You are on page 1of 33

PCA R&D Serial No.

3013

CORROSION OF REINFORCING BARS


IN CONCRETE
by C. M. Hansson; A. Poursaee, and S. J. Jaffer

Portland Cement Association 2007


All rights reserved

KEYWORDS
cathodic protection, chlorides, corrosion, deicers, ECE, electrochemical chloride
extraction, galvanic corrosion, reinforcing materials, reinforcing steels

ABSTRACT
Concrete provides an ideal environment for steel providing both physical and chemical
protection from atmospheric attack. Nevertheless, corrosion of embedded steel does
occur and is a major factor adversely influencing the durability of reinforced concrete
structures. The primary focus of this report is chloride-induced corrosion of steel
reinforcement: the factors affecting it and its influence on durability. This review
describes the corrosion process, the lack of a single chloride threshold concentration for
initiation of corrosion and the relative contributions of micro-cell and macro-cell
corrosion in sound concretes. Carbonation-induced corrosion is also briefly explained.
Concrete parameters: water/cementitious materials ratio; the use of supplementary
cementing material; adequate curing and adequate cover depth obviously play major roles
in extending the time to corrosion initiation and the rate of corrosion once initiated. The
type of reinforcement is the other half of the equation: the effectiveness of coatings (for
example: epoxy and zinc) and alloying (for example: MMFX and stainless steels) on
corrosion are discussed.
The advantages and disadvantages of various corrosion monitoring techniques for
use in reinforced concretes are examined. The report also discusses corrosion prevention
and protection methods such as inhibitors, non-chloride de-icing agents, electrochemical
chloride extraction and cathodic protection.

REFERENCE
Hansson, C. M.; Poursaee, A., and Jaffer, S. J., Corrosion of Reinforcing Bars in
Concrete, R&D Serial No. 3013, Portland Cement Association, Skokie, Illinois, USA,
2007, 33 pages.

CORROSION OF REINFORCING
BARS IN CONCRETE
by C. M. Hansson*, A. Poursaee and S. J. Jaffer
INTRODUCTION
The causes and mechanisms of corrosion of reinforcing bars in concrete are described in
terms of the practical issues as well as the electrochemistry. The parameters affecting
corrosion are (i) the ingress of aggressive species, such as chlorides, which break down
the protective film on the reinforcing bar and (ii) the amount of those species necessary to
do so. The former is largely controlled by the concrete properties while the latter is a
function of the type and condition of the reinforcing bar. Thus, the influence of the
cementitious components of concrete, the water/cementitious materials ratio and the
presence of cracks in the concrete cover, on the ingress of aggressive species is
considered and the various currently available reinforcing bar materials and their merits
are reviewed.

BACKGROUND
Although corrosion of reinforcing steel is now recognized as the major cause of
degradation of concrete structures in many parts of the world, it should be understood
that this is not due to any intrinsic property or limitation of the concrete itself whether it
be normal quality portland cement concrete (OPCC) or high performance concrete
(HPC). In fact, all sound portland cement concretes provide an ideal environment for
corrosion protection of even the poorest quality steel such as has sometimes been used for
reinforcement.
Steel is thermodynamically unstable in the earths atmosphere and will always
tend to revert to a lower energy state such as an oxide or hydroxide by reaction with
oxygen and water. The question of interest in the use of steel is not whether this process
will occur (it will!) but how fast it will occur in practice. Fortunately, only the surface
atoms of the steel are exposed to the atmosphere and, therefore, are available to react. In
the case of a 15 mm diameter bar, this amounts to only about 1 in every 40 million atoms.
Any coating on the steel will reduce this number even further. For steel embedded in
concrete, the concrete itself provides a coating limiting the access of water and oxygen to
the steel surface. A second beneficial aspect of concrete is that the solution in the pores
of the cement paste has a very high alkalinity and, as indicated in the Pourbaix diagram in
Figure 1, at the pH levels typical of concrete, the corrosion products which do form are
insoluble. They produce a very thin (~ few nm) protective coating on the steel (a passive
film) which limits the metal loss from the steel surface due to corrosion to about 0.1 - 1.0
*

Professor of Materials Engineering, University of Waterloo, Waterloo, Ontario, 519-888-4538,


chanson@uwaterloo.ca.

Ph.D. Candidates, Department of Materials Engineering, University of Waterloo, Waterloo, Ontario,


Canada

m/year [1-3]. It is generally considered that, at these passive corrosion rates, the steel
embedded in concrete would not be noticeably degraded within a 75 year lifetime and the
volume of corrosion products would not be sufficient to cause any damaging stresses
within the concrete. The passive film does not form immediately but starts as soon as the
pH of the mixing water rises in the concrete when the cement begins to hydrate and
stabilizes over the first week to protect the steel from active corrosion [4].

Figure 1. The Pourbaix diagram for iron showing regions of electrochemical potential and
pH where metallic iron is stable (grey region); where active corrosion occurs (white areas)
and where the metal is passivated (green and orange areas) [5].

Unfortunately, however, the passive film is not stable in solutions containing


chloride ions or at pH levels below about 9, as indicated in Figure. 1 [6].
Also
unfortunately, concrete is both permeable, allowing the ingress of chlorides from de-icing
salts or marine atmospheres, and reactive, allowing acidic gases, particularly CO2, to
neutralize the pore solution. When the passive film is broken down either by chlorides or
by carbonation of the concrete, active corrosion occurs at rates as high as several
mm/year and it is this process which is responsible for much of the structural degradation
occurring in reinforced concrete, such as that illustrated in Figure 2.

Figure 2. (a) Retaining wall of bridge and (b) underside of bridge deck.

Corrosion, whether at the negligible passive rate or the damaging active rate, is an
electrochemical process, involving the establishment of anodic and cathodic half-cell
reactions on the microscopic and/or macroscopic levels. In high pH solutions and in the
absence of chloride ions, the anodic dissolution reaction of iron:
Fe Fe2+ + 2e-

Eq. 1

is balanced by the cathodic reaction:


O2 + H2O + 2e- 2OH-

Eq . 2

and the Fe2+ ions combine with the OH- ions to produce the stable passive film. The
electrochemical process is illustrated schematically in Figure 3(a).

Chloride-Induced Corrosion
The mechanism by which chloride ions break down the passive film is not fully
understood [1], largely because the film is too thin to be examined and because the events
occur inside the concrete. One hypothesis is that the chloride ions become incorporated
into the passive film and reduce its resistance. This incorporation is not uniform and,
where it occurs, it allows a more rapid reaction and the establishment of an anodic area
where corrosion continues while the remaining steel remains passive, Figure 3(b).
A second hypothesis is that the Cl- ions "compete" with the OH- anions for
combining with Fe2+ cations and, because the Cl- ions form soluble complexes with the
Fe2+ ions, a passive film is not formed and the process stimulates further metal
dissolution. The soluble iron-chloride complexes diffuse away from the steel and
subsequently break down, resulting in the formation of expansive corrosion products and,
simultaneously freeing the Cl- ions, which are then able to migrate back to the anode and
react further with the steel. In this overall process, hydroxyl ions are continuously
consumed, locally decreasing the pH (i.e. making the solution acidic in that localized
3

region) and, thereby, enhancing further metal dissolution. The Cl- ions, on the other
hand, are not consumed and the attack then becomes autocatalytic. Ultimately, the
reinforcement cross-section and its structural resistance are seriously compromised.

Figure 3. Schematic representation of (a) passive corrosion and (b) chloride-induced


active corrosion of steel in concrete.

Either of these hypothesized mechanisms would explain the local nature of the
attack often observed. The local actively corroding areas behave as anodes while the
remaining passive areas become cathodes where reduction of dissolved oxygen takes
place. The galvanic cells may be "macro" or "micro" in scale depending on a number of
factors, as described below. Thus, the anode and cathode may be widely separated or they
may be adjacent on an atomic scale.

Chloride threshold concentration for corrosion initiation


Many studies [7-11] have been made to determine the threshold or critical value of
chloride concentration below which active corrosion will not occur. The interest in
knowing this value is twofold: (i) to specify limits of chloride contamination in
aggregates or water used in the concrete mixture and (ii) to permit prediction of the
incubation period between first exposure to chlorides and the onset of active corrosion
and, thereby, allow for scheduled maintenance and rehabilitation. Unfortunately, the idea
of a universal value applicable to all structures is not realistic because it will be a function
of many variables including:
mixture proportions of the concrete;
type and specific surface area of the cement;
use of any supplementary cementing materials (SCMs)
w/c ratio;
sulfate content;
curing conditions, age and environmental history of the concrete;
degree of carbonation of the concrete;
temperature and relative humidity of the environment;
roughness and cleanliness of the reinforcement.

