Anda di halaman 1dari 20

SP 170-27

A Critical Review of Deterioration of


Concrete Due to Corrosion of
Reinforcing Steel
by Kimberly E. Kurtis and Kumar Mehta
Synopsis: Throughout the world, sizable portions of national budgets must be
appropriated for the repair and rehabilitation of concrete structures which have
suffered damage due to corrosion of reinforcing steel. This paper presents a
critical review of the current state-of-the-art on the mechanisms responsible for
deterioration of concrete and reinforcing steel. The protective effect of the
passive film and the depassivation of reinforcing steel by carbonation and
chloride ingress are discussed. In addition to the passive film, some researchers
believe that a mineral scale may contribute to the protection of embedded steel
against corrosion.
The following ambiguities in the current understanding of the
deterioration of concrete due to corrosion of the reinforcing steel are discussed:
mechanisms of passivation of steel in the concrete environment; stoichiometry of
the passive film; applicability of Fick's second law for the prediction of chloride
diffusion in concrete; mechanisms of depassivation of steel by chloride ions;
threshold Cl-/OH ratio; the composition and mechanism of protection provided
by mineral scales; mechanisms of expansion of the steel corrosion product in
concrete; and the detrimental effects of aggressive ions on concrete properties.
Keywords: Carbonation; chlorides; concretes; corrosion; durability; permeability;
reinforced concrete; reinforcing steels.
535
536 Kurtis and Mehta
Kimberly E. Kurtis is a graduate student researcher in the Civil Engineering
Department at the University of California at Berkeley. She received her Master
of Science degree in Structural Engineering and Materials at Berkeley and her
Bachelor of Science in Civil Engineering at Tulane University.
P. Kumar Mehta is Professor Emeritus in the Civil Engineering Department at the
University of California at Berkeley. He is the author or co-author of numerous
papers on cementitious materials and properties of concrete including a textbook
on the subject. A Fellow of the American Concrete Institute, he has received
several awards including the ACI's Wason Medal for materials research,
CANMET/ACI award for outstanding contributions to research on performance
of concrete in the marine environment, and the Mohan Malhotra Award for
research on supplementary cementing materials. He has held the Roy Carlson
Distinguished Professorship in Civil Engineering at Berkeley and has received the
highest campus honor, the Berkeley Citation, for exceptional contribution to his
field and .to the university.
INTRODUCTION
Steel corrosion and the resulting deterioration of reinforced concrete
structures is one of the primary causes of increasing damage to infrastructure.
Design of concrete structures based on material strength rather than durability, the
growing use of deicing salts and chloride-containing admixtures, and construction
in increasingly aggressive environments are some of the factors that have lead to
the surge of concrete structures experiencing reinforcement corrosion in recent
decades. The concrete durability crisis, particularly involving the corrosion of
embedded steel, is extensive, and published literature is replete with case histories
from all parts of the globe, such as those described below.
United States
The Federal Highway Administration (FHW A) recently reported that
226,000 U.S. reinforced concrete bridges, or 39% of the total, were
deficient. In addition, 134,000 bridges, or 23% of the total, were
classified as "structurally deficient", meaning that they can support only
light traffic loads or that they require immediate repair [ 1]. In 1991, the
United States Department of Transportation estimated rehabilitation costs
for these damaged bridges at $90.9 billion [2]. Corrosion of the
reinforcing steel is implicated in the majority of these damaged bridges
[3].
Arabian Gulf
In a survey of more than 100 reinforced concrete structures in the Arabian
Gulf, Matta [4] found the typical service life to be 10-20 years,
considerably less than the 50+ year design life. This reduction in service
life resulted from the widespread corrosion of reinforcing steel that was
caused by either chloride contamination of the concrete through mix
ingredients or ingress, or by carbonation of concrete.
Concrete Durability 537
Canada
According to Khanna et al. [5], thermal cracking, frost action and
corrosion of reinforcing steel caused severe damage to more than 400
concrete piles of the 7 year-old Rodney Terminal at Saint John harbor in
Eastern Canada.
Norway
According to Gjorv [6], inspection of coastal bridges in Norway indicated
that approximately 25 percent have experienced some reinforcing steel
corrosion, mainly caused by chloride ingress. Many of these structures are
less than 25 years old.
Germany
In May 1980, the roof of the Berlin Congress Hall, constructed in 1957,
collapsed. Poor construction finishing at joints allowed moisture ingress,
and hydrogen induced stress corrosion cracking in the prestressing tendons
led to structural failure [7].
Mexico
Castro et al. [8] assert that Mexico's infrastructure is corroding largely
unchecked. This has resulted not only in economic losses but also the loss
of human life. In the Yucatan Peninsula, where aggregates are rough and
angular and the water/cement ratios are consequently high, there is much
corrosion due to the high permeability of concrete.
Pacific Rim
In Hong Kong, the Kanmon Tunnels, which had been designed for 120
year life, leaked so profusely within 4 to 10 years of service that tunnels
liners, rails, and equipment severely corroded. As a result, tunnel
operation had to be interrupted to perform major repairs and
reconstruction [9].
When internal steel reinforcement corrodes, the strength of the reinforced
concrete member is undermined in several ways. Since corrosion products have a
greater volume than the parent steel, internal tensile stresses will develop in the
cement mortar at the steel/mortar interface. As a result, the surrounding concrete
cracks and will eventually spall away as corrosion of the internal steel advances.
In addition, under tensile stress developed during corrosion, existing fine cracks
and microcracks in the surrounding concrete tend to enlarge and coalesce to form
a network of interconnected cracks, providing increased ionic transport between
the surface of the concrete and the surface of the reinforcing steel, effectively
promoting the corrosion process. Crack growth decreases concrete stiffness and
tensile strength, while the formation of a network of cracks increases concrete
permeability. Thus, the structural integrity of the reinforced concrete member is
increasingly compromised as cracking progresses. As the steel is progressively
lost to corrosion, the reinforcing bar cross-section is reduced, causing a decrease
the member's tensile strength. Furthermore, as corrosion advances, the bond
between the steel and surrounding concrete is weakened, adversely affecting the
load transfer between the two materials. To ensure that reinforced concrete
members perform according to their design capacity and design service life, it is
important to prevent or delay the occurrence of corrosion.
