxide film. When a potential is applied to an iron electrode, the rate of current flow depends on the state of passivity. Iron in concrete is generally passive, such that little current flows when a potential is progressively increased, eventually the current will flow. This is because, at this point, oxygen is evolved and the electrode reaction involves the electrolysis of water (see Curve I, Fig. 2.5). 2H2O -> 4H+ + O2 + 4e(2-4)

This phenomenon is called transpassivity. The use of a corrosion testing procedure involving an impressed potential has been reported.2.17 Grimes et al.2.18 found that when a high anodic voltage was applied to steel in concrete, the pH around the steel was changed to values ranging between almost 0 and 4. He also found that a liquid pool formed quickly around the reinforcing steel. Neither steel nor concrete is durable in a low pH (acidic) environment. 2.2.9 Types of corrosion-controlling mechanismsAs previously mentioned, it is necessary to have both a cathodic and an anodic reaction for a corrosion process to occur. However, different corrosion situations can be visualized by plotting, on the same graph, the log of the absolute current (I) versus potential (E) curves which would be obtained by polarizing each half-cell with an auxiliary counter electrode in absence of the other half-cell. For this discussion, we assume (a) no IR drop between anodic and cathodic areas and (b) simple algebraic addition of the anodic and cathodic currents. Under these conditions, the rest potential or open circuit potential of the corroding sample will be that at which the two i versus E curves intercept. At this point no net, external current flows, and the absolute value of current at the intercept is equal to the corrosion current. If the cathodic process is the slower process (the one with the larger polarization), the corrosion rate is considered to be cathodically controlled (Fig. 2.6). Conversely, if the anodic process is slower, the corrosion rate is said to be anodically controlled (Fig. 2.7). In concrete, one or two types of corrosion-rate-controlling mechanisms normally dominate. One is cathodic diffusion, where the rate of oxygen diffusion through the concrete

CORROSION OF METALS

222R-7

Fig. 2.6Cathodic control Fig. 2.8Cathodic diffusion control

Fig. 2.7Anodic control

Fig. 2.9Resistance control voltage than that of the anodic surface. The current then flows to the original anodic surface resulting in cathodic reactions occurring there. The difficulties in using this method, however, are to determine the correct potential to apply to the system and to make sure that it is applied uniformly. 2.2.12 Effects of other salt ionsIt has been reported that sulfate2.20 and carbonate2.21 salts can also cause reinforcing steel to corrode. However, this has not been well documented. Although the concrete may have cracked in certain cases and the exposed steel may have appeared rusted, these salts may not have been the primary cause. Their low solubility in a high calcium ion environment would reduce their availability. However, certain soluble salts, such as perchlorates, acetates, and salts of halogens other than chlorine may be corrosive to steel embedded in concrete. Hydrogen sulfide has also been cited as a cause of corrosion.2.22 2.2.13 Stress corrosion crackingStress corrosion2.23 is defined as the process in which the damage caused by stress and corrosion acting together greatly exceeds that produced when they act separately.

determines the rate of corrosion. In Fig. 2.8, cathodic diffusion control is shown for two different rates of oxygen diffusion. The other type of controlling mechanism involves the development of a high resistance path. When steel corrodes in concrete, anodic and cathodic areas may be as much as several feet apart; therefore, the resistance of the concrete may be of great importance. Fig. 2.9 illustrates two cases of resistance control where the potential available for corrosion is the difference in the potential between anode and cathode minus the relevant IR loss. 2.2.10 Stray current corrosionStray electric currents are those that follow paths other than the intended circuit. They can greatly accelerate the corrosion of reinforcing steel. The most common sources of these are electric railways, electroplating plants, and cathodic protection systems. Kondo et al.2.9 reported corrosion of reinforcing steel embedded in concrete used in an electric railroad. 2.2.11 Cathodic protectionThe principle of cathodic protection is to change the potential of a metal to reduce the current flow and thereby the rate of corrosion. This is accomplished by the application of a protective current at a higher

222R-8

ACI MANUAL OF CONCRETE PRACTICE

Fig. 2.10Rate of reaction of gypsum with tricalcium aluminate in portland cement2.29 In stressed steel, a small imperfection caused by corrosion can lead to a serious loss in tensile strength as the corrosion continues at the initial anode area. Another form of corrosion that is related to stress corrosion cracking is intergranular corrosion. In this case, a gas, usually hydrogen, is absorbed in the iron, causing a loss of ductility and cracking. Hydrogen cracking in connection with cathodic protection will be discussed in Chapter 5. Other materials that may cause intergranular corrosion are hydrogen sulfide and high concentrations of ammonia and nitrate salts. The mechanism of how this type of corrosion proceeds is not fully understood; however, it is believed that it involves the reduction in the cohesive strength of the iron. Documented examples of stress corrosion cracking of steel in concrete have not been found in the literature. 2.3Effects of the concrete environment on corrosion 2.3.1 Portland cementWhen portland cement hydrates, the silicates react with water to produce calcium silicate hydrate and calcium hydroxide. The following simplified equations give the main reactions of portland cement with water 2(3CaO SiO2) + 6H2O 3CaO 2SiO2 3H20 + 3Ca(OH)2 (2-5) 2(2CaO SiO2) + 4H2O 3CaO 2SiO2 3H2O + Ca(OH)2 (2-6) As previously mentioned, the high alkalinity of the chemical environment normally present in concrete protects the embedded steel because of the formation of a protective oxide film on the steel. The integrity and protective quality of this film depends on the alkalinity (pH) of the environment. Differences in the types of cement are a result of variation in composition or fineness or both, and as such, not all types of cement have the same ability to provide protection of

embedded steel. According to Pressler et al.2.24 a well-hydrated portland cement may contain from 15 to 30 percent calcium hydroxide by weight of the original cement. This is usually sufficient to maintain a solution at a pH about 13 in the concrete independent of moisture content. Rosenqvist2.25 has described an example of exceedingly rapid steel corrosion in a tropical concrete wharf where a pozzolan was added to the portland cement. Corrosion of the reinforcement was observed shortly after construction was completed. Measurements of pH in the extract from the concrete adjacent to the steel gave values from 5.7 to 8.5. When the pozzolan was omitted, an increase in pH was obtained and no damage was observed. Other research2.26 showed that a pozzolan did not adversely affect the performance of steel in concrete that was partially immersed in a saturated salt solution. In this latter study, the pozzolan studied was a calcined volcanic tuff that was added to the concrete at dosages of 16 and 32 percent by weight of the cement. It may be that the protective qualities of the pozzolan are related to its source or generic type. The use of blended cements might, under certain circumstances, be detrimental because of a reduction in alkalinity. However, blended cements can give a substantial reduction in permeability and also an increase in electrical resistivity especially where a reduction in the water-cement ratio is made possible. Also, such blended cements may give concrete as much as two to five times higher resistance to chloride penetration than concrete made with portland cements.2.27 The effects would be beneficial as far as corrosion is concerned and in some circumstances the benefits associated with blended cements more than offset the adverse effects. Even for cements with the highest reserve basicity, the alkalinity may be reduced in a number of ways. Reduction of alkalinity by leaching of soluble alkaline salts with water is an obvious process. Partial neutralization by reaction with carbon dioxide (carbonation), as present either in air or dissolved in water, is another common process. The silicates are the major components in portland cement imparting strength to the matrix. No reactions have been detected between chloride ions and silicates. Calcium chloride accelerates the hydration of the silicates when at least 1 percent by weight is added.2.28 Calcium chloride seems to act as an accelerator in the hydration of tricalcium silicate as well as to promote the corrosion of steel. Also present in portland cement are C3A and an aluminoferrite phase reported as C4AF. The C3A reacts rapidly in the cement system to cause flash set unless it is retarded. Calcium sulfate is used as the retarder. Calcium sulfate forms a coating of ettringite (C3A 3CaSO4 32H2O) around the aluminate grains thereby retarding their reactivity (Fig. 2.10). Calcium chloride also forms insoluble reaction products with the aluminates in cement (see Fig. 2.11). The most commonly noted complex is C3A CaCl2 xH2O, Friedels salt (Table 2.1). The rate of formation of this material is slower than that of ettringite (compare Fig. 2.10 with Fig. 2.11). The chloridealuminate complex forms after ettringite and prevents further reactions of sulfate with the remaining aluminates.2.29

CORROSION OF METALS

222R-9

Table 2.1Comparison of producers analysis of eight different portland cements with total amount of calcium chloride that reacted with each cement2.29
1/ SO content* 3 3

0.20 0.05 0.30 0.07 0.72 0.25 0.26 0.23 Average

C3A + C4AF - 1/3SO3 1.43 1.42 1.10 1.62 0.69 1.03 1.05 1.15 1.18

Amount of calcium chloride reacted 1.19 1.37 1.18 1.45 0.63 1.08 1.07 1.05 1.13

*Data expressed in molar equivalents rather than usual percent for comparison purposes. Wide differences in SO3 content obtained intentionally.

The chemical combining of C3A with chlorides is frequently referred to as a beneficial effect in that it will reduce the rate of chloride penetration into concrete. According to Mehta,2.30 a chemical binding of penetrating chlorides cannot be expected unless the C3A content is higher than about 8 percent. However, recent experimental observations have shown that as much as 8.6 percent C3A does not effectively reduce chloride penetration when concrete is immersed in seawater2.31 and other studies have found more corrosioninduced distress associated with high C3A contents.2.32 2.3.2 AggregateThe aggregate generally has little effect on the corrosion of steel in concrete. There are exceptions. The most serious problems arise when the aggregates contain chloride salts. This can happen when sand is dredged from the sea or taken from seaside or arid locations. Porous aggregates can absorb considerable quantities of salt. Care should be exercised when using admixtures containing chloride in combination with lightweight aggregates. Helms and Bowman2.33 found that reinforcing steel in lightweight concrete was particularly susceptible to corrosion when exposed to moisture and changes in temperature. Lightweight aggregate containing sulfides can be damaging to high-strength steel under stress. 2.3.3 WaterThe effect of moisture content or degree of water saturation on the rate of oxygen diffusion into concrete has already been discussed. A high moisture content will also substantially reduce the rate of diffusion of carbon dioxide and hence the rate of carbonation of the concrete. An important effect of the moisture content of concrete is its effect on the electrical resistivity of the concrete. Progressive drying of initially water-saturated concrete results in the electrical resistivity increasing from about 7 103 ohm cm to about 6000 103 ohm cm (Fig. 2.12).2.34 Field observations indicate that when the resistivity exceeds a level of 50 to 70 103 ohm cm, steel corrosion would be negligible.2.35 Other authors quote values of resistivity of 10 x 103 ohm cm2.36 and 12 103 ohm cm2.37 above which corrosion induced damage is unlikely even in the presence of chloride ions, oxygen, and moisture. 2.3.4 Corrosion inhibiting admixturesNumerous chemical admixtures, both organic and inorganic, have been suggested as specific inhibitors of steel corrosion.2.13 Some of the

Fig. 2.11Rate of reaction of calcium chloride with portland cement2.29

Fig. 2.12Effect of water saturation on the resistivity of concrete2.34 admixtures, however, may retard time of setting of the cement or be detrimental at later ages. Many would be subject to leaching and hence less effective in concrete that has lost soluble material by leaching. Among those compounds reported as inorganic inhibitors are potassium dichromate, stannous chloride, zinc and lead chromates, calcium hypophosphite,

222R-10

ACI MANUAL OF CONCRETE PRACTICE

concrete, in situations where the rate of carbonation is extremely slow, carbonation is normally not a problem unless cracking of the concrete has occurred or the concrete cover is defective or very thin. Carbonation is not a problem in very dry concrete or in water-saturated concrete. Maximum carbonation rates are observed at about 50 percent water saturation. A more complete discussion of carbonation and the corrosion of steel in carbonated concrete is given in References 2.17 and 2.21. 2.4References
2.1. Uhlig, Herbert H., Corrosion and Corrosion Control, 2nd Edition, John Wiley & Sons, New York, 1971, 419 pp. 2.2. Gjrv, O. E.; Vennesland, ; and El-Busaidy, A. H. S., Diffusion of Dissolved Oxygen through Concrete, Paper No. 17, NACE Corrosion 76, National Association of Corrosion Engineers, Houston, Mar. 1976, 13 pp. 2.3. Shalon, R., and Raphael, M., Inuence of Sea Water on Corrosion of Reinforcement, ACI JOURNAL, Proceedings V. 55, No. 8, Feb. 1959, pp. 1251-1268. 2.4. Grifn, Donald F., and Henry, Robert L., Effect of Salt in Concrete on Compressive Strength, Water Vapor Transmission, and Corrosion of Reinforcing Steel, Proceedings, ASTM, V. 63, 1963, pp. 1046-1078. 2.5. Stratfull, R. F.; Jurkovich, W. J.; and Spellman, D. L., Corrosion Testing of Bridge Decks, Transportation Research Record No. 539, Transportation Research Board, 1975, pp. 50-59. 2.6. Browne, Roger D., Mechanisms of Corrosion of Steel in Concrete in Relation to Design, Inspection, and Repair of Offshore and Coastal Structures, Performance of Concrete in Marine Environment, SP-65, American Concrete Institute, Detroit, 1980, pp. 169-204. 2.7. Aziz, M. A., and Mansur, M. A., Deterioration of Marine Concrete Structures with Special Emphasis on Corrosion of Steel and Its Remedies, Corrosion of Reinforcement in Concrete Construction, Ellis Horwood Limited, Chichester, 1983, pp. 91-99. 2.8. Gjrv, Odd E., Durability of Concrete Structures in the Ocean Environment, Proceedings, FIP Symposium on Concrete Sea Structures (Tbilisi, Sept. 1972), Federation Internationale de la Precontrainte, London, 1973, pp. 141-145. 2.9. Foley, R. T., Complex Ions and Corrosion, Journal, Electrochemical Society, V. 122, No. 11, 1975, pp. 1493-1549. 2.10. Pourbaix, M., Atlas of Electrochemical Equilibrium in Aqueous Solutions, Pergamon Press Limited, London, 1976. 2.11. Monfore, G. E., and Verbeck, G. J., Corrosion of Prestressed Wire in Concrete, ACI JOURNAL, Proceedings V. 57, No. 5, Nov. 1960, pp. 491-516. 2.12. Monfore, G. E., and Ost, Borje, Corrosion of Aluminum Conduit in Concrete, Journal, PCA Research and Development Laboratories, V. 7, No. 1, Jan. 1967, pp. 10-22. 2.13. Boyd, W. K., and Tripler, A. B., Corrosion of Reinforcing Steel Bars in Concrete, Materials Protection, V. 7, No. 10, 1968, pp. 40-47. 2.14. Baker, E. A.; Money, K. L.; and Sanborn, C. B., Marine Corrosion Behavior of Bare and Metallic-Coated Steel Reinforcing Rods in Concrete, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 30-50. 2.15. Stern, M., and Geary, A. L., Electrochemical Polarization No. 1 Theoretical Analysis of the Shape of Polarization Curves, Journal, Electrochemical Society, V. 104, No. 1, 1957, pp. 56-63. 2.16. Rosenberg, A. M., and Gaidis, J. M., The Mechanism of Nitrite Inhibition of Chloride Attack on Reinforcing Steel in Alkaline Aqueous Environments, Materials Performance, V. 18, No. 11, 1979, pp. 45-48. 2.17. Corrosion of Reinforcement and Prestressing Tendons: A State of the Art Report, Materials and Structures/Research and Testing (RILEM, Paris), V. 9, No. 51, May-June 1976, pp. 187-206. 2.18. Grimes, W. D.; Hartt, W. H.; and Turner, D. H., Cracking of Concrete in Sea Water Due to Embedded Metal Corrosion, Corrosion, V. 35, No. 7, 1979, pp. 309-316. 2.19. Kondo, Yasuo; Takeda, Akihiko; and Hideshima, Setsuji, Effect of Admixtures on Electrolytic Corrosion of Steel Bars in Reinforced Concrete, ACI JOURNAL, Proceedings V. 56, No. 4, Oct. 1959, pp. 299-312.