An illustrative representation of the interrelation of some of these factors is given by CEB


(Comit Euro-International du Beton) [12] and is shown in Figure 4. Different national
standards consider different limits for chlorides depending on the experience within that
country but a value of 0.4% Cl- by weight of the dry cement is the most common value.
In the United States, the Federal Highway Administration (FHWA) has stated that
a chloride ion concentration of 0.15% by weight of the cement can be tolerated but that
0.3% is considered dangerous [13]. More recently, McDonald et al. [14] have determined
a value of 0.2%. However, when chlorides diffuse into concrete, some react with the
aluminate phases and become chemically bound and some may become physically
trapped in closed pores or in the hydrate phases. Only the free chlorides will be available
to attack the passive film on the steel and, therefore, the use of total chloride content as
measured per ASTM C 1152 [15] as a measure of the threshold value for corrosion can
be misleading.
The limits on chloride content of constituents have recently been questioned in
view of the increasing use of supplementary cementing materials and their effect on the
pH of the pore solution. For example, the maximum chloride content at the
reinforcement level for steel which did not exhibit any active corrosion after exposure to
natural sea water is given in Table 1. for steel in concretes with different fly ash contents
[16]. The threshold for steel in mortar with either fly ash or silica fume was also found to
be lower than that of steel in mortars without supplementary cementitious materials [8].

Figure 4. The critical chloride content of concrete for corrosion of reinforcing steel
according to CEB recommendations [12].

Two factors should be noted. First, that this is a potential problem only for
chloride contamination of the constituents; chlorides penetrating into hardened concretes
with SCMs exhibit lower diffusion rates and lower active corrosion rates, which are
believed to more than offset the lower chloride threshold value. Second, that, in both of
these investigations, the chloride thresholds were significantly higher than those
recommended by the FHWA, suggesting that the latter are conservative relative to those
measured in laboratory experiments. Nevertheless, suppliers of potentially contaminated
materials should be cognizant of the effects of pozzolans on the amount of chloride
needed to initiate active corrosion.
Table 1. Effect of Fly Ash Replacement on the Chloride Threshold Concentration for
Corrosion [16].

Fly-Ash content (% Chloride


threshold
replacement of cement) concentration (% by
mass of dry cement)
0
0.70
15
0.65
30
0.50
50
0.20
Very high levels of chlorides can accumulate in concrete, particularly in the
splash zone of marine structures or, for example, in columns embedded in ground which
are exposed to de-icing salt run off or the substructure of bridges exposed to either runoff from above or splash from the roadway below. As moisture evaporates from the
exposed surface of the concrete, the salts remain behind. The subsequent wicking of
the saline solution from the sea or soil by capillary suction into the dry concrete
replenishes the water and builds up the chloride levels and the process repeats itself.
There is little field information on the effect of temperature and relative humidity
on the rate of chloride-induced corrosion but laboratory tests have shown that, if the
internal humidity in concrete is less than ~85%, high active corrosion rates cannot be
sustained [17].

Carbonation-Induced Corrosion
Carbon dioxide from the atmosphere reacts with the calcium hydroxide (and other
hydroxides) in the cement paste by the following reaction:
Ca(OH)2 + CO2 CaCO3 + H2O

Eq. 3

effectively neutralizing the pore solution.


Carbonation is detected as a reduction in the pH of the pore solution in the surface
regions of the concrete and appears as a fairly sharp front, parallel to the surface. Behind
the front, the Ca(OH)2 has completely reacted and the pH is ~8 whereas ahead of the

front, the pH is >12.5. The depth of carbonation increases with time and the rate at
which it advances is a function of relative humidity (RH): the penetration of the CO2 into
the concrete is highest at low RH but the reaction with the Ca(OH)2 takes place in
solution and is, therefore, highest in saturated concrete. The net result of these two
factors is that carbonation is most rapid in the 50% - 70% RH range, Figure 5 [7, 18].
The carbonation front penetrates the concrete at an ever decreasing rate because of three
factors. Firstly, the gas has to penetrate further into the concrete and, secondly, the
concrete continues to hydrate and becomes more impermeable as it ages. Finally, the
carbonation itself decreases the permeability both by the precipitation of the carbonate in
the existing pores and because the reaction releases water, which could result in increased
hydration [19].
When the carbonation front reaches the reinforcement, the passive film is no
longer stable and active corrosion initiates. Unlike chloride-induced corrosion, the
corrosion process is generalised and relatively homogeneous. Moreover, the corrosion
products tend to be more soluble in the neutral carbonated concrete and may diffuse to
the surface appearing as rust stains on the concrete, rather than precipitating in the
concrete cover and causing stresses and cracking. The corrosion rates are lower than
those caused by chlorides but, over a long period, the cross-section of the reinforcement
can be reduced significantly while there is little visible damage to the concrete. The
active corrosion process is described schematically by Fig. 3(a) but with a much higher
rate than the passive corrosion.
Although an intermediate RH provides the highest rate of carbonation, active
corrosion of any significance does not occur in that humidity range [1]. Consequently,
the most aggressive environment for carbonation-induced corrosion is alternate semi-dry
and wet cycles [18]. Carbonation can, therefore, be a major factor in the durability of
concrete in hot climates where the concrete is easily dried out and periodically subjected
to saturation by rainstorms. Chloride attack and carbonation can act synergistically and
are responsible for major problems in hot coastal areas. Carbonation-induced corrosion
is not found to be a major problem in northern North America where adequate concrete
cover over steel is used.
1.0

Rate of Carbonation

0.8
0.6
0.4

0.2
0
20

40

60

80

100

Relative H umidity, %

Figure 5. The influence of relative humidity on the rate of carbonation of concrete [18].

Corrosion Products
The most detrimental consequence of chloride-induced reinforcement corrosion is the
build-up of voluminous, insoluble corrosion products in the concrete which leads to
internal stresses and, eventually, to cracking and spalling of the concrete cover.
Obviously once such damage is visually apparent, the reinforcement is prone to very
rapid further corrosive attack because access to oxygen and moisture is no longer limited
by diffusion through the concrete cover.
All forms of iron oxide and hydroxide have specific volumes greater than that of
steel but their volumes vary by a factor of more than five as indicated in Figure 6 [19].
Thus, the degree of damage to the concrete produced by a certain amount of corrosion
will depend on the specific corrosion products formed and their distribution within the
concrete cover as well as on the porosity and strength of the concrete itself.
Fe
FeO
Fe3O4
Fe2O3
Fe2O3
FeOOH
FeOOH
FeOOH
FeOOH
Fe(OH)2
Fe(OH)3
Fe2O3 . 3H2O
0

3
4
5
Unit Volume

Figure 6. Specific volume of the corrosion products from iron [19].

In many reports it is assumed that the corrosion products are rust, i.e. Fe2O3.3H2O
because this is the orange colored product observed on damaged concrete. Consequently,
in models, it is also assumed that the corrosion products are more than six times as
voluminous as the steel from which they are formed and the predicted stresses in the
concrete are based on this conclusion. In fact, analysis of the products formed indicates
that they are Fe3O4, -Fe2O3 -Fe2O3, -FeOOH and -FeOOH and have a specific
volume between 2.2 and 3.3 times that of the steel. It is only after cracking and spalling
and, thus, exposure to the atmosphere, that these products convert to the familiar rust.
Moreover, results which have been obtained to date [20, 21] suggest the quality of the
concrete, the use of supplementary cementing materials and the presence of macrocracks
do have an influence of the specific corrosion products and on the corrosion-induced
deterioration of the cover once active corrosion has been initiated.
Theoretical models of the quantity of corrosion [22] needed to cause cracking of
the concrete cover assume the corrosion products are (a) of a uniform specific volume
and (b) all form at the interface between the reinforcement and the concrete.
8

Observations [20, 21] of the products formed during the active corrosion in real concrete
with normal shrinkage and loading cracks shows that this is not the case. The corrosion
products vary in composition within the same concrete and they precipitate within the
cover, as illustrated in Figure 7 not just at the steel/concrete interface.
existing crack
B

corrosion
further
products
cracking
developed

10 mm
40 m

(b)

(a)

Figure 7. (a) Precipitate of magnetite in the concrete cover of HPC. A is a particle of


Akaganeite (- FeOOH) containing approx. 40% Fe in HPC. The surrounding material
labelled B contains approximately 18% Fe as corrosion product intimately embedded in
the cement [20]; (b) corrosion products in a crack penetrate into the concrete causing
further cracking [23].