538 Kurtis and Mehta
The growing incidence of reinforced concrete corrosion and the resulting
repair costs have led the construction industry to acknowledge that changes are
necessary in the design and construction for new reinforced concrete structures.
Consequently, new products and techniques are being developed and made
available for the construction of new structures. Techniques currently used to
guard against corrosion include: sealants, epoxy-coated reinforcing bars,
galvanized steel, fiber-reinforced plastic reinforcement, cathodic protection, and
protective overlays and membranes for bridge decks. Many of these protective
measures increase construction costs greatly, and some have met with mixed
success. A more traditional, and typically less expensive, means of decreasing
corrosion involves reducing the permeability of concrete. This can be
accomplished by a combination of several techniques, such as providing adequate
concrete cover thickness, using mineral admixtures and water-reducing agents in
the concrete mixture, and proper curing. With sophisticated structures, a
synergistic application of the traditional approach and newer methods and
materials is required to ensure durability in aggressive environments.
Despite the plenitude of publications on the subject of corrosion and its
prevention, adequate fundamental information concerning the mechanisms of
deterioration of reinforced concrete from corrosion is lacking. The purpose of
this paper is to present an overview of the current state-of-the-art and to address
deficiencies in our knowledge with the aim of providing a research impetus for
better understanding of the underlying mechanisms. After a general review of the
corrosion process, the nature of the passive film, the effects of carbonation and
chloride ingress, and the role of mineral scale will be discussed. The
electrochemical process of corrosion in concrete will be described, and a holistic
damage model for reinforced concrete exposed to aggressive environmental
conditions will be presented.
THE CORROSION PROCESS
In general, the corrosion process occurs when metals revert back to lower
energy states. Independent of the metal involved, five specific conditions are
required for corrosion to occur in an aerobic environment. These requirements
are:
1) the presence of an anode to produce electrons,
2) the presence of a cathode to accept electrons,
3) the availability of oxygen at the cathode site,
4) the availability of water at the cathode site, and
5) an electrical connection between the anode and cathode sites to transfer
electrons.
If any of the above listed conditions is absent, corrosion will not occur.
The anodic and cathodic regions may form on the metal embedded in the
concrete. If dissimilar metals are present, one metal can act as a cathode and the
other as an anode. For example, when aluminum conduits and steel reinforcing
bars are embedded in concrete, the steel, being more noble than the aluminum,
will be cathodic to the conduit. Even when only one type of metal is embedded,
anodic and cathodic sites may form on the surface due to metal defects, uneven
formation of films, presence of different microstructural phases, different grain
orientations, or localized differences in the surrounding pore solution chemistry.
Concrete Durability 539
Thus, requirements 1 and 2 are dependent on the characteristics of the metal or
metals present.
Requirements 3-5, listed above, are governed by the properties of the
cement paste. To understand the effect of cement paste on corrosion, it is helpful
to consider it as a two-phase material, composed of hydrated solid minerals and
the pore fluid. The structure, size, distribution, and interconnection of the pores
in the cement paste, in conjunction with the presence of cracks and microcracks,
control the permeability of concrete. The more permeable the concrete, the
greater the availability of oxygen and water at the cathode for electrochemical
reaction. Gonzalez et al. have shown that higher concentrations of water and
oxygen in the vicinity of the cathode will increase the rate of corrosion [ 10]. This
is supported by the Nernst equationt, a fundamental concept used in the study of
corrosion which demonstrates that the corrosion rate increases as the
concentration of reactants, such as oxygen, increases. A more permeable
concrete also allows greater ingress of aggressive chloride ions and carbon
dioxide. Both of these substances contribute to the breakdown of an protective
iron oxide film that is typically present on the steel surface when exposed to a
highly alkaline environment. This film, termed a 'passive film', will be discussed
further in the following sections.
While the permeability of uncracked concrete is largely determined by the
pore structure of the hydrated cement paste, it is the pore fluid and the hydrated
paste that together act as the electrolyte needed for the transfer of electrons during
the corrosion process. As the resistivity of the pore solution is much less than that
of the hydrated cement paste, the solution is the more conductive electrolyte.
Therefore, the transfer of electrons during the corrosion process can be considered
to be dependent on the size, distribution, and interconnection of pores and their
degree of saturation. Typically, unsaturated concrete is a high resistance material,
and, as a result, the corrosion rate of embedded steel is limited to some
degree[l1]. Although the electrical resistivity of ordinary concrete is high
(typically 10-20Qm), electron transfer from the cathode to the anode during the
corrosion process does occur. However, according to Wolsiefer [12], when the
electrical resistivity of concrete lies above 60Qm corrosion will not occur even in
chloride contaminated concrete.
From the above discussion, it is evident that corrosion requirements 1, 2, 4
and 5 - the presence of an anode, a cathode, water at the cathode, and an
electrolyte - are readily met by most reinforced concrete members. The
embedded steel contains anodic and cathodic sites, water present in the pores us
available for the cathodic reaction, and both the surrounding pore fluid and
cement paste provide a conductive medium for charge transfer. Requirement 3,
the presence of oxygen at the cathodic site, is dependent on the permeability of
the concrete. As will be explained further, the corrosion process cannot occur
without the availability of oxygen at the cathodic site. However, it has been
established that even very dense concrete is fairly permeable to oxygen [13, 14].
t E =Eo- (0.0592/n) log ([products]/[reactants]), where E represents the potential difference
betweeen the metal and solution, E
0
is the equilibrium potential, n represents the number of
equivalents exchanged, and [] denotes concentration in mol/L for a reduction reaction at room
temperature.
540 Kurtis and Mehta
Up to this point, the five general requirements for corrosion, independent
of environment, have been examined. However, an important sixth requirement -
a requirement specific to corrosion of steel embedded in portland cement concrete
- has been neglected. This sixth requirement for corrosion of embedded steel in
concrete is the penetration or removal of a thin oxide layer which forms a
protective, passive film on the steel surface. When this film is intact, corrosion of
the steel cannot occur because requirement 1 is not satisfied. The destruction of
this film is dependent on both the permeability of the concrete and the chemistry
of the cement paste as discussed in detail in the following sections.
