Carbonation-Induced and Chloride-Induced Corrosion in Reinforced

Carbonation-Induced and Chloride-Induced Corrosion
in Reinforced Concrete Structures
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Y. Zhou 1; B. Gencturk, A.M.ASCE 2; K. Willam, F.ASCE 3; and A. Attar 4
Abstract: Corrosion is one of the most critical problems that impair the durability of RC structures. Both carbonation-induced and chlorideinduced corrosion widely prevail in civil infrastructure around the globe. Expansive products are formed due to corrosion at the interface
between concrete and reinforcing bar (rebar). The cracking and spalling in concrete due to expanding corrosion products and the reduction in
the cross-sectional area of rebar jeopardize the safety and serviceability of RC structures. From an outsider perspective, this literature review
summarizes the state of the art on the mechanisms of the two types of corrosion, mechanical degradation in RC structures as a result of these
mechanisms, the analytical methods to predict the basic parameters most related to corrosion, and the available laboratory and field corrosion
measurement techniques. DOI: 10.1061/(ASCE)MT.1943-5533.0001209. © 2014 American Society of Civil Engineers.
Author keywords: Corrosion; Reinforced concrete; Chloride; Carbonation; Review.
Introduction
Corrosion widely exists in various types of civil infrastructure.
Particularly, long-term durability of RC structures is threatened by
chloride-induced and carbonation-induced corrosion. In 2012, one
out of nine bridges in the United States was rated as structurally
deficient, while the average age of 607,380 bridges was 42 years
(ASCE 2013). The Federal Highway Administration (FHWA) estimated that to eliminate the nation’s bridge deficient backlog by
2028, $20.5 billion/year would have to be invested. Additionally,
24.9% of the bridges in the United States are marked as either
structurally deficient or functionally obsolete. A highway bridge
is classified as structurally deficient if the deck, superstructure,
substructure, or culvert is rated in poor condition, i.e., 0 to 4 on
the National Bridge Inventory (NBI) rating scale (FHWA 1995).
A bridge can also be classified as structurally deficient if its loadcarrying capacity is significantly below current design standards or
if a waterway below frequently overtops the bridge. Functionally
obsolete is used to define the highway bridges that are not structurally deficient, but whose design is outdated. To develop approaches to service life design beyond 100 years for existing and
new bridges, the Transportation Research Board (TRB) conducted
extensive research and published the “Design Guide for Bridges
for Service Life” (Azizinamini et al. 2013). Despite the fact that
corrosion of the reinforcing steel is not the sole cause of all the
structural deficiencies, it is a significant contributor and has, therefore, become a subject of major concern (Carpenter Technology
1
Postdoctoral Research Fellow, Dept. of Civil and Environmental
Engineering, Univ. of Houston, Houston, TX 77204-4003.
2
Assistant Professor, Dept. of Civil and Environmental Engineering,
Univ. of Houston, N107 Engineering Building 1, Houston, TX 77204-4003
(corresponding author). E-mail: [email protected]
3
Professor, Dept. of Civil and Environmental Engineering, Univ. of
Houston, Houston, TX 77204-4003.
4
Ph.D. Candidate, Dept. of Civil and Environmental Engineering, Univ.
of Houston, Houston, TX 77204-4003.
Note. This manuscript was submitted on June 11, 2014; approved on
September 29, 2014; published online on November 7, 2014. Discussion
period open until April 7, 2015; separate discussions must be submitted for
individual papers. This paper is part of the Journal of Materials in Civil
Engineering, © ASCE, ISSN 0899-1561/04014245(17)/$25.00.
© ASCE
Corporation 2002). In the United States, the annual cost of repair
of bridge deck corrosion due to deicing salts was estimated to be
$50 to $200 million with another $100 million needed for repair
of substructure and other bridge components (Transportation
Research Board 1991). Corrosion is also an important problem
for the nuclear industry (Sindelar et al. 2011). Since 1986, there
have been more than 32 reported occurrences of corrosion in steel
containments or liners of RC containments. Two instances have
been reported in which corrosion has completely penetrated the
liner of the RC containments (Dunn et al. 2011; Petti et al.
2011; Naus and Graves 2013).
Today, the necessity to monitor and mitigate corrosion in order
to maintain and extend the service life of RC structures is widely
accepted in the United States and worldwide. In European countries
and North America, it has become a required practice to limit
the tolerable chloride content to approximately 0.4% by weight of
cement (RILEM 1994). ACI 318 [American Concrete Institute
(ACI) 2011] permits a maximum water soluble chloride ion (Cl− )
content in concrete of 1% by weight of cement. For the most severe
exposure in the splash zone, the Fédération Internationale de la
Précontrainte (FIP) (1996) requires that the water-to-cement ratio
should not exceed 0.45, and preferably be 0.40 or less. A minimum
cement content of 400 kg=m3 should be applied, and ordinary
reinforcement and prestressing tendons should be protected by a
nominal concrete cover of 75 and 100 mm, respectively.
A large body of literature is available on corrosion in RC
structures. However, there is no recent paper that synthesizes the
existing literature on this topic. As presented in Table 1, this review
paper introduces the two main degradation mechanisms associated
with corrosion, discusses different forms of corrosion-induced mechanical degradation, summarizes analytical methods for predicting the basic corrosion mechanisms, and compares and contrasts
various corrosion measurement and detection techniques. Purely
numerical studies were left outside the scope of this paper due to
length limitations.
Mechanisms of Corrosion in RC
Concrete is alkaline in nature with a pore solution pH of
12–13 that naturally passivizes embedded reinforcing bars (rebar).
04014245-1
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Table 1. Main Contents Reviewed and Discussed in the Paper
Mechanical degradation in RC
structures due to corrosion
Analytical methods for predicting
basic corrosion mechanisms in RC
Carbonation-induced corrosion
Cracking and spalling of concrete
Chloride-induced corrosion
Reduction in rebar property
Corrosion as a result of direct
chloride additives into concrete
Loss in interfacial bond strength
Prediction of concrete carbonation
depth
Prediction of chloride penetration
into concrete
Other models for service life
prediction and mechanical
degradation of RC
—
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Mechanisms of corrosion in RC
Effect of binders on the
carbonation-induced and
chloride-induced corrosion of RC
—
The passivation of steel is broken down by the presence of chloride
ions or a reduction in alkalinity of concrete caused by carbonation.
While chloride-induced corrosion is generally more pernicious and
more expensive to repair, carbonation-induced corrosion of rebar
may affect a wider range of RC structures at a larger scale. Both
mechanisms of corrosion may result in significant reduction in
the load-bearing capacity of the structure by reducing the crosssectional area of steel rebar, degrading steel elongation capacity and
causing severe cracking to concrete.
Evans (1960) defined the corrosion types as general corrosion,
localized corrosion, and pitting. Either or both of the microcell
and macrocell corrosion mechanisms may exist in RC structures.
Microcell corrosion occurs when anodic and cathodic half-cell reactions take place at adjacent parts of the same metal. On the contrary, macrocell corrosion takes place when the actively corroding
rebar is coupled to another rebar, which is passive, either because
of its different composition or because of different environment.
Fig. 1 illustrates the microcell and macrocell corrosion mechanisms
in RC. Through a 3-year monitor program, Hansson et al. (2006)
concluded that microcell corrosion is the major mechanism of corrosion in RC.
Carbonation-Induced Corrosion
Carbonation of concrete has started to attract more attention
recently as a result of climate change (Yoon et al. 2007). This type
of corrosion occurs naturally in RC structures at a rather slow yet
Laboratory and field studies to
determine carbonation depths
and chloride penetration
Laboratory and field studies on
concrete carbonation
Methods to determine chloride
penetration in concrete
—
—
invasive rate. The process of carbonation of concrete is presented
in the following form by Leber and Blakey (1956) and Papadakis
et al. (1989):
CaðOHÞ2 → Ca2þ ðaqÞ þ 2OH− ðaqÞ
ð1Þ
Ca2þ ðaqÞ þ 2OH− ðaqÞ þ CO2 → CaCO3 þ H2
ð2Þ
3CaO · 2SiO2 · 3H2 O þ 3CO2 → 3CaCO3 · 2SiO2 · 3H2 O ð3Þ
3CaO · SiO2 þ μH2 O þ 3CO2 → 3CaCO3 þ SiO2 · μH2 O ð4Þ
2CaO · SiO2 þ μH2 O þ 3CO2 → 2CaCO3 þ SiO2 · μH2 O ð5Þ
Carbonation of concrete is the chemical reaction of portlandite,
CaðOHÞ2 , in the cement matrix with carbon dioxide (CO2 ) gas
leading to calcite (CaCO3 ), as shown in Eqs. (1)–(5). Carbonation
takes place as a result of the interaction of carbon dioxide with the
calcium hydroxide in concrete. The carbon dioxide gas dissolves in
water to form carbonic acid (H2 CO3 ), which reacts with calcium
hydroxide and precipitates mainly as calcium carbonate (CaCO3 ),
which lines the pores. Depletion of hydroxyl ions (OH−1 ) lowers
the pore water pH from above 12.5 to below 9.0 where the passive
layer becomes unstable, allowing general corrosion to occur if sufficient oxygen and water are present in the vicinity of the rebar
(Heiyantuduwa et al. 2006).
Fig. 1. Schematic illustrations of corrosion mechanisms [reprinted from Hansson et al. (2006), with permission from Elsevier]: (a) microcell
corrosion; (b) macrocell corrosion
© ASCE
04014245-2
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
The statistical analysis by Das et al. (2012) showed that the
carbonation potential of concrete decreases with an increase in the
compressive strength of the concrete. The results indicated that
using a decrease in the charge passed through concrete as a measure
of carbonation could lead to misleading results in evaluation of the
service life of concrete structures. It was also observed that a low
water-to-cement ratio concrete with portland pozzolana cement has
higher resistance to carbonation and rapid chloride ion permeability
compared with ordinary portland cement.