Microcell and Macrocell / Galvanic Corrosion


Microcell corrosion is the term given to the situation where active dissolution and the
corresponding cathodic half-cell reaction (the reduction of dissolved oxygen) take place
at adjacent parts of the same bar, as illustrated in Figure 8 (a). This process always
occurs in practice and, in most cases, is the dominant corrosion process. Macrocell or
galvanic corrosion can occur when the actively corroding bar is coupled to another bar,
which is passive, either because of its different composition or different environment. For
example, the former situation might occur when black steel is in contact with stainless
steel and the latter situation when a top mat in chloride-contaminated concrete is coupled
to a bottom mat in chloride-free concrete, as in Figure 8 (b). Macrocells can also form on
a single bar exposed to different environments within the concrete or where part of the
bar extends outside the concrete. The process is the same in all cases and, in all cases,
the corrosive action of the macrocell is added to that of the microcells.
It should be emphasized that the simplified view of the active steel becoming the
anode and the passive steel becoming the cathode is not actually correct. In each of
these cases, the anodic and cathodic reactions occur on both metal surfaces; when the two
metals are coupled, the anodic corrosion of the active metal increases and the anodic
corrosion of the passive metal decreases.
While macrocell corrosion can be measured directly, the same is not true of
microcell corrosion and most investigators choose to neglect the microcell component.
This has led to the general assumption that macrocell corrosion is always the dominant
component. On the other hand, Trejo and Monteiro [24] concluded that the difference
between the corrosion mass loss measured gravimetrically and that calculated from

macrocell corrosion rate measurements must be due to the microcell corrosion. Andrade
et al. [25] have analyzed the relative contributions of microcells and macrocells and
concluded that: (i) the influence of the latter only becomes significant, i.e. of the same
order as the microcell corrosion, when the ratio of surface areas of the passive:active
regions is greater than approximately 50:1 and (ii) the theoretical maximum effect (with
an infinitely large cathode and infinitely small anode) would be an increase in active
corrosion rate of only 2 - 5 times that of the microcells alone. This has been confirmed in
laboratory studies [26]. Suzuki et al. [27] concluded that the maximum anodic dissolution
rate of the steel in concrete was the limiting rate controlling factor not the anode/cathode
area ratio.

a) Microcell
corrosion

b) Macrocell
corrosion

Figure 8. Schematic illustration of (a) microcell corrosion and (b) macrocell corrosion [26].

THE INFLUENCE OF CONCRETE PARAMETERS ON


CORROSION OF REINFORCEMENT
Concrete Mixture Design
Chloride ions from de-icing salts and/or marine environments penetrate the concrete
cover depth to reach the surface of the reinforcing steel by a number of mechanisms. As
illustrated in Figure 9, the surface of the concrete may be dry, allowing the dissolved
chlorides to be absorbed by capillary action together with moisture through the
interconnected pores in the cement paste. At deeper levels, concrete rarely dries out in
10

the atmosphere [28] and so continued penetration of the chlorides is by diffusion through
the pores, which is a much slower process than absorption. A third mechanism is via
cracks in the concrete cover, which is discussed below.
Porosity in cement paste consists of capillary pores, gel pores and calcium silicate
hydrate (C-S-H) interlayers [29]. Capillary pores are the remains of originally watercontaining spaces between cement particles that have not been filled up by products of
hydration [30]. They are the largest (diameter > 5 nm [31]), and their number and
interconnectivity control the ingress of chloride ions, oxygen and moisture into concrete
[32]. Gel pores and interlayer spaces are believed to be too small and disconnected to
contribute to transport.
Two factors that significantly influence capillary porosity in concrete are the
water to cementitious materials (w/cm) ratio [33] and the use of supplementary
cementing materials (SCMs) [34]
Theoretically, a w/cm ratio of 0.42 is required for the complete hydration of cement.
However, hydration is a gradual process and the unused mixing water is retained in the
capillary pores [22]. Higher w/cm ratios, traditionally used to give a workable mixture,
increase the amount and interconnectivity of capillary porosity in the cement paste
allowing greater diffusion. With the advent of high range water reducing agents, much
lower w/cm ratios are now possible and significantly limit the penetration of chloride
ions.
SCMs, such as fly ash (FA) silica fume (SF) and ground granulated blast furnace
slag (GGBFS) react with the potassium, sodium and calcium hydroxides in the pore
solution of the cement paste, thereby reducing the pH of the solution. As has been shown
above, this can reduce the amount of chlorides necessary to initiate active corrosion. On
the other hand, the additional C-S-H produced as a result of the pozzolanic reactions
between the SCMs and Ca(OH)2 can block the capillary pores. Because of its small
particle size (<0.1m), un-reacted SF also reduces the capillary porosity. This retards the
ingress of chloride ions, thereby increasing the initiation time for corrosion. It also
decreases the corrosion rate, once initiated, by decreasing the mobility of OH-, which is
required to complete the macrocell and microcell corrosion circuits, as shown in Figure 8.
Concrete with low w/cm ratios ( 0.35) and incorporated SCMs provide high
strength and low permeability. It should be noted here that these mixtures must be
accompanied by adequate curing to prevent the concrete from cracking due to selfdesiccation or autogenous shrinkage. These concretes, exhibit long reinforcement
corrosion initiation times [26, 31] and low corrosion rates [35]

The Influence of Cracks in the Concrete on the Corrosion of


Embedded Steel
While interconnected capillary porosity provides a tortuous path for the ingress of
chloride ions, oxygen and moisture into concrete, cracks provide a more direct path.
Cracks in concrete can be classified as macrocracks and microcracks or, from the
viewpoint of their effects on corrosion as those transverse to the reinforcement and those
parallel to the bar (longitudinal cracks). The sources of cracks in concrete include
shrinkage [22], chemical reactions (e.g. alkali aggregate reaction, [31]), weathering

11

processes (e.g. freezing and thawing [36]), reinforcement corrosion [37] and mechanical
loading.
Concrete always contains cracks and codes on concrete structural design such as
ACI 318 [38] take this into account and relate permissible crack widths to exposure
conditions. However, an understanding of the effects of cracks on corrosion is limited
[39-41]. For concrete with multiple cracks, corrosion at one crack appears to protect the
steel at the other cracks by forming a galvanic cell or there is a low corrosion rate at all
the cracks [42]. Chloride ingress is significantly enhanced by cracks because the ions
penetrate the concrete cover from the walls of the crack as well as from the outer surface
of the concrete [43], as illustrated schematically in Figure 9. Thus, while the chlorides
reach the steel very rapidly directly through the crack, they also reach adjacent areas of
steel more rapidly than in uncracked concrete.

Figure 9. Schematic illustration of chloride diffusion in cracked concrete.

Low w/cm ratio concrete containing SCMs has been found to provide
goodprotection for steel exposed to transverse cracks, in large part because of its
resistance to chloride penetration from the walls of the crack. However, the benefits are
not as great as they are for sound (uncracked) concrete [40]. Moreover, the higher the
quality of the concrete, the more spatially localized the corrosion along the
reinforcement. Unfortunately, however, this can be accompanied by a greater depth of
corrosion, leading to the possibility of the bar being severed [21].
In the case of cracks parallel to the reinforcement, low w/cm ratio concrete
containing SCMs does not appear to have any beneficial influence on the corrosion of
bars [44], which is not surprising in view of the fact that the whole length of the bar is
directly exposed to the environment via the crack.

ALTERNATIVE REINFORCING MATERIALS


Epoxy coated reinforcing bars
Epoxy provides a barrier coating to the steel and, thereby, prevents the chlorides from
breaking down the passive layer on the steel providing a longer service life. While many
structures with epoxy-coated steel have performed well, some have not. The latter cases
have been attributed either to flaws in the coating, most likely due to defects introduced

12

during construction, or to absorption of moisture by the epoxy leading to swelling and


debonding from the steel [45, 46]. It is generally concluded that good quality epoxy
coatings will increase the time to initiate corrosion but, once initiated, the corrosion rate
will be about the same as that of uncoated black steel.

Stainless steel reinforcing bars


With increasing service life requirements, stainless steel is being regarded as a viable
alternative reinforcement despite its higher cost. The most common grades of stainless
steel for reinforcement are 316LN and 2205, both of which have excellent corrosion
resistance [47, 48] and are commercially available. They should provide service lives
well in excess of 100 years. Research shows that grade 304 is less corrosion resistant
than the other two grades [49] but, in fact has the longest successful field record of any
stainless steel [50]. The cost of the stainless steels is more than five times that of black
steel [51] but, on a life cycle cost basis, they are considered to be cost-effective [52].

Corrosion resistant reinforcing bars MMFX and 2201LDX


In order to provide better corrosion resistance than black steel but at lower cost than the
traditional stainless steels, efforts have been made in recent years to develop corrosion
resistant alloys for reinforcement. Three examples of these steels are (i) a low carbon,
low chromium microcomposite steel designated as MMFX-2 (ASTM A615 Grade 75)
[53], (ii) a 1.5% Ni 21% Cr alloy designated 2101LDX (ASTM A955-98) [51], and
(iii) a 3% Ni, 12% Cr alloy designated 3Cr12 [53]. In macrocell corrosion tests [54], the
corrosion initiation times for these alloys were measured to be 2, 7 and 1.6 times that of
black steel, whereas the 304 and 316LN grades had initiation times in excess of 12 times
that of the black steel (i.e. they had not begun to corrode during the period of the
research). The corrosion rates of the MMFX and 2102LDX after initiation were less than
half of that exhibited by the black steel. Their cost is between 2 and 4 times that of black
steel, whereas the cost of the stainless steels is more than five times that of black steel
[55].
Other investigators [56] have concluded that high performance alloys
outperformed black steel from a corrosion resistance standpoint. Unlike the various
grades of black steel however, a relatively wide range of corrosion performance was
apparent for the high performance counterparts depending on the alloy and surface
condition.