THE PASSIVE FILM
Any metal or alloy is said to be in a "passive" state when it resists
corrosion in an environment where corrosion is thermodynamically favorable
[15]. Passivity is not a characteristic of a metal or alloy, but instead it is
characteristic of a material with respect to a given environment. For steel
embedded in concrete, the formation of a passive film is a function of
environmental pH.
Typically, the pH of the pore solution surrounding the steel lies between
12.5 and 13.5, depending on the concentration of alkalis in the cement. Sodium
and potassium, derived from raw materials used in the production of cement
clinker, are the source of much of the high pore solution alkalinity. Calcium
hydroxide, one of the solid hydration products, acts as a buffer by maintaining
the high pH of the pore solution [16].
As the Pourbaix diagram (Figure 1) for iron in an aqueous environment
indicates, steel subjected to high pH environment remains passive over a wide
range of potentials. Hausmann [18] demonstrated that a very limited supply of
oxygen was sufficient to maintain potentials in this range. Many authors cite the
Pourbaix diagram as a proof that a protective passive ferric or ferrous oxide film
forms on reinforcing steel in concrete. Borgard et al. [ 11] caution that Pourbaix
diagrams should not be used to predict whether a certain metal-oxide species will
form a film that will provide corrosion protection. It is also important to note that
these diagrams describe conditions of thermodynamic stability of metal species
and do not consider the kinetics of the reactions implied. Even more, the
Pourbaix diagram depicted in Figure I is applicable to the behavior of iron or
steel in an aqueous solution. Yet, many researchers have wrongly used this
diagram to explain the passivity of steel in concrete, assuming that pH is the only
relevant parameter for passivation.* Because of these limitations, the Pourbaix
diagram should not be used to predict the stoichiometry of the passive film or the
amount of corrosion protection imparted by the film that forms on surface of steel
embedded in concrete.
Despite the confusion in the literature over the applicability of Pourbaix
diagrams for predicting the behavior of steel in reinforced concrete, it is generally
accepted that when steel is exposed to a highly alkaline environment a
submicroscopic oxide film forms on the steel surface. The film is thought to form
rapidly during the initial stages of cement hydration [ 19]. Adding to the
uncertainties surrounding the understanding of this film, various stoichiometries
* see "Mineral Scale" discussion for more details
Concrete Durability 541
are reported in the literature. Kerkar et al. [20] state that the film consists of
hydrated ferric oxide (FeOOH) or ferric oxide (Fe203), while others [21, 22]
report that the film consists of an inner layer of ferrous oxide (Fe304) under an
outer layer of some form of ferric oxide. Sagoe-Crentsil and Glasser [23] state
that the "chemical and mineralogical compositions (of the passive film) are yet to
be determined ... it is, however, feasible that the passive film consists of several
phases, perhaps intimately mixed or zoned in response to changing oxygen
gradients across the interface". In addition, the authors imply that calcium ions
from the cement paste may be incorporated into the passive film protecting
reinforcing steel embedded in concrete. Indeed, Devine** has suggested that the
passive film may be composed of a single or a combination of previously
unknown oxides, hydroxides, or oxyhydroxides.
Because the passive film forms when the steel is subjected to the alkaline
environment provided by concrete, any study of the film must take place within
this environment. This constraint has posed a great challenge for the study and
characterization of passive films. It has been determined, however, that the film
is thinner than 2nm which is equivalent to a few layers of oxygen. As a result, the
film is optically transparent. The strongly adherent passive film is thought to act
as a barrier, inhibiting the anodic dissolution of iron, and, thus, protecting it from
corrosion. While corrosion may not be completely absent during this period of
passivity, its rate is greatly limited by the presence of the film. Hansson [22] has
shown that the corrosion rate of the passivated steel in concrete is about 0.1
J..Lrnlyear, an amount which can be considered negligible.
The passive film has been studied for over two centuries, and various
theories have been developed to explain its protective effect on metals. Yet, little
research has been conducted on the mechanisms of steel passivation in concrete.
Borgard et al. [ 11] state that this concept of protective passive film formation on
reinforcing steel in concrete was "introduced (by Cornet et al. in 1968 [24]) with
no experimental evidence to support the passivation mechanism". That is, the
concept of passivity of embedded reinforcing steel has been embraced with little
research to explain its protective mechanisms.
Currently, the most plausible and accepted explanations for the passivation
of steel in alkaline aqueous solutions are the adsorption theory and the film
theory. The adsorption theory was first proposed by Faraday and presumes that
the metal surface has such a great affinity for oxygen, that a thin, chemisorbed
oxygen layer builds up on the metal surface over time. Because the adsorption of
this layer is in some way "equivalent to oxidation" of the metal, the oxygen layer
inhibits further chemical reaction of metallic iron with the surrounding
electrolyte. The film theory is based on the presence of a hydrated metal oxide
surface film as a result of environmental interaction. This film is assumed to be
thicker than the oxygen layer adsorbed to the metal surface in the alternate theory
presented here. By the film theory, it is postulated that impervious oxide films
offer protection against corrosion by decreasing the diffusivity of metal ions into
solution. Chawla and Gupta [15] state that these two theories may not be
contradictory, but that instead they may complement one another. When the
passive film is first forming and is thin, the adsorption theory may describe its
protective behavior. But as the film thickens, the oxide theory may be more
applicable. Again, these are theories only, and little experimental work has been
**Personal communication, T.M. Devine, University of California, Berkeley, 1996.
542 Kurtis and Mehta
done to determine the mechanisms of protection of steel in the concrete
environment.
It is known that the film formed at the metal surface is relatively unstable,
and the passivity provided is temporary. For steel embedded in concrete, the two
known means of depassivation are:
a reduction in concrete alkalinity and
attack by aggressive ions.