A variety of interrelated factors influence the carbonation
depth in concrete, including cover thickness, carbonation resistivity, effective CO2 diffusion coefficient, binding capacity for CO2,
curing condition, age, cement type, cement composition, calcium
oxide (CaO) content in cement, surface concentration of carbon
dioxide, time of wetness, ambient temperature, and relative humidity. Environmental conditions, such as sheltered versus exposed and
underground versus atmospheric, also have an important impact on
concrete carbonation process.
with 40°C temperature conditions was seen. It was stated that this
observation could be due to the reduction of oxygen solubility in
the pore solution at high temperature and blockage of concrete
pores at high relative humidity (85%). The gradient of corrosion
profile with temperature was observed to become significantly
steep with the increase in chloride concentration.
As discussed previously, similar to carbonation depth, a variety
of material and environmental related factors influence the chloride
concentration around reinforcement, including cover thickness,
chloride resistivity, and chloride binding capacity of concrete,
water-binder ratio of the concrete mixture, curing conditions, age
of concrete, cement type, cement composition, surface chloride
concentration, ambient temperature, and relative humidity. The
geographic environment of the structure, such as inland zone,
coastal region, or marine, also has a significant impact.
Chloride-Induced Corrosion
Many RC structures are either located in a close vicinity of the
sea or subjected to frequent deicing chemical attack. It has been
recognized since the mid-1970s that the corrosion of the rebar
in concrete is induced by the presence of even small amounts of
chlorides (FHWA 1998). Above a critical chloride concentration,
the passive layer on the rebar is destroyed and the corrosion at the
steel-concrete interface is initiated (Treadaway et al. 1989; Glass
and Buenfeld 1995; Bertolini et al. 1996; Breit 1998; Trejo and
Pillai 2004; Ann and Song 2007; Yu and Hartt 2007; Angst 2011).
As mentioned previously, in European countries as well as in
North America it has become common practice to limit the tolerable chloride content to or around 0.4% by weight of cement
(RILEM 1994).
For countries surrounded by seas, it is convenient to use seawater as concrete mixing water. The seawater, obviously, introduces chlorides into concrete. Fukute et al. (1990) found that the
long-term effects of the use of sea water as mixing water were
negligible in marine environments. Experimental results obtained
after 20 years of exposure showed no long-term effects on concrete strength and corrosion of rebar embedded in concrete. On
the contrary, Wegian (2010) reported that the concrete compressive strength, splitting tensile strength, flexural strength, and pullout bond strength increased for specimens mixed and cured in
seawater at early ages up to 14 days, while a clear decrease was
observed for ages more than 28 days and up to 90 days. No data
were collected after 90 days of casting specimens. The reduction
in strength increased with an increase in seawater exposure time
and this was attributed to salt crystal formation. Islam et al. (2012)
reported that seawater is not suitable for mixing or curing of
concrete. Concrete specimens made and cured with seawater exhibited compressive strength loss of approximately 10% compared
with concrete mixed and cured concrete using tap water. Shigeru
et al. (2012) claimed that in terms of compressive strength, typical
concrete could be replaced with seawater mixed concrete. Concrete compressive strength at 91 days was investigated and it was
found that the service life of structures constructed with seawater
mixed concrete and carbon fiber, epoxy-coated steel, and rustproof coated steel rebar was approximately 100, 70, and 30 years,
respectively.
Table 2 summarizes some of the key studies on the performance of RC with different corrosive additives. By adding corrosive additives into the concrete mixtures, these studies intended to
simulate the accidental addition of chlorides or their long-term penetration from the environment. It is concluded from Table 2 that
the presence of CaCl2 generates a more corrosive environment for
Chloride-induced corrosion is a concern for RC structures that are
located in a marine environment or subjected to deicing chemicals.
The process of chloride-induced corrosion is described by the following reactions as presented by Bentur et al. (1997):
Fe2þ þ 2Cl− → FeCl2
ð6Þ
Fe → Fe2þ þ 2e−
ð7Þ
O2 þ 2H2 O þ 4e− → 4OH−
ð8Þ
Fe2þ þ 2OH− → FeðOHÞ2
ð9Þ
1
2FeðOHÞ2 þ O2 þ H2 O → 2FeðOHÞ3 → Fe2 O3 · 3H2 O ð10Þ
2
The products given by the reactions in Eqs. (7)–(9) combine
together to produce a stable film that passivizes the rebar. The stability of this film depends on the oxygen (O2 ) availability and the
pH of the interstitial solution at the interface between rebar and
concrete (Montemor et al. 2003). As shown in Eqs. (6) and (10),
the passivized film can be disrupted and the corrosion process is
initiated at the presence of sufficient chloride ions (Cl− ), oxygen
(O2 ), and water.
Li et al. (2011) did a salt spray test (5% NaCl solution and temperature equal to 47 3°C) to study the chloride ion penetration in
stressed concrete. For given values of water-to-cement ratio and age
of concrete, chloride content was found to be highest for concrete
stressed in tension, and lowest for concrete stressed in compression.
The chloride content is also affected by the stress level. A higher
level of compressive stress increased the penetration of chlorides
due to formation of transverse microcracks. Resistance of concrete
to chloride ion penetration could be improved by reducing the
water-to-cement ratio. Wet and dry cycles and high temperature
also stimulate chloride penetration rate in concrete (Castel et al.
2000a; Alhozaimy et al. 2012). It was found that the cement
deterioration due to sulphate ions is retarded in the presence of
chlorides (Hossain 2006). Temperature level had no impact on
the chloride binding capacity of concrete during immersion tests
in NaCl solution (Hossain 2006). Alhozaimy et al. (2012) tested
chloride-induced corrosion under three different temperature conditions (30, 40, and 50°C) and 85% relative humidity. A decrease in
corrosion potential and corrosion mass loss at 50°C in comparison
© ASCE
Corrosion As a Result of Direct Chloride Additives into
Concrete
04014245-3
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
© ASCE
04014245-4
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
0.56
0.1, 0.5, 1, 2% by weight of
water
0.5, 1.5 mol=kg cement; 0.25,
0.75 mol=kg cement
1, 2, 3% by weight of water
33% by weight of water
NaCl
NaCl
CaCl2
Potassium acetate
Agricultural product
NaCl
CaCl2
MgCl2
NaCl
CaCl2
MgCl2
Calcium magnesium
acetate (CMA)
NaCl
NaCl
CaCl2
MgCl2
Noninhibited NaCl
NaCl
—
3, 6, 20% by weight of water
10, 20, 30% by weight of water
1, 5, 10% by weight of cement
0.40
0.46
0.39 and 0.38
0.33
15-25% by weight of water
3% by weight of water
0.45
3, 15% by weight of water
0.48
0.53
0.69
0.5
0.5
1, 2% by weight of cement
NaCl
CaCl2
Corrosion inhibitor added
deicing salts (CIADS)
NaCl
NaCl
CaCl2
NaCl
CaCl2
0.5
Water-to-cement
ratio
0.5, 2% by weight of cement
Dosage
NaCl
CaCl2
Additive
Table 2. Rebar Corrosion in the Presence of Corrosive Additives
NaCl was the most chemically benign of the principal deicers but it has the highest
sorptivity rate. MgCl2 reaction was temperature and concentration dependent. Effect
of CMA is not noticeable on concrete or mortar. Exposure of concrete and mortar to
NaCl results in little to no chemical interaction or related distress. Exposure of
concrete and mortar to MgCl2 , CaCl2 and deicing chemicals based on these salts
results in significant chemical interaction and related distress. If MgCl2 -based and
CaCl2 -based deicing chemicals were to be used, they should be used at the lowest
possible concentration
Noninhibited NaCl additive resulted in the most severe corrosion in RC and the lowest
compressive strength in concrete. While the corrosion inhibitors in deicer products
provide some benefits in delaying the chloride ingress and subsequent corrosion
initiation, such benefits seem to diminish once the corrosion of the rebar is initiated
Loss of stiffness and strength developed rapidly at an accelerated rate for 10% NaCl.
Cracking behavior of beams with 1% NaCl at 18 months were concentrated in the pure
bending region of the beam and they formed over a narrow space; on the other hand,
the cracking of beams with 5 and 10% NaCl were more invasive, distributed over the
pure bending region, and extended outside at an inclined angle. The 10% NaCl
addition resulted in a loss of structural integrity and stiffness of tested beams, and the
loss was approximately 30% at 18 months age
For 30% NaCl, salt solution corrosion starts in less than a day
NaCl and MgCl2 solution can cause measurable damage to concrete at low and high
concentrations, 3 and 15% solutions, respectively, within 95 weeks of exposure
Corrosion of steel in concrete was more severe in the presence of CaCl2 than NaCl.
The 2% case showed the highest corrosion
Corrosivity was primarily tested by measurements of electrical resistivity and acid
capacity. NaCl as chloride source yielded a higher pH value and a lower ðCl− Þ=ðOH− Þ
ratio (less corrosive) in comparison with CaCl2 . Therefore, the corrosive effect of
NaCl added to fresh cement mortar was significantly less than that of CaCl2
The corrosion rate of rebar increased with increasing NaCl content
in the mix water. However, the corrosion rate of steel in the samples decreased
with the addition of silica fume
Compared with NaCl and several other deicing chemicals, CaCl2 solutions with and
without corrosion inhibitor displayed the most severe damage to concrete under both
wetting-drying and freezing-thawing conditions
Concentration of chloride ions dropped rapidly but reached an equilibrium level
within approximately 10 days. Measured values of CI− =OH− ratios in the equilibrium
pore solutions increased with increasing dosage. NaCl incorporated as an admixture
resulted in a response similar to that for CaCl2
Ratio of Cl− to OH− increased rapidly with salt addition from 1 to 2%. Corrosion of
rebar was more serious in concrete containing CaCl2 than that containing NaCl
Compared with CIADS, NaCl created a more aggressive environment for rebar
Major findings
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
References
Anacta (2013)
El-Salam et al.