Fig. 2.13Gradient of total chloride concentration depth depends on whether chemical reaction occurs with cement2.41

sodium nitrite, and calcium nitrite. Organic inhibitors suggested have included sodium benzoate, ethyl aniline, and mercaptobenzothiazole. With some inhibitors, inhibition occurs only at addition rates sufficiently high to counteract the effects of chlorides. Sodium nitrite has been used with apparent effectiveness in Europe.2.38 Calcium nitrite2.39 is being studied in the United States since several side effects of the sodium salt would be avoided. Some of the side effects are low strength, erratic times of setting, efflorescence, and enhanced susceptibility to the alkali-aggregate reaction.2.40 2.3.5 Concrete qualityConcrete will offer more protection against corrosion of embedded steel if it is of a high quality. A low water-cement ratio will slow the diffusion of chlorides, carbon dioxide, and oxygen and also the increase in strength of the concrete may extend the time before corrosioninduced stresses cause cracking of the concrete. The pore volume and permeability can be reduced by lowering the watercement ratio. The type of cement or use of superplasticizing and mineral admixtures may also be an important factor in controlling the permeability and the ingress of chlorides.2.31 2.3.6 Thickness of concrete cover over steelThe amount of concrete cover over the steel should be as large as possible, consistent with good structural design, the severity of the service environment, and cost. The effect of cover thickness is more than a simple linear relationship. Considering the normal diffusion of an electrolyte into a porous solid without chemical reaction, a relationship such as shown in Fig. 2.13 would be anticipated. However, in the case of cement paste, the diffusion of chloride ions into the paste is accompanied by both physical adsorption and chemical binding. These effects reduce the concentration of chloride ion at any particular site and hence the tendency for inward diffusion is further reduced. This is also illustrated in Fig. 2.13. 2.3.7 CarbonationCarbonation occurs when the concrete reacts with carbon dioxide from the air or water and reduces the pH to about 8.5. At this low pH the steel is no longer passive and corrosion may occur. For high-quality

CORROSION OF METALS 2.20. Eickemeyer, J., Stress Corrosion Cracking of a High-Strength Steel in Saturated Ca(OH)2 Solutions Caused by Cl and SO4 Additions, Corrosion Science, V. 18, No. 4, 1978, pp. 397-400. 2.21. Hamada, M., Neutralization (Carbonation) of Concrete and Corrosion of Reinforcing Steel, Proceedings, 5th International Symposium on the Chemistry of Cement (Tokyo, 1968), Cement Association of Japan, Tokyo, 1969, V. 3, pp. 343-360. 2.22. Treadaway, K. W. J., Corrosion of Prestressed Steel Wire in Concrete, British Corrosion Journal (London), V. 6, 1971, pp. 66-72. 2.23. Sluijter, W. L., and Kreijger, P. C., Potentio Dynamic Polarization Curves and Steel Corrosion, Heron (Delft), V. 22, No. 1, 1977, pp. 13-27. 2.24. Weise, C. H., Determination of the Free Calcium Hydroxide Contents of Hydrated Portland Cements and Calcium Silicates, Analytical Chemistry, V. 33, No. 7, June 1961, pp. 877-882. Also, Research Department Bulletin No. 127, Portland Cement Association. 2.25. Rosenqvist, I. T., Korrosjon av Armeringsstal i Betong, Teknisk Ukeblad (Oslo), 1961, pp. 793-795. 2.26. Spellman, D. L., and Stratfull, R. F., Concrete Variables and Corrosion Testing, Highway Research Record No. 423, Highway Research Board, 1973, pp. 27-45. 2.27. Page, C. L.; Short, N. R.; and El Tarras, A., Diffusion of Chloride Ions in Hardened Cement Pastes, Cement and Concrete Research, V. 11, No. 3, May 1981, pp. 395-406. 2.28. Ramachandran, V. S., Calcium Chloride in Concrete, Applied Science Publishers, London, 1976, 216 pp. 2.29. Rosenberg, Arnold M., Study of the Mechanism through Which Calcium Chloride Accelerates the Set of Portland Cement, ACI JOURNAL, Proceedings V. 61, No. 10, Oct. 1964, pp. 1261-1270. 2.30. Mehta, P. K., Effect of Cement Composition on Corrosion of Reinforcing Steel in Concrete, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 12-19. 2.31. Gjrv, O. E., and Vennesland, ., Diffusion of Chloride Ions from Seawater into Concrete, Cement and Concrete Research, V. 9, No. 2, Mar. 1979, pp. 229-238. 2.32. Stratfull, R. F., Discussion of Long-Time Study of Cement Performance. Chapter 12Concrete Exposed to Sea Water and Fresh Water, by I. L. Tyler, ACI JOURNAL, Proceedings V. 56, Part 2, Sept. 1960, pp. 1455-1458. 2.33. Helms, S. B., and Bowman, A. L., Corrosion of Steel in Lightweight Concrete Specimens, ACI JOURNAL, Proceedings V. 65, No. 12, Dec. 1968, pp. 1011-1016. 2.34. Gjrv, O. E.; Vennesland, .; and El-Busaidy, A. H. S., Electrical Resistivity of Concrete in the Oceans, OTC Paper No. 2803, 9th Annual Offshore Technology Conference, Houston, May 1977, pp. 581-588. 2.35. Tremper, Bailey; Beaton, John L.; and Stratfull, R. F., Causes and Repair of Deterioration to a California Bridge Due to Corrosion of Reinforcing Steel in a Marine Environment II: Fundamental Factors Causing Corrosion, Bulletin No. 182, Highway Research Board, Washington, D.C., 1958, pp. 18-41. 2.36. Browne, R. D., Design Prediction of the Life for Reinforced Concrete in Marine and Other Chloride Environments, Durability of Building Materials (Amsterdam), V. 1, 1982, pp. 113-125. 2.37. Cavalier, P. G., and Vassie, P. R., Investigation and Repair of Reinforcement Corrosion in a Bridge Deck, Proceedings, Institution of Civil Engineers (London), Part 1, V. 70, 1981, pp. 461-480. 2.38. Woods, Hubert, Durability of Concrete Construction, ACI Monograph No. 4, American Concrete Institute/Iowa State University Press, Detroit, 1968, p. 102. 2.39. Rosenberg, A. M.; Gaidis, J. M.; Kossivas, T. G.; and Previte, R. W., A Corrosion Inhibitor Formulated with Calcium Nitrite for Use in Reinforced Concrete, Chloride Corrosion of Steel in Concrete, STP629, ASTM, Philadelphia, 1977, pp. 89-99. 2.40. Craig, R. J., and Wood, L. E., Effectiveness of Corrosion Inhibitors and Their Inuence on the Physical Properties of Portland Cement Mortars, Highway Research Record No. 328, Highway Research Board, 1970, pp. 77-88. 2.41. Verbeck, George J., Mechanisms of Corrosion of Steel in Concrete, Corrosion of Metals in Concrete, SP-49, American Concrete Institute, Detroit, 1975, pp. 21-38.

222R-11

CHAPTER 3PROTECTION AGAINST CORROSION IN NEW CONSTRUCTION 3.1Introduction Measures which can be taken in reinforced concrete construction to protect the steel against corrosion can be divided into three categories: (a) Design and construction practices that maximize the protection afforded by the portland cement concrete. (b) Treatments that penetrate or are applied on the surface of the reinforced concrete member to exclude chloride ion from the concrete. (c) Techniques that prevent corrosion of the reinforcement directly. In the last case, two approaches are possible: to use corrosion-resistant reinforcing steel or to nullify the effects of chloride ions on unprotected reinforcement. 3.2Design and construction practices Through careful design and good construction practices, the protection provided by portland cement concrete to embedded reinforcing steel can be optimized. It is not the sophistication of the structural design that determines the durability of a concrete member in a corrosive environment but the detailing practices.3.1 The provision of adequate drainage and a method of removing drainage water from the structure are particularly important. In reinforced concrete members exposed to chlorides and subjected to intermittent wetting, the degree of protection against corrosion is determined primarily by the depth of cover to the reinforcing steel and the permeability of the concrete.3.2-3.6 Estimates of the increase in corrosion protection provided by an increase in cover have ranged between slightly more than a linear relationship3.3,3.7 to as much as the square of the cover.3.8 The time to spalling is a function of the ratio of cover to bar diameter,3.8 the reinforcement spacing, and the concrete strength. Although conventional portland cement concrete is not impermeable, concrete with a very low permeability can be made through the use of good quality materials, a minimum water-cement ratio consistent with placing requirements, good consolidation and finishing practices, and proper curing. In concrete that is continuously submerged, the rate of corrosion is controlled by the rate of oxygen diffusion that is not significantly affected by the concrete quality or the thickness of cover.3.9 However, as noted in Chapter 2, corrosion of embedded steel is a rare occurrence in continuously submerged concrete structures. Placing limits on the allowable amounts of chloride ion in concrete is an issue still under active debate. On the one side are the purists who would like to see essentially no chlorides in concrete. On the other are the practitioners, including those who must produce concrete under cold weather conditions, precast concrete manufacturers who wish to minimize curing times, producers of chloride-bearing aggregates, and some admixture companies, who would prefer the least restrictive limit possible. Since chlorides are present naturally in most concrete-making materials, specifying a zero chloride content for any of the mix ingredients is unrealistic.

222R-12

ACI COMMITTEE REPORT

However, it is also known that wherever chloride is present in concrete, the risk of corrosion increases as the chloride content increases. When the chloride content exceeds a certain value (termed the chloride corrosion threshold), unacceptable corrosion may occur provided that other necessary conditions, chiefly the presence of oxygen and moisture, exist to support the corrosion reactions, It is a difficult task to establish a chloride content below which the risk of corrosion is negligible, which is appropriate for all mix ingredients and under all exposure conditions, and which can be measured by a standard test. Three different analytical values have been used to designate the chloride content of fresh concrete, hardened concrete, or any of the concrete mixture ingredients: (a) total, (b) acid-soluble, and (c) water-soluble. The total chloride content of concrete is measured by the total amount of chlorine. Special analytical methods are necessary to determine it, and acid-soluble chloride is often mistakenly called total chloride. The acid-soluble method is the test method in common use and measures that chloride that is soluble in nitric acid. Water-soluble chloride is chloride extractable in water under defined conditions. The result obtained is a function of the analytical test procedure, particularly with respect to particle size, extraction time, and temperature, and to the age and environmental exposure of the concrete. It is also important to distinguish clearly between chloride content, sodium chloride content, calcium chloride content, or any other chloride salt content. In this report, all references to chloride content pertain to the amount of chloride ion (Cl-) present. Chloride contents are expressed in terms of the mass of cement unless stated otherwise. Lewis3.10reported that, on the basis of polarization tests of steel in saturated calcium hydroxide solution and water extracts of hydrated cement samples, corrosion occurred when the chloride content was 0.33 percent acid-soluble chloride, or 0.16 percent water-soluble chloride based on a 2 hr extraction in water. However, it has been shown that the porewater in many typical portland cement concretes made using relatively high-alkali cements is a strong solution of sodium and potassium hydroxides with a pH approaching 14. Since the passivity of embedded steel is determined by the ratio of the hydroxyl concentration to the chloride concentration,3.11 the amounts of chloride that can be tolerated in concrete are higher than those that will cause pitting corrosion in a saturated solution of calcium hydroxide.3.12 Work at the Federal Highway Administration laboratories3.5 showed that for hardened concrete subject to externally applied chlorides, the corrosion threshold was 0.20 percent acid-soluble chlorides. The average content of water-soluble chloride in concrete was found to be 75 to 80 percent of the content of acid-soluble chloride in the same concrete. This corrosion threshold value was subsequently confirmed by field studies of bridge decks including those in California3.13 and New York.3.14 These investigations show that, under some conditions, a chloride content of as little as 0.15 percent water-soluble chloride (or 0.20 percent acidsoluble chloride) is sufficient to initiate corrosion of embedded

steel in concrete exposed to chlorides in service. However, in determining a limit on the chloride content of the mix ingredients, several other factors need to be considered. As noted in the figures already given, the water-soluble chloride content is not a constant proportion of the acidsoluble chloride content. It varies with the amount of chloride in the concrete,3.10 the mix ingredients, and the test method. All the materials used in concrete contain some chlorides, and, in the case of cement, the chloride content in the hardened concrete varies with cement composition. Although aggregates do not usually contain significant amounts of chloride,3.15 there are exceptions. There are reports of aggregates with an acid-soluble chloride content of more than 0.1 percent of which less than one-third is water-soluble when the aggregate is pulverized.3.16 The chloride is not soluble when the aggregate is placed in water over an extended period and there is no difference in the corrosion performance of structures in southern Ontario made from this aggregate when compared to other chloride-free aggregates in the region. However, this is not always the case. Some aggregates, particularly those from arid areas or dredged from the sea, may contribute sufficient chloride to the concrete to initiate corrosion. A limit of 0.06 percent acid-soluble chloride ion in the combined fine and coarse aggregate (by mass of the aggregate) has been suggested with a further proviso that the concrete should not contain more than 0.4 percent chloride (by mass of the cement) derived from the aggregate.3.17 There is thought to be a difference in the chloride corrosion threshold value depending on whether the chloride is present in the mix ingredients or penetrates the hardened concrete from external sources. When chloride is added to the mix, some will chemically combine with the hydrating cement paste. The amount of chloride forming chloroaluminates is a function of the C3A content of the cement.3.18 Chloride added to the mix will also tend to be distributed relatively uniformly and, therefore, has less tendency to cause the creation of concentration cells. Conversely, when chloride permeates from the surface of hardened concrete, uniform chloride contents will not exist around the steel because of differences in the concentration of chlorides on the concrete surface resulting from poor drainage, for example, local differences in permeability, and variations in the depth of cover to the steel. All these factors promote differences in the environment (oxygen, moisture, and chloride content) along a given piece of reinforcement. Furthermore, most structures contain reinforcement at different depths, and, because of the procedures used to fix the steel, the steel is electrically connected. Thus, when chloride penetrates the concrete, some of the steel is in contact with chloride-contaminated concrete while other steel is in chloride-free concrete. This creates a macroscopic corrosion cell that can possess a large driving voltage and a large cathode to small anode ratio which accelerates the rate of corrosion. In seawater, it has been suggested that the permeability of the concrete to chloride penetration is reduced by the precipitation of magnesium hydroxide.3.19 In laboratory studies3.20 in which sodium chloride was added to the mix ingredients, it was found that a substantial increase in

CORROSION OF METALS

222R-13

corrosion rate occurred between 0.4 and 0.8 percent chloride, although the moisture conditions of the test specimens were not clearly defined. Other researchers have suggested3.21 that the critical level of chloride in the mix ingredients to initiate corrosion is 0.3 percent and that this has a similar effect to 0.4 percent chloride penetrating the hardened concrete from external sources. In studies in which calcium chloride was added to portland cement concrete, it was found3.22 that the chloride ion concentration in the pore solution remained high during the first day of hydration. It declined gradually, but it appeared that substantial concentrations of chloride ion remained in solution indefinitely. Chloride limits in national codes vary widely. ACI 318 allows a maximum water-soluble chloride ion content of 0.06 percent in prestressed concrete, 0.15 percent for reinforced concrete exposed to chloride in service, 1.00 percent for reinforced concrete that will be dry or protected from moisture in service, and 0.30 percent for all other reinforced concrete construction. The British Code, CP 110, allows an acid-soluble chloride ion content of 0.35 percent for 95 percent of the test results with no result greater than 0.50 percent. These values are largely based on an examination of several structures in which it was found there was a low risk of corrosion up to 0.4 percent chloride added to the mixture.3.23 However, corrosion has occurred at values less than 0.4 percent,3.24 particularly where the chloride content was not uniform. The Norwegian Code, NS 3474, allows an acid-soluble chloride content of 0.6 percent for reinforced concrete made with normal portland cement but only 0.002 percent chloride ion for prestressed concrete. Both these codes are under revision. Other codes have different limits, though the rationale for these limits is not well established. Corrosion of prestressing steel is generally of greater concern than corrosion of conventional reinforcement because of the possibility that corrosion may cause a local reduction in cross section and failure of the steel. The high stresses in the steel also render it more vulnerable to stresscorrosion cracking and, where the loading is cyclic, to corrosion fatigue. However, most examples of failure of prestressing steel that have been reported3.24-3.26 have resulted from macrocell corrosion reducing the load carrying area of the steel. Nevertheless, because of the greater vulnerability and the consequences of corrosion of prestressing steel, chloride limits in the mix ingredients are lower than for conventional concrete. Based on the present state of knowledge, the following chloride limits in concrete used in new construction, expressed as a percentage by weight of portland cement, are recommended to minimize the risk of chloride-induced corrosion.
Category Test method Prestressed concrete Reinforced concrete in wet conditions Reinforced concrete in dry conditions
*

Chloride limit for new construction Acid-soluble Water-soluble ASTM C 1152 ASTM C 1218 Soxhlet* 0.08 0.10 0.20 0.06 0.08 0.15 0.06 0.08 0.15

The Soxhlet Method of Test is described in the appendix to this report.