Galvanized steel reinforcing bars


Galvanized steel reinforcement has been used in reinforced concrete structures since
1930s [57]. The advantages of hot-dipped galvanizing are two-fold: unlike most other
forms of coatings, a metallurgical bond is formed between the steel and the zinc which
means that the coating is not susceptible to flaking or other forms of separation from the
substrate. Secondly, zinc not only forms a barrier coating but acts as a sacrificial anode.
Thus, any scratches or other flaws in the coating are not critical and do not lead to active
corrosion of the underlying steel. Zinc has the advantage over black steel that it can
13

tolerate more chlorides (approx 2.5 times [58-60] and lower pH levels [pH~8] before
significant active corrosion is initiated. Thus, it would provide better protection than
black steel to both chloride-induced and carbonation-induced corrosion.
The galvanized layer has the disadvantage that it corrodes very rapidly in the wet
cement but this reaction rate ceases once the concrete hardens [61, 62]. While
chromating the galvanized bars has been performed to minimize this initial corrosion, the
use of chromates is not recommended because of adverse health effects of the hexavalent
chromate ions. However, recent research has indicated the effect of chromating is
minimal at best [62]. Because of its passivation in neutral solutions and its sacrificial
anode role when in contact with steel, galvanized steel is ideally suited for parts which
are to be partially embedded in concrete and partially exposed to the atmosphere.

Non-metallic reinforcement
The carbon-fiber reinforcements currently being marketed [63] do not suffer from
corrosion and will not be discussed here. The long term performance of these materials in
concrete has not yet been evaluated. Whether the epoxy matrix for these bars will suffer
the moisture absorption that was observed in epoxy coatings is not yet known.

THE CORROSION BEHAVIOR OF OTHER METALS


IN CONCRETE
Aluminum and its alloys
Aluminum develops a passivating layer on its surface at pH 4-9 [6] which protects it from
further corrosion. However, the passive layer is unstable in alkaline environments and
aluminum reacts with sodium and potassium hydroxides in concrete, a reaction which is
accompanied by the evolution of hydrogen gas. This reaction occurs at high rates in wet
concrete causing plastic concrete to develop a very porous structure as it hardens. This
may affect the bond between aluminum and concrete, as well as, increase the ingress of
chloride ions into concrete which exacerbates the corrosion of aluminum. The use of
chloride admixtures also greatly increases aluminum corrosion and spalling. The
corrosion products of aluminum are extremely expansive and cause cracking and spalling
of the surrounding concrete [64, 65].
Aluminum elements should never be embedded in concrete exposed to moist
service environments, unless the aluminum is prevented from direct contact through use
of a protective barrier.

Lead
Although lead is deemed to have superior corrosion properties to those of steel and
aluminium under atmospheric conditions, it reacts with calcium hydroxide in concrete to
form soluble lead oxides and hydroxides, which do not protect the metal from further
corrosion. As with aluminium, the reaction occurs at high rates in wet concrete which
decrease as concrete hardens. If lead is to be placed in wet concrete, it should be
protected using bituminous or plastic coatings [64, 65].
14

Copper and its alloys


Copper and its alloys do not corrode in non-chloride contaminated concrete provided it
does not contain ammonia and nitrates, which can induce stress corrosion cracking in the
metal. The copper oxide formed in alkaline environments such as concrete is not soluble
and, hence, forms a protective layer that limits further corrosion. However, the presence
of chlorides in concrete perforates copper [64, 65]. Macdonald et al. [14] tested copperclad reinforcement and found it to have significantly higher resistance to chlorideinduced corrosion than did black steel.

SERVICE LIFE OF REINFORCED CONCRETE EXPOSED


TO CHLORIDES
The service life of black steel reinforcement in OPC concrete is often depicted as
consisting of two phases: an incubation period during which chlorides or carbonation
penetrates the cover, and a propagation period where active corrosion causes degradation
of the concrete. However, it is illustrative to assess the behavior of the reinforcement
independently of that of the concrete, as has been shown schematically Figure 10.

Figure 10. Schematic representation of the variation of corrosion rate of the reinforcement
as a function of time (shown in blue) and the corresponding damage produced in the
concrete cover (shown in red) [18].

The blue curve in Figure 10 represents the steel corrosion and shows an initial,
relatively high, corrosion rate until the passive film on the steel is developed (a matter of
several days to a few weeks). Thereafter, the corrosion rate will be low (~10-4 A/m2
0.116 m/yr) until the chlorides reach the steel in sufficient quantities to initiate active
corrosion. The corrosion rate will jump very rapidly, possibly by as much as three orders
of magnitude, and will remain high until all the dissolved oxygen in the cement paste
pores adjacent to the steel has been consumed by the cathodic half cell reaction:
O2 + H2O + 2e- 2OH15

The corrosion rate will then drop to a level determined by the


diffusion/permeation of oxygen from the environment. The specific corrosion products
formed during these processes and their distribution within the concrete cover will be
determined by: (i) the porosity and the density of micro- and macro-cracks in the
concrete, i.e. the available space for precipitation of the products and (ii) the availability
of oxygen and moisture, and the concentration of chlorides. Once the volume of
corrosion products causes stresses in excess of the tensile strength of the concrete, cracks
will develop and propagate out to the surface. At this point, oxygen is made readily
available at the reinforcement and the corrosion rate increases as shown.
Simultaneously, the damage in the cover can be described by the red curve in
Figure 10. There will be some initial damage as the structure comes into service and
loading and settling cracks develop but do not propagate further. After initiation of
active corrosion and the precipitation of corrosion products, cracks will begin to develop
around the corrosion products, as illustrated in Figure 10. As these grow, they will
intersect and eventually propagate to the surface, causing spalling. For the purposes of
this report, this is considered the end of service life.
Many models [66-72], based on Ficks Second Law, have been developed to
predict the length of the incubation period under different circumstances. Only a few
attempts [37, 73] have been made, however, to model the propagation phases of the
deterioration process including the cracking.

CORROSION MONITORING TECHNIQUES


One of the major problems associated with reinforcement corrosion is that its initiation
and early stages of propagation cannot be detected visibly. Yet early detection of
corrosion in a reinforced concrete structures can provide the opportunity to schedule
appropriate maintenance procedures, thereby ensuring the safety of the structure. If
corrosion remains undetected until cracking and spalling occur, then the costs of repair
are significantly higher because all of the concrete cover and much of the reinforcement
must be replaced. Moreover, patch repair at this time can enhance active corrosion in the
surrounding original reinforcement and lead to a continuous cycle of repairs.
The most commonly used techniques for evaluating the condition of embedded
reinforcement are based on the electrochemical nature of the process. The electrons
released by the iron as it corrodes (Equation 1) and consumed by the dissolved oxygen
(Equation 2) constitute a current, illustrated in Figure 3, which can be indirectly
measured. This current, representing dissolution of one iron atom for each two electrons
in the current, can be converted to thickness of steel dissolved by the equation:
1mA/m2 = 1.16 mm/year = 0.0456 mils/year (mpy)

(Eq. 3)

The currently available electrochemical monitoring techniques are described


below. Unfortunately only a few of these can be readily used in the field.

16

Half-cell potential mapping


Half-cell potential measurements are the most widely used method of detection of
corrosion of steel reinforcement in concrete. The method was introduced in the 1970s by
Richard F. Stratfull in North America [74, 75] and by the Danish Corrosion Centre in
Europe [75-77] and was approved in 1980 as a standard by ASTM [78]. ASTM C 876
involves measuring the electrochemical potential of the steel reinforcement with respect
to a copper/copper sulfate reference electrode by connecting one wire of a high
impedance voltmeter to the reinforcement and the other to the reference electrode placed
on the surface of the concrete. The voltage can provide an indication of the probability of
active corrosion of the steel.
The recommended ASTM interpretation of the
measurements is as follows:
Table 2. Probability of Corrosion of Carbon Steel According to Half-Cell Potential Reading
[78].

Half-cell potential reading vs. Cu/CuSO4

Corrosion activity

More positive than -200 mV

90% probability of no corrosion

Between -200 and -350 mV

An increased probability of corrosion

More negative than -350 mV

90% probability of corrosion

It is important to note that the interpretations are of probabilities of corrosion only


and that there are other possible causes of high negative potentials. For example, in
deaerated concrete, such as that submerged in deep water, potentials as low as -700 mV
are normal despite the fact that the steel remains passive. It should also be noted that
half-cell potentials give no indication of the rate of corrosion, nor of how long the steel
has been corroding.

Linear Polarization Resistance (LPR)


When a very small electrical potential (of the order of 10 mV) is applied to the corroding
steel, the relationship between potential and current is linear. The polarization resistance,
Rp, is the ratio of the applied potential to resulting current and is inversely proportional to
the corrosion rate [79, 80]. In order to calculate a corrosion rate with the LPR technique,
the following fundamental assumptions must be made [5]:
uniform corrosion damage;
the rate controlling step in corrosion is activation polarization (i.e. the ease of

stripping electrons from the iron);


there is a single anodic and a single cathodic reaction;
the proportionality constant between corrosion rate and Rp must be known;

17

the electrical resistance of the solution (i.e. the concrete) is negligible;


the half-cell potential is stable.