(2012)
Shi et al. (2010)
Sutter et al.
(2008)
Darwin et al.
(2007)
Wang et al.
(2006)
Kelestemur and
Yildiz (2006)
Pakshir and
Esmaili (1998)
Pruckner and
Gjørv (2004)
Helmuth et al.
(1993)
Jang et al. (1995)
Diamond (1986)
rebar inside the concrete compared with NaCl. Corrosion rate
increases with the increase of chloride concentration, but it
can be retarded by the presence of another additive such as silica
fume.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Effect of Binders on the Carbonation-Induced and
Chloride-Induced Corrosion of RC
Different types of binders have been used in concrete to increase its
resistance to corrosion. The most commonly used binders are pulverized fuel ash, granulated blast furnace slag, and silica fume. The
introduction of these binders affects the carbonation and chloride
intrusion processes. Generally, the inclusion of these binders stimulates the electrical resistivity in concrete, which is usually used as
an indicator of corrosion rate. The following threshold values are
recommended to correlate the electrical resistivity and the likelihood of corrosion: if resistivity ≥120 Ω · m, corrosion is unlikely;
if resistivity = 80 to 120 Ω · m, corrosion is possible; if resistivity
≤80 Ω · m, corrosion is fairly certain (Broomfield 1997). These
threshold values should be used with caution because chloride diffusion and surface electrical resistivity are also dependent on other
factors such as temperature, humidity in concrete pore structure,
mix composition, age, and whether the samples are sealed or not
if laboratory testing is conducted (Kessler et al. 2008; Spragg et al.
2013). The electrical resistivity of concrete increases as the waterto-binder ratio decreases and binder content increases. Based on the
information available in the literature, Fig. 2 summarizes the effects
three binders mentioned previously as well as the effects of waterto-binder ratio and increase in binder content on carbonation and
chloride penetration.
Mechanical Degradation in RC Structures due to
Corrosion
Corrosion, induced by both carbonation and chloride, causes degradation in the mechanical properties of RC structures. While the
corrosion damage was found not to significantly impact the ductility of RC members (Stanish 1997), it results in concrete spalling
and cracking, reduction in rebar properties, and interfacial bond
loss, which are discussed in detail in the following.
Cracking and Spalling of Concrete
Cracking of concrete occurs because corrosion products (oxides)
have a higher volume than basic metal. These cracks reduce the
load-bearing capacity, shorten the service life, and increase the rate
of ingress of aggressive elements (Alonso et al. 1998). Castel et al.
(2000a) conducted three-point bending tests on 3,000-mm-long,
14-year-old, salt fog–aged beam specimens. Crack widths were
monitored and concrete cores were examined (after the flexural
test) for compressive strength and elastic modulus. It was found
that cracking in the compression region had no effect on the flexural behavior of the corroded beams; however, significant changes
to the service behavior were observed due to degradations in the
tensile zone, namely, loss of bending stiffness and asymmetrical
behavior (due to asymmetrical distribution of the cracking pattern).
Castel et al. (2000b) noted that rebar area reduction and bond
strength loss are coupled. Simulation of pitting corrosion showed
that a local steel cross-section reduction, located between flexural
transverse cracks, has no influence on the global flexural behavior
when the bond strength is not modified. However, the global flexural behavior of the beams seems to be greatly affected when the
rebar area reduction and bond strength loss are coupled because of
the local increase in the steel stresses due to the reduction in the
steel cross section and due to no contribution of concrete to tensile
resistance.
Webster (2000) reported that the amount of corrosion observed
to cause cracking is typically an order of magnitude greater than the
amount of radial expansion theoretically required to induce cracking. Additionally, ignoring the effects of concrete spalling led to
an overestimation of the flexural capacity. Opening of the first longitudinal crack and complete breakdown of the bond had a direct
relationship with the concrete cover-to-rebar diameter ratio. By
studying the concrete stress contribution versus steel reinforcement
average strain, it was found that the tension stiffening in RC is also
sensitive to the degree of reinforcement corrosion (Shayanfar
et al. 2007).
Chloride penetration
Carbonation
Change in
mixture
composition
diffusion
Binding
capacity
Overall effect
on
carbonation
Chloride
diffusion
Chloride
migration
Critical
chloride
content
Concrete
electrical
resistivity
(relates to
corrosion rate)
Inclusion of
pulverized fuel ash
Inclusion of
ground granulated
blast furnace slag
Inclusion of
silica fume
Reduction of
water-binder ratio
Increase in
binder content
: increase;
: decrease.
Related References: Jaegermann (1990), Al-Amoudi (1995), Polder (1996), Dhir and Jones (1999), Leng et al.
(2000), Chi et al. (2002), Otsuki et al. (2003), Collepardi et al. (2004), Richardson (2006), Hooton et al. (2010),
Eguchi et al. (2013), Turk et al. (2013).
Fig. 2. Effects of binders on carbonation and chloride ingress
© ASCE
04014245-5
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Loss in Interfacial Bond Strength
Use of groups of rebar (bundled rebar) to replace a rebar of
larger diameter (by conserving total area) was not recommended
if there is a chance of corrosion because bundled rebar generate
higher stresses over the concrete due to the larger perimeter, thus
a larger surface susceptible to corrosion (Ortega et al. 2011).
Corrosion significantly influences the interface bond behavior
between concrete and rebar. It was found that the bond
strength first increases and then decreases with increasing level
of corrosion (Al-Sulaimani et al. 1990; Jin and Zhao 2001; Bajaj
2012).
Al-Sulaimani et al. (1990) observed a sharp jump in the value
of the free-end slip with the opening of a longitudinal crack indicating a sudden loss of rebar confinement in the rebar pullout tests.
It was found by Almusallam et al. (1996) that in the precracking
stage (0–4% corrosion, measured as gravimetric loss in weight of
rebar) the ultimate bond stress increases, whereas the slip at the
ultimate bond stress decreases with increasing corrosion level. The
degradation of bond results from the crushing of the concrete near
the lugs of the rebar. When reinforcement corrosion is in the range
of 4–6%, the bond failure occurs suddenly at a very low free-end
slip. At this level of corrosion, a large slip was noted as the ultimate
failure of the bond, occurring due to the splitting of the specimens.
Beyond 6% rebar corrosion, the bond failure resulted from a continuous slippage of the rebar. The ultimate bond stress initially
increased with an increase in the degree of corrosion until the
corrosion reached a maximum value of 4%, after which there was
a sharp reduction in the ultimate bond stress up to 6% rebar corrosion. Beyond the 6% rebar corrosion level, the ultimate bond
stress did not vary much, even up to 80% corrosion. Bajaj (2012)
confirmed that the bond strength increases with an increase in the
corrosion level up to a critical percentage (2% for plain concrete,
Rebar area loss is a direct result of corrosion and it is usually quantified in terms of mass loss. It was found that the yield and ultimate
stress and strain at ultimate stress of rebar deteriorated, and the
yield plateau became narrower or even disappeared with the development of corrosion (Webster 2000; Zhang et al. 2006). These reductions could lead to a premature fracture of the rebar before
yielding is observed. The inelastic constitutive behavior of corroded rebar was investigated by Kashani et al. (2013). It was found
that a corrosion level above 15% mass loss significantly affected
the ductility and plastic deformation of rebar in tension. Additionally, corrosion changed the buckling mechanism of the rebar in
compression from classical inelastic buckling to inelastic buckling
with unsymmetrical plastic hinges or inelastic buckling with multiple intermediate plastic hinges. A 10% mass loss was observed to
produce approximately a 20% reduction in the buckling capacity
of the corroded rebar. The distribution of the corrosion pits along
the length of corroded rebar was found to be the most important
parameter affecting the stress-strain response in both tension and
compression.
120
40
0%
4%
9%
100
0%
3.3%
6.8%
35
30
Load (kN)
Load (kN)
80
60
40
25
20
15
10
20
0
5
0.0
0.2
0.4
0.6
0.8
0
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
1.0
Slip (mm)
(a)
Slip (mm)
(b)
90
120
80
100
70
80
60
60
0%
3.8%
6.0%
40
Load (kN)
Load (kN)
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Reduction in Rebar Property
50
30
20
20
10
0
0
0.0
(c)
3.0%
4.7%
5.2%
40
0.2
0.4
0.6
0.8
1.0
Slip (mm)
1.2
1.4
1.6
(d)
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Slip (mm)
Fig. 3. Relationship between load and slip [reprinted from Fang et al. (2004), with permission from Elsevier]: (a) deformed rebar in unconfined
concrete; (b) smooth rebar in unconfined concrete; (c) deformed rebar in confined concrete; (d) smooth rebar in confined concrete
© ASCE
04014245-6
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
3.5% for concrete with polypropylene fiber as an additive, and
4.5% for concrete with basalt fiber as an additive) and then
decreases. The bond strength of corroded rebar increased with increasing concrete cover-to-rebar diameter ratio and concrete tensile
strength. Additionally, the bond strength of corroded, ribbed rebar
was enhanced due to the presence of extruded ribs that act as
shear links.
Fang et al. (2004) tested a series of specimens having 0–9%
corrosion with and without stirrups. Bond strength was found to
be very sensitive to the corrosion level and generally decreased with
increasing corrosion level for deformed rebar in unconfined concrete. Corrosion had no substantial influence on the bond strength
of deformed rebar in confined concrete. For smooth rebar in unconfined concrete when the corrosion level was low, bond strength
increased as corrosion level increased, with the ultimate bond
strength being as much as 2.5 times that of noncorroded rebar;
while the bond strength decreased rapidly at higher corrosion
levels. The break point was found to be approximately 2–4% corrosion. For smooth rebar in confined concrete, bond strength
increased as corrosion level increased, up to a relatively high degree
of corrosion. The increase in bond strength could even be observed
at a corrosion level of more than 5%. Fig. 3 illustrates the relations
between load and slip for deformed and smooth rebar in unconfined and confined concretes at different corrosion levels (Fang
et al. 2004).