It is reported that when the pH of the surrounding environment is higher than
11.5, the passive film will remain intact and will provide protection to the steel
against corrosion [25]. Once the pH falls below this level, the film becomes
unstable and the protection of the steel will be lost. Carbonation of concrete is
the most common mechanism for the loss of alkalinity in concrete. In addition,
aggressive ions, such as chloride, can dissolve or penetrate the passive film at any
pH. Most often, this occurs when chlorides diffuse through the concrete and
attack the passive barrier. The two most common means of depassivation,
carbonation and chloride ion attack, are presented in the following sections.
EFFECT OF CARBONATION
Decaying organic matter often imparts significant concentrations of
dissolved carbon dioxide (C02) into mineral, groundwater, and seawater [25].
The solution of this gas in water, carbonic acid (H2C03), reacts with the cement
paste- a process termed "carbonation". Carbonation of calcium hydroxide
(Ca(OH)2) present in the hydrated cement paste is represented by the reactions
below:
Ca(OH)2 + H2C03 -+ CaC03 + 2H20
CaC03 + C02 + H20 <-+ Ca(HC03)2
(1)
(2)
From these equations, it can been seen that calcium hydroxide in the hydrated
cement paste is consumed through reaction with carbon dioxide or carbonic acid.
This results in a gradual loss of buffering capacity which is essential for
maintaining alkalinity in concrete. As stated previously, the passive film becomes
unstable and no longer provides protection to iron at a pH below 11.5 which
implies that carbonation can result in depassivation of reinforcing steel.
The kinetics of the carbonation reaction in concrete have been widely
studied, and a parabolic model is generally accepted to describe the depth of
carbonation expected over a period of time. According to this model, the depth of
carbonation can be predicted as follows:
X= k t J/2 (3)
In this equation, xis the thickness of the carbonation layer, tis exposure time, and
k is a constant dependent on the permeability of the concrete cover. In good
quality concrete, the rate of carbonation is typically on the order of lmrnlyear,
indicating that carbonation should not present a problem during the design service
life of a reinforced concrete structure which has 50mm or more concrete cover
over the reinforcement [14].
Concrete Durability 543
Carbonation and the resulting loss in alkalinity typically begins at the
concrete surface and progresses in the form of a "carbonation front" toward the
center of the concrete section. Whereas the carbonated concrete will have pH of
less than 8, the uncarbonated concrete maintains its high alkalinity. Thus, the
uncarbonated concrete continues to provide a protective cover for the embedded
steel. In permeable concrete, concrete with low cover depth, or in concrete
containing microcracks connected to surface cracks, the carbonation front may
reach the reinforcing steel and thus leads to the onset of corrosion [26].
Therefore, adequate cover depth and a concrete mixture with low permeability
and low potential for microcracking is crucial for the protection of embedded steel
from corrosion.
Typically, corrosion induced by carbonation is a uniform attack. That is,
the presence of very localized corrosion -or pitting - is notably absent. It is
important to remember that depassivation of the steel by carbonation is not
sufficient to cause corrosion. Water and oxygen must also be present at the
cathode site. Because concrete structures are often subjected to cycles of drying
and rewetting, an interesting interaction of carbonation and corrosion is found.
Venuat [27] has shown that carbon dioxide diffusion is most rapid in concrete at
intermediate humidity contents (50-80% relative humidity). When the concrete
is completely dry, the carbonation reaction is very slow, and when the concrete is
completely saturated, the diffusion of carbon dioxide is very slow. When the
concrete is semi-dry, carbonation occurs. Upon wetting, the cathodic reaction of
the corrosion process is enhanced. Thus, alternating cycles of drying and wetting
provide the most aggressive environment for the carbonation-induced corrosion
[16].
EFFECT OF CHLORIDE IONS
As compared to corrosion resulting from concrete carbonation, corrosion
of reinforcing steel caused by chloride ion ingress is much more prevalent.
Chloride ions can be introduced into the concrete during its manufacture. This
can occur when seawater or water with a high chloride concentration is used as
mixing water, when chloride-contaminated coarse or fine aggregates are used, or
when chloride-containing admixtures such as calcium chloride are used. Chloride
ions present in the surrounding environment may penetrate the concrete. Deicing
salts, seawater, and chloride-contaminated soils are the primary sources of
external chloride. When concrete is dry, chlorides can penetrate several
millimeters in a few hours by the capillary draw of salt water into the concrete
[16]. When concrete is partly or fully saturated, chloride ions penetrate by
diffusion through the pore solution. Typical diffusion rate for a fully saturated
cement paste are on the order of w-8 cm2/s [28].
Different models exist to describe the ingress of fluids and ions, including
chlorides, in concrete. For steady-state flow, Darcy's law relates the rate of flow
dqldt to the hydraulic pressure gradient dhldl by:
dqldt = K F (dhldl) (4)
544 Kurtis and Mehta
with the permeability coefficient K and the cross section exposed to flow F: For
the diffusion of gases, Fick's laws are typically applied. For the steady state,
Pick's first law
m = -D (fx:lox) (5)
applies, and for the unsteady state, Pick's second law
(6)
applies, where c is concentration and D is the diffusion coefficient. Recent
research by Chatterji [29] has shown that Fick's second law, in particular, is not
necessarily applicable to concrete as it assumes that the material is non-ionic and
completely saturated. Based upon field experience with concrete structures
exposed to chloride environments such as seawater, Sandberg [30] asserted that
difficulties in the application of Fick's laws may arise from a pore-blocking effect
that results when hydrated cement paste interacts with seawater. Much research
has been conducted in recent years to predict the time required for migration of
chlorides through concrete. In most cases, some form of Pick's second law has
been used to calculate the penetration depth over long periods of time, and, as
discussed above, the predicted behavior may be unreliable for use in service life
design of reinforced concrete structures.