In fact, of these assumptions, only the third one is applicable to the case of chlorideinduced corrosion of embedded reinforcement. Despite this, the LPR technique has
become increasingly popular for measuring corrosion in the field because: (i) it is a nondestructive technique; (i) it is simple to apply and (iii) it usually needs only a few minutes
for corrosion rate determination [81].
Consequently, an increasing number of commercial instruments are available for
field measurements of LPR. Measurements are performed by applying a potential in the
range of 10 to 20 mV about the Ecorr, either as a constant pulse (potentiostatic), or a
potential sweep (potentiodynamic), and measuring the current response. Alternatively, a
current pulse (galvanostatic) or a current sweep (galvanodynamic) can be applied, and
potential response is measured. The RP and, in turn, the corrosion rate is calculated by
the instrument.
In addition to the assumptions mentioned above, the major limitation of these
techniques is that it is impossible to know the area of steel which is being polarized by
the applied potential, or the area which is actively corroding. Therefore, the corrosion is
generally considered to be uniform over the polarized area and the measured corrosion
current is divided by a guesstimated polarized area to give an average corrosion rate.
An additional limitation is that this is an instantaneous corrosion rate and (a) gives no
indication of how long corrosion has been going on or (b) how the corrosion rate varies
with time and ambient conditions. An example of how much corrosion rates vary is
given in Figure 11 together with the corresponding temperature and relative humidity
variations inside the concrete over the same time period. It should be noted corrosion rate
increases or decrease with increasing or decreasing temperature, as expected. The
relative humidity, on the other hand, remains approximately constant, at ~100%, over this
period, despite significant changes in the external temperature and relative humidity. This
illustrates that the concrete is not dried out at all at depths of approximately 25 mm, and
that only the surface layers are affected by drying.

18

Current density (A/m2)

0.03
0.025

(a)

0.02
0.015
0.01
0.005
0
40 60 80 100 120 140 160 180 200 220 240 260

Temperature (C)

Time (days after casting)

40
35
30
25
20
15
10
5
0
-5
-10
-15
-20
-25

(b)

40 60 80 100 120 140 160 180 200 220 240 260

Relative Humidity (%)

Time (days after casting)


120

(c)

100
80
60
40
20
0
40 60 80 100 120 140 160 180 200 220 240 260
Time (days after casting)

Figure 11. (a) Corrosion rate of steel, (b) internal temperature and (c) internal relative
humidity variations in OPC concrete [41].

Attempts to overcome the uncertainty in the polarized area have resulted in the
development of the guard ring. The objective of the guard ring is to concentrate the
polarization to within a specified length of the steel reinforcement. Guard ring electrodes
19

have been incorporated in several commercial corrosion measurement instruments [82,


83]. Studies have shown, however, that all of instruments have some limitations [84-86]
and that the use of a guard ring does not solve all of the deficiencies of the technique..

Cyclic potentiodynamic polarization


Cyclic polarization is relatively non-destructive technique that provides the corrosion
rate, corrosion potential and susceptibility to pitting corrosion of the metal in the test
environment, as well as giving information about the expected behavior of the steel
should its potential be changed by, for example, exposure to stray currents, coupling with
other metals or the surrounding concrete becoming anaerobic. Like most electrochemical
techniques, cyclic polarization is carried out with three electrodes: a working electrode
(the reinforcing steel), a counter electrode and a reference electrode. The potential of the
specimen is changed continuously or in steps, while the resulting current is monitored.
From a plot of the applied potential versus the logarithm of the resulting current density,
the condition of the steel in the present environmental, as well as its potential behavior
under other conditions, can be assessed. Cyclic polarization is most useful in the
laboratory, for example, to evaluate the behavior of steel in new concrete mixes or the
behavior of alternative reinforcing materials in normal concrete. While such tests have
been successfully performed in the field by the authors, they suffer the same limitation of
LPR that of knowing the polarized area of the steel, as well as taking a longer time to
perform.

Electrochemical Impedance Spectroscopy (EIS)


Analyzing the response of corroding electrodes to small-amplitude alternating potential
signals of widely varying frequency is the basis of EIS [81]. EIS has become popular as
a tool of corrosion measurement and monitoring applications for which traditional DC
techniques have been unsuitable. It has been used extensively for coated and painted
materials to detect the impact of holidays (flaws) in the coating. Thus it is a useful
technique for evaluating epoxy-coated reinforcement [87]. It has also been employed to
evaluate the effectiveness of cathodic protection [86] and corrosion inhibitors [88] as
well as chloride diffusivity in concrete [89] and electrical resistance of concrete. Most of
these studies have been carried out in the laboratory and, based on the authors
experience, this technique is unsuitable for field measurements because the
measurements take too long and because the system is susceptible to electrical
interference.

PREVENTION AND PROTECTION METHODS


Cathodic Protection (CP)
The first example of cathodic protection applied to concrete structures was reported in
1957 [90] and widespread use of the technique for protecting bridge decks contaminated
by deicing salts began in 1973 in North America [91, 92].

20

Cathodic protection is based on changing the potential of the steel to more


negative values. This potential change can be obtained by connecting an external anode
to the steel and impressing an electrical DC current through the reinforcement using a
rectified power supply. Activated titanium expanded mesh with a surface coating of
titanium oxide is the most widely used anode in practice. This external anode is mounted
on the surface of concrete. The positive terminal of low voltage DC current source is
connected to the mesh and the negative terminal is connected to the steel bars. In order
to halt corrosion of the reinforcement completely, the potential would have to be lowered
to values more negative than ~ -1200 mV CSE (immunity region in Pourbaix diagram for
iron) [6]. However, at those values, the cathodic reaction would result in hydrogen
evolution which could cause hydrogen embrittlement of any prestressed steel in the
structure. Consequently, the potential is carefully controlled to a level (typically -1000
mV CSE) at which the corrosion is negligible but not actually stopped.
Alternatively, the process can rely on the galvanic effect and a less noble metal
can be used as sacrificial anodes. In this case, metals such as zinc and zinc-aluminium
alloys, are applied to the concrete surface by flame spray, as a mesh or sheet adhered via
a conductive gel [93].
CP must be applied for the remaining lifetime of the structure and must be
monitored regularly. For this purpose, the potential of the steel should be monitored with
respect to a reference electrode before and after disconnecting the system to make sure
that the system is depressing the potential from the open-circuit potential. The limits and
standard practices are outlined in [94]. Resistance of the concrete should be considered.
Electrochemical Chloride Extraction (ECE)
The goal is to repel the (negatively charged) chloride ions from the reinforcement and out
of the surface of the concrete by applying a negative charge to the reinforcement, as
illustrated schematically in Figure 12. ECE is similar to CP but a higher current is
applied for a short period of time (a matter of weeks). During the treatment, any
corrosion products are electrochemically reduced at the reinforcing steel [95].

21

Figure 12. Schematic illustration of the application of the electrochemical chloride


extraction treatment to a corroding reinforced concrete structure [95].

ECE is being applied successfully in North America but the long term effects of
such treatment have not yet been evaluated. In laboratory tests on steel in mortar, it has
been shown that the applied current can significantly alter the composition and
morphology of the mortar at the steel/mortar interface and the steel/concrete bond
strength [96, 97]. ECE is best applied when the chlorides are almost reaching the
reinforcing steel but have not yet initiated active corrosion. It cannot stop the damage if
the process of deterioration is advanced. Therefore, the extraction treatment may be
considered an appropriate preventive method but, strictly speaking, not a rehabilitation
method [98]. Thus, while the electrochemical extraction process is obviously successful
in removing the chlorides, it should be used with caution and practical limits of current
density should be established to minimize the harmful effects [97].

Inhibitors
Corrosion inhibitors are defined as chemical compounds that reduce the corrosion
rate by affecting the anodic or cathodic half-cell reactions or both. It should be noted that
the materials which affect the transportation mechanism of the aggressive species to the
reinforcing steel bars are not considered inhibitors [99].
A corrosion inhibiting admixture to concrete can function by [100]:
Increasing the resistance of the passive film on the steel to breakdown by chlorides
Creating a barrier film on the steel;
Increasing the degree of chloride binding in the concrete;
Scavenging the oxygen dissolved in the pore solution; and
Blocking the ingress of oxygen.
Numerous compounds have been investigated in the laboratory as potential
corrosion inhibiting admixtures to concrete and the most widely used ones are
summarized in Table 3.

22

Table 3. The Most Widely used Inhibitors for Applying in Steel Reinforced Concretes.