Factor
or
Fig. 4 summarizes the overall behavior of RC affected
by the factors discussed in this section. Shaded cells are used in
Fig. 4 if there is no relationship between the two variables or
if the relationship is not pertinent to the discussions presented
here.
Analytical Methods for Predicting Basic Corrosion
Mechanisms in RC
As mentioned previously, the two most important causes for corrosion of RC are chloride attack due to deicing salts and seawater,
and the carbonation of concrete due to carbonic acid from carbon
dioxide. Several studies presented analytical models to predict
the main parameters in these two types of corrosion mechanisms.
The following sections present some of the key analytical contributions to predict basic parameters related to corrosion including
the carbonation depth, chloride penetration rate, crack initiation
time, crack width, and other parameters related to RC mechanical
degradation.
Prediction of Concrete Carbonation Depth
A simple equation for carbonation rate was proposed by Comité
Euro-International du Béton (CEB) (1997) based on the CO2 diffusion model. By considering the effect of microclimatic condition
ELm
Reference
or
Villagrán-Zaccardi et al. (2008), Li et al.
(2011)
w/c
W-D
Costa and Appleton (1999)
F-T
Wang et al. (2006)
Tstress
Li et al. (2011)
Cstress
Li et al. (2011)
Temp
Alhozaimy et al. (2012)
ccp
Das et al. (2012)
Corr
Al-Sulaimani et al. (1990), Castel et al.
(2000a, b), Webster (2000), Jin and Zhao
(2001), Li (2003), Fang et al. (2004),
Zhang and Lounis (2006), Bajaj (2012)
Webster (2000)
Webster (2000)
Zhang and Ba (2001)
Wegian (2010)
Notes:
: concrete compressive strength;
: concrete crack width;
: rebar yield strength;
: rebar ultimate strength;
: rebar ultimate strain;
: rebar elongation at maximum load;
: interface bond strength;
: mass loss;
: flexural member stiffness;
: flexural member strength;
: chloride penetration rate;
: seawater;
w/c: water cement ratio;
W-D: wet and dry cycles;
F-T: freeze-thaw cycles;
Tstress: stressed in tension;
Cstress: stressed in compression;
Temp: temperature;
ccp: concrete carbonation potential;
Corr: corrosion;
: concrete tensile strength;
: concrete cover to rebar diameter ratio;
: chloride concentration;
: equivalent apparent chloride diffusion coefficient;
: first increase and then decrease;
: increase;
: decrease;
shaded cell: factor influence is not closely related to the scope of this review paper on corrosion.
Fig. 4. Factors affecting the RC behavior
© ASCE
04014245-7
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Fig. 5. Analytical models to predict carbonation depth in concrete
on carbonation, the equation was modified by Yoon et al. (2007).
Valcuende and Parra (2010) proposed an estimation for carbonation
rate in concrete using the volume of pores that are more than
0.065 μm in size and the threshold diameter at which continuous
mercury intrudes rapidly. Fig. 5 summarizes these equations to predict carbonation depth in concrete.
Prediction of Chloride Penetration into Concrete
Capillary absorption, hydrostatic pressure, and diffusion are the
means by which chloride ions penetrate into concrete. Among
these, diffusion is the principal one. However, it is important to note
that other transport mechanisms might overwhelm diffusion under
certain circumstances, especially in the surficial convection zone of
the concrete cover (McCarter et al. 1992; Neville 1995; Broomfield
1997; Medeiros and Helene 2009). A good example is the rapid
initial absorption of chloride-laden deicing compounds when they
are applied on the very dry bridge deck concrete in expectation of an
upcoming icing or snow. If the cover or concrete quality is low, and
if wet and dry cycles occur, then other transport mechanisms such as
absorption and capillary action might overwhelm diffusion at least
to rebar depth. Concrete must have a continuous liquid phase and
there must be a chloride ion concentration gradient for diffusion to
occur. In most studies, the chloride concentration inside the concrete was predicted based on Fick’s second law of diffusion. Fig. 6
summarizes the equations proposed to predict the chloride penetration into concrete through either diffusion or absorption.
Costa and Appleton’s original equation (Costa and Appleton
1999b) is only valid for constant diffusion coefficients and surface
chloride contents; however, these parameters are strongly time dependent. Xi et al. (1999) developed mathematical and numerical
models for binding capacity and chloride diffusivity, which have
a dominant effect on the chloride diffusion process. Bioubakhsh
(2011) reported that curing time and conditioning temperature have
significant effects on sorptivity and chloride penetration at early
ages, but this effect reduces with increasing number of wet-dry
cycles. Song et al. (2006) reported that the diffusion coefficient is
time dependent because the process of cement hydration results
in connection and condensation of concrete pore structures. Chloride
diffusion coefficients vary significantly with the concrete quality and
© ASCE
exposure conditions. Particularly, the water-to-cement ratio has a major impact on the pore structure of the concrete, and thereby on the
diffusivity of concrete. Larger water-to-cement ratios tend to create a
more porous concrete with higher coefficients of diffusion.
The surface chloride content is considerably affected by the
exposure conditions, reducing from a tidal to an atmospheric zone.
In a tidal zone (having wet and dry cycles) penetration rates were
found to be high (Costa and Appleton 1999a). The specific coefficients and parameters for the modified chloride penetration model,
which are dependent on different exposure conditions, can be found
in Costa and Appleton (1999a, b). It is noteworthy that there is a
disagreement among researchers on the selection of realistic values
for critical chloride content, Ccr (Treadaway et al. 1989; Glass and
Buenfeld 1995; Bertolini et al. 1996; Breit 1998; Trejo and Pillai
2004; Ann and Song 2007; Yu and Hartt 2007; Angst 2011).
Other Models for Service Life Prediction and
Mechanical Degradation of RC
Corrosion has a direct impact on the serviceability of RC structures.
Although there is a large body of literature on this issue, more
efforts were spent on numerical simulations and probabilistic assessment of service life, while few studies were accompanied by
experimental verification and calibration.
RC flexural members were found to deteriorate at different rates
with stiffness deteriorating faster than strength (Li 2003). Life-365
(Bentz and Thomas 2013), which is a model developed for predicting
the service life and life-cycle cost of RC exposed to chlorides, assumes that initially the diffusion coefficient decreases with increasing time of exposure, but remains almost constant after a period of
time. Zhang and Ba (2011) used the time to reach a critical chloride
concentration, which could lead to rebar corrosion, as an indicator
for the service life of concrete. Accordingly, the predicted service life
for structures with 10-mm concrete cover in the chloride environment was approximately 12 years. Vu and Stewart (2000) used a
reliability-based deterioration model to calculate the probabilities
of structural failure. They found that the application of deicing salts
causes significant long-term deterioration and reduction in structural
safety for poor durability design. A reduced cover or increased waterto-cement ratio increases failure probabilities. Spatial variability was
04014245-8
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Fig. 6. Analytical models to predict chloride penetration rate, chloride diffusion coefficient, and water capillary absorption in concrete
found to be very important in the reliability analysis for deteriorating
structures, particularly for corroding RC beams in flexure (Stewar
and Al-Harthy 2008). The maximum corrosion loss in a rebar conditional on beam collapse was no more than 16%. The probabilities of
failure considering spatial variability of pitting corrosion were up to
200% higher than the probabilities of failure obtained from a nonspatial analysis after 50 years of corrosion.
Fig. 7 summarizes analytical models that have been proposed
to predict various important parameters related to service life and
mechanical degradation of RC structures. Some of these models
have also been reviewed and summarized by Ahmad (2003).
Bazant’s model (1979) and Morinaga’s model (1988) were
based on the steady-state corrosion process and both models
underestimate the time to cracking of concrete. Corrosion can
© ASCE
be described as a stochastic process, and this underestimation
was attributed to the relatively low density used for rust, a simple
linear function to describe the relationship between growth of rust
products and time, and not accounting for the environmental factors
(e.g., temperature, relative humidity, rainfall) (Liu and Weyers
1998). Wang and Zhao’s model (1993) can only be used in conjunction with a finite-element model to determine the thickness
of corrosion products. Val and Chernin (2012) developed an
alternative approach to account for the diffusion of corrosion products into concrete and recommended the use of 0.70 as a tentative
estimate for the ratio between the amount of corrosion products
diffused into concrete and the total amount of corrosion products
formed since the time of crack initiation. Melchers (2003) showed
a schematic representation of the changing phase in immersion
04014245-9
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
relationship with the electrochemical potential. Papakonstantinou
and Shinozuka (2013) developed a probabilistic model for predicting the corrosion of steel considering the effect of concrete cracking
on the corrosion process itself.
Laboratory and Field Studies to Determine
Carbonation Depths and Chloride Penetration
Laboratory and Field Studies on Concrete Carbonation
It usually requires special chambers in the laboratory to provide
a certain atmospheric concentration of CO2 to accelerate the
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
corrosion loss as a function of exposure time. Bossio et al. (2011)
proposed analytical models, which are based on a nonlinear system
of equations of equilibrium (rebar-oxide-concrete) and compatibility (boundaries of the layers), regarding corrosion effects. Zhang
and Ba (2011) reported that the experimental value is in agreement
with the Life-365 model (Bentz and Thomas 2013) prediction of
the initiation period, which is defined as the time for chlorides
to penetrate from the external environment through the concrete
cover, accumulate at the embedded steel in sufficient quantity to
break down the protective passive layer on the steel, and thereby
initiate an active state of corrosion. Additionally, the negative
logarithm of chloride concentration was shown to follow a linear
Fig. 7. Analytical models to predict parameters related to service life and mechanical degradation of RC
© ASCE
04014245-10
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Fig. 7. (Continued.)
concrete carbonation. The gas diffusion test (Jung et al. 2010) is
used to measure CO2 diffusivity in concrete following ASTM
E104 (ASTM 2012). Temperature and relative humidity have significant influence on the carbonation rate (De Ceukelaire and
Van Nieuwenburg 1993; Yoon et al. 2007; Matsuzawa et al.