While it has been generally agreed that chloride ions act as catalysts for
the loss of the protection offered by the passive film, the exact mechanisms of this
process are not well understood. Two theories appear most often in the literature:
the adsorption theory and the oxide film theory (note that the same nomenclature
is used to the describe the mechanisms for passivation). According to the
adsorption theory, the chloride ions replace oxygen atoms held within the passive
film [16]. This causes differences in electrochemical potential across the film,
and the film becomes unstable. The high reaction rate of steel and chloride, in
areas where chloride ions have replaced oxygen, is thought to explain the
occurrence of pitting on the steel surface which is typical of chloride-induced
corrosion [23]. According to the oxide film theory, the passive film contains
inherent defects and pores. By a selective dissolution of more reactive
components of the passive film, chloride ions penetrate the film at these sites
more readily than other anions present in the pore solution. The theory suggests
that these localized attacks by chloride ions are responsible for pitting corrosion.
Both theories suggest that attack of the passive film by chloride ions is a
localized phenomenon. This form of attack causes microgalvanic cells to form on
the reinforcing steel as shown in Figure 2. In regions where the depassivation has
occurred, iron will be lost by oxidation. The areas that remain protected by the
passive film will become cathodic and, thus, sites of oxygen reduction. These
processes are described more fully in a subsequent section detailing the chemical
reactions involved in the corrosion process.
Because of the damaging effects of chloride in reinforced concrete, much
research has been performed to determine a threshold chloride ion content below
which corrosion will not occur in concrete. The ions exist in concrete in two
forms -bound and free. Only the free chloride ions, those dissolved in the pore
fluid, participate in the corrosion process, and as a result, it is the free chloride
concentration, not the total chloride content, that is critical when determining
Concrete Durability 545
threshold levels at the steel concrete interface. It is well-known that the chloride-
binding capacity of a concrete depends on the cement composition.
Since chloride ions react with calcium aluminate (C3A) present in the
cement paste to form Friedel's salt (C3ACaCI215H20), concrete made from
cement with a high C3A content will have a greater potential for binding chloride
ions. Hussain et al. [31] demonstrated the beneficial effects of such cement. By
raising the C3A content of cement from 2.43% to 14%, with all other factors
remaining constant, the chloride threshold was increased by a factor of 2.85.
Kayyali and Haque [32] found that the use of superplasticizing admixtures
increased the levels of free chlorides in concrete, but that the addition of fly ash
resulted in an increase in chloride binding capacity. Hussain et al. reported
moderate increases in the chloride threshold value with higher levels of concrete
alkalinity. Also, it was found that the presence of sulfates either moderately
increased or decreased the threshold value depending on the cement composition.
Hausmann [18] was the first to suggest a relationship between chloride
content, concrete alkalinity, and the onset of corrosion. Subsequently, much
research has been conducted to establish a CI-JOH threshold value for corrosion
initiation. As shown in Table 1, a range of values are reported in the literature.
For some time, the CI-JOH ratio of 0.6 reported by Hausmann was accepted by
many, but it should be noted that the study that produced this value was
conducted in an alkaline solution meant to simulate concrete pore fluid, not within.
concrete itself. Recent research suggests that the source of chloride
contamination has a significant influence on the threshold values. Diamond [33]
reported a threshold of 0.3 when the chloride is introduced into the concrete as an
admixture or in the mixing water. Lambert et al. [34] reported a much higher
threshold ratio, 3.0, when the chloride is introduced from an external source. It is
important to recognize that the threshold CI-/OH ratio is not easily defined and
that no value has been accepted as a unique limit for all concrete mixes.
Threshold values are highly dependent on such parameters as the water/cement
ratio, cement content, cement surface area, cement composition, ambient
temperature, and alkalinity. As a result, a well-defined threshold relationship
universally applicable to all concrete mixtures remains elusive. Consequently,
service life predictions based on certain assumed threshold values are not reliable.
ROLE OF MINERAL SCALES
While it is generally accepted that the presence of a passive film on the
steel surface provides some corrosion protection, some researchers, believe that
an additional phenomenon may offer further protection. Many authors consider
the pore fluid to be the environment of exposure in concrete. As a result of this
assumption, much reinforced concrete corrosion research has been conducted in
aqueous solutions that are meant to simulate the concrete pore fluid.
Consequently, the corrosion behavior of steel in concrete has been extrapolated
from its behavior in aqueous solutions of similar pH. This approach neglects the
effect of the interfacial steel-cement paste zone which consists of hydrated cement
paste minerals.
546 Kurtis and Mehta
Borgard et al. [11] postulate that mineral scales, consisting of precipitates
of cementitious compounds, form on metal surfaces in concrete and act in
conjunction with the metal oxide passive film to protect reinforcing steel. The
authors suggest that this phenomenon has been ignored in much of the literature
as a consequence of the use of pore solution models, which do not account for the
presence of the hydrated cement paste minerals, in corrosion studies. Page [35]
has also proposed that passivation may result from a lime-rich layer of hydration
products at the interface between steel and concrete.
Borgard et al. state that calcium carbonate scales are used for protection
of oil well production tubing even in the presence of chloride concentrations
much greater than in seawater. They assert that the formation of high calcium
scales on steel embedded in concrete is likely since portland cement has a high
calcium content. The authors reason that corrosion in concrete may be
prevented by a similar mechanism and that corrosion will occur only after the
calcium scales have been removed.
It should be recognized that chemical attack on the calcium hydroxide
scale by carbon dioxide, chloride, or sulfate ions can result in eventual removal of
the scale. However, the process of removal can provide additional protection to
the passive film on the steel surface because of the dissociation of the calcium
hydroxide by reaction with such species as chloride ions and carbon dioxide.
Such reactions produce free hydroxyl ions which act to buffer the alkalinity of the
pore solution. If this reaction takes place in the vicinity of the steel, passivity can
be maintained by ensuring that the pH remains above 11.5. However, once the
calcium hydroxide scale has been depleted, further ingress of chloride ions or
carbon dioxide damage the passive film.
CORROSION CHEMISTRY
As stated previously, corrosion is an electrochemical process involving the
transfer of ions. In reinforced concrete, electrons are produced at the anodic site
by the oxidation of the steel. That is, metal ions from the parent metal go into
solution at the anode, and the loss of metal indicates that this is the site of
corrosion. The rate of steel corrosion in concrete is dependent on the rate of the
anodic reaction:
Fe ---+ Fe++ + 2e- (7)
Electrons released at the anodic site are consumed at the cathodic site.