Normally concrete materials are tested for chloride content using either the acid-soluble test described in ASTM C 1152, Acid-Soluble Chloride in Mortar and Concrete, or water-soluble test described in ASTM C 1218, WaterSoluble Chloride in Mortar and Concrete. If the materials meet the requirements given in either of the relevant columns in the previous table they are acceptable. If they meet neither of the relevant limits given in the table then they may be tested using the Soxhlet test method. Some aggregates, such as those discussed previously,3.16 contain a considerable amount of chloride that is sufficiently bound that it does not initiate or contribute towards corrosion. The Soxhlet test appears to measure only those chlorides that contribute to the corrosion process, thus permitting the use of some aggregates that would not be allowed if only the ASTM C 1152 or ASTM C 1218 tests were used. If the materials fail the Soxhlet test, then they are not suitable. For prestressed and reinforced concrete that will be exposed to chlorides in service, it is advisable to maintain the lowest possible chloride levels in the mix to maximize the service life of the concrete before the critical chloride content is reached and a high risk of corrosion develops. Consequently, chlorides should not be intentionally added to the mix ingredients even if the chloride content in the materials is less than the stated limits. In many exposure conditions, such as highway and parking structures, marine environments, and industrial plants where chlorides are present, additional protection against corrosion of embedded steel is necessary. Since moisture and oxygen are always necessary for electrochemical corrosion, there are some exposure conditions where the chloride levels may exceed the recommended values and corrosion will not occur. Concrete which is continuously submerged in seawater rarely exhibits corrosioninduced distress because there is insufficient oxygen present. Similarly, where concrete is continuously dry, such as the interior of a building, there is little risk of corrosion from chloride ions present in the hardened concrete. However, interior locations that are wetted occasionally, such as kitchens or laundry rooms or buildings constructed with lightweight aggregate that is subsequently sealed (e.g., with tiles) before the concrete dries out, are susceptible to corrosion damage. Whereas the designer has little control over the change in use of a building or the service environment, the chloride content of the mix ingredients can be controlled. Estimates or judgments of outdoor dry environments can be misleading. Stratfull3.27 has reported the case of approximately 20 bridge decks containing 2 percent calcium chloride built by the California Department of Transportation. The bridges were located in an arid area where the annual rainfall was about 5 in., most of which fell at one time. Within 5 years of construction, many were showing signs of corrosion-induced spalling and most were removed from service within 10 years. For these reasons, the committee believes a conservative approach is necessary. The maximum chloride limits suggested in this report differ from those suggested by ACI Committee 2013.2 and those contained in the ACI Building Code. When making a comparison between the figures, it should be noted that the

222R-14

ACI COMMITTEE REPORT

Fig. 3.1Effect of water-cement ratio on salt penetration3.5

Fig. 3.2Effect of inadequate consolidation on salt penetration3.5

Fig. 3.3Effect of water-cement ratio and depth of cover on relative time to corrosion3.5 figures in this report refer to acid-soluble chlorides while the other documents make reference to water-soluble chlorides. As noted previously, Committee 222 has taken a more conservative approach because of the serious consequences of corrosion, the conflicting data on corrosion threshold values, and the difficulty of defining the service environment throughout the life of a structure. Various nonferrous metals and alloys will corrode in damp or wet concrete. Surface attack of aluminum occurs in the

presence of alkali-hydroxide solutions. Anodizing provides no protection. Much more serious corrosion can occur if the concrete contains chloride ions, particularly if there is electrical (metalto-metal) contact between the aluminum and steel reinforcement, because a galvanic cell is created. Serious cracking or splitting of concrete over aluminum conduits has been reported.3.28,3.29 Certain organic protective coatings have been recommended3.30 when aluminum must be used and it is impracticable to avoid contamination by chlorides. Other metals such as zinc, nickel, and cadmium, which have been evaluated for use as coatings for reinforcing steel, are discussed elsewhere in this chapter. Additional information is contained in Reference 3.31. Where concrete will be exposed to chloride, the concrete should be made with the lowest water-cement ratio consistent with achieving maximum consolidation and density. The effects of water-cement ratio and degree of consolidation on the rate of ingress of chloride ions are shown in Fig. 3.1 and 3.2. Concrete with a water-cement ratio of 0.40 was found to resist penetration by deicing salts significantly better than concretes with water-cement ratios of 0.50 and 0.60. A low water-cement ratio is not, however, sufficient to insure low permeability. As shown in Fig. 3.2, a concrete with a watercement ratio of 0.32 but with poor consolidation is less resistant to chloride ion penetration than a concrete with a watercement ratio of 0.60. The combined effect of water-cement ratio and depth of cover is shown in Fig. 3.3, which illustrates the number of daily applications of salt before the chloride content reached the critical value (0.20 percent acid-soluble) at the various depths. Thus, 40 mm (1.5 in.) of 0.40 water-cement ratio concrete was sufficient to protect embedded reinforcing steel against corrosion for 800 applications of salt. Equivalent protection was afforded by 70 mm (2.75 in.) of concrete with a water-cement ratio of 0.50 or 90 mm (3.5 in.) of 0.60 water-cement ratio concrete. On the basis of this work, ACI 201.2R recommends a minimum of 50 mm (2 in.) cover for bridge decks if the water-cement ratio is 0.40 and 65 mm (2.5 in.) if the water-cement ratio is 0.45. Even greater cover, or the provision of additional corrosion protection treatments, may be required in some environments. These recommendations can also be applied to other reinforced concrete components exposed to chloride ions and intermittent wetting and drying. Even where the recommended cover is specified, construction practices must insure that the specified cover is achieved. Conversely, placing tolerances, the method of construction, and the level of inspection should be considered in determining the specified cover. The role of cracks in the corrosion of reinforcing steel is controversial. Two schools of thought exist.3.32,3.33 One viewpoint is that cracks reduce the service life of structures by permitting more rapid penetration of carbonation and a means of access of chloride ions, moisture, and oxygen to the reinforcing steel. The cracks thus accelerate the onset of the corrosion processes and, at the same time, provide space for the deposition of the corrosion products. The other viewpoint is that while cracks may accelerate the onset of corrosion, such

CORROSION OF METALS

222R-15

corrosion is localized. Since the chloride ions eventually penetrate even uncracked concrete and initiate more widespread corrosion of the steel, the result is that after a few years service there is little difference between the amount of corrosion in cracked and uncracked concrete. The differing viewpoints can be partly explained by the fact that the effect of cracks is a function of their origin, width, intensity, and orientation. Where the crack is perpendicular to the reinforcement, the corroded length of intercepted bars is likely to be no more than three bar diameters.3.33 Cracks that follow the line of a reinforcement bar (as might be the case with a plastic shrinkage crack, for example) are much more damaging because the corroded length of bar is much greater and the resistance of the concrete to spalling is reduced. Studies have shown that cracks less than about 0.3 mm (0.012 in.) wide have little influence on the corrosion of reinforcing steel.3.8 Other investigations have shown that there is no relationship between crack width and corrosion.3.34-3.36 Furthermore, there is no direct relationship between surface crack width and the internal crack width. Consequently, it has been suggested that the control of surface crack widths in design codes is not the most rational approach.3.37 For the purposes of design, it is useful to differentiate between controlled and uncontrolled cracks. Controlled cracks are those which can be reasonably predicted from a knowledge of section geometry and loading. For cracking perpendicular to the main reinforcement, the necessary conditions for crack control are that there be sufficient steel so that it remains in the elastic state under all loading conditions, and that at the time of cracking, the steel is bonded (i.e., cracking must occur after the concrete has set). Examples of uncontrolled cracking are cracks resulting from plastic shrinkage, settlement, or an overload condition. Uncontrolled cracks are frequently wide and usually cause concern, particularly if they are active. However, they cannot be dealt with by conventional design procedures, and measures have to be taken to avoid their occurrence or, if they are unavoidable, to induce them at places where they are unimportant or can be conveniently dealt with, by sealing for example. 3.3Methods of excluding external sources of chloride ion from concrete 3.3.1 Waterproof membranesWaterproof membranes have been used extensively to minimize the ingress of chloride ions into concrete. Since external sources of chloride ions are waterborne, a barrier to water will also act as a barrier to any dissolved chloride ions. The requirements for the ideal waterproofing system are straightforward;3.38 it should: (a) Be easy to install, (b) Have good bond to the substrate, (c) Be compatible with all the components of the system including the substrate, prime coat, adhesives, and overlay (where used),

(d) Maintain impermeability to chlorides and moisture under service conditions, especially temperature extremes, crack bridging, aging, and superimposed loads. The number of types of products manufactured to satisfy these requirements makes generalization difficult. Any system of classification is arbitrary, though one of the most useful is the distinction between the preformed sheet systems and the liquid-applied materials.3.38 The preformed sheets are manufactured under factory conditions but are often difficult to install, usually require adhesives, and are highly vulnerable to the quality of the workmanship at critical locations in the installation. Although it is more difficult to control the quality of the liquid-applied systems, they are easier to apply and tend to be less expensive. Given the different types and quality of available waterproofing products, the differing degrees of workmanship, and the wide variety of applications, it is not surprising that laboratory3.39-3.41 and field3.14,3.42,3.43 evaluations of membrane performance have also been variable and sometimes contradictory. Sheet systems generally perform better than liquid-applied systems in laboratory screening tests because quality of workmanship is not a factor. Although there has been little uniformity in both methods of test or acceptance criteria, permeability (usually determined by electrical resistance measurements) has generally been adopted as the most important criterion. However, some membranes do offer substantial resistance to chloride and moisture intrusion, even when pinholes or bubbles occur in the membrane.3.41 Field performance has been found to depend not only on the type of waterproofing material used, but also on the quality of workmanship, weather conditions at the time of installation, design details, and the service environment. Experience has ranged from satisfactory3.43 to failures that have resulted in the membrane having to be removed.3.44,3.45 Blistering, which affects both preformed sheets and liquidapplied materials, is the single greatest problem encountered in applying waterproofing membranes.3.46 It is caused by the expansion of entrapped gases, solvents, and moisture in the concrete after application of the membrane. The frequency of blisters occurring is controlled by the porosity and moisture content of the concrete3.47 and the atmospheric conditions. Water or water vapor is not a necessary requirement for blister formation, but is often a contributing factor. Blisters may also result from an increase in concrete temperature or a decrease in atmospheric pressure during or shortly after application of membranes. The rapid expansion of vapors during the application of hot-applied products sometimes causes punctures (which are termed blowholes) in the membrane. Membranes can be placed without blisters if the atmospheric conditions are suitable during the curing period. Once cured, the adhesion of the membrane to the concrete is usually sufficient to resist blister formation. To insure good adhesion, the concrete surface must be carefully prepared and be dry and free from curing membranes, laitance, and contaminants such as oil drippings. Sealing the concrete prior to applying the membrane is possible, but rarely practical.3.48 Where the membrane is to be covered (for example, with insulation or a protective layer), the risk of blister for-

222R-16

ACI COMMITTEE REPORT

mation can be reduced by minimizing the delay between placement of the membrane and the overlay. Venting layers have been used in Europe, but rarely in North America, to prevent blister formation by allowing the vapor pressures to disperse beneath the membrane. The disadvantages of using venting layers are that they require controlled debonding of the membrane, leakage through the membrane is not confined to the immediate area of the puncture, and they increase cost. 3.3.2 Polymer impregnationPolymer impregnation consists of filling some of the voids in hardened concrete with a monomer and polymerizing in situ. Laboratory studies have demonstrated that polymer impregnated concrete (PIC) is strong, durable, and almost impermeable. The properties of PIC are largely determined by the polymer loading in the concrete. Maximum polymer loadings are achieved by drying the concrete to remove nearly all the evaporable water, removing air by vacuum techniques, saturation with a monomer under pressure, and polymerizing the monomer in the voids of the concrete while simultaneously preventing evaporation of the monomer. In the initial laboratory studies on PIC, polymerization was accomplished by gamma radiation.3.49 However, this is not practical for use in the field. Consequently, chemical initiators, which decompose under the action of heat or a chemical promoter, have been used exclusively in field applications. Multifunctional monomers are often used to increase the rate of polymerization. The physical properties of PIC are determined by the extent to which the ideal processing conditions are compromised. Since prolonged heating and vacuum saturation are difficult to achieve, and increase processing costs substantially, most field applications have been aimed toward producing only a surface polymer impregnation, usually to a depth of about 25 mm (1 in.). Such partially impregnated slabs have been found to have good resistance to chloride penetration in laboratory studies, but field applications have not always been satisfactory.3.5,3.50 There have been a few full-scale applications of PIC to protect reinforcing steel against corrosion and it must still be considered largely an experimental process. Some of the disadvantages of PIC are that the monomers are expensive and that the processing is lengthy and costly. The principal deficiency identified to date has been the tendency of the concrete to crack during heat treatment. 3.3.3 Polymer concrete overlaysPolymer concrete overlays consisting of aggregate in a polymer binder have been placed experimentally. Most monomers have a low tolerance to moisture and low temperatures; hence the substrate must be dry and in excess of 4 C (40 F). Improper mixing of the two (or more) components of the polymer has been a common source of problems in the field. Aggregates must be dry so as not to inhibit the polymerization reaction. Workers should wear protective clothing when working with epoxies and some other polymers because of the potential for skin sensitization and dermatitis.3.51 Manufacturers rec-

ommendations for safe storage and handling of the chemicals must be followed. A bond coat of neat polymer is usually applied ahead of the polymer concrete. Blistering, which is a common phenomenon in membranes, has also caused problems in the application of polymer concrete overlays. A number of applications were reported in the 1960s.3.52,3.53 Many lasted only a few years. More recently, experimental polymer overlays based on a polyester-styrene monomer have been placed, using heavy-duty finishing equipment to compact and finish the concrete.3.54 3.3.4 Portland cement concrete overlaysPortland cement concrete overlays for new reinforced concrete are applied as part of a two-stage construction process. The overlay may be placed before the first-stage concrete has set or several days later, in which case a bonding layer is used between the two lifts of concrete. The advantage of the first alternative is that the overall time of construction is shortened and costs minimized. In the second alternative, cover to the reinforcing steel can be assured, and small construction tolerances achieved because dead load deflections from the overlay are very small. No matter which sequence of construction is employed, materials can be incorporated in the overlay to provide superior properties, such as resistance to salt penetration and wear and skid resistance, than possible using single stage construction. Where the second stage concrete is placed after the first stage has hardened, sand or water blasting is required to remove laitance and to produce a clean, sound surface. Resin curing compounds should not be used on the firststage construction because they are difficult to remove. Etching with acid was once a common means of surface preparation.3.55,3.56 but is now rarely used because of the possibility of contaminating the concrete with chlorides and the difficulty of disposing of the runoff. Several different types of concrete have been used as concrete overlays including conventional concrete,3.57 concrete containing steel fibers,3.57 and internally sealed concrete.3.57,3.60 However, two types of concrete, low-slump and latex-modified concrete, each designed to offer maximum resistance to penetration by chloride ions, have been used most frequently. 3.3.5 Low-slump concrete overlaysThe performance of low-slump concrete is dependent solely on the use of conventional materials and good quality workmanship. The water-cement ratio is reduced to the minimum practical (usually about 0.32) through the use of a high cement content (over 470 kg/m3 or 800 lb/yd3) and a water content sufficient to produce a slump less than 25 mm (1 in.). The concrete is air-entrained and a water-reducing admixture or mild retarder is normally used. The use of such a high cement factor and low workability mixture dictates the method of mixing, placing, and curing the concrete. Following preparation of the first-stage concrete, a bonding agent of either mortar or cement paste is brushed into the base concrete just before the application of the overlay. The base concrete is not normally prewetted. The overlay concrete is mixed on site, using either a stationary paddle mixer