Inhibitor
Calcium nitrite

Calcium nitrate

Sodium
monofluorophosphate
(MFP)
Hydroxyalkylamines

Application
Mechanism
Added to the mixing Due to the negative charge of
water of concrete
nitride, it migrated into the pit and
enhance passivation by its oxidative
properties [3, 101]
Added to the mixing In the presence of corrosion
water of concrete
products, nitrate is reduced to nitrite.
This
reaction
in
alkaline
environment is fast enough to
provide required nitrite to inhibit
corrosion [99]
Added to the mix or None of the studies on MFP
applied to the surface
determined the mechanism by
which MFP inhibits the corrosion of
steel in concrete (anodic, cathodic or
mixed) [102-104].
Added to the mix or Generally, is produced in a gas
applied to the surface
phase and migrate relatively fast to
the surface of concrete and reduces
the corrosion rate. No detailed
information available regards to
protection mechanism [99].

The efficiency of the inhibitors in steel reinforced concrete structures has not yet
been proved and there are many discussions in this area. Lack of understanding the
mechanism, environmental and safety aspects are some on the major limitations of using
inhibitors in reinforced concrete structures properly [99, 101]. Consequently, while
calcium nitrite is used in many jurisdictions and is required in some, it is not permitted in
others. Corrosion inhibitors should conform to ASTM C 1582.

Non-chloride de-icing agents


Sodium chloride, calcium chloride and magnesium chloride have traditionally been used
for snow and ice removal operations. However, these de-icers cause durability problems
in concrete structures. Environmental concerns are also involved. In this regard, several
non-chloride de-icers have been used in recent years. De-icers based on potassium
acetate, sodium acetate, sodium formate, calcium magnesium acetate (CMA), urea,
ammonium nitrate and ammonium sulfate are the most common ones. None of these
compounds contains chlorides and so they are not expected to corrode the embedded
steel. However, there are reports of deterioration of concrete due to application of some
non-chloride de-icers. The cement matrix of concrete is reported to be severely attacked
by CMA solutions [105]; potassium acetate is found to exacerbate alkali aggregate
reactions [106]; ammonium nitrate and ammonium sulfate rapidly attack and disintegrate

23

concrete [107] and urea decreases the resistance of concrete to freezing and thawing
[108] and salt scaling [109].

SUMMARY AND CONCLUSION


Portland cement concretes provide excellent protection for embedded steel in the
absence of chloride contamination or carbonation by (a) acting as a physical barrier and
(b) by chemically passivating the steel surface. Nevertheless, reinforcing bar corrosion is
a major cause of the degradation of reinforced concrete structures particularly in northern
North America, because of the high quantities of chloride de-icing salts used. In addition
the increase in construction near coastal marine environments may increase the potential
for corrosion deterioration in these areas. As a result, some structures of ordinary
portland cement concrete with black steel reinforcement are requiring repair and
remediation long before their current specified service lives (typically 40 50 years) are
reached. Therefore, easier, faster and more reliable condition analysis techniques are
required than those currently available, and described above, to allow corrosion detection
at an earlier stage and, thus, permit remedial action to be taken before major repairs are
required.
At the same time, with the current emphasis on sustainability, building codes are
now requiring longer service lives, of the order of 75 to 100 years. Consequently, for
new structures, there must be a greater understanding of the reinforcement corrosion
process and of materials and structural designs aimed at minimizing the risk of corrosion.
A two-fold approach to corrosion resistant structures should include:
The use of high performance concrete (HPC) to lower concretes permeability
and reduce the rate of ingress of chlorides or carbonation and, thereby,
increase the effectiveness of the physical barrier.
The use of more resistant reinforcing bar materials to provide better chemical
resistance. In those parts of structures exposed to very severe chloride
environments, stainless steel is recommended. Despite the initial expense, it
is a cost effective solution in these circumstances when both direct and
indirect costs (such as user costs) are taken into account. In the somewhat less
severe chloride environments, corrosion resistant alloys such as MMFX or
2101LDX, which are more resistant to chlorides than black steel - but less
corrosion resistant and much less costly than stainless steel - should be
considered. Galvanized reinforcement is recognized as having greater
resistance to chlorides than black steel and is significantly more resistant to
carbonation-induced corrosion or combinations of chlorides and carbonation
than black steel.

ACKNOWLEDGEMENTS
The research reported in this paper (PCA R&D Serial No. 3013) was conducted by
Hansson, C. M.; Poursaee, A. and Jaffer, S. J with the sponsorship of the Portland
Cement Association. The contents of this paper reflect the views of the author, who is
responsible for the facts and accuracy of the data presented. The contents do not
necessarily reflect the views of the Portland Cement Association.
24

REFERENCES:
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.

15.
16.

17.

Rosenberg, A.; Hansson, C. M., and Andrade, C., Mechanisms of Corrosion of


Steel in Concrete, in The Materials Science of Concrete, J. Skalny, Editor, 1989,
The American Ceramic Society, pages 285-313.
Hansson, C. M., Comments on Electrochemical Measurements of the Rate of
Corrosion of Steel in Concrete, Cement and Concrete Research, 1984, 14, pages
547 to 584.
ACI Committee 222, 222R-96, Corrosion of Metals in Concrete, 1996.
Poursaee, A. and Hansson, C.M., Reinforcing steel passivation in mortar and
pore solution, Cement and Concrete Research, Accepted for publication.
M. Tullmin, Linear Polarization Resistance, 2001, Corrosion-Club,
http://www.corrosion-club.com/lpr.htm.
M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions 1974,
Houston, Texas, USA, National Association of Corrosion Engineers.
Tuutti , K., Corrosion of Steel in Concrete, 1982, Swedish Cement and Concrete
Research Institute.
Hansson, C. M.and Srensen, B., The Threshold Concentration of Chloride in
Concrete for the Initiation of Reinforcement Corrosion. in Corrosion Rates of
Steel in Concrete, 1990, Baltimore, Maryland, USA, ASTM STP 1065.
Srensen, B.and Maahn, E, Nordic Concrete Research, 1982, 1, (Paper 24).
Gouda, V. K., Corrosion and Corrosion Inhibition of Reinforcing Steel I:
Immersed in Alkaline Solutions. British Corrosion Journal, 1970, 5, pages 198 to
203.
Gouda , V. K. and Halaka, W. Y., Corrosion and Corrosion Inhibition of
Reinforcing Steel II: Embedded in Concrete, British Corrosion Journal, 1970, 5
pages 204 to 208.
CEB, Guide to Durable Concrete Structures, 1985, Comit Euro-International du
Beton, Lausanne, Switzerland.
Locke, C. E., Corrosion of Steel in Portland Cement Concrete: Fundamental
Studies, 8th International Conference on Metallic Corrosion, 1986, Toronto,
Ontario, Canada.
McDonald, D.; Pfeifer, D., and P. Virmani, Corrosion-Resistant Reinforcing Bars
- Findings of a 5-Year FHWA Study, International Conference on Corrosion and
Rehabilitation of Reinforced Concrete Structures, 1998, Orlando, Florida, USA,
FHWA.
ASTM C1152, Standard Test Method for Acid-Soluble Chloride in Mortar and
Concrete, 1997, ASTM, West Conshohocken, Pennsylvania, USA.
Thomas, M. D. A., and Evans, C. M., Chloride Penetration in High-Performance
Concrete Containing Silica Fume and Fly Ash, in Proceedings of the Second
International Conference on Concrete under Severe Conditions, 1998, Tromso,
Norway, E & FN Spon.
Enevoldsen, J. N., Hansson, C. M., and Hope, B. B., The Influence of Internal
Relative Humidity on the Rate of Corrosion of Steel Embedded in Concrete and
Mortar, Cement and Concrete Research, 1994, 24, pages 1373 to 1382.

25

18.

19.
20.

21.
22.
23.
24.

25.
26.
27.
28.
29.
30.
31.
32.
33.