Typically, at the cathode, either oxygen or hydrogen is reduced. The
thermodynamic driving force for the reduction of oxygen is much greater than
that for hydrogen. As a result, the oxygen reaction is more likely to occur, and
the reduction of hydrogen in concrete corrosion is rarely a factor. The cathodic
reactions are:
(8)
(9)
Concrete Durability 547
Ferrous ions (Fe++) produced in the anodic reaction (eq. 7) combine with
the hydroxyl ions produced in the cathodic reaction (eq. 8) to form the corrosion
product ferrous hydroxide (Fe(OH)2). The corrosion rate can be increased by the
presence of other ions in the concrete. According to Hime and Erlin [36], the
reaction of ferrous ions and hydroxyl ions in the presence of chloride ions can
also result in the formation of the corrosion product Fe(OH)2 (eqs. 10-12). The
reaction represented in equation 16 demonstrates that chloride ions are produced
by this process, increasing the corrosion rate of the steel reinforcing.
Fe+++ 6Cl- +-> FeC16-4 (10)
Fe+3 + 6Cl- +-> FeCl6-3 (11)
FeC16-3 + 20H- +-> Fe(OH)2 + 6Cl- (12)
The process of corrosion for steel reinforcement bars embedded in
concrete is illustrated in Figure 2. The diagram shows that the electrons are
released at the anodic site and travel through the steel to the cathodic site, where
they are consumed. The reduction of oxygen at the cathode produces hydroxyl
ions (OH-). The production of hydrogen gas, caused by the reduction of
hydrogen, is less likely, but possible. Equations 7 and 8 demonstrate that the
corrosion process may be limited by the rate of diffusion of the hydroxyl ions, as
well as by the availability of oxygen (02), water (H20), or other aggressive ions
(Cl- in Fig. 2 and eq. 10-12).
Glasser and Sagoe-Crensil [37] have examined a 27 year old reinforced
concrete post to determine the morphology and characteristics of the corrosion
product (i.e. rust) present. In the specimen studied, the corrosion product was
found to be composed of two layers: an inner, highly crystalline and dense region
of Fe304 and y-Fe203 and an outer more porous region of primarily o:-FeOOH
(goethite). The inner layer was found to have preferential orientation related to
that of the substrate metal microstructure that resulted in epitaxial growth of the
inner layer corrosion product. Wang and Monteiro [39] confirmed that the
corrosion product tends to form a continuous band poorly bound to the reinforcing
steel and that this band is composed of multiple layers. It has been suggested that
this structure results from variations in oxygen availability and pore solution
conductivity across the region of corrosion.
MECHANISMS OF EXPANSION AND CRACKING OF CONCRETE
The solid-state transformation of iron to the ferrous hydroxide corrosion
product resulting in a volumetric expansion is generally assumed to be the
mechanism by which concrete expands and cracks by the corrosion of embedded
steel. However, the corrosion chemistry shows that the corrosion of iron to
ferrous hydroxide occurs as a result of a through-solution process. That is, iron
present in steel must oxidize to soluble ferrous ions. The ferrous ions combine, in
solution, with hydroxyl ions generated at the cathode to produce ferrous
hydroxide. Consequently, Mehta [39] and Figg [40] postulate that poorly
548 Kurtis and Mehta
crystalline or gel-like ferrous hydroxide tends to swell by water adsorption and
that the hydraulic pressure generated is responsible for expansion and cracking.
Recently, Wang and Monteiro [38] studied the mechanisms by which
corrosion of embedded steel reinforcement undermines the strength of concrete
structures. Once the depassivation of the steel occurs and corrosion begins, they
found that oxidation products form as a band around the steel. Loss of reinforcing
steel by corrosion leads directly to loss in the strength of the reinforced concrete
member. While the presence of water is necessary for the corrosion of reinforcing
steel, Wang and Monteiro found that cycles of relative wet and dry conditions at
the concrete and steel interface lead to weakening of the corroding reinforced
concrete member. When wet, adsorbed water will force corrosion product
particles further apart, resulting in expansion and cracking of the surrounding
concrete. Upon drying, the corrosion product will tend to shrink and crack. In
this way, cycles of wetting and drying will cause the steel and concrete to debond
at the interface. In addition, Wang and Monteiro found that corrosion products
diffuse away from the interface into the microcracks and pores of the surrounding
concrete. The researchers reason that 'plugging' of existing voids in the concrete
by diffused corrosion product will decrease the probability of crack arrest,
effectively embrittling the concrete and weakening the reinforced member. Loss
of steel, debonding at the steel/concrete interface, and cracking and embrittlement
of surrounding concrete is detrimental to the strength and integrity of reinforced
concrete.
Mehta [41] has proposed a holistic model, Figure 3, which illustrates all
the physico-chemical changes occurring in the reinforcing steel and concrete in
response to all environmental effects including weathering and ingress of
aggressive chloride ions and carbon dioxide. Most descriptions of the effects of
corrosion of steel embedded in concrete center on the volume expansion
associated with the corrosion products and the resulting cracking and debonding
of the concrete. As discussed previously, corrosion of reinforcing steel damages
the surrounding concrete as well as the steel. The mechanisms depicted in Figure
3 are holistic. That is, no part of the system is overlooked. The mechanisms that
lead to the corrosion of steel also affect the surrounding concrete, and these
relationships are considered in this model.
For instance, since C-S-H is the primary source of strength in the cement
paste and since the stability of C-S-H is dependent on the concentration of
hydroxyl ions in the pore fluid, a decrease in alkalinity has a detrimental effect on
the strength and elastic modulus [41]. Free hydroxyl ions in the pore fluid can
combine with other species, such as carbonate, sulfate, and chloride. This type of
reaction, which results in a decrease of the pore fluid alkalinity, causes loss of
strength and elastic modulus in the concrete. The mechanism of expansion and
cracking of concrete shown here takes into consideration the combined effect of
weakened microstructure of the cement paste and the development of hydraulic
pressure in the pores of the water-saturated system. Thus, Figure 3 demonstrates
an integration of all environmental effects on both components of the reinforced
concrete system.