CORROSION OF METALS

222R-17

or a mobile continuous mixer, because truck mixers are not suited to producing either the quantity or consistency of concrete required. The concrete must be compacted and screeded to the required surface profile using equipment specially designed to handle stiff mixtures. Such machines are much heavier and less flexible than conventional finishing machines and have considerable vibratory capacity. The permeability of the concrete to chloride ions is controlled by its degree of consolidation, which is often checked with a nuclear density meter as concrete placement proceeds. Wet burlap is placed on the concrete as soon as practicable without damaging the overlay (usually within 20 min of placing), and the wet curing is continued for at least 72 hr. Curing compounds are not used, since not only is externally available water required for more complete hydration of the cement, but the thin overlay is susceptible to shrinkage cracking and the wet burlap provides a cooling effect by evaporation of the water. Low-slump concrete was originally proposed as a repair material for concrete pavements3.55 and was developed for patches and overlays on bridges3.61-3.63 in the early 1960s. More recently, concrete overlays have been used as a protection against reinforcing steel corrosion in new bridges. In general, the performance of low-slump overlays has been good.3.64-3.66 Local bond failures have been reported, but these have been ascribed to inadequate surface preparation3.64 or premature drying of the grout.3.66 The overlays are susceptible to cracking, especially on continuous structures, though this is a characteristic of all rigid overlays. 3.3.6 Latex-modified concrete overlaysLatex-modified concrete is conventional portland cement concrete with the addition of a polymeric latex emulsion. The water of suspension in the emulsion hydrates the cement and the polymer provides supplementary binding properties to produce a concrete having a low water-cement ratio, good durability, good bonding characteristics, and a high degree of resistance to penetration by chloride ions, all of which are desirable properties in a concrete overlay. The latex is a colloidal dispersion of synthetic rubber particles in water. The particles are stabilized to prevent coagulation, and antifoaming agents are added to prevent excessive air entrapment during mixing. Styrene-butadiene latexes have been used most widely. The rate of addition of the latex is approximately 15 percent latex solids by weight of the cement. The construction procedures for latex-modified concrete are similar to those for low-slump concrete with minor modifications. The principal differences are: 1. The base concrete must be prewetted for at least 1 hr prior to placing the overlay, because the water aids penetration of the base and delays film formation of the latex. 2. A separate bonding agent is not always used, because sometimes a portion of the concrete itself is brushed over the surface of the base. 3. The mixing equipment must have a means of storing and dispensing the latex. 4. The latex-modified concrete has a high slump so that conventional finishing equipment can be used.

5. Air entrainment of the concrete is believed not required for resistance to freezing and thawing. 6. A combination of moist curing to hydrate the portland cement and air drying to develop the film forming qualities of the latex are required. Typical curing times are 24 hr wet curing, followed by 72 hr of dry curing. The film-formation property of the latex is temperature sensitive and film strengths develop slowly at temperatures below 13 C (55 F). Curing periods at lower temperatures may need to be extended and application at temperatures less than 7 C (45 F) is not recommended. Hot weather causes rapid drying of the latex-modified concrete, which makes finishing difficult and promotes shrinkage cracking. Some contractors have placed overlays at night to avoid these problems. The entrapment of excessive amounts of air during mixing has also been a problem in the field. Most specifications limit the total air content to 6.5 percent. Higher air contents reduce the flexural, compressive, and bond strengths of the overlay. Furthermore, the permeability to chloride ions increases significantly at air contents greater than about 9 percent. Where a texture is applied to the concrete as, for example, grooves to impart good skid resistance, the time of application of the texture is crucial. If applied too soon, the edges of the grooves collapse because the concrete flows. If the texturing operation is delayed until after the latex film forms, the surface of the overlay tears and, since the film does not reform, cracking often results. High material costs and the superior performance of latexmodified concrete in chloride penetration tests have led to latex-modified concrete overlays being thinner than most low-slump concrete overlays. Typical thicknesses are 40 and 50 mm (1.5 and 2 in.). Although latex-modified overlays were first used in 1957,3.67 the majority of installations were placed after 1975. Performance has been generally satisfactory, though extensive cracking and some debonding have been reported,3.68 especially in overlays 20 mm (0.75 in.) thick that were not applied at the time of the original deck construction. The most serious deficiency reported has been the widespread occurrence of shrinkage cracking in the overlays. Many of these cracks have been found not to extend through the overlay and it is uncertain whether this will impair long-term performance. 3.4Methods of protecting reinforcing steel from chloride ion The susceptibility to corrosion of mild steel reinforcement in common use is not thought to be significantly affected by its composition, grade, or the level of stress.3.69 Consequently, to prevent corrosion of the reinforcing steel in a corrosive environment, either the reinforcement must be made of a noncorrosive material, or conventional reinforcing steel must be coated to isolate the steel from contact with oxygen, moisture, and chlorides. 3.4.1 Noncorrosive steelsNatural weathering steels commonly used for structural steelwork do not perform well in concrete containing moisture and chloride3.2 and are not suitable for reinforcement. Stainless steel reinforcement has been used in special applications, especially as hardware for

222R-18

ACI COMMITTEE REPORT

attaching panels in precast concrete construction, but is much too expensive to replace mild-steel reinforcement in most applications. Stainless-steel-clad bars have been evaluated in the FHWA time-to-corrosion studies, and found to reduce the frequency of corrosion-induced cracking compared to black steel in the test slabs, but did not prevent it. However, it was not determined whether the cracking was the result of corrosion of the stainless steel or corrosion of the basis steel at flaws in the cladding. 3.4.2 CoatingsMetallic coatings for steel reinforcement fall into two categories, sacrificial and noble or nonsacrificial. In general, metals with a more negative corrosion potential (less noble) than steel, such as zinc and cadmium, give sacrificial protection to the steel. If the coating is damaged, a galvanic couple is formed in which the coating is the anode. Noble coatings such as copper and nickel protect the steel only as long as the coating is unbroken, since any exposed steel is anodic to the coating. Even where steel is not exposed, macrocell corrosion of the coating may occur in concrete through a mechanism similar to the corrosion of uncoated steel. Nickel,3.70,3.71 cadmium,3.72 and zinc3.70,3.73,3.74 have all been shown to be capable of delaying, and in some cases preventing, the corrosion of reinforcing steel in concrete, but only zinc-coated (galvanized) bars are commonly available. Results of the performance of galvanized bars have been conflicting, in some cases extending the time-to-cracking of laboratory specimens,3.75 in others reducing it3.76 and sometimes giving mixed results.3.77 It is known that the zinc will corrode in concrete3.71,3.78 and that pitting can occur under conditions of nonuniform exposure in the presence of high chloride concentrations.3.79 Field studies of galvanized bars in service for many years in either a marine environment or exposed to deicing salts have failed to show any deficiencies in the concrete.3.74 However, in these studies, chloride ion concentrations at the level of the reinforcing steel were low, such that the effectiveness could not be established conclusively. Marine studies3.80 and accelerated field studies3.81 have shown that galvanizing will delay the onset of delaminations and spalls but will not prevent them. In general, it appears that only a slight increase in life will be obtained in severe chloride environments.3.82 When galvanized bars are used, all bars and hardware in the structure should be coated with zinc to prevent galvanic coupling between coated and uncoated steel.3.82 Numerous nonmetallic coatings for steel reinforcement have been evaluated,3.83-3.86 but only fusion-bonded epoxy powder coatings are produced commercially and widely used. The epoxy coating isolates the steel from contact with oxygen, moisture, and chloride. The process of coating the reinforcing steel with the epoxy consists of electrostatically applying finely divided epoxy powder to thoroughly cleaned and heated bars. Many plants operate a continuous production line and many have been constructed specifically for coating reinforcing steel. Integrity of the coating is monitored by a holiday detector installed directly on the production line. The use of epoxy-

coated reinforcing steel has increased substantially since it was first used in 1973. The chief difficulty in using epoxy-coated bars has been preventing damage to the coating in transportation and handling. Cracking of the coating has also been observed during fabrication where there has been inadequate cleaning of the bar prior to coating or the thickness of the coating has been outside specified tolerances. Padded bundling bands, frequent supports, and nonmetallic slings are required to prevent damage during transportation. Coated tie wires and bar supports are needed to prevent damage during placing. Accelerated time-to-corrosion studies have shown that nicks and cuts in the coating do not cause rapid corrosion of the exposed steel and subsequent distress in the concrete.3.87 It should be noted, however, that the damaged bars were not electrically connected to uncoated steel in the early accelerated tests. Subsequent tests3.88 showed that even in the case of electrical coupling to large amounts of uncoated steel, the performance of damaged and nonspecification bars was good, but not as good as when all the steel was coated. Consequently, for long life in severe chloride environments, consideration should be given to coating all the reinforcing steel. If only some of the steel is coated, precautions should be taken to assure that the coated bars are not electrically coupled to large quantities of uncoated steel. A damaged coating can be repaired using a two-component liquid epoxy, but it is more satisfactory to adopt practices that prevent damage to the coating and limit touch-up only to bars where the damage exceeds approximately 2 percent of the area of the bar. Studies have demonstrated that epoxy-coated, deformed reinforcing bars embedded in concrete can have bond strengths and creep behavior equivalent to those of uncoated bars.3.89,3.90 Another study3.91 reported that epoxy-coated reinforcing has less slip resistance than normal mill scale reinforcing, although, for the particular specimens tested, the epoxy-coated bars attained stress levels compatible with ACI development requirements. 3.5Corrosion control methods 3.5.1 Chemical inhibitorsA corrosion inhibitor is an admixture to the concrete used to prevent the corrosion of embedded metal. The mechanism of inhibition is complex and there is no general theory applicable to all situations. The effectiveness of numerous chemicals as corrosion inhibitors for steel in concrete3.69,3.92-3.94 has been studied. The compound groups investigated have been primarily chromates, phosphates, hypophosphites, alkalies, nitrites, and fluorides. Some of these chemicals have been suggested as being effective; others have produced conflicting results in laboratory screening tests. Many inhibitors that appear to be chemically effective produce adverse effects on the physical properties of the concrete, such as a significant reduction in compressive strength. More recently, calcium nitrite has been reported to be an effective corrosion inhibitor3.95 and studies are continuing.3.88

CORROSION OF METALS

222R-19

Admixtures used to prevent corrosion of the steel by waterproofing the concrete, notably silicones, have been found to be ineffective.3.92 3.5.2 Cathodic protectionAlthough cathodic protection is a viable method of protecting reinforcing steel against corrosion in new construction, most installations to date have been to arrest corrosion in existing structures. Consequently, the principles and performance of cathodic protection systems are described in Chapter 5. It should be noted, however, that the reinforcement in many offshore structures is connected to the cathodic protection system used on the exposed steel. This results in protection of the reinforcement and current densities of 0.5 to 1.0 mA/m2 (0.05 to 0.1 mA/ft2) have been reported.3.96 Thus cathodic protection of the reinforcement, though unintentional, has been applied in several of the largest offshore structures. 3.6References
3.1. Durability of Concrete Bridge Decks, NCHRP Synthesis No. 57, Transportation Research Board Washington, D.C., 1979, 60 pp. 3.2. ACI Committee 201, Guide to Durable Concrete (ACI 201.2R77) (Reafrmed 1982), American Concrete Institute, Detroit, 1977, 37 pp. 3.3. Beaton, J. L., and Stratfull, R. F., Environmental Inuence on Corrosion of Reinforcing in Concrete Bridge Substructures, Highway Research Record No. 14, Highway Research Board, 1963, pp. 60-78. 3.4. Ost, Borje, and Monfore, G. E., Penetration of Chloride into Concrete, Journal, PCA Research and Development Laboratories, V. 8, No. 1, Jan. 1966, pp. 46-52. 3.5. Clear, K. C., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 3: Performance after 830 Daily Salt Applications, Report No. FHWA-RD-76-70, Federal Highway Administration, Washington, D.C., 1976, 64 pp. 3.6. Stark, David, Studies of the Relationships between Crack Patterns, Cover over Reinforcing Steel and the Development of Surface Spalls in Bridge Decks, Special Report No. 116, Highway Research Board, Washington, D.C., 1971, pp. 13-21. Also, Research and Development Bulletin No. RD020E, Portland Cement Association. 3.7. Cornet, I., Protection with Mortar Coatings, Materials Protection, V. 6, No. 3, 1967, pp. 56-58. 3.8. Atimay, E., and Ferguson, P. M., Early Chloride Corrosion of Reinforced ConcreteA Test Report, Materials Performance, V. 13, No. 12, 1974, pp. 18-21. 3.9. Gjrv, O. E.; Vennesland, .; and El-Busaidy, A. H. S., Diffusion of Dissolved Oxygen through Concrete, Paper No. 17, NACE Corrosion 76, National Association of Corrosion Engineers, Houston, Mar. 1976, 13 pp. 3.10. Lewis, D. A., Some Aspects of the Corrosion of Steel in Concrete, Proceedings, 1st International Congress on Metallic Corrosion, Butterworths, London, 1962, pp. 547-552. 3.11. Gouda, V. K., Corrosion and Corrosion Inhibition of Reinforcing Steel I: Immersed in Alkaline Solutions, British Corrosion Journal (London), V. 5, 1970, pp. 198-203. 3.12. Arup. H., Recent Progress Concerning Electrochemistry and Corrosion of Steel in Concrete, ARBEM Symposium, Paris, Oct. 1982. 3.13. Stratfull, R. F.; Jurkovich, W. J.; and Spellman. D. L., Corrosion Testing of Bridge Decks, Transportation Research Record No. 539, Transportation Research Board, 1975, pp. 50-59. 3.14. Chamberlin, William P.; Irwin, Richard J.; and Amsler, Duane E., Waterproong Membranes for Bridge Deck Rehabilitation, Research Report No. 52 (FHWA-NY-77-59-1), New York State Department of Transportation/Federal Highway Administration, Washington, D.C., 1977, 43 pp. 3.15. Hime, William D., and Erlin, Bernard, Chloride Free Concrete, ACI JOURNAL, Proceedings V. 74, No. 10, Oct. 1977, p. N7. 3.16. Rogers. C., and Woda, G., The Chloride Ion Content of Concrete Aggregate from Southern Ontario, Report No. EM-17, Ontario Ministry of Transportation and Communications, Downsview, 1977. 3.17. Impurities in Aggregates for Concrete, Advisory Note No. 18, Cement and Concrete Association, London, 1970, 8 pp.