2010; Valcuende and Parra 2010; Benzarti et al. 2011; Rostami et al.
2011; Chen and Ho 2013). It was observed that corrosion occurs
when the distance between carbonation front and rebar surface (the
uncarbonated depth) is less than 5 mm (Yoon et al. 2007). Table 3
summarizes the major findings regarding concrete carbonation tests
in the laboratory.
Various surface treatments have also been tried to reduce the rate
of carbonation. Huang et al. (2012) investigated the effect of plastering on the carbonation of a 35-year-old RC building. A significant reduction in carbonation was discovered for the columns and
beams of the building that was plastered. There was no carbonation
of the concrete when the plaster thickness exceeded 50 mm.
Surface coatings, such as tile with high compact and impermeable
material, clearly delayed the carbonation of concrete. Aguiar and
Junior (2013) reported that the use of epoxy resin on concrete surface showed better protection than the use of acrylic and siloxane
resins. It was reported by Han et al. (2013) that the influence of
carbonation on the durability of harbors is not as significant as
© ASCE
for other structural types such as buildings or subway tunnels. They
evaluated the effect of carbonation in harbors and quantified its influence based on in situ tests performed at 436 locations in 80 harbors. The results of these in situ tests showed that the ratio of
carbonation-to-cover depth was less than 0.2 for harbors. In most
cases, the probability of failure (defined as the initiation of corrosion) due to carbonation was less than 10%.
Contrary to the detrimental carbonation process for mature concrete, combined steam and CO2 curing at an early stage is favorable
for enhancing the durability of concrete (Rostami et al. 2011).
Carbonated concretes also exhibit improved resistance to sulfate
attack, water absorption, and chloride ion penetration.
Methods to Determine Chloride Penetration in
Concrete
In most cases, diffusion is the main process for chloride ingress;
however, as mentioned previously, other transport mechanisms
such as absorption might overwhelm diffusion in certain cases. Diffusion requires that the concrete has a continuous liquid phase and a
chloride ion concentration gradient. The rate of ingress of chlorides
depends on the pore structure of the concrete, which is affected by
factors including materials ingredients, construction practices, and
04014245-11
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Table 3. Summary of Laboratory Studies on Carbonation of Concrete
Temperature
(°C)
Relative humidity
(RH) (%)
CO2
Major findings
References
35, 45, 60, 80
5%
60.8
349.6 ppm
Matsuzawa et al.
(2010)
Valcuende and
Parra (2010)
27
65
10%
25
70, 90, 70–90,
50–90
50%
25
70
20%
20
66
50%
The lower the RH, the higher the carbonation rate at 60°C.
Concrete carbonation rate is strongly affected by RH
The difference between self-compacting concrete and normally
vibrated concrete tends to disappear as their fines content
becomes similar. Under the test conditions, for pore sizes under
0.065 μm CO2 diffusion in the interior of the concrete is not
significant
Carbonation potential of concrete decreases with an increase in
the compressive strength of the concrete. A significant decrease
in charge passed through concrete owing to carbonation can
result in misleading results in evaluation of the service life of the
concrete structures. A low water-cement ratio concrete with
portland pozzolana cement has discernible resistance to
carbonation and rapid chloride ion permeability compared with
its counterpart ordinary portland cement
Carbonation of concrete is influenced by the ambient humidity,
humidity cycling, and surface geometry of concrete. The largest
coefficient of carbonation occurred at a constant 70% relative
humidity, and the coefficient of carbonation decreases with
humidity cycling
Carbonation depth of the samples containing Grade II fly ash
and steel slag powder are higher than that of the reference sample
while the pH of the pore solution is lower. When cement was
replaced by a large amount of other types of binders (Grade II fly
ash, pulverized fly ash, granulated blast furnace slag, and steel
slag), the carbonation of lightweight aggregate concrete
accelerated with the increase of binder content
Carbonation depth was found to be a decreasing function of the
compressive strength at 28 days and increases with the initial
water absorption
20, 60
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
20.2
age. Stanish et al. (1997) summarized and compared a family of
methods to measure chloride penetration. Based on the work by
Stanish et al. (1997), Table 4 is produced here to summarize the
existing methods to determine chloride penetration.
Other than the methods listed in Table 4, there lie a variety
of electrochemical and nondestructive methods that can be used
in the field to determine in situ corrosion rate for RC structures.
These methods include open circuit potential (OCP) measurements,
surface potential measurements, Wenner array probe (a resistivity
test), linear polarization resistance (LPR) measurement, galvanostatic pulse transient method, electrochemical impedance spectroscopy (EIS), harmonic analysis, noise analysis, embeddable
corrosion monitoring sensor, ultrasonic pulse velocity technique,
X-ray, gamma radiography measurement, and infrared thermograph electrochemical method.
Although various methods are available to detect chloride penetration in concrete, there is no consensus as to which test gives the
most reliable results. Bagheri and Zanganeh (2012) compared rapid
tests for evaluation of chloride resistance of concretes with supplementary cementitious materials. Chloride penetration, rapid chloride migration, and electrical resistance tests were compared.
The results of all three methods showed a considerable decrease
in chloride permeability of mixes containing silica fume and other
supplementary cementitious materials at 28 and 90 days, and at
90 days only, respectively. However, the rapid chloride permeability
test (RCPT) considerably overestimated the improvement compared with the other two methods, mainly as a result of the temperature rise effect in this test. Despite the simplicity and speed of the
electrical resistance test, its results correlate well with those of the
rapid chloride migration test (CTH). Others have proposed the use
© ASCE
Das et al. (2012)
Chen et al. (2013)
Gao et al. (2013)
Rabehi et al.
(2013)
of the formation factor, which is defined as the ratio of the pore
solution conductivity to the bulk conductivity (solid microstructure
and pore solution together), as a measure of microstructural diffusion coefficient (Tumidajski and Schumacher 1996; Snyder 2001).
The advantage of formation factor lies in the fact that it is a material
property, and unlike resistivity measurements it is not affected by
environmental conditions such as temperature and humidity. The
bulk diffusion test [Nordtest NT Build 443 (Nordtest 1995) and
ASTM C1556 (ASTM 2012)] was considered to be a better
long-term method than the salt ponding test [AASHTO T259
(AASHTO 2006) and ASTM C1543 (ASTM 2012)].
Conclusions
The following main conclusions are drawn from this literature
review:
• CaCl2 generates a more corrosive environment in the concrete
mixture compared with NaCl. Corrosion potential increases
with the increase of chloride concentration, but can be retarded
by the presence of other additive, such as silica fume. Inclusion
of pulverized fuel ash, ground granulated blast furnace slag, and
silica fume increases concrete electrical resistivity. If seawater is
used as mixing water, concrete compressive strength, splitting
tensile strength, flexural strength, and pullout bond strength are
expected to increase at early ages up to 14 days and decrease for
ages more than 28 days.
• Corrosion results in deteriorating rebar area, yield strength,
ultimate strength, and ultimate strain. Concrete crack width increases with an increase in corrosion level. Mass loss is higher
under freeze-thaw cycles and high temperatures. Increase in
04014245-12
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
© ASCE
04014245-13
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Propan-2-ol
counterdiffusion
Gas diffusion
14 days with thin
paste sample
2–3 h
30 min
1 week including
conditioning
Depends on pressure
and concrete
Pressure
penetration
Sorptivity
L/ND
F/ND
30 min
L/ND
L/ND
F/ND
L/ND
L/ND
L/D
L/ND
L/ND
8h
Depends on voltage
and concrete
L/D
40–120 days after
curing and
conditioning
6h
L/ND
L/D
90 days after curing
and conditioning
Test
condition
Rapid migration
test (chloride
migration
test)
Resistivity
Electrical
migration
AASHTO T259,
ASTM C1543
(salt ponding)
Nordtest, ASTM
C1556 (bulk
diffusion)
AASHTO T277
rapid chloride
permeability test
(RCPT), ASTM
C1202
Duration of test
procedure
Pros
Diffusion coefficient obtained comparable
to chloride ponding test
Measures capillary forces exerted by the
pore structure; useful if steel is very
shallow
Relatively easy to conduct
Chloride penetration profile can be
obtained
Relatively easy to conduct
Relatively easy to conduct; measures
capillary forces exerted by the pore
structure
Low voltage applied for a short time;
addresses temperature rise
Addresses temperature rise; considers
chloride migration
Addresses temperature rise; considers
chloride migration
Chloride penetration profile can be
obtained; capable of modeling chloride
diffusion
Relatively easy to conduct
Chloride penetration profile can be
obtained
Cons
Only thin cement paste can be tested; chloride
binding effect is ignored
Difficult to adequately seal the sides of concrete;
mathematics involved in determining useful values
are complex; chloride binding effect is ignored
Only chloride penetration depth can be obtained
Highly dependent on the moisture content of the
sample in a field test; evaluates only concrete surface
and cannot give any information on bulk properties;
practicing engineers are less interested in sorptivity
Requires chloride profile grinding
Difficult to determine the conductivity of the
pore solution
Conducting material present in the concrete sample
will bias the results
Current passed is related to all ions in the pore
solution not just chloride ions; measurements are
taken before steady-state migration is achieved; the
high voltage applied leads to an increase in
temperature; any conducting material present in the
concrete sample will bias the results, causing them
to be too high
Conducting material present in the concrete sample
will bias the results
Overemphasizes the importance of sorption, and to
a lesser extent wicking; takes longer if high-quality
concrete is tested
Takes longer if high-quality concrete is tested
Standard and reference
Sharif et al. (1997)
Feldman (1987)
Deutsches Institu für Normung
(DIN) (1987), Hall (1989), Martys
and Ferraris (1997), Medeiros and
Helene (2009), British Standards
Institute (BSI) (2011), ASTM
(2013)
DeSouza et al. (1995)
Monfore (1968), Andrade et al.