Concrete Durability 549
CONCLUSIONS
Upon critical review of the state-of-the-art understanding of expansion and
cracking of reinforced concrete due to corrosion of the embedded steel, several
areas of uncertainty or ambiguity emerge. These are identified as:
Mechanisms of steel passivation in the concrete environment.
The stoichiometry of the passive film on steel and the mechanisms of its
formation.
The relative significance of the passive film and mineral scales in
providing corrosion protection to steel.
The composition of the mineral scales formed in the vicinity of the steel
and the mechanism of corrosion protection.
Applicability of Fick's second law for the prediction of chloride
diffusion in concrete.
Mechanisms of depassivation of reinforcing steel by chloride ions and
carbonation.
The threshold Cl-JOH ratio for corrosion to occur.
Composition, morphology, and mechanisms of expansion of the steel
corrosion products.
Detrimental effects of aggressive species on the material properties of
concrete itself.
The study of corrosion is limited to some extent by the inherent difficulties
associated with observing the corrosion of steel within concrete. However, some
of the gaps in our understanding of the steel corrosion process in concrete seem to
stem from the reductionist or fragmentary approach taken by researchers on this
subject. Corrosion engineers and material scientists tend to study the corrosion of
steel in alkaline aqueous solutions and, as a result, neglect the possible effects of
the cement paste microstructure on the corrosion process and changes in the
microstructure during corrosion. For instance, the effect of chloride attack and
carbonation on the surface chemistry of steel is considered, whereas their
detrimental effect on the strength and elastic modulus of the concrete is often
completely overlooked in our corrosion damage models. Widespread
misapplication of Pourbaix diagrams for prediction of passivity of steel embedded
in concrete is an example of the narrow approach applied to the study of corrosion
of steel embedded in concrete. Only when researchers take a holistic approach for
the study of combined response of both the reinforcing steel and the surrounding
concrete to all environmental effects will these ambiguities be resolved. To
address the deficiencies and advance the state-of-the-art in concrete science, we
obviously need a paradigm shift in our approach to corrosion research.
REFERENCES
[1] Federal Highway Administration, 1991 Report to Congress, June, 1991.
[2] Federal Highway Administration. "1991 Status of the Nation's Highways and
Bridges: Conditions, Performance, and Capital Investment Requirements."
July 2, 1991.
550 Kurtis and Mehta
[3] ELTECH Research Corp. , "Cathodic Protection of Reinforced Concrete
Bridge Elements: A State-of-the-Art Report.", SHRP-S-337, 1993.
[4] Matta, Z.G., "More Deterioration of Reinforced Concrete in the Arabian
Gulf", Concrete International, Nov. 1993, p. 50-51.
[5] Khanna, J.; Seabrook, P.; Gerwick, B.C. ; and Bickley, J. "Investigation of
distress in precast concrete piles at Rodney Tenninal", Performance of
Concrete in Marine Environment ed. V. Malhotra, ACI SP-109, 1988, p.
277-320.
[6] Gjorv, O.E. "Steel Corrosion in Concrete Structures Exposed to Norwegian
Environment", Concrete International, April, 1994, p. 35-39.
[7] Isecke, B. "Collapse of the Berlin Congress Hall Prestressed Concrete Roof",
Materials Performance, Dec. 1982, p. 36-39.[13] Gjorv, O.E.;
Vennesland, 0. and El Busaidy, A.H.S., Proc. Corrosion, 76, NACE,
Houston, 1976.
[8] Castro, P.; Maldonado; and deCoss, R. "Study of Chloride Diffusion as a
Corrosive Agent in Reinforced Concrete for a Tropical Marine
Environment", Corrosion Science, vol. 35, no.5-8, 1993, p. 1557-1562.
[9] Gerwick, B.C. "Pressing Needs and Future Opportunities in Durability of
Concrete in the Marine Environment", International Experience with
Durability of Concrete in the Marine Environment, Ed. P. K. Mehta, Jan.
16-17, 1989, p. 1-5.
[10] Gonzalez, J.A.; Lopez, W. ;and Rodriguez, P. "Effects of Moisture
Availability on Corrosion Kinetics of Steel Embedded in Concrete",
Corrosion, December 1993, p. 1004-1010.
[11] Borgard, B.; Warren, C.; Somayaji, S.; and Heidersbach, R. "Mechanisms of
Corrosion of Steel in Concrete", Corrosion Rates of Steel in Concrete,
STP1065, ASTM, 1990.
[12] Wolsiefer, J.T. "Silica Fume: A Solution to Steel Reinforcement Corrosion
in Concrete", ACI SP-126, 1991.
[13] Gjorv, O.E.; Vennesland, 0.; and El Busaidy, A.H.S., Proc. Corrosion, 76,
NACE Houston, 1976.
[14] Page, C.L. and Treadaway, K.W.J., "Aspects of the Electrochemisty of Steel
in Concrete", Nature, 297, May 1982, p. 109-115.
[15] Chawla, S.L. and Gupta, R.K. Materials Selection for Corrosion Control.
ASM International, 1993.
[16] Rosenberg, A.; Hansson, C.M.; and Andrade, C. "Mechanisms of Corrosion
of Steel in Concrete", Materials Science of Corrosion I, The American
Ceramic Society, 1989.
[17] Pourbaix, M. Atlas of Electrochemical Equilibrium in Aqueous Solutions.
Pergamom, London, 1976.
Concrete Durability 551
[18] Hausmann, D.A. "Steel Corrosion in Concrete: How Does it Occur?",
Materials Protection, November 1967, p.19-23.
[19] ACI Committee 222. "Corrosion of Metals in Concrete", ACI Journal,
Jan./Feb. 1985, p. 3-32.
[20] Kerkar, J.; Robinson, A.J.; and Forty, A.J. Faraday Discussion, Chemical
Society, 89, 1990, p.31.
[21] Sato, N., Passivity of Metals, Electrochemical Society, Princeton, NJ, 1978,
p. 29.
[22] Hansson, C.M., "Comments on the Electrochemical Measurements of the
Rate of Corrosion of Steel in Concrete", Cement and Concrete Research,
Vol. 14, 1984, p. 574-584, Pergamon Press Ltd.