3.18. Roberts, M. H., Effect of Calcium Chloride on the Durability of Pretensioned Wire in Prestressed Concrete. Magazine of Concrete Research (London), V. 14, No. 42, Nov. 1962, pp. 143-154. 3.19. Haynes, Harvey H., Permeability of Concrete in Sea Water, Performance of Concrete in Marine Environment, SP-65, American Concrete Institute, Detroit. 1980, pp. 21-38. 3.20. Locke, C. E., and Siman, A., Electrochemistry of Reinforcing Steel in Salt-Contaminated Concrete, Corrosion of Reinforcing Steel in Concrete, STP-713, ASTM, Philadelphia, 1980, pp. 3-16. 3.21. Browne, Roger D., Mechanisms of Corrosion of Steel in Concrete in Relation to Design, Inspection, and Repair of Offshore and Coastal Structures, Performance of Concrete in Marine Environment, SP-65, American Concrete Institute, Detroit, 1980, pp. 169-204. 3.22. Diamond, S., and Lopez-Flores, F., Fate of Calcium Chloride Dissolved in Concrete Mix Water, Journal, American Ceramic Society, V. 64, No. 11, Nov. 1981, pp. C 162-164. 3.23. The Role of Calcium Chloride in Concrete, Concrete Construction, V. 21, No. 2, Feb. 1976, pp. 57-61. 3.24. Cornet, I., Corrosion of Prestressed Concrete Tanks, Materials Protection, V. 3, No. 1, Jan. 1964, pp. 9-100. 3.25. Peterson, Carl A., Survey of Parking Structure Deterioration and Distress, Concrete International: Design & Construction, V. 2, No. 3, Mar. 1980. pp. 53-61. 3.26. Monfore, G. E., and Verbeck, G. J., Corrosion of Prestressed Wire in Concrete, ACI JOURNAL, Proceedings V. 57, No. 5, Nov. 1960, pp. 491-516. 3.27. A Manual for the Corrosion Control of Bridge Decks, Contract No. DTFH-61-C-00016, Stratfull, R. F., ed., U.S. Department of Transportation, Washington, D.C. 3.28. Conduit in Concrete, Engineering Journal (Montreal), V. 38, No. 10, 1955, pp. 1357-1362. 3.29. Spalled Concrete Traced to Conduit, Engineering News-Record, V. 172, No. 11, Mar. 12, 1964, pp. 28-29. 3.30. McGeary, Frank L., Performance of Aluminum in Concrete Containing Chlorides, ACI JOURNAL, Proceedings V. 63, No. 2, Feb, 1966. pp. 247-266. 3.31. Erlin, B., and Woods, H., Corrosion of Embedded Materials Other than Reinforcing Steel, Signicance of Tests and Properties of Concrete and Concrete-Making Materials, STP-169B, ASTM, Philadelphia, 1978, pp. 300-319. 3.32. Manning, D. G., Corrosion Resistant Design of Concrete Structures, Proceedings, Canadian Structural Concrete Conference, University of Toronto, 1981, pp. 199-223. 3.33. Beeby, A. W., Corrosion of Reinforcing Steel in Concrete in Its Relation to Cracking, The Structural Engineer (London), V. 56A, No. 3, 1978, pp. 77-81. 3.34. Tremper, Bailey, The Corrosion of Reinforcing Steel in Cracked Concrete, ACI JOURNAL, Proceedings V. 43, No. 10, June 1947, pp. 1137-1144. 3.35. Martin, H., and Schiessel, P., The Inuence of Cracks on the Corrosion of Steel in Concrete, Preliminary Report, RILEM International Symposium on the Durability of Concrete, Prague, 1969, V. 2. 3.36. Raphael, M., and Shalon, R., A Study of the Inuence of Climate on Corrosion of Reinforcement, Proceedings, RILEM Symposium on Concrete and Reinforced Concrete in Hot Countries, Building Research Station, Haifa, 1971, pp. 77-96. 3.37. Beeby, A. W., Concrete in the OceansCracking and Corrosion, Technical Report No. 1, CIRIA/UEG, Construction Industry Research and Information Association/Department of Energy, London, 1978. 3.38. Manning, David G., and Ryell, John, Durable Bridge Decks, Report No. RR203, Ontario Ministry of Transportation and Communications, Downsview, 1976, 67 pp. 3.39. Boulware, R. L., and Elliott, A. L., California Seals Salt-Damaged Bridge Decks, Civil Engineering-ASCE, V. 41, No. 10, Oct. 1971, pp. 42-44. 3.40. Van Til, C. J.; Carr, B. J.; and Vallerga, B. A., Waterproof Membranes for Protection of Concrete Bridge Decks: Laboratory Phase, NCHRP Report No. 165, Transportation Research Board, Washington, D.C., 1976, 70 pp. 3.41. Frascoia, R. I., Vermonts Experience with Bridge Deck Protective Systems, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 69-81. 3.42. Spellman, D. L., and Stratfull, R. F., Bridge Deck Membranes Evaluation and Use in California, Report No. CA-DOT-TL-5116-9 73-38, California Department of Transportation, Sacramento, 1973.

222R-20

ACI COMMITTEE REPORT 3.67. Cardone, S. M.; Brown, M. G.; and Hill, A. A., Latex-Modied Mortar in the Restoration of Bridge Structures, Bulletin No. 260, Highway Research Board, Washington, D.C., 1960, pp. 1-13. 3.68. Bishara, A. G., Latex Modied Concrete Bridge Deck Overlays, Field Performance Analysis, Report No. ODOT 2895, Ohio Department of Transportation, Columbus, 1978. 3.69. Verbeck, George J., Mechanisms of Corrosion of Steel in Concrete, Corrosion of Metals in Concrete, SP-49, American Concrete Institute, Detroit, 1975, pp. 21-38. 3.70. Tripler, Arch B.; White, Earl L.; Haynie, F. H.; and Boyd, W. K., Methods of Reducing Corrosion of Reinforcing Steel, NCHRP Report No. 23, Highway Research Board, Washington, D.C., 1966, 22 pp. 3.71. Baker, E. A.; Money, K. L.; and Sanborn, C. B., Marine Corrosion Behavior of Bare and Metallic-Coated Reinforcing Rods in Concrete, Chloride Corrosion of Steel in Concrete, STP-692, ASTM, Philadelphia, 1977, pp. 30-50. 3.72. Bird, C. E., and Strauss, F. J., Metallic Coating for Reinforcing Steel, Materials Protection, V. 6, July 1967, pp. 48-52. 3.73. Cornet, I.; Ishikawa, T.; and Bresler, B., The Mechanism of Steel Corrosion in Concrete Structures, Materials Protection, V. 7, No. 3, Mar. 1968, pp. 44-47. 3.74. Cook, A. R., and Radtke, S. F., Recent Research on Galvanized Steel for Reinforcement of Concrete, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 51-60. 3.75. Cornet, I., and Bresler, B., Corrosion of Steel and Galvanized Steel in Concrete, Materials Protection, V. 5, Apr. 1966, pp. 69-72. 3.76. Grifn, Donald F., Effectiveness of Zinc Coating on Reinforcing Steel in Concrete Exposed to a Marine Environment, Technical Note No. N-1032, Naval Civil Engineering Laboratory, Port Hueneme, 1969, 42 pp. 3.77. Hill, George A.; Spellman, D. L.; and Stratfull, R. F., Laboratory Corrosion Tests of Galvanized Steel in Concrete, Transportation Research Record No. 604, Transportation Research Board, 1976, pp. 25-30. 3.78. Pourbaix, M., Atlas of Electrochemical Equilibria in Aqueous Solutions, translated from the French by J. A. Franklin, Pergamon Press, New York, 1966, pp. 409-410. 3.79. Unz, M., Performance of Galvanized Reinforcement in Calcium Hydroxide Solution, ACI JOURNAL, Proceedings V. 75, No. 3, Mar. 1978, pp. 91-99. 3.80. Sopler, B., Corrosion of Reinforcement in ConcretePart Series D, Report No. FCB 73-4, Norwegian Institute of Technology, University of Trondheim, 1973. 3.81. Arnold, C. J., Galvanized Steel Reinforced Concrete Bridge Decks: Progress Report, Report No. FHWA-MI-78-R1033, Federal Highway Administration, Washington, D.C., 1976. 3.82. Clear, K. C., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 4: Galvanized Reinforcing Steel, Report No. FHWA-RD82-028, Federal Highway Administration, Washington, D.C., 1981. 3.83. Castleberry, J. R., Corrosion Prevention for Concrete and Metal Reinforcing in the Construction Industry, Materials Protection, V. 7, Mar. 1968, pp. 21-28. 3.84. Clifton, J. R.; Beeghley, H. F.; and Mathey, R. G., Nonmetallic Coatings for Concrete Reinforcing Bars, Report No. FHWA-RD-74-18, Federal Highway Administration, Washington, D.C., 1974, 87 pp. 3.85. Backstrom, T. E., Use of Coatings on Steel Embedded in Concrete, Corrosion of Metals in Concrete, SP-49, American Concrete Institute, Detroit, 1975, pp. 103-113. 3.86. Pike, R. G.; Hay, R. E.; Clifton, J. R.; Beeghly, H. F.; and Mathey, R. G., Nonmetallic Protective Coatings for Concrete Reinforcing Steel, Transportation Research Record No. 500, Transportation Research Board, 1974, pp. 36-44. 3.87. Clear, K. C., FCP Annual Progress ReportYear Ending Sept. 30, 1978, Project 4B, Federal Highway Administration, Washington, D.C., 1978. 3.88. Virmani, Y. P.; Clear, K. C.; and Pasko, T. J., Jr., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 5: Calcium Nitrite Admixture or Epoxy-Coated Reinforcing Bars as Corrosion Protection Systems, Report No. FHWA-RD-83-012, Federal Highway Administration, Washington, D.C., 1983, 71 pp. 3.89. Mathey, Robert G., and Clifton, James R., Bond of Coated Reinforcing Bars in Concrete, Proceedings, ASCE, V. 102, ST1, Jan. 1976, pp. 215-229.

3.43. Frascoia, R. I., Evaluation of Bridge Deck Membrane Systems and Membrane Evaluation Procedures, Report No. 77-2, Vermont Department of Highways, Montpelier, 1977. 3.44. Corkill, J. T., A Field Study of the Performance of Bridge Deck Waterproong Systems in Ontario, Proceedings, 1975 Annual Conference, Roads and Transportation Association of Canada, 1975, V. 2, pp. 79-100. 3.45. Membrane Waterproong for Bridge Decks, Final Report, Oregon Department of Transportation, Salem, 1977. 3.46. MacDonald, M. D., Waterproong Concrete Bridge Decks: Materials and Methods, TRRL Report No. LR636, Transport and Road Research Laboratory, Crowthorne, Berkshire, 1974, 32 pp. 3.47. Legvold, T. L., Bridge Deck Waterproong Membrane Study, Report No. R-262, Iowa State Highway Commission, Ames, 1974. 3.48. Meader, A. L., Jr.; Schmitz, C. G.; and Henry, J. E., Development of a Cold Poured Bridge Deck Membrane System, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 164-177. 3.49. Steinberg, M., et al., Concrete-Polymer MaterialsFirst Topical Report, BNL 50134 (T-509) and USBR General Report No. 41, Brookhaven National Laboratory/U.S. Bureau of Reclamation, Denver, 1968. 3.50. Smoak, W. G., Development and Field Evaluation of a Technique for Polymer Impregnation of New Concrete Bridge Deck Surfaces, Report No. FHWA-RD-76-95, Federal Highway Administration, Washington, D.C., 1976. 3.51. Tremper, Bailey, Repair of Damaged Concrete with Epoxy Resins, ACI JOURNAL, Proceedings V. 57, No. 2, Aug. 1960, pp. 173-182. 3.52. McConnell, W. R., Epoxy Surface Treatments for Portland Cement Concrete Pavements, Epoxies with Concrete, SP-21, American Concrete Institute, Detroit, 1968, pp. 9-17. 3.53. Santucci, L. E., Polyester Overlays for Portland Cement Concrete Surfaces, Highway Research Record No. 14, Highway Research Board, 1963, pp. 44-56. 3.54. Jenkins, J. C.; Beecroft, G. W.; and Quinn, W. J., Polymer Concrete Overlays: Interim Users Manual, Report No. FHWA-TS-78-218, Federal Highway Administration, Washington, D.C., 1977. 3.55. Felt, Earl J., Resurfacing and Patching Concrete Pavements with Bonded Concrete, Proceedings, Highway Research Board, V. 35, 1956, pp. 444-469. 3.56. Westall, William G., Bonded Resurfacing and Repairs of Concrete Pavements, Bulletin No. 260, Highway Research Board, Washington, D.C., 1960, pp. 14-24. 3.57. McKeel, W. T., Jr., and Tyson, S. S., Two-Course Bonded Construction and Overlays, ACI JOURNAL, Proceedings V. 72, No. 12, Dec. 1975, pp. 708-713. 3.58. Tyson, S. S., Two-Course Bonded Concrete Bridge Deck ConstructionInterim Report No. 2: Concrete Properties and Deck Condition Prior to Opening to Trafc, Report No. VHTRC-R3, Virginia Highway and Transportation Research Council, Charlottesville, 1976. 3.59. Jenkins, G. H., and Butler, J. M., Internally Sealed Concrete, Report No. FHWA-RD-75-20, Federal Highway Administration, Washington, D.C., 1975, 106 pp. 3.60. Clear, K. C., and Forster, S. W., Internally Sealed Concrete: Material Characterization and Heat Treating Studies, Report No. FHWARD-77-16, Federal Highway Administration, Washington, D.C., 1977, 73 pp. 3.61. Hilton, N., A Two-Inch Bonded Concrete Overlay for the Port Mann Bridge, Engineering Journal (Montreal), May 1964, pp. 39-44. 3.62. OConnor, E. J., Iowa Method of Partial-Depth Portland Cement Resurfacing of Bridge Decks, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 116-123. 3.63. Bukovatz, J. E.; Crumpton, C. F.; and Worley, H. E., Bridge Deck Deterioration Study, Final Report, State Highway Commission of Kansas, Topeka, 1973. 3.64. Manning, D. G., and Owens, M. J., Ontarios Experience with Concrete Overlays for Bridge Decks, RTAC Forum (Ottawa), V. 2, No. 1, 1979, pp. 31-37. 3.65. Bergen, J. V., and Brown, B. C., An Evaluation of Concrete Bridge Deck Resurfacing in Iowa, Special Report, Iowa State Highway Commission, Ames, 1975. 3.66. Tracy, R. G., Bridge Deck Deterioration and Restoration Investigation No. 639, Interim Report, Minnesota Department of Transportation, St. Paul, Dec. 1976.