(1993), Streicher and Alexander
(1995), Morris et al. (1996)
Valenta (1969), Freeze and Cherry
(1979)
Detwiler et al. (1991), Zhang and
Gjorv (1991), Andrade (1993),
Andrade and Sanjuan (1994),
Streicher and Alexander (1995),
Delagrave et al. (1996), Jacobsen
et al. (1996), McGrath (1996),
McGrath and Hooton (1996), Lu
(1997)
Collepardi et al. (1970), Tang and
Nilsson (1991), Otsuki et al. (1992)
Whiting (1981), Samaha and Hover
(1992), Saito and Ishimori (1995),
Goodspeed et al. (1996), Thomas
and Jones (1996), AASHTO
(2005), ASTM (2012)
NT Build 443 (Nordtest 1995),
ASTM (2012)
AASHTO (2006),
ASTM (2012)
Note: D = destructive test; F = field test; L = laboratory test; ND = nondestructive test; wicking = vapor transmission from the wet front in the concrete to the drier atmosphere at the external face, causing more water
and chloride ions along with it to be drawn into the concrete.
Others
Short term
Long term
Test method
Table 4. Summary of Test Methods to Determine Chloride Penetration
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
•
•
•
•
•
water-cement ratio, wet and dry cycles, and tensile stresses accelerate chloride penetration rate, while compressive stresses
have the reverse effect. Use of rebar bundles to replace a larger
diameter rebar is not recommended. Bond strength increases
with increase in corrosion level up to a critical percentage (approximately 2–4% of mass loss) and then decreases. The bond
strength of corroded rebar increases with increasing concrete
cover-to-rebar diameter ratio and concrete tensile strength.
Carbonation depth of concrete can be estimated by the classical
(CEB 1997) and modified (Yoon et al. 2007) diffusion models.
Composition of the concrete is an important factor affecting the
diffusion of CO2 . Carbonation rate is also strongly affected by
humidity. Surface tiles with high compact and impermeable
material may clearly delay the carbonation of concrete.
Chloride concentration inside the concrete can be calculated
by diffusion models (Costa and Appleton 1999b; Roa-Rodriguez
et al. 2013). Different relationships have been proposed to take
into account the time dependence of diffusion coefficient for
concretes under various conditions (Leber and Blakey 1956;
Kwon et al. 2009; Zhang et al. 2011; Bentz and Thomas
2013; Roa-Rodriguez et al. 2013). One should consider the
variations in the diffusion coefficient due to cracking of concrete
and changing water-to-cement ratio.
Other transport mechanisms such as absorption and capillary
action might overwhelm the diffusion to rebar depth, especially
if the cover is low, the concrete is subjected to wetting and
drying cycles, and the concrete quality is low. In these cases,
the prediction models for sorptivity might be more relevant as
opposed to those for diffusivity.
For service life prediction, the Life-365 model (Bentz and
Thomas 2013) was found to be a good estimation. Among
the available models to predict time to crack initiation, Bazant’s
model (1979) and Morinaga’s model (1988) tend to underestimate the time to cracking of corrosion of steel in the concrete.
One should exercise caution in using these models because
different laboratory and field environments may cause bias in
the results.
A variety of experimental methods exists to detect carbonationinduced and chloride-induced corrosion in RC structures. It
usually requires a special chamber to provide a certain atmospheric concentration of CO2 to accelerate the concrete carbonation in the laboratory. Temperature and relative humidity have
been found to have a significant influence on the carbonation
rate. Although various methods are available to detect chloride
penetration in concrete, there is no consensus on which method
yields the most reliable results. The bulk diffusion test is considered to be a more reliable long-term method than the salt
ponding test. Among short-term tests, electrical migration and
rapid migration are considered to yield reliable results if there
is no electrically conductive material present in the concrete.
Electrical resistivity may be preferred, for ease of measurement,
to determine diffusivity; however, formation factor, determined
through conductivity tests, could be a better alternative for being
a material property and not being affected by environmental
conditions such as temperature.
Acknowledgments
The financial support for this project was provided by the U.S.
Department of Energy through the Nuclear Energy University Program under the Contract No. 00128931. The findings presented
herein are those of the authors and do not necessarily reflect the
views of the sponsor.
© ASCE
References
AASHTO. (2005). “Standard method of test for rapid determination of the
chloride permeability of concrete.” T277-05, Washington, DC.
AASHTO. (2006). “Standard method of test for resistance of concrete to
chloride ion penetration.” T259-06, Washington, DC.
Aguiar, J. B., and Júnior, C. (2013). “Carbonation of surface protected
concrete.” Constr. Build. Mater., 49, 478–483.
Ahmad, S. (2003). “Reinforcement corrosion in concrete structures,
its monitoring and service life prediction––A review.” Cem. Concr.
Compos., 25(4-5), 459–471.
Al-Amoudi, O. S. B. (1995). “Durability of reinforced concrete in aggressive sabkha environments.” ACI Mater. J., 92(3), 236–245.
Alhozaimy, A., Hussain, R. R., Al-Zaid, R., and Al-Negheimish, A. (2012).
“Coupled effect of ambient high relative humidity and varying temperature marine environment on corrosion of reinforced concrete.” Constr.
Build. Mater., 28(1), 670–679.
Almusallam, A. A., Al-Gahtani, A. S., Aziz, A. R., and Rasheeduzzafar,
(1996). “Effect of reinforcement corrosion on bond strength.” Constr.
Build. Mater., 10(2), 123–129.
Alonso, C., Andrade, C., Rodriguez, J., and Diez, J. M. (1998). “Factors
controlling cracking of concrete affected by reinforcement corrosion.”
Mater. Struct., 31(7), 435–441.
Al-Sulaimani, G. J., Kaleemullah, M., Basunbul, I. A., and Rasheeduzzafar,
(1990). “Influence of corrosion and cracking on bond behavior and
strength of reinforced concrete members.” ACI Struct. J., 87(2), 220–231.
American Concrete Institute (ACI). (2005). “Building code requirements
for structural concrete (ACI 318-05) and commentary.” ACI 318-05,
Farmington Hills, MI.
American Concrete Institute (ACI). (2011). “Building code requirements
for structural concrete (ACI 318-11) and commentary.” ACI 318-11,
Farmington Hills, MI.
Anacta, E. T. (2013). “Effect of salt-contaminated mixing water and aggregates on time-to-initiate rebar corrosion in concrete.” Int. J. Sci. Eng.
Res., 4(4), 1524–1527.
Andrade, C. (1993). “Calculation of chloride diffusion coefficients in concrete
from ionic migration measurements.” Cem. Concr. Res., 23(3), 724–742.
Andrade, C., Alonso, C., and Goni, S. (1993). “Possibilities for electrical
resistivity to universally characterise mass transport processes in concrete.” Concrete 2000: Economic and durable construction through
excellence, Vol. 2, R. K. Dhir and M. R. Jones, eds., CRC Press,
Cambridge, 1639–1652.
Andrade, C., and Sanjuan, M. A. (1994). “Experimental procedure for the
calculation of chloride diffusion coefficients in concrete from migration
tests.” Adv. Cem. Res., 6(23), 127–134.
Angst, U. (2011). “Chloride induced reinforcement corrosion in concrete.”
Ph.D. thesis, Dept. of Structural Engineering, Norwegian Univ. of
Science and Technology, Trondheim, Norway.
Ann, K. Y., and Song, H.-W. (2007). “Chloride threshold level for corrosion
of steel in concrete.” Corros. Sci., 49(11), 4113–4133.
ASCE. (2013). “2013 report card for America’s infrastructure.” 〈http://
www.infrastructurereportcard.org/〉 (Jul. 3, 2013).
ASTM. (2012). ASTM annual book, West Conshohocken, PA.
ASTM. (2013). “Standard test method for specific gravity, absorption, and
voids in hardened concrete.” C642-13, West Conshohocken, PA.
Azher, S. A. (2005). “A prediction model for the residual flexural strength
of corroded reinforced concrete beams.” M.S. thesis, Dept. of Civil
Engineering, King Fahd Univ. of Petroleum and Minerals, Dhahran,
Saudi Arabia.
Azizinamini, A., et al. (2013). “Design guide for bridges for service life.”
SHRP 2 Renewal Project R19A, Transportation Research Board,
Washington, DC.
Bagheri, A. R., and Zanganeh, H. (2012). “Comparison of rapid tests
for evaluation of chloride resistance of concretes with supplementary
cementitious materials.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT.1943
-5533.0000485, 1175–1182.
Bajaj, S. (2012). “Effect of corrosion on physical and mechanical properties
of reinforced concrete.” M.S. thesis, Univ. of Akron, Akron, OH.
Bazant, Z. P. (1979). “Physical model for steel corrosion in concrete sea
structures—Application.” J. Struct. Div., 105(6), 1155–1166.
04014245-14
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Bentur, A., Diamond, S., and Berke, N. S. (1997). Steel corrosion in
concrete, E and FN Spon, New York.
Bentz, E. C., and Thomas, M. D. A. (2013). “Life-365 service life prediction model.” Life-365 Consortium III, The Silica Fume Association,
Lovettsville, VA.
Benzarti, K., Chataigner, S., Quiertant, M., Marty, C., and Aubagnac, C.
(2011). “Accelerated ageing behaviour of the adhesive bond between
concrete specimens and CFRP overlays.” Constr. Build. Mater.,
25(2), 523–538.
Bertolini, L., Bolzoni, F., Pastore, T., and Pedeferri, P. (1996). “Behaviour
of stainless steel in simulated concrete pore solution.” Br. Corros. J.,
31(3), 218–222.
Bioubakhsh, S. (2011). “The penetration of chloride in concrete subject
to wetting and drying: Measurement and modeling.” Ph.D. thesis, Dept.
of Civil, Environmental and Geomatic Engineering, Univ. College
London, London.
Bossio, A., et al. (2011). “Analytical modeling of reinforcement corrosion
effects on RC members degradation.” 2nd Workshop the New Boundaries of Structural Concrete, Università Politecnica delle Marche, Italy.
Breit, W. (1998). “Critical corrosion inducing chloride content—State of the
art and new investigation results.” Concr. Technol. Rep., 7, 145–167.
British Standards Institute (BSI). (2011). “Testing concrete. Method for
making and curing no-fines test cubes.” BS 1881-113:2011, London.
Broomfield, J. P. (1997). Corrosion of steel in concrete: Understanding,
investigation and repair, E & FN Spon, New York.
Carpenter Technology Corporation. (2002). “New bridge using stainless
steel rebar to last 120 years in corrosive marine and earthquake environment.” 〈https://www.cartech.com/news.aspx?id=1424〉 (Jun. 10, 2014).
Castel, A., Francois, R., and Arliguie, G. (2000a). “Mechanical behaviour
of corroded reinforced concrete beams—Part 1: Experimental study of
corroded beams.” Mater. Struct., 33(9), 539–544.
Castel, A., Francois, R., and Arliguie, G. (2000b). “Mechanical behaviour
of corroded reinforced concrete beams—Part 2: Bond and notch
effects.” Mater. Struct., 33(9), 545–551.
Chen, C.-T., and Ho, C.-W. (2013). “Influence of cyclic humidity on
carbonation of concrete.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT
.1943-5533.0000750, 1929–1935.
Chi, J. M., Huang, R., and Yang, C. C. (2002). “Effects of carbonation
on mechanical properties and durability of concrete using accelerated
testing method.” J. Marine Sci. Technol., 10(1), 14–20.
Collepardi, M., Collepardi, S., Skarp, U., and Troli, R. (2004). “Optimization of silica fume, fly ash and amorphous nano-silica in superplasticized high-performance concretes.” ACI Spec. Publ., 221, 495–506.
Collepardi, M., Marcialis, A., and Turrizuani, R. (1970). “Kinetics of penetration of chloride ions into the concrete.” Cemento, 67(4), 157–164.
Comité Euro-International du Béton (CEB). (1997). “New approach to
durability design—An example for carbonation induced corrosion.”
Bulletin No. 238, Lausanne, Switzerland.
Costa, A., and Appleton, J. (1999a). “Chloride penetration into concrete
in marine environment—Part I: Main parameters affecting chloride
penetration.” Mater. Struct., 32(4), 252–259.
Costa, A., and Appleton, J. (1999b). “Chloride penetration into concrete
in marine environment—Part II: Prediction of long term chloride
penetration.” Mater. Struct., 32(5), 354–359.
Darwin, D., Browning, J., Gong, L., and Hughes, S. R. (2007). “Effects
of deicers on concrete deterioration.” SL Rep. 07-3, Structural Engineering and Materials Laboratory, Univ. of Kansas Center for Research,
Lawrence, KS.
Das, B. B., Singh, D. N., and Pandey, S. P. (2012). “Rapid chloride ion
permeability of OPC- and PPC-based carbonated concrete.” J. Mater.
Civ. Eng., 10.1061/(ASCE)MT.1943-5533.0000415, 606–611.
De Ceukelaire, L., and Van Nieuwenburg, D. (1993). “Accelerated carbonation of a blast-furnace cement concrete.” Cem. Concr. Res., 23(2),
442–452.
Delagrave, A., Marchaud, J., and Samson, E. (1996). “Prediction of diffusion coefficients in cement-based materials on the basis of migration
experiments.” Cem. Concr. Res., 26(12), 1831–1842.
DeSouza, S. J., Hooton, R. D., and Bickely, J. A. (1995). “A field test for
evaluating high performance concrete covercrete quality.” Can. J. Civ.
Eng., 25(3), 551–556.
© ASCE
Detwiler, R. J., Kjellsen, K. O., and Gjorv, O. E. (1991). “Resistance
to chloride intrusion of concrete cured at different temperatures.” ACI
Mater. J., 88(1), 19–24.
Deutsches Institut für Normung (DIN). (1987). “Determination of the water
absorption coefficient of building materials.” DIN 52617-87, Berlin.
Dhir, R. K., and Jones, M. R. (1999). “Development of chloride-resisting
concrete using fly ash.” Fuel, 78(2), 137–142.
Diamond, S. (1986). “Chloride concentrations in concrete pore solutions
resulting from calcium and sodium chloride admixtures.” J. Cem.
Concr. Aggregates, 8(2), 97–102.
Dunn, D. S., Pulvirenti, A. L., and Hiser, M. A. (2011). “Containment
liner corrosion operating experience summary technical letter report—
Revision 1.” ML112070867, U.S. Nuclear Regulatory Commission,
Office of Nuclear Regulatory Research, Washington, DC.
Eguchi, K., Takewaka, K., Yamaguchi, T., and Ueda, N. (2013). “A study
on durability of blast furnace slag cement concrete mixed with
metakaolin-based artificial pozzolan in actual marine environment.”
3rd Int. Conf. on Sustainable Construction Materials and Technologies,
Kyoto Univ., Kyoto, Japan.
El-Salam, S. A., Eldin, H. S., El-Shamy, E. A., and Shawky, S. M. M.
(2012). “Effect of high percentage of sodium chloride (Nacl) on the
behavior of reinforced concrete beams.” Life Sci. J., 9(3), 315–320.
Evans, U. R. (1960). The corrosion and oxidation of metals: Scientific
principles and practical applications, Edward Arnold LTD, London.
Fang, C., Lundgren, K., Chen, L., and Zhu, C. (2004). “Corrosion influence
on bond in reinforced concrete.” Cem. Concr. Res., 34(11), 2159–2167.
Federal Highway Administration (FHWA). (1995). “Recording and coding
guide for the structure inventory and appraisal of the nation’s bridges.”
FHWA-PD-96-001, U.S. Dept. of Transportation, Washington, DC.
Federal Highway Administration (FHWA). (1998). “Corrosion protection—
Concrete bridges.” FHWA-RD-98-088, U.S. Dept. of Transportation,
McLean, VA.
Féderation Internationale de la Précontrainte (FIP). (1996). “Durability of
concrete structures in the North Sea, state-of-the-art report.” London.
Feldman, R. F. (1987). “Diffusion measurements in cement paste by water
replacement using propan-2-ol.” Cem. Concr. Res., 17(4), 602–612.
Freeze, R. A., and Cherry, J. A. (1979). Groundwater, Prentice-Hall,
Englewood Cliffs, NJ.
Fukute, T., Yamamoto, K., and Hamada, H. (1990). Study on durability of
concrete mixed with sea water, Port and Harbour Research Institute,
Ministry of Transport, Nagase, Yokosuka, Japan (in Japanese).
Gao, Y., Cheng, L., Gao, Z., and Guo, S. (2013). “Effects of different
mineral admixtures on carbonation resistance of lightweight aggregate
concrete.” Constr. Build. Mater., 43, 506–510.
Glass, G. K., and Buenfeld, N. R. (1995). Chloride threshold levels for
corrosion induced deterioration of steel in concrete, RILEM Publications, Saint-Rémy-lès-Chevreuse, France.
Goodspeed, C. H., Vanikar, S., and Cook, R. A. (1996). “High performance
concrete defined for highway structures.” ACI Concr. Int., 18(2), 62–67.
Hall, C. (1989). “Water sorptivity of mortars and concretes: A review.”
Mag. Concr. Res., 41(147), 51–61.
Han, S.-H., Park, W.-S., and Yang, E.-I. (2013). “Evaluation of concrete
durability due to carbonation in harbor concrete structures.” Constr.
Build. Mater., 48, 1045–1049.
Hansson, C. M., Poursaee, A., and Laurent, A. (2006). “Macrocell and
microcell corrosion of steel in ordinary portland cement and high performance concretes.” Cem. Concr. Res., 36(11), 2098–2102.
Heiyantuduwa, R., Alexander, M. G., and Mackechnie, J. R. (2006).
“Performance of a penetrating corrosion inhibitor in concrete affected
by carbonation-induced corrosion.” J. Mater. Civ. Eng., 10.1061/
(ASCE)0899-1561(2006)18:6(842), 842–850.
Helmuth, R., Stark, D., Diamond, S., and Moranville-Regourd, M. (1993).
“Alkali-silica reactivity: An overview of research.” SHRP-C-342,
Strategic Highway Research Program, National Research Council,
Washington, DC.
Hooton, R. D., Bentz, E., and Kojundic, T. (2010). “Long-term chloride
penetration resistance of silica fume concretes based on field exposure.”
2nd Int. Symp., Service Life Design for Infrastructure, K. van Breugel,
G. Ye, and Y. Yuan, eds., RILEM Publications, Delft, Netherlands.
04014245-15
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
Hossain, K. M. A. (2006). “Performance of volcanic ash and pumice-based
blended cements in sulphate and sulphate–chloride environments.” Adv.
Cem. Res., 18(2), 71–82.
Huang, N. M., Chang, J. J., and Liang, M. T. (2012). “Effect of plastering
on the carbonation of a 35-year-old reinforced concrete building.”
Constr. Build. Mater., 29, 206–214.
Islam, M. M., Islam, M. S., Al-Amin, M., and Islam, M. M. (2012). “Suitability of sea water on curing and compressive strength of structural
concrete.” J. Civ. Eng., 40(1), 37–45.