Tuutti , K., Service Life of Structures with Regard to Corrosion of Embedded


Steel, International Conference on Performance of Concrete in Marine
Environment, 1980, St. Andrews by-the-Sea, Canada, American Concrete
Institute.
Nielsen, A., Concrete Durability (in Danish), The Concrete Book (Beton Bogen),
A.D. Herholdt, et al., Editors, 1985, Aalborg Portland.
Marcotte, T. D. and Hansson, C. M., A Comparison of the Chloride-Induced
Corrosion Products from Steel-Reinforced Industrial Standard versus High
Performance Concrete exposed to Simulated Sea Water, International Symposium
on High Performance and Reactive Powder Concretes, 1998, University of
Sherbrooke, Sherbrooke, Canada,.
Marcotte, T. D. and Hansson, C. M., The Influence of Silica Fume on the
Corrosion Resistance of Steel in High Performance Concrete exposed to
Simulated Sea Water, J. Materials Science, 2003, 38, pages 4765 to 4776.
Atcin, P.C.; Neville, A.M., and Acker, P., Integrated View of Shrinkage
Deformation, in Concrete International, 1997, pages 35 to 41.
Jaffer, S. J. and Hansson, C. M., Corrosion of steel in dynamically loaded
reinforced concrete: a macro- and micro-structural perspective, unpublished.
Trejo, D. and Montiero, P. J., Corrosion performance of conventional (ASTM
A615) and low-allow (ASTM A706) reinforcing bars embedded in concrete and
exposed to chloride environments, Cement and Concrete Research, 2005, 35:
pages 562 to 571.
Andrade, C. M.; Feliu, I. R.; Gonzalez , S.; Feliu, J. A., and Feliu, J. S.,The Effect
of Macrocells Between Active and Passive Areas of Steel Reinforcements,
Corrosion Science, 1992, 33 (2), pages 237 to 249.
Hansson, C. M; Poursaee, A., and Laurent, A., Macrocell and microcell corrosion
of steel in ordinary Portland cement and high performance concretes, Cement
and Concrete Research, 2006, 36, pages 2098 to 2102.
Suzuki, K.; Ohno, Y.; Praparntanatorn, S. and Tamura, H.. Some Phenomena of
Macrocell Corrosion, Corrosion of Reinforcement in Concrete, 1990, Wishaw,
England, Elsevier Applied Science.
Hong, K. and Hooton, R.D. Effects of cyclic chloride exposure on penetration of
concrete cover, Cement and Concrete Research, 1999, 29, pages 1379 to 1386.
Soroka, I., Portland Cement Paste and Concrete, 1979, Surrey, England, The
Macmillan Press Ltd.
Powers, T.C., Properties of Cement Paste and Concrete. in Chemistry of Cement:
Fourth International Symposium, 1960, Washington, D.C., USA, National Bureau
of Standards, U.S. Department of Commerce.
Mehta, P. K., Concrete Structure, Properties and Materials, 1986, Englewood
Cliffs, New Jersey, USA, Prentice-Hall Inc.
Jensen, O. M., Hansen, P. F., Coats, A. M., and Glasser,. F. P., Chloride Ingress
in Cement Paste and Mortar, Cement and Concrete Research, 1999, 29 (9), pages
1497 to 1504.
Kosmatka, S. H.; Kerkhoff, B., Panarese, W. C., MacLeod, N. F., and McGrath,
R. J., Design and Control of Mixtures, Seventh Canadian Edition ed. Vol.

26

34.
35.

36.
37.
38.
39.
40.
41.
42.

43.
44.
45.
46.
47.

Engineering Bulletin 101, 2002, Ottawa, Ontario, Canada, Cement Association of


Canada.
Oh, B.H.; Cha, S. W.; Jang, B. S. and Jang, S. Y. Development of HighPerformance Concrete Having High Resistance to Chloride Penetration, Nuclear
Engineering and Design, 2002, 212, pages 221 to 231.
Ismail, M. E. and Soleymani, H.R., Monitoring corrosion rate for ordinary
Portland concrete (OPC) and high performance concrete (HPC) specimens
subject to chloride attack, Canadian Journal of Civil Engineering, 2002, 29, pages
863 to 874.
ACI Committee 224, Causes, Evaluation and Repair of Cracks in Concrete
Structures, 1998, American Concrete Institute, Farmington Hills, Michigan, USA.
Andrade,C.; Alonzo,C., and Molina, F. J., Cover Cracking as a Function of Rebar
Corrosion: Part II - Numerical Model, Materials and Structures, 1993, 26, pages
532 to 548.
ACI 318, Building Code Requirements for Structural Concrete, American
Concrete Institute, Farmington Hills, Michigan, USA.
Francois, R. and Arliguie, G., Effect of Microcracking and Cracking on the
Development of Corrosion in Reinforced Concrete Members, Magazine of
Concrete Research, 1999, 51 (2), pages 143 to150.
Okulaja, S. A. and Hansson, C. M. Corrosion of Reinforcing Steel in Cracked
High Performance Concrete, Advances in Cement and Concrete IX, 2003, Copper
Mountain, Colorado, USA, Engineering Conferences International.
Poursaee, A., The Effect of Concrete Composition, Wet Curing Period and
Ambient Condition on the Internal Environment in Concrete, MASc. thesis,
University of Waterloo, Canada, 2004.
Suzuki, K.; Ohno,Y.; Praparntanatorn, S., and Tamura, H.. Mechanism of Steel
Corrosion in Cracked Concrete, International Symposium on Corrosion of
Reinforcement in Concrete Construction (3rd, 1990, Wishaw, England, Elsevier
Applied Science.
Win, P. P.; Watanabe, M., and Machida, A., Penetration profile of chloride ion in
cracked reinforced concrete, Cement and Concrete Research, 2004, 34 (7), pages
1073 to 1079.
Poursaee, A. and Hansson, C., Does the type of concrete and curing time affect
the corrosion of reinforcing steel in cracked concrete?, Cement and Concrete
Research, to be submitted, June 2007.
Weyers, R. E.; Pyc, W., and Sprinkel, M. M., Estimating the Service Life of
Epoxy-Coated Reinforcing Steel, ACI Materials Journal, 1998, 95 (5), pages 546
to 557.
Sags, A. A., Mechanism of Corrosion of Epoxy-Coated Reinforcing Steel in
Concrete, 1991, Florida Department of Transportation & Federal Highway
Administration.
McDonald, D. B.; Sherman, M. R.; Pfeifer, D. W., and Virmani, Y. P., Stainless
Steel Reinforcing as Corrosion Protection, Concrete International, May 1995,
pages 65 to 70.

27

48.
49.
50.
51.
52.
53.
54.

55.

56.

57.
58.
59.
60.
61.
62.

Hansson, C.M.; Tullmin, M.; Hunt, S., and Johnson, A. Corrosion Resistance of
Stainless Steel Reinforcing Bars for Concrete, ACI Spring Meeting, 1996, Denver,
Colorado, USA.
Trejo, D. and Pillai, R.G., Accelerated Chloride Threshold Testing Part II:
Corrosion Resistant Reinforcement, ACI Materials Journal, January to February
2004.
Rambll, Pier in Progresso - Mexico Inspection Report - Evaluation of the
Stainless Steel Reinforcement, 1999, Arminox, Viborg, Denmark.
ASTM, A955M-96 Standard Specification for Deformed and Plain Stainless Steel
Bars for Concrete Reinforcement [Metric], 1996, American Society for Testing
and Materials, Philadelphia, Pennsylvania, USA.
Hansson, C. M.; Haas, R.; Green, R., Gepraegs, O. K., and Asaar, R. Al,
Corrosion Protection Strategies for Ministry Bridges, 2000, Ministry of
Transportation Ontario, Canada.
ASTM, A240/A240M-06b: Standard Specification for Chromium and ChromiumNickel Stainless Steel Plate, Sheet, and Strip for Pressure Vessels and for General
Applications, 2006, Philadelphia, Pennsylvania, USA.
Clemea, G. G. and Virmani, Y. P.. Comparison of the Corrosion Resistance of
Selected Metallic Reinforcing Bars for Extension of Service Life of Future
Concrete Bridges in Outdoor Concrete Blocks, CORROSION, March 16 to 20,
2003, San Diego, California, USA.
Hart, W.; Powers, T. C.; Lysogorski, D. K.; Paredes, M., and Virmani, Y. P., Job
Site Evaluation of Corrosion Resistant alloys for use as Reinforcement in
Concrete, 2006, US Federal Highway Administration Publication No. FHWAHRT-06-078
Hart, W.; Powers, R. G.; Leroux, V., and Lysogorski, D. K., A Critical Literature
Review of High Performance Corrosion Reinforcements in concrete Bridge
Applications, 2004, US Federal Highway Administration Publication No. FHWA
FHWA-HRT-04-093.
Yeomans, S. R., Galvanized Steel in Concrete: An Overview, in Galvanized Steel
Reinforcement in Concrete, S. R. Yeomans, Editor, 2004, International Lead Zinc
Research Organization, North Carolina, USA.
Tonini, D. E. and Dean, S. W., Chloride corrosion of steel in concrete, 1976,
ASTM-STP 629.
Cornet, I. and Bresler, B., Corrosion of steel and galvanized steel in concrete,
Materials Protection, 1966, 5 (5), pages 69 to 72.
Yeomans, S.R., Performance of Black, Galvanized and Epoxy-Coated
Reinforcing Steels in Chloride-Contaminated Concrete, Corrosion, 1994, 50 (1),
pages 72 to 81.
Macias, A. and Andrade, C., Corrosion Rate of Galvanized Steel Immersed in
Saturated Solutions of Ca(OH)2 in the pH Range 12-13.8, British Corrosion
Journal, 1983, 18 (2), pages 82 to 87.
Tan, Q. Z. and Hansson, C. M., The effect of different surface treatment of
galvanised steel on its corrosion behaviour in concrete, to be published.

28

63.
64.
65.
66.
67.

68.

69.

70.

71.
72.
73.
74.
75.
76.
77.
78.