CORROSION OF METALS 3.90. Clifton, James R.; Mathey, Robert G.; and Anderson, Erik D., Creep of Coated Reinforcing Bars in Concrete, Proceedings, ASCE, V. 105, ST10, Oct. 1979, pp. 1935-1947. 3.91. Johnston, D. W., and Zia, P., Bond Characteristics of Epoxy Coated Reinforcing Bars, Report No. FHWA-NC-82-002, Federal Highway Administration, Washington, D.C., 1982. 3.92. Clear, K. C., and Hay, R. E., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 1: Effect of Mix Design and Construction Parameters, Report No. FHWA-RD-73-32, Federal Highway Administration, Washington, D.C., 1973, 103 pp. 3.93. Craig, R. J., and Wood, L. E., Effectiveness of Corrosion Inhibitors and Their Inuence on the Physical Properties of Portland Cement Mortars, Highway Research Record No. 328, Highway Research Board, 1970, pp. 77-88. 3.94. Grifn, D. F., Corrosion Inhibitors for Reinforced Concrete, Corrosion of Metals in Concrete, SP-49, American Concrete Institute, Detroit, 1975, pp. 95-102. 3.95. Rosenberg, A. M.; Gaidis, J. M.; Kossivas, T. G.; and Previte, R. W., A Corrosion Inhibitor Formulated with Calcium Nitrite for Use in Reinforced Concrete, Chloride Corrosion of Steel in Concrete, STP-629, ASTM, Philadelphia, 1977, pp. 89-99. 3.96. Fidjestol, P.; Askheim, N. E.; and Roland, B., Location of Potential Corrosion Areas in Concrete Marine Structures, Concrete in the Oceans, Phase II, Plc, Final Report.

222R-21

CHAPTER 4PROCEDURES FOR IDENTIFYING CORROSIVE ENVIRONMENTS AND ACTIVE CORROSION IN CONCRETE 4.1Introduction There are many approaches available for identifying corrosive environments and active corrosion of steel in concrete. Generally, a visual inspection of the structure and the environment in which it serves is the first step in any examination. Visual inspections may range from a simple cursory examination to those that are very detailed, wherein all cracks and other visual evidences of physical deterioration are plotted on scaled diagrams of the structure and specific information is gathered on environmental exposure. This type of inspection may also include taking a limited number of cores to be examined visually for evidence of deterioration due to corrosion. The detailed type of visual inspection is time-consuming and costly, and generally only useful for research studies of structure performance. It does not develop the type of information that is required for scheduling of maintenance. There are several techniques and tools that can be used to more specifically delineate areas of deteriorated concrete and areas of potential or active corrosion of steel.4.1-4.6 However, an inspection program can become expensive if it is necessary to survey more than a nominal number of structures. For purposes of planning a maintenance or rehabilitation program, techniques such as suggested by Stratfull, Jurkovich, and Spellman,4,5 or Manning,4.7,4.8 should be used to minimize expenses. These techniques have been derived from experience and include judicious use of visual examinations together with collection of specific information on the extent of physical deterioration, active corrosion, chloride ion contamination, and depth of cover over reinforcing steel. The references noted previously should be studied for more detailed information on these techniques.

4.2Methods of evaluation Certain tools are used for identifying and quantifying corrosive environments, extent of active corrosion, and concrete deterioration.4.9 Following is a brief description of these tools, together with their purpose and limitations. 4.2.1 PachometerThis tool is used to locate reinforcing steel embedded in concrete, and to determine the amount of cover over the steel. It is battery-operated and contains a transistorized oscillator that establishes an elecromagnetic field in a search coil. In the presence of a steel reinforcing bar, the magnetic field is distorted. By calibration, the distance from the bar may be read from the meter dial. Two styles of equipment are available. The first is a handheld device4.10 and the second is an automated device4.11 mounted on wheels. Automatic data recording equipment is added to faciliate the speed with which a survey can be conducted. The knowledge of cover depth is essential if it is desired to obtain samples of the concrete at the level of the reinforcing steel for chloride ion analysis. It is also useful in determining the potential for corrosion and subsequent concrete deterioration since it has been well established that structures in corrosive environments with inadequate concrete cover are subject to early deterioration. 4.2.2 Delamination detectorsThere are many tools that may be used to detect delaminations or subsurface fracture planes parallel to the concrete surface. These devices range from simple chain drags or lightweight hammers to more sophisticated devices such as the Delamtect.4.3,4.9 Almost any sounding device can be used to locate hollow areas or delaminations caused by corrosion of the reinforcing steel. The automated Delamtect is useful for surveying large numbers of bridge decks or other horizontal surfaces such as parking garage floors if a record of the area of delamination is desired. However, the simpler chain drag is adequate for locating delaminated areas during repair operations. 4.2.3 Electrical potential measuring equipmentThis equipment, adapted for use with reinforced concrete by Stratfull,4,2 consists of a copper-copper sulfate half-cell (CSE), a high-impedance voltmeter, and lead wires to connect the half-cell and the reinforcing steel to the voltmeter.4.9 The equipment is used to determine if the reinforcing steel is in a passive or active state relative to active corrosion.4.2 Details of the test method are given in ASTM C 876. The accuracy of the method is good when proper concrete prewetting is used. The significance of the measurements can be summarized as follows for structures exposed to air: Potentials more negative than -0.35 V CSE: Very high probability of the presence of active corrosion. Potentials more positive than -0.20 V CSE: Very high probability of no corrosion. Potentials in the range of -0.20 V to -0.35 V CSE: Uncertainty as to whether or not corrosion is present. It is in the uncertain range that potential differences across a structure, and other detection methods, must often be relied on to deduce the probable condition. In Federal Highway Administration studies, potential differences rarely exceed 100 mV when corrosion was not active, or was active only at extremely low rates. In reinforced concrete undergoing

222R-22

ACI COMMITTEE REPORT

significant corrosion, the potential differences were commonly over 200 mV.4.12,4.13 It should also be noted that a potential value more negative than -0.35 V CSE on actively corroding field structures (with many electrically interconnected anodes and cathodes) generally provides information on only the presence or absence of corrosion and not on corrosion rate. It is the potential difference between the anode and cathode that most closely relates to corrosion rate rather than simply the magnitude of the anode potential. A common example in which highly negative potentials are not indicative of high corrosion rates is a totally water-saturated reinforced concrete structure. In such a structure, oxygen availability to the noncorroding steel is severely restricted and cathodic polarization results. This drives both the anode and the cathode potentials to very negative values, and yet corrosion rate is most often quite low. By careful measurement of potentials on a closely spaced grid pattern, high versus low corrosion rate situations can be identified by studying potential differences. Large potential differences generally indicate high corrosion rates. 4.2.4 Chloride analysisMeasurements of chloride ion concentration at the level of reinforcing steel in concrete are made to determine if any environment exists that is conducive to corrosion of the steel. Two wet chemical analysis techniques are used to isolate chloride from the concrete, one to determine acid-soluble chloride and the other to determine water-soluble chloride. As discussed in Chapter 3, the chloride ion content measured by the water-soluble test is very sensitive to the test procedures. Consequently, all values of chloride ion content in this report are referenced to the acid-soluble test described in ASTM C 114 and also in AASHTO T 260. The preferred method of sample procurement for chloride measurement is to obtain concrete in powdered form without the aid of liquid coolants that could leach out water-soluble chloride. This can be done by using impact drilling equipment and collecting the pulverized material. Alternatively, a 3-in. (80 mm) diameter or larger core can be obtained by wet coring and then extracting samples from the interior by dry sawing or other pulverizing methods. Measurements for acid-soluble and water-soluble chloride can be made on this type of sample using the standard test procedures. Additional guidance is given in References 4.14 and 4.15. In many existing concrete structures, the exact cement content is not known. Thus, chloride levels can be expressed in terms of percent by weight of concrete, or, sometimes, pounds of chloride per cubic yard of concrete. The latter requires an assumed or measured unit weight of the concrete. A table in Reference 4.7 gives the conversion formulas for the various methods of expressing chloride in concrete. For greater precision, nonevaporable water contents (water chemically combined through cement hydration) can be measured on each powdered sample from a given concrete and used as correction factors for aggregate induced distortions in measured chloride levels. In any interpretation of chloride data, sound engineering judgment must be used to assess the actual potential for corrosion. As stated earlier, free moisture and oxygen as well as chloride must be available to induce corrosion. If it can be

concluded that either moisture or oxygen is not available, there would be no corrosion threshold. Such conditions may prevail, for example, in concrete that is continuously submerged or in internal members in buildings where air conditioning units maintain constantly low humidities. However, the difficulty of assessing the possibility of corrosion in the service environment is discussed more fully in Chapter 3. 4.2.5 Rate of corrosion probesTwo basic types of probes are available for embedment into concrete to provide an indication of rate of corrosion. One type involves the use of two or three electrically isolated short sections of steel wire or reinforcing steel and the use of linear polarization techniques to estimate instantaneous corrosion rates.4.17 The second, more widely used device is the electrical resistance rate-of-corrosion probe which provides cumulative rate of corrosion data from periodic measurements of the electrical resistance of a steel wire or hollow cylinder embedded in the concrete. Experiences to date with use of these probes have been conflicting. However, based on recent Federal Highway Administration studies,4.12 this appears to be related more to the rate of corrosion process of steel in concrete than to inaccuracies of the devices themselves. The Federal Highway Administration studies indicate that current flow within physically separated, macroscopic corrosion cells, such as the case of large quantities of steel in chloride-free moist concrete in close proximity and electrically coupled to steel in chloride-bearing concrete, are primarily responsible for the very high rates of corrosion on bridge decks. In contrast, microscopic self-corrosion of a small section of steel in chloride-contaminated concrete most often results in only relatively low corrosion rates. Since the electrically isolated linear polarization devices only simulate this latter process, valid predictions of the overall effect of corrosion on the structure are not possible. The electrical resistance probe, on the other hand, can be electrically coupled with the reinforcing steel in the structure and thus, potentially, can indicate macrocell activity. However, this is possible only if the probe becomes the anode of a macrocell and current flow within this macrocell is typical of that of a macrocell active on the reinforcing steel. These are very important uncertainties and have generally limited the use of resistance probes to research and field evaluation efforts in which special installation procedures are required, and electrical potential and current measurements can be made to define the characteristics of the probe-reinforcing bar-macrocell. Continued study of these devices is needed, as well as lower cost options such as short sections of reinforcing steel installed in a specific manner, and current flow measurements. To date, use of such procedures in the field to aid in studies on the effect of rehabilitation procedures on corrosion rate have provided excellent results. 4.2.6 Electrical resistance measuring equipmentThe primary use of these measurements is to determine the resistance of waterproof membranes that are made from dielectric materials. This equipment consists of a copper contact plate, sponges, ohmmeter, and lead wire.4.1 One terminal of the ohmmeter is connected to the reinforcing steel, while the other terminal is connected to the copper plate. The wetted

CORROSION OF METALS

222R-23

sponges are fastened to the bottom of the copper plate to facilitate contact. The electrical resistance can be measured at any point by moving the copper contact plate to that point. The resistance has been related to the number of holes in the membrane and its permeability to water. This technique provides a nondestructive method to evaluate the environment and potential for corrosion in concrete decks protected by a membrane system. Advantages and cautions needed when using these measurements are discussed in Reference 4.7. The technique has not proven satisfactory for evaluation of epoxy resin coatings on reinforcing steel embedded in concrete. Some work has been done on measurement of the resistance of these coatings using alternating current meters. However, the state of the art is not sufficiently advanced for general use as an evaluation technique. Other techniques and tools to determine the condition of the concrete and to detect the presence of a corrosive environment or active corrosion have been studied to varying degrees. Ultrasonic methods have been used successfully to investigate the quality and condition of concrete. To a lesser extent, infrared and radar scans have been used to record the condition of concrete. Trial uses of these techniques on bridge decks have been encouraging,4.18 but more development work is needed, especially to determine their value in assessing the condition of asphalt-covered decks. A study is underway to develop nondestructive procedures for direct measurement of the rate of corrosion of reinforcing steel in concrete.4.12 Three electrode-linear-polarization measurements are encouraging. Because the measurements are made directly on the in situ reinforcing steel, many of the problems discussed under the section on corrosion probes should not be present. 4.3References
4.1. Spellman, Donald L., and Stratfull, Richard F., An Electrical Method for Evaluating Bridge Deck Coatings, Highway Research Record No. 357, Highway Research Board, 1971, pp. 64-71. 4.2. Stratfull, R. F., Half-Cell Potentials and the Corrosion of Steel in Concrete, Research Report No. CA-HY-MR-51 16-7-72-42, California Department of Transportation, Sacramento, 1972. 4.3. Moore, William M.; Swift, Gilbert; and Milberger, Lionel J., An Instrument for Detecting Delamination in Concrete Bridge Decks, Highway Research Record No. 451, Highway Research Board, 1973, pp. 44-52. 4.4. Clear, K. C., Evaluation of Portland Cement Concrete for Permanent Bridge Deck Repair, Report No. FHWA-RD-74-5, Federal Highway Administration, Washington, D.C., 1974, 48 pp. 4.5. Stratfull, R. F.; Jurkovich, W. J.; and Spellman, D. L., Corrosion Testing of Bridge Decks, Research Report No. CA-DOT-TL-5116-12-75-10, California Department of Transportation, Sacramento, 1975. 4.6. Ross, Joseph E., Bridge Deck Deterioration Study, Research Report No. 85, Louisiana Department of Highways, Baton Rouge, 1975. 4.7. Durability of Concrete Bridge Decks, NCHRP Synthesis No. 57, Transportation Research Board, Washington, D.C., 1979, 60 pp. 4.8. Manning, David G., and Holt, Frank B., Detecting Delamination in Concrete Bridge Decks, Concrete International: Design & Construction, V. 2, No. 11, Nov. 1980, pp. 34-41. 4.9. Clear, K. C., Permanent Bridge Deck Repair, Public Roads, V. 39, No. 2, 1975, pp. 53-62. 4.10. Durability of Concrete Bridge DecksA Cooperative Study, Report No. 2 (EBO44E), Michigan State Highway Department/Bureau of Public Roads/Portland Cement Association, Skokie, 1965, 107 pp. 4.11. Rolling Pachometer, Operating Manual and Specication, Ofce of Development, Federal Highway Administration, Washington, D.C., 1975.

4.12. Clear, K. C., FCP Annual Progress ReportYear Ending Sept. 30, 1981, Project 4K: Cost Effective Rigid Concrete Construction and Rehabilitation in Adverse Environments, Federal Highway Administration, Washington, D.C., 1979. 4.13. Clear, K. C., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 4: Galvanized Reinforcing Steel, Report No. FHWA-RD-82-028, Federal Highway Administration, Washington, D.C., 1981. 4.14. Berman, H. A., Determination of Chloride in Hardened Cement Paste, Mortar, and Concrete, Report No. FHWA-RD-72-12, Federal Highway Administration, Washington, D.C., 1972, 22 pp. 4.15. Clear, K. C., and Harrigan, E. T., Sampling and Testing for Chloride Ion in Concrete, Report No. FHWA-RD-77-85, Federal Highway Administration, Washington, D.C., 1977, 25 pp. 4.16. Clear, K. C., and Hay, R. E., Time-to-Corrosion of Reinforcing Steel in Concrete Slabs, V. 1: Effect of Mix Design and Construction Parameters, Report No. FHWA-RD-73-32, Federal Highway Administration, Washington, D.C., 1973, 103 pp. 4.17. Lankard, D. R.; Slater, J. E.; Hedden, W. A.; and Niesz, D. E., Neutralization of Chloride in Concrete, Report No. FHWA-RD-76-60, Federal Highway Administration, Washington, D.C., 1975, 143 pp. 4.18. Holt, F. B., and Manning, D. G., Detecting Concrete Bridge Deck Delaminations with Infrared Thermography, Public Works, Mar. 1979, pp. 66-69.