Jacobsen, S., Marchaud, J., and Boisvert, L. (1996). “Effect of cracking
and healing on chloride transport in OPC concrete.” Cem. Concr. Res.,
26(6), 869–881.
Jaegermann, C. (1990). “Effect of water-cement ratio and curing on chloride penetration into concrete exposed to mediterranean sea climate.”
ACI Mater. J., 87(4), 333–339.
Jang, J.-W., Iwasaki, I., and Weiblen, P. (1995). “Effect of salt additives on
concrete degradation (phase II).” MN/RC-96/10, Minnesota Dept. of
Transportation, Office of Research Administration, St. Paul, MN.
Jin, W.-L., and Zhao, Y. (2001). “Effect of corrosion on bond behavior and
bending strength of reinforced concrete beams.” J. Zhejiang Univ., 2(3),
298–308.
Jung, S. H., Lee, M. K., Lee, S. L., and Oh, B. H. (2010). “Experimental
investigation on diffusion coefficient of carbon dioxide for sustainable
construction materials.” 2nd Int. Conf. on Sustainable Construction
Materials and Technologies, Università Politecnica della Marche,
Ancona, Italy.
Kashani, M. M., Crewe, A. J., and Alexander, N. A. (2013). “Nonlinear
stress–strain behaviour of corrosion-damaged reinforcing bars including inelastic buckling.” Eng. Struct., 48, 417–429.
Kelestemur, O., and Yildiz, S. (2006). “Effect of various NaCl concentration on corrosion of steel in concrete produced by addition of
styrofoam.” Gazi Univ. J. Sci., 19(3), 163–172.
Kessler, R. J., Powers, R. G., Vivas, E., Paredes, M. A., and Virmani, Y. P.
(2008). “Surface resistivity as an indicator of concrete chloride penetration resistance.” 2008 Concrete Bridge Conf., National Concrete Bridge
Council, Skokie, IL.
Kwon, S. J., Na, U. J., Park, S. S., and Jung, S. H. (2009). “Service life
prediction of concrete wharves with early-aged crack: Probabilistic
approach for chloride diffusion.” Struct. Safety, 31(1), 75–83.
Leber, I., and Blakey, F. A. (1956). “Some effects of carbon dioxide on
mortars and concrete.” ACI Mater. J., 28(3), 295–308.
Leng, F., Feng, N., and Lu, X. (2000). “An experimental study on the
properties of resistance to diffusion of chloride ions of fly ash and blast
furnace slag concrete.” Cem. Concr. Res., 30(6), 989–992.
Li, C. Q. (2003). “Life-cycle modeling of corrosion-affected concrete structures: propagation.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2003)
129:6(753), 753–761.
Li, C. Q., Lawanwisut, W., and Zheng, J. J. (2005). “Time-dependent reliability method to assess the serviceability of corrosion-affected concrete
structures.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2005)131:
11(1674), 1674–1680.
Li, G., Hu, F., and Wu, Y. (2011). “Chloride ion penetration in stressed
concrete.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT.1943-5533
.0000281, 1145–1153.
Liu, T., and Weyers, R. W. (1998). “Modeling the dynamic corrosion process in chloride contaminated concrete structures.” Cem. Concr. Res.,
28(3), 365–379.
Liu, Y. (1996). “Modeling the time-to-corrosion cracking of the cover
concrete in chloride contaminated reinforced concrete structures.”
Ph.D. thesis, Dept. of Civil Engineering, Virginia Polytechnic Institute
and State Univ., Blacksburg, VA.
Lu, X. (1997). “Application of the Nernst-Einstein equation to concrete.”
Cem. Concr. Res., 27(2), 293–302.
Martys, N. S., and Ferraris, C. F. (1997). “Capillary transport in mortars and
concrete.” Cem. Concr. Res., 27(5), 747–760.
Matsuzawa, K., Kitsutaka, Y., and Tsukagoshi, M. (2010). “Effect of humidity on rate of carbonation of concrete exposed to high-temperature
environment.” Int. Symp. on the Ageing Management & Maintenance of
Nuclear Power Plants, Mitsubishi Research Institute, Japan, 109–114.
© ASCE
McCarter, W. J., Ezirim, H., and Emerson, M. (1992). “Absoprtion of water
and chlorides into concrete.” Mag. Concr. Res., 44(158), 31–37.
McGrath, P. (1996). “Development of test methods for predicting chloride
penetration into high performance concrete.” Ph.D. thesis, Dept. of
Civil Engineering, Univ. of Toronto, Toronto, Canada.
McGrath, P., and Hooton, R. D. (1996). “Influence of voltage on chloride
diffusion coefficients from chloride migration tests.” Cem. Concr. Res.,
26(8), 1239–1244.
Medeiros, M. H. F., and Helene, P. (2009). “Surface treatment of reinforced
concrete in marine environment: Influence on chloride diffusion coefficient and capillary water absorption.” Constr. Build. Mater., 23(3),
1476-1484.
Melchers, R. E. (2003). “Modeling of marine immersion corrosion for mild
and low alloy steels—Part 1: Phenomenological model.” Corrosion,
59(4), 319–334.
Monfore, G. E. (1968). “The electrical resistivity of concrete.” J. Portland
Cem. Assoc., 10(2), 35–48.
Montemor, M. F., Simões, A. M. P., and Ferreira, M. G. S. (2003).
“Chloride-induced corrosion on reinforcing steel: From the fundamentals
to the monitoring techniques.” Cem. Concr. Compos., 25(4–5), 491–502.
Morinaga, S. (1988). “Prediction of service lives of reinforced concrete
buildings based on the corrosion rate of reinforcing steel.” Special
Rep. of Institute of Technology, No. 23, Shimuzu Corporation, Tokyo.
Morris, W., Moreno, E. I., and Sagues, A. A. (1996). “Practical evaluation
of resistivity of concrete in test cylinders using a Wenner array probe.”
Cem. Concr. Res., 26(12), 1779–1787.
Naus, D. J., and Graves, H. L., III (2013). Detection of aging of nuclear
power plant structures (draft), U.S. Nuclear Regulatory Commission,
Office of Nuclear Regulatory Research, Washington, DC.
Neville, A. (1995). “Chloride attack of reinforced concrete: An overview.”
Mater. Struct., 28(2), 63-70.
Nordtest. (1995). “Accelerated chloride penetration.” NT Build 443, Espoo,
Finland.
O’Reilly, M., Darwin, D., Browning, J., and Locke, C. E. (2011). “Evaluation of multiple corrosion protection systems for reinforced concrete
bridge decks.” SM Rep. No. 96, Structural Engineering and Materials
Laboratory, Univ. of Kansas Center for Research, Lawrence, KS.
Ortega, N. F., Rivas, I. E., Avelda˜no, R. R., and Peralta, M. H. (2011).
“Beams affected by corrosion influence of reinforcement placement
in the cracking.” Struct. Eng. Mech., 37(2), 163–175.
Otsuki, N., Miyazato, S.-i., and Yodsudjai, W. (2003). “Influence of
recycled aggregate on interfacial transition zone, strength, chloride
penetration and carbonation of concrete.” J. Mater. Civ. Eng., 10.1061/
(ASCE)0899-1561(2003)15:5(443), 443–451.
Otsuki, N., Nagataki, S., and Nakashita, K. (1992). “Evaluation of AgNO3
solution spray method for measurement of chloride penetration into hardened cementitious matrix materials.” ACI Mater. J., 89(6), 587–592.
Pakshir, M., and Esmaili, S. (1998). “The effect of chloride ion concentration on the corrosion of concrete.” Scientia Iranica, 4(4), 201–205.
Papadakis, V. G., Vayenas, C. G., and Fardis, M. N. (1989). “A reaction engineering approach to the problem of concrete carbonation.”
AIChE J., 35(10), 1639–1650.
Papakonstantinou, K. G., and Shinozuka, M. (2013). “Probabilistic model
for steel corrosion in reinforced concrete structures of large dimensions
considering crack effects.” Eng. Struct., 57, 306–326.
Petti, J. P., Naus, D., Sagues, A., Weyers, R. E., Erler, B. A., and
Berke, N. S. (2011). “Nuclear containment steel linear corrosion workshop: Final summary and recommendation report.” SAND2010-8718,
Sandia National Laboratories, Albuquerque, NM.
Polder, R. B. (1996). “The influence of blast furnace slag, fly ash and silica
fume on corrosion of reinforced concrete in marine environment.”
Heron, 41(4), 287–300.
Pruckner, F., and Gjørv, O. E. (2004). “Effect of CaCl2 and NaCl additions
on concrete corrosivity.” Cem. Concr. Res., 34(7), 1209–1217.
Rabehi, M., Mezghiche, B., and Guettala, S. (2013). “Correlation between
initial absorption of the cover concrete, the compressive strength and
carbonation depth.” Constr. Build. Mater., 45, 123–129.
Richardson, D. N. (2006). “Strength and durability characteristics of a 70%
ground granulated blast furnace slag (GGBFS) concrete mix.”
RI99-035/RI99-035B, Univ. of Missouri-Rolla, Rolla, MO.
04014245-16
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.
Downloaded from ascelibrary.org by University of Houston on 03/31/15. Copyright ASCE. For personal use only; all rights reserved.
RILEM. (1994). “Draft recommendation for repair strategies for concrete
structures damaged by reinforcement corrosion.” Mater. Struct., 27(27),
415–436.
Roa-Rodriguez, G., Aperador, W., and Delgado, A. (2013). “Calculation of
chloride penetration profile in concrete structures.” Int. J. Electrochem.
Sci., 8(4), 5022–5035.
Rostami, V., Shao, Y., and Boyd, A. J. (2011). “Durability of concrete pipes
subjected to combined steam and carbonation curing.” Constr