ACI committee 440R-96, State-of -the -Art Report on Fibre Reinforced Plastic
(FRP) Reinforcement for Concrete Structures, American Concrete Institute,
Farmington Hills, Michigan, USA.
Woods, H., Corrosion of Embedded Material Other Than Reinforcing Steel, 1966,
ASTM STP No. 169-A, pages 230 to 238.
Nrnberger, U., Corrosion of Metals in Contact With Mineral Building Materials,
Otto-Graf-Journal, 2001, 12.
Weyers, R. E., Service Life Model for Concrete Structures in Chloride Laden
Environments, ACI Materials Journal, 1998, 95 (4), pages 445 to 453.
Weyers, R. E.. Utilization of a Service Life Model for the Selection of Corrosion
Protection Methods for Reinforced Concrete Structures, Understanding
Corrosion Mechanisms of Steel in Concrete, 1997, MIT, Cambridge,
Massachusetts, USA.
MacDonald, K. A. Effect of Assumed Boundary Conditions on the Unsteady State
Determination of Diffusivity of Chloride Ion in Concrete, Understanding
Corrosion Mechanism of Steel in Concrete, 1997, MIT, Cambridge,
Massechusetts, USA.
Kranc, S. C. and Sags, A. A.. Development of Damage Functions to Predict the
Durability of Steel Reinforced Concrete Structural Elements, Understanding the
Corrosion Mechanism of Steel in Concrete, 1997, MIT, Cambridge,
Massechusetts, USA.
Martin-Perez, B., Thomas, M. D. A., and Pantazopoulou, S. J., Service Life
Modelling of Highway Structures Exposed to Chloride Environments,
Understanding the Corrosion Mechanisms of Steel in Concrete, 1997, MIT,
Cambridge, Massechuttes, USA.
Maage, M.; Poulsen, E.; Vennesland, ., and Carlsen, J. E., Service Life Model
for Concrete Structures Exposed to Marine Environments, Initiation Period,
1995, SINTEF, Trondheim, Norway.
Frederiksen, J. M.; Nilsson, L. O.; Sandberg, P.; Poulsen, E., and Tang, L.,
HETEK - A System for Estimation of Chloride Ingress into Concrete, Theoretical
Background, 1997, Danish Road Directorate.
Andrade, C.; Alonso, C., and Molina, F. J., Cover Cracking as a Function of Bar
Corrosion Part I - Experimental Test, Materials and Structures, 1993, 26, pages
453 to 464.
Stratfull, R. F., Laboratory corrosion test of steel in concrete, 1968, California
Division of Highways, State of California, Sacramento, California, USA.
Stratfull, R. F., Half-cell potential and the corrosion of steel in concrete, 1972,
California Division of Highways, State of California, Sacramento, California,
USA.
Stratfull, R. F., S., Laboratory corrosion test of steel in concrete, 1968, California
Division of Highways, State of California, Sacramento, California, USA.
FORCE Technology, Evolution of corrosion in reinforced concrete structures,
www.force.dk/NR/rdonlyres/7A1681DB-71A2-4424-B948D44EEDC38C35/204/17861en3.pdf, 2004.
ASTM, C876-91: Standard Test Method for Half-cell Potentials of Uncoated
Reinforcing Steel in Concrete, 1999, pages 446 to 451.

29

79.
80.
81.
82.
83.
84.
85.
86.

87.
88.

89.
90.
91.
92.
93.
94.
95.

Princeton Applied Research, Electrochemistry and corrosion overview and


techniques, Application Note corr-4, http://new.ametek.com/contentmanager/files/PAR/088.pdf.
Stern, M. and Geary, A.L., Electrochemical polarisation: I. A theoretical analysis
of the shape of polarisation curves, Journal of the Electrochemical Society, 1957,
104 (1), pages 56 to 63.
Jones, D. A., Principles and Prevention of Corrosion, 1992, Macmillan
Publishing Company.
Germann Instruments Inc., http://www.germann.org, Products: Galvapulse. 2004.
James Instruments Inc.,
http://ndtjames.com/catalog/corrosionTesting/gecor8.html, Gecor 8, 2005.
Gepraegs, O. K., Comparison and Evaluation of Electrochemical Techniques and
Monitoring Instruments to determine the Corrosion Rate of Steel in Concrete,
2002, University of Waterloo, Canada.
Wojtas, H., Determination of corrosion rate of reinforcement with a modulated
guard ring electrode; analysis of errors due to lateral current distribution,
Corrosion Science, 2004, 46, pages 1621 to 1632.
Pruckner, F.; Theiner, J.; Eri, J. and Nauer, E.G., In-situ monitoring of the
efficiency of the cathodic protection of reinforced concrete by electrochemical
impedance spectroscopy, Electrochimica Acta, 1996, 41 (7/8), pages 1233 to
1238.
Havdahl, J.; Hope, B. B., and Hansson, C. M. Corrosion Activity Determination
of Epoxy-Coated Steel, 2nd, Latin American Corrosion Conference, 1996, Rio de
Janiero, Brazil, South America, NACE.
Dhouibi, L;, Triki, E.,and Raharinaivo, A., The application of electrochemical
impedance spectroscopy to determine the long-term effectiveness of corrosion
inhibitors for steel in concrete, Cement & Concrete Composites, 2002, 24, pages
35 to 43.
Shi, M.; Chen, Z., and Sun, J., Determination of chloride diffusivity in concrete by
AC impedance spectroscopy, Cement and Concrete Research, 1999, 29 (7), pages
1111 to 1115.
Stratfull, R. F., Corrosion of Steel in a Reinforced Concrete Bridge, Corrosion,
1957, 1 3 (3), pages 43 to 48.
Stratfull, R. F., Preliminary investigation of cathodic protection of a bridge deck,
1973, California Department of Transportation, Sacramento, California, USA.
Stratfull, R. F., Experimental cathodic protection of a bridge deck, 1974,
California Department of Transportation, Sacramento, California, USA.
Polder, R. B., Electrochemical techniques for corrosion protection and
maintenance, in Corrosion in reinforced concrete structures H. Bohni, Editor,
2005, Woodhead publishing Ltd. and CRC Press LLC, pages 215 to 241.
NACE Standard, RP0290-90: Cathodic Protection of Reinforcing Steel in
Atmospherically Exposed Concrete Structures, 1990, NACE standard, Houston,
Texas, USA.
Marcotte, T. D.; Hansson, C. M., and Hope, B.B., The effect of the
electrochemical chloride extraction treatment on steel-reinforced mortar-Part I:

30

96.
97.

98.
99.
100.
101.
102.

103.
104.
105.
106.

107.
108.
109.

Electrochemical measurements, Cement and Concrete Research, 1999, 29, pages


1555 to 1560.
Marcotte, T. D., The Effect of Electrochemical Chloride Extraction on the
Microstructure of Steel-Reinforced Mortar, 1996, Queen's University, Kingston,
Ontario, Canada.
Marcotte, T. D.; Hansson, C. M., and Hope, B.B., The effect of the
electrochemical chloride extraction treatment on steel-reinforced mortar-Part II:
Microstructural characterization, Cement and Concrete Research, 1999, 29,
pages 1561 to 1568.
Miranda, J.M.; Cobo, A.; Otero, E., and Gonzlez , J.A., Limitations and
advantages of electrochemical chloride removal in corroded reinforced concrete
structures, Cement and Concrete Research, accepted 2006.
Buchler, M., Corrosion inhibitors for reinforced concrete, in Corrosion in
reinforced concrete structures, H. Bohni, Editor, 2005, Woodhead publishing Ltd.
and CRC Press LLC pages 190 to 214.
Hansson, C. M.; Mammoliti, L., and Hope, B.B., Corrosion inhibitors in
concrete-Part I: The principles, Cement and Concrete Research, 1998, 28 (12),
pages 1775 to 1781.
Mammoliti, L., Examination of the Mechanism of Corrosion Inhibition by
Ca(NO2)2- and Ca(NO3)2- based Admixtures in Concrete, PhD thesis, University
of Waterloo, Canada, 2001.
Chaussadent, T.; Nobel-Pujol, V.; Farcas, F.; Mabille, I., and Fiaud, C.,
Effectiveness conditions of sodium monofluorophosphate as a corrosion inhibitor
for concrete reinforcements, Cement and Concrete Research, 2006, 36, pages 556
to 561.
Elsener, B., Corrosion Inhibitors for Steel in Concrete-State of the Art, 2001,
Report, European Federation of Corrosion Publication, London, England.
Ngala, V. T.; Page, C. L., and Page, M. M., Corrosion inhibitor systems for
remedial treatment of reinforced concrete, Part 2: sodium monofluorophosphate,
Corrosion Science, 2003, 45, pages 1523 to 1537.
Santagata, M.C. and Collepardi, M., The effect of CMA deicers on concrete
properties, Cement and Concrete Research, 2000, 30, pages 1389 to 1394.
Rangaraju, P. R.; Sompura, K.; Olek, J.; Diamond, S., and Lovell, J. Potential for
Development of Alkali-silica Reaction (ASR) in the Presence of Airfield Deicing
Chemicals, 8th International Conference on Concrete Pavements, 2005, Colorado
Springs, Colorado, USA.
What effects do deicers have on concrete?, Portland Cement Association,
Concrete Technology FAQs, http://www.cement.org/tech/faq_deicers.asp.
De-icers, donnan.com, http://www.donnan.com/de-icers.htm, 2006.
Kirchner, H., Using deicers correctly, 2001, Salt Institute,
http://www.saltinstitute.org/kirchner-1.html.

31

You might also like