CHAPTER 5REMEDIAL MEASURES 5.1 Introduction This chapter discusses measures available to arrest, control, or minimize corrosion activity on an existing reinforced concrete structure after it is found unnecessary to completely replace the structure. Some of the normally accepted and commonly implemented remedial measures for controlling corrosion on steel structures are not applicable to the reinforced concrete composite structure. For example, application of protective coatings is an effective tool for controlling corrosion on an existing steel structure. Obviously, coating of the steel surfaces to be embedded in concrete would have to be accomplished prior to or during the construction phase. Thus the use of protective coatings for steel reinforcement is a design measure and has been discussed in Chapter 3. Chemical inhibitors are commonly used to control corrosion in a closed system. Inhibitors would need to be introduced as an admixture during the concrete batching and mixing phase and are not applicable to the existing reinforced concrete structure. 5.2General Remedial measures for controlling corrosion of steel embedded in portland cement concrete use sound corrosion engineering principles directed toward: 1. Insulating the concrete surfaces from the corrosive environment. 2. Modifying the environment to make it less corrosive. 3. Actively controlling the electron flow within the environment so that no metal is lost from the structure. 4. Applying a combination of the previous techniques. The reinforced concrete structure may be insulated from the corrosive environment through application of an impermeable, dielectric barrier between the structure and the corrosive environment. The barrier may be a coating or membrane applied to the surface of the concrete, may be formed as an integral part of the concrete matrix through polymer impregnation, or may be an overlay of polymer concrete, low-slump concrete, latex-modified concrete, or

222R-24

ACI COMMITTEE REPORT

internally-sealed concrete. These have been discussed in detail in Chapter 3 and will only be mentioned briefly in this chapter. The environment may be altered to reduce corrosion either by removing detrimental constituents (such as the chloride ion), or by removing or neutralizing stray current sources. Cathodic protection can be used to control the direction of electron flow within the steel-environment electrochemical circuit to stop corrosion of the reinforcing steel. 5.3Applicability All reinforced concrete structures are susceptible to corrosion. Although bridge decks5.1-5.3 are perhaps the most notorious examples today, the literature contains many references5.4-5.6 to other types of reinforced concrete structures where corrosion of the reinforcing steel is being or has been experienced. These include buildings, caissons, foundations, parking garages, piers, piles, pipes, silos, tower footings, and water tanks. Some of these structures may be totally or partially buried in soil. Others, such as offshore platforms, water tanks, and the internal surfaces of pipe, are exposed to aqueous solutions. Bridge decks, parking garages, and buildings are exposed to the atmosphere. If the structure is buried or permanently underwater such that the concrete surfaces are not accessible for treatment, and it is impractical to expose them, treatment of the surfaces with surface coatings or membranes or by application of overlays is not applicable. Similarly, if the structure to be maintained is a buried pipeline or an offshore platform exposed in a large body of water, modifying the environment to make it less corrosive would not be a practical solution. Thus, not all the remedies discussed here are applicable to all types of reinforced concrete structures in the various environments. Cathodic protection is by far the most versatile method of corrosion control since it is applicable to any electrically continuous structure within a suitable electrolyte. In as much as the steel embedded in concrete, and not the concrete, requires the protection from metallic corrosion, damp concrete serves as a suitable electrolyte and even structures exposed to the atmosphere, such as bridge decks, can be protected cathodically. 5.4The remedies and their limitations 1. Insulative remedies: The insulative methods currently employed to isolate reinforced concrete structures from the corrosive environment include surface coatings and membranes, polymer impregnation and overlays of polymer concrete, low-slump concrete, or latex-modified concrete. These methods are suitable only when the surfaces of the concrete structure are exposed for treatment. Ideally, these barriers would prevent continued intrusion of harmful contaminants and the availability of oxygen or moisture to sustain the corrosion reactions. However, in existing structures active corrosion is already underway and harmful species have contaminated the concrete. Insulative methods used after active corrosion is experienced do not stop corrosion but may mitigate the effects of the corrosion processes. However, if insulative methods are used without initially decontaminating the concrete, sufficient amounts of the contaminants, oxygen, and moisture may be entrapped such that corrosion

progresses until structural integrity is threatened. These insulative measures, when used without initial decontamination, should be considered as only a temporary remedy.5.7 Also with insulative remedies an erroneous expectation is that a perfect seal is attainable in practice. All barriers will contain discontinuities such as pinholes, breaks, cracks, poor seams, or other defects that will allow intrusion of contaminants in localized areas. Nevertheless, these insulative methods can substantially reduce the rate of intrusion of harmful contaminants and retard the corrosion processes. In many cases, they can successfully extend the useful life of a structure. 2. Modification of the environment: Methods available for rendering the environment less corrosive include the removal or elimination of harmful constituents from the electrolyte. These constituents could be in the form of chemicals such as water and chlorides, gases such as oxygen or hydrogen sulfide, or electrical currents. Water can often be eliminated by facilitating drainage away from rather than through a structure. Chlorides can be eliminated by a process known as electrochemical chloride removal. This process has been used to decontaminate structures such as bridge decks5.8 in research studies. The removal is accomplished electrochemically, by using a suitable electrolyte, an ion exchange resin, and a noble anode. The reinforcing steel is the cathode (negatively charged) in this electrochemical circuit. The negatively charged chloride ion (Cl-) is attracted to the positively charged anode where it is trapped by the exchange resin. Although a field test at Marysville, Ohio, on a bridge deck displayed promising results, the method remains in its initial development stage. In this test, up to 90 percent of the chlorides present in the concrete above the top mat of reinforcement was removed. Disadvantages are that the method is expensive and timeconsuming and requires the application of high direct current voltage that generates heat (around 200 F [90 C]) which, in turn, can induce cracking of the concrete. In addition, the permeability of the concrete was increased. Harmful gases such as oxygen and hydrogen sulfide can be stripped by chemical processes from the electrolyte, thus making it less corrosive. This method is applicable predominantly for structures exposed in aqueous solution. Deep polymer impregnation5.9 of the critically contaminated concrete around the reinforcing steel is being evaluated. The method ties up the existing contaminants and prevents intrusion of additional contaminants. Although the theory is realistic, the practicality and economic feasibility of deep polymerization are not established. Corrosion of steel in concrete can be caused by environmental factors other than chemical constituents such as chlorides, moisture, or oxygen. Stray electrical currents can result in corrosion by electrolysis, i.e., cathodic interference.5.5,5.10 In corrosion by electrolysis, direct current strays from an exterior source and is collected by the steel in a reinforced concrete structure. Inasmuch as the collected current must return to its source to complete the electrochemical circuit, the current is discharged from the structure at some locations. At the point of current discharge from the structure

COR ROSION OF ME TALS

222R-25

Fig. 5.1Underground pipe crossing Layout Line A is cathodically protected

Fig. 5.3Mitigation with galvanic anodes (current returns via low resistance anode)

Fig. 5.2Elevation showing current ow patterns to the electrolyte, metal loss is experienced. The most common sources of stray currents include cathodic protection systems, electrified railways, and electroplating plants. This type of corrosion is most commonly manifested in grounded structures, i.e., those in contact with the earth. A method of mitigating this type of corrosion, implemented for many years in the buried pipe industry, is the installation of resistance bonds. In resistance bonding, the structure being affected is electrically connected through a resistor to the source of the interference currents. In this manner, the current returns to its source via a metallic path such that no metal loss from the affected structure occurs. Another method uses galvanic anodes to drain the collected current. Collected current is passed on to the electrolyte and back to its source from the surface of the anode which corrodes, rather than the structure. Examples of stray current intrusion and mitigation methods are shown in Fig. 5.1 to 5.4. 3. Active control of electron flowFrom the Pourbaix diagram for iron (Fig. 2.4), it can be seen that steel embedded in concrete is normally passivated due to the highly alkaline (high pH) concrete environment. The diagram shows another area wherein no steel corrosion occurs. This area at the lower portion of the diagram is labeled immunity. In this domain, the potential of the steel is more negative than in any naturally occurring condition, regardless of pH. The method of providing the highly negative steel potentials required for immunity is referred to as cathodic protection. In cathodically protecting a structure, a favorable electrochemical circuit is established by installing an external electrode in the electrolyte and passing current (conventional) from that electrode through the electrolyte to the structure to be protected. This current polarizes the potential of the cathodic surfaces (relatively positive) on the steel to that of the anodic

Fig. 5.4Mitigation with resistance bond (current returns via metallic path) (more negative) surfaces. When this is accomplished, there is no current flow between the formerly anodic and cathodic surfaces and corrosion is arrested. This represents a balanced or equilibrium condition. In normal practice, sufficient current is passed to the surfaces so that the formerly anodic areas will be receiving current from the electrolyte and their potential will be shifted to the more negative direction. There are two ways in which the protective electrochemical circuit can be established. One method uses an electrode made of a metal or alloy more negative than the structure to be protected. For example, if the structure to be protected is constructed of steel, either magnesium, zinc, or aluminum may be coupled with the structure. Inasmuch as a protective galvanic cell is set up between the steel and the alloy selected, this method is known as the galvanic anode method of cathodic protection. Also, since the galvanic anodes pass current to the electrolyte, they corrode or sacrifice themselves to protect the structure. Hence, magnesium, zinc, and aluminum are termed sacrificial anodes. Sacrificial anodes corrode at relatively high rates. Corrosion rates for magnesium, zinc, and aluminum are of the order of 17, 26, and 12 lb per amp year, respectively.5.11,5.12 The high consumption rates, as well as low-driving voltage, are the primary disadvantages of the galvanic anode method of cathodic protection. The open circuit potential between steel and magnesium is on the order of 1 V, while zinc and aluminum are somewhat less.5.11,5.12 Thus, with this method, it is imperative that a low-resistance circuit be established by installation of many anodes in a low-resistance media. The anodes installed should also be sized in accordance with their respective consumption rates to provide the necessary design life.

222R-26

ACI COMMITTEE REPORT

Fig. 5.5Galvanic cathodic protection (buried pipe) Fig. 5.7Cathodic protection circuit using conductive mix as secondary anode These anodes are coupled to the structure via the external source of electrical power. This source can be in the form of batteries, thermoelectric generators, generators, or photovoltaic cells. Most commonly, however, alternating current line voltage is converted to direct current by a rectifier. The two ways in which cathodic protection can be achieved are shown in Fig. 5.5 and 5.6. A comparison of the respective advantages and disadvantages of the system are shown in Table 5.1. Successful application of cathodic protection to buried pipe were documented as early as 1946.5.13,5.14 The initial application of cathodic protection to bridge decks was in 1974 and other applications have subsequently been made with encouraging results. The cathodic protection of reinforced concrete structures is thus proven technology and the problems being currently encountered deal with criteria of protection, design, and inspection of the installation. In protecting buried structures or structures exposed in water or in soils, low-resistance electrochemical circuits can normally be established. However, on other structures such as bridge decks, a highly conductive overlay consisting of a coke breeze-asphalt mixture or closely spaced anodes to reduce the circuit resistance and to promote uniform distribution of current to all reinforcement is required.5.15,5.16 A typical arrangement is shown in Fig. 5.7. The criteria for protection of steel embedded in concrete are not clearly defined. Most commonly, corrosion engineers use the potential compared to a standard reference cell as the sole criterion. The criterion for steel that is buried or submerged is normally accepted as -0.85 V, or more negative than a copper/CSE (copper-copper sulfate reference electrode). However, steel embedded in concrete exhibits more noble potentials than exposed steel in the order of 0.2 to 0.3 V more positive. Therefore, some investigators claim that protection is provided at lower potentials, in the order of -0.5 V with CSE reference.5.17 For steel embedded in concrete exposed to the atmosphere, research has indicated that the -0.85-V criterion may not be attainable. Quite possibly the result may be sufficient current to cause concern about lack of bond.5.18 The possibility of the loss of bond of the reinforcing steel is related to high current densities, at least 25 mA/ft2.

Fig. 5.6Impressed current cathodic protection (buried pipe)

Table 5.1Comparison of electrochemical circuit characteristics


1 2 3 4 5 6 7 Galvanic or sacricial anode No external power required Fixed, small driving voltage Limited, small current output Interference of adjacent structure not likely Overprotection not likely Anodes rapidly consumed Sensitive to temperature and moisture conditions Impressed current External power required Voltage variable over wide range Current variable over wide range Interference can result Overprotection can result Anodes slowly consumed

The other way in which the favorable electrochemical circuit can be established is by introducing electrical current from an external source. Because an outside source of current is used, this method is termed impressed current cathodic protection. This method also requires the installation of an external electrode in the electrolyte with the structure to be protected. However, since the current flow is not dependent on the favorable potential difference between the electrode and the structure to be protected, more noble materials can be selected for the anode. These materials include high-silicon cast iron, graphite, and even more noble materials such as platinized-titanium or platinized-niobium. These metals corrode or are consumed very slowly, less than 1 lb per amp year.5.11,5.12