[23] Sagoe-Crentsil, K.K. and Glasser, F.P. "Steel in Concrete: Part I. A Review
of the Electrochemical and Thermodynamic Aspects", Magazine of
Concrete Research, Dec. 1989, p. 205-212.
[24] Cornet, I.; Ishikawa, T.; and Bresler, B. Materials Protection, Vol. 7, No.3,
March 1968, p.44-47.
[25] Mehta, P.K. and Monteiro, P.J.M., Concrete: Structure. Properties. and
Materials. Second Edition, Prentice Hall, 1993.
[26] Parrott, L.J. "A Study of Carbonation Induced Corrosion", Magazine of
Concrete Research, vol. 46, no. 166, 1994, p. 23-29.
[27] Venuat, M. "Relationship between Concrete Carbonation and the Corrosion
of Reinforcement", Recentres CEFRACOR-77, JTBTP, Oct. 1977.
[28] Ngala, V.T.; Page, C.L.; Parrott, L.J.; and Yu, S.W. "Diffusion in
Cementious Materials: Further Investigations of Chloride and Oxygen
Diffusion in Well-Cured OPC and OPC/30%PFA Pastes", Cement and
Concrete Research, Vo1.24, No.4, 1995, p.819-826.
[29] Chatterji, S. "On the Applicability of Pick's Second Law to Chloride Ion
Migration through Portland Cement Concrete", Cement and Concrete
Research, Vol. 25, No.2, 1995, p. 299-303.
[30] Sandberg, P. "Critical Evaluation of Factors Affecting Chloride Initiated
Reinforcement Corrosion in Concrete", Report TVBM-3068, University of
Lund, Sweden, 1995.
[31] Hussain, S.E.; Rasheeduzzafar; Al-M.usallam, A.; and Al-Gahtani, A.S.
"Factors Affecting Threshold Chloride for Reinforcement Corrosion in
Concrete", Cement and Concrete Research, Vol. 25, No.7, 1995, p. 1543-
1555.
552 Kurtis and Mehta
[32] Kayyali, O.A. and Haque, M.N. "The Cl-/OH Ratio in Chloride-
Contaminated Concrete- a Most Important Criterion", Magazine of
Cement and Concrete Research, 47, No. 172, Sept. 1995, p. 235-242.
[33] Diamond, S. "Chloride Concentrations in Concrete Pore Solutions Resulting
from Calcium and Sodium Chloride Admixtures", Cement, Concrete, and
Aggregates, Vol. 8, No.2, Winter 1986, p. 97-102.
[34] Lambert; Page, C.L.; and Vassie, P.R.W. "Investigations of Reinforcement
Corrosion 2: Electrochemical Monitoring of Steel in Chloride-
Contaminated Concrete", Structures and Materials, 24, 1991, p. 351-358.
[35] Page, C.L. "Mechanisms of Corrosion Protection in Reinforced Concrete
Marine Structures", Nature, Dec. 1975, p. 514-515.
[36] Hime, W. and Erlin, B., "Some Chemical and Physical Aspects of
Phenomena Associated with Chloride-Induced Corrosion". Corrosion.
Concrete. and Chlorides, ACI SP-102, p. 1-12, Ed. F.W. Gibson, 1987.
[37] Glasser, F.P. and Sagoe-Crentsil, K.K. "Steel in Concrete: Part II. Electron
Microscopy Analysis", Magazine of Concrete Research, Dec. 1989, p.213-
220.
[38] Wang, K. and Monteiro, P.J.M. "Corrosion Products of Reinforcing Steel
and Their Effects on the Concrete Deterioration", currently in press.
[39] Mehta, P.K. "A Disscussion of the Paper on 'Osmotic Pressure and Swelling
of Gels' by L.S. Dent Glasser", Cement and Concrete Research, vol.lO,
n.l, Jan. 1980, p. 123.
[40] Figg, J. "Salt, Sulfate, and Other Chemical Effects", Intemational
Experience with Durability of Concrete in the Marine Environment, Ed. P.
K. Mehta, Jan. 16-17, 1989, p.49-69.
[41] Mehta, P.K. "Concrete Technology at the Crossroads- Problems and
Opportunities", Concrete Technology: Past, Present, and Future. ACI SP-
144, 1994, p. 1-30.
TABLE 1-RANGE OF REPORTED C 1'/0H THRESHOLD VALUES
Researcher Experimental Method CI-/OH Ratio
Hausmann [18] Steel bars in chloride-contaminated alkaline
solution 0.6
Diamond [33] Steel bars embedded in concrete subjected to
internal chloride attack. (made from Type I slightly greater
cement (9.1% C3A);w/c=0.50; with 0.5% by than 0.3
weight CaC!z admixture)
Lambert, Page Steel bars embedded in concrete subjected to
and Vassie [34] external chloride attack; no details on mix approximately
proportions reported 3
Concrete Durability 553
i.li'
::r:
Cl)
"' ---
;>
2: Fe20J
(;l
0
-G-- Fe++
0
'+=I
-==
--
2 corrosic;'r1------ _
0
0....
immunity
Fe
2 2
0 4 8 12 16
pH
Fig. 1-Pourbaix diagram (potential vs. pH) for iron in an aqueous solution [171
Fig. 2-Schematic illustration of the corrosion process of reinforcing steel in
concrete (adapted from [22])
554 Kurtis and Mehta
Reinforced Concrete
I
Weathering and
Loading Effects
l
Increase in the Permeability of Cover I

Penetration of Water,
J
Oz, COz, and Cl-

Loss of OH- Ions from I
Cement Paste
1
A: (i.) Depassivation of the Steel Reinforcement
(ii.) Formation of a Colloidal Rust
B. Gradual Loss of Adhesion by C-S-H
A: Expansion of Rust Increases the
Hydraulic Pressure in Pores
B: Reduction in Concrete Strength and Stiffness
-
Expansion and Cracking of Concrete
I
Fig. 3-A model of reinforced concrete damage from exposure to aggressive
environments [42]

Anda mungkin juga menyukai