CORROSION OF METALS

222R-27

However, it would be most unusual for a cathodic protection system, typically designed to operate at 2 mA/ft2 of steel surface, to operate in excess of 25 mA/ft2 for sufficient time (several years) to cause deterioration in bond strength unless the potential criterion was applied inappropriately. Corrosion of steel in concrete is controlled by oxygen access. Polarization of the steel is controlled by cathodic protection. As mentioned in Chapter 2, concrete is a very alkaline medium and cathodic reaction is the reduction of oxygen to hydroxides. The same is true for the cathodic protection currents. If the current reduces the oxygen faster than it can be replenished, when cathodically protecting steel in concrete, the steel will polarize to a more negative value. If the oxygen supply is great, then to obtain greater degrees of polarization, the current supply must be increased. In some bridge decks, the current required to obtain the criterion of -0.85 V would be such that there would be fear for disbondment even though the half-cell potential was not even close to that value. Thus, cathodic protection of concrete embedded steel is not necessarily a standard procedure. For concrete that is buried or submerged, probably moisture-saturated, the -0.85 V CSE criterion is easily obtained at current densities as low as 25 A/ft2. For bridge decks, where the concrete is comparatively dry and oxygen is abundant, the criteria may be -0.85 V if obtainable with reasonable, current density (probably a maximum of 3 mA/ft2 of deck surface). If not, a shift of 400 mV for all bridge deck half-cell potentials is a criterion developed from the statistical distribution of half-cell potentials that could change the least negative potential to equal or exceed the most negative half-cell potential, or to provide the current required to achieve the cathodic protection requirements as determined by the E-log I curve.5.18 In this latter case, the halfcell potential of the bridge deck can vary wildly depending on the moisture content and temperature of the concrete. It also eliminates the need of a permanent half-cell which would be required in the case of half-cell potential control. The theory behind the constant current determined by the E-log I criteria is that the potential of steel is directly related to oxygen reduction. The corrosion of the steel is directly related to the corrosion current. Thus, with a constant current, variations in oxygen reduction will cause variations in the half-cell potential of the steel. Conversely, if the cathodic protection criteria is a constant half-cell potential, then the current output of the system will vary as the oxygen supply varies. When using the half-cell potential criterion as developed through the E-log I method, there is a risk that there will be times when the cathodic system will not completely control the corrosion of the steel. For example, if the concrete is near saturation, the steel can usually be polarized with relatively small current densities. Then if the rectifier is regulated by a half-cell potential and the concrete dries so that oxygen becomes abundant, and the polarized potential drifts significantly less negative, it is likely that there will be insufficient current capacity to raise the potential to the protective potential value. Corrosion is caused by the flow of electrons or current. The differences in half-cell potentials is the voltage that causes the

current to flow. Once the steel is made cathodic in that it receives current, the current causes oxygen to be reduced. This same amount of current may reduce oxygen faster than it is being replenished and result in polarization with an associated potential change. If the oxygen is replenished at the same rate as it is reduced, no additional polarization will result. Thus, if the amount of current for cathodic protection will make all of the steel cathodic and oxygen reduction is taking place, any greater amount of cathodic protection current will simply be wasted on reducing oxygen. In addition to disbondment, overprotection can result in hydrogen embrittlement.5.6,5.19 In Chapter 2 it was shown that in acid environments hydrogen ions are reduced at the cathode to atomic hydrogen which, in turn, combine to form gaseous hydrogen. When overprotection results, hydrogen gas is formed at a faster rate than can be diffused through the coating, in this instance, concrete. When this occurs, gaseous pressure is developed at the steel-coating interface which tends to either spall the coating (disbondment) or to diffuse as atomic hydrogen into the metal. When hydrogen diffuses into the metal, it further strains the metal lattice, resulting in reduced ductility and toughness. These phenomena are referred to as hydrogen embrittlement. Normally, hydrogen embrittlement affects high-strength steels only, generally those having yield strengths of 90 ksi (620 MPa) or higher5.20,5.21 and is consequently a potential problem in applying cathodic protection to prestressed concrete elements. Because of the adverse effects possible from overprotection, polarized potentials (determined immediately after the current has been interrupted) are normally limited to -1.10 V CSE to avoid the possibility of disbonding and hydrogen embrittlement problems.5.22 In addition, protection above that level would require more current and a costlier installation without achieving additional protection from corrosion. 5.5Summary Remedies for controlling corrosion on existing reinforced concrete structures use sound corrosion engineering principles directed at insulation of the reinforced concrete from the corrosive environment, alteration of the environment, or control of electrical current flow within the environment. In dealing with the reinforced concrete composite, when corrosion is detected, the deleterious contaminants are already within the concrete matrix. Insulative measures, although they minimize the rate of corrosion or the intrusion of additional contaminants, entrap the existing quantities of these corrosive contaminants. Their effectiveness can be improved by removal of contaminants prior to sealing such as by electrochemical chloride removal. Many of the proposed remedies are in the early development stage. Such approaches as deep polymer impregnation and corrosion inhibitors have not been proven as practicable methods of corrosion control on existing reinforced concrete structures. Of the remedies discussed, only cathodic protection has proven to be capable of stopping corrosion on an existing structure. The technology is proven and has been found to be cost-effective. The design procedures for such structures as

222R-28

ACI COMMITTEE REPORT

buried pipe and water tanks are well established, but the design criteria for structures exposed to the atmosphere, such as bridge decks and parking structures, are still in the developmental stages. 5.6References
5.1. Godfrey, Kneeland A., Bridge Decks, Civil EngineeringASCE, V. 45, No. 8, Aug. 1975, pp. 60-65. 5.2. Manning, David G., and Ryell, John, Durable Bridge Decks, Report No. RR203, Ontario Ministry of Transportation and Communications, Downsview, 1976, 67 pp. 5.3. Arnold, C. J., Bridge Decks in Michigan: A Summary of Research and Performance, Conference on Federally Coordinated Program for Research and Development, Pennsylvania State University, University Park, Sept. 1976. 5.4. Grifn, D. F., Corrosion of Reinforced Concrete in Marine Environments, Materials Protection, V. 4, No. 11, Nov. 1965, pp. 8-11. 5.5. Ellis, W. J., Corrosion of Cement Mortar Coated Pipelines, American Water Works Association Meeting, Honolulu, Oct. 1974. 5.6. Philips, E., Survey of Corrosion of Prestressing Steel in Concrete Water-Retaining Structures, Technical Paper No. 9, Australian Water Resources Council, Canberra, 1975. 5.7. Frascioia, R. I., Waterproong MembranesAre Their Problems Insurmountable, FCP Project 4B Review, State College, Pennsylvania, Sept. 1976. 5.8. Slater, John E.; Lankard, David R; and Moreland, Peter J., Electrochemical Removal of Chlorides from Concrete Bridge Decks, Materials Protection, V. 15, No. 11, Nov. 1976, pp. 21-26. 5.9. Hay, R. E., The Bridge Deck ProblemAn Analysis of Potential Solutions, Public Roads, V. 39, No. 4, Mar. 1976, pp. 142-147. 5.10. Mudd, O. C., Control of Pipe-Line Corrosion, Corrosion, V. 1, No. 12, Dec. 1945, pp. 192-218, and V. 2, No. 3, Mar. 1946, pp. 25-58. 5.11. Corrosion Prevention and Control Manual, Navdocks MO-306, Department of the Navy, Bureau of Yards and Docks, Washington, D.C., June 1964. 5.12. Peabody, A. W., Control of Pipeline Corrosion, National Association of Corrosion Engineers, Houston, 1967. 5.13. Henderson, D., Coated Pipe and Cathodic Protection, Consulting Engineer, Mar. 1962. 5.14. Deskins, R. L., Cathodic Protection of a Mortar Coated Steel Water Distribution System, Materials Protection, V. 5, No. 9, Sept. 1966, pp. 35-37. 5.15. Stratfull, R. F., Experimental Cathodic Protection of a Bridge Deck, Transportation Research Record No. 500, Transportation Research Board, 1974, pp. 1-15. 5.16. Fromm, H. J., Cathodic Protection of Rebar in Concrete Bridge Decks, Materials Performance, V. 16, No. 11, Nov. 1977, pp. 21-29. 5.17. Robinson, R. C., Cathodic Protection of Steel in Concrete, Corrosion of Metals in Concrete, SP-49, American Concrete Institute, Detroit, 1975, pp. 83-93. 5.18. Stratfull, Richard F., Criteria for the Cathodic Protection of Bridge Decks, Corrosion of Reinforcement in Concrete Construction, Ellis Horwood, Chichester, 1983, pp. 287-331. 5.19. Dykmass, M. J., Corrosion of Prestressing Steel in Concrete and How This Can Be Minimized or Prevented, National Association of Corrosion Engineers Western Regional Conference, San Diego, Sept. 1976. 5.20. Uhlig, Herbert H., Corrosion and Corrosion Control, John Wiley & Sons, New York, 1963. 5.21. Fontana, M. G., and Greene, N. D., Corrosion Engineering, 2nd Edition, McGraw-Hill Book Co., New York, 1978, 448 pp. 5.22. Scott, G. N., The Corrosion Inhibitive Properties of Cement Mortar Coatings, National Association of Corrosion Engineers Annual Convention, Kansas City, Mar. 1962.

American Association of State Highway and Transportation Officials T 260 Sampling and Testing for Total Chloride Ion Content in Concrete and Concrete Raw Materials American Concrete Institute 201.2R Guide to Durable Concrete 318 Building Code Requirements for Reinforced Concrete ASTM C 114 C 876

C 1152 C 1218

Standard Method for Chemical Analysis of Hydraulic Cement Standard Test Method for Half-Cell Potentials of Reinforcing Steel in Concrete Standard Test Method for Acid-Soluble Chloride in Mortar and Concrete Standard Test Method for Water-Soluble Chloride in Mortar and Concrete

These publications may be obtained from the following organizations: American Association of State Highway and Transportation Officials 444 North Capitol St., NW Suite 225 Washington, D.C. 20001 American Concrete Institute P.O. Box 9094 Farmington Hills, MI 48333-9094 ASTM 100 Barr Harbor Drive West Conshohocken, PA 19428-2959
This report was submitted to letter ballot of the committee on an item-by-item basis. The committee consists of 17 members. All items were approved by the necessary two-thirds vote.

Appendix AACI 222.1-96 Provisional Standard Test Method for WaterSoluble Chloride Available for Corrosion of Embedded Steel in Mortar and Concrete Using the Soxhlet Extractor Reported by ACI Committee 222
Some water-soluble chlorides, primarily in certain aggregates, do not induce corrosion of embedded reinforcing steel since these chlorides are bound within the aggregate. Currently, available test methods cannot distinguish between the water-soluble chlorides that support corrosion and those that do not. This test method detects only water-soluble chlorides that contribute to the corrosion of the reinforcing steel.

CHAPTER 6REFERENCES TO DOCUMENTS OF STANDARD-PRODUCING ORGANIZATIONS The documents of the various standards-producing organizations referred to in this document are listed below with their serial designation.

CORROSION OF METALS Keywords: Water-soluble chlorides; corrosion; steel; mortar; concretes; Soxhlet Extractor.

222R-29

1Scope 1.1This test method provides procedures for the sampling and analysis of hydraulic-cement mortar, concrete, or aggregate for chloride that is water-soluble and available for the corrosion reaction under the conditions of the test. 1.2This test method does not purport to address all of the safety problems, if any, associated with its use. It is the responsibility of the user of this test method to establish appropriate safety and health practices and determine the applicability of regulatory limitations prior to use. 2Signicance and Use 2.1Water-soluble chloride, when present in sufficient amounts, may initiate or accelerate the corrosion of metals such as steel embedded in or contacting a cement system such as mortar, grout, or concrete. Other test methods exist for the determination of water-soluble chloride in a cement system.* However, some aggregates contain a considerable amount of chloride that is bound in the aggregate and is not available for the corrosion reaction. The test method described in ASTM C 1218 measures a portion of the chloride contained in these aggregates. However, the amount of chloride measured is very dependent on the degree of fineness to which the aggregates are ground during sample preparation. The problem with the ASTM C 1218 test method is therefore twofold: the test measures chlorides that are not generally available for the corrosion reaction, and the test gives widely variable results. The test method described herein should be used when chloride-bearing aggregates influence the results obtained using ASTM C 1218. 2.2Sulfides are known to interfere with the determination of chloride content. Blast-furnace slag aggregates and cement contain sulfide sulfur in concentrations high enough to cause significant interference and produce erroneous test results. Treatment with hydrogen peroxide, as discussed in ASTM C 114, shall be used to eliminate such interference. 3Apparatus 3.1Sampling Equipment: The apparatus required for obtaining samples by coring or sawing is described in ASTM C 42. Sampling by drilling is not applicable for this test and shall not be used. 3.2Sample Processing Apparatus: 3.2.1 Samples too large to fit in the sample holder of the Soxhlet shall be reduced in size by means of a jaw crusher or by hammering. 3.2.2 Extract chlorides from the sample using a Soxhlet extractor, a schematic of which is shown in Fig. A1. The Soxhlet
* ASTM Standard Test Method C 1218-92, Standard Test Method for Water-Soluble Chloride in Mortar and Concrete. For more information see The Determination of the Chloride Content of Concrete by Brian B. Hope, John A. Page, and John S. Poland, Cement and Concrete Research, V. 15, No. 5, Pergamon Press, New York, Sept. 1985, pp. 863-870. ASTM Test Method C 114, Test Methods for Chemical Analysis of Hydraulic Cement. ASTM C 42, Test Method for Obtaining and Testing Drilled Cores and Sawn Beams of Concrete.

Fig. A1Soxhlet extraction apparatus

extractor consists of a heater, a lower flask to hold water, the sample compartment, and a condenser. The extractor contains approximately 100 ml of distilled water in the lower flask. Heat is applied to this flask; vapor from the boiling water passes to the condenser; and the condensate collects in the sample compartment. The sample is contained in a porous holder and the hot condensate collects around the sample. When the condensate reaches a critical height, the liquid is siphoned back into the lower flask and the process repeats. The nonvolatile components extracted from the sample accumulate in the lower flask, while each extraction involves fresh hot distillate. The heat input shall be sufficient to give an extraction cycle about every 20 min. For convenience, suitable commercial equipment is available.** 4Reagents 4.1The reagents required for the chloride determination are given in the test method for chloride of ASTM C 114. 5 Sampling 5.1Reduce the size of a minimum 300 gm sample as specified in Section 6, and divide this sample to a minimum 30 gm representative sample for use in the chloride determination. If the sample is taken from concrete or mortar then the concrete or mortar shall be at least 7 days old before sampling. Note 1Concrete cores taken in accordance with ASTM C 42 or concrete cylinders cast from the proposed mix may
** Suitable Soxhlet extraction equipment is available from Fisher Scientic (catalog No. 09-551A) and other manufacturers.

222R-30

ACI COMMITTEE REPORT

be cut longitudinally or laterally to provide the required 300 gm sample representative of the core or cylinder. Experience has shown that the cooling water from core cutting will not dissolve a significant amount of chloride. 6 Sample Preparation 6.1Using the jaw crusher or hammer, reduce the sample so that it fits the sample holder using the minimum crushing necessary. The sample shall not be crushed to a powder since this would release chloride bound in some aggregates which, as previously discussed, are known not to contribute to corrosion. 7Procedure 7.1A single test shall consist of determination of chloride contents of three individual 30 gm samples. 7.2Weigh each sample (30 g 5 g) to the nearest 0.01 g and place in the porous sample holder of a Soxhlet extractor. Add a wad of glass wool. Place approximately 100 ml of deionized water in the lower flask. 7.2.1 Assemble the condenser complete with cooling water supply pipes to the extractor and place on the heater. Turn on both the heater and condenser cooling water and allow extraction to continue for 24 hr; adjust the heating rate to give a cycle about every 20 min. 7.2.2 At the conclusion of the extraction stage, transfer the solution to a 500 ml volumetric flask. Rinse the Soxhlet flask three times with distilled water, transferring the washings to the 500 ml volumetric flask; add distilled water to produce a volume of 500 ml. With a pipette transfer a 25 ml aliquot to a 250 ml conical flask. Add 3 drops of methyl orange indicator (prepared in accordance with ASTM C 114) and add

dilute (1+1) nitric acid until the solution is acidified. Add 3.0 0.1 ml of hydrogen peroxide (30 percent solution) to the solution. Proceed in accordance with the reference ASTM C 114, starting with the procedure specified in Section 19.5.3 and continuing to the end of Section 19.5.8. 7.2.3 Make a blank determination by using the Soxhlet, complete with thimble and glass wool, but containing no sample of cementitious material. 8Calculation 8.1Calculate percent of chloride to the nearest 0.001 percent as the average chloride content of the triplicate samples, each calculated as follows N 500 - -------- (1) Chloride, percent = 3.5453 ( V 1 V b ) ---M V2 where V1 = ml of 0.05 N AgNO3 solution used for titration of the sample (equivalence point) Vb = ml of 0.05 N AgNO3 solution used for titration of the blank (equivalence point) N = normality of 0.05 N AgNO3 solution, calculated to 0.001 M = mass of concrete or mortar sample, g V2 = volume of the 25 ml aliquot determined to 0.1 ml (larger or smaller aliquots may be used depending on the chloride concentrations present) 8.2Sufficient data are not available at this time to provide precision and bias statements.

Você também pode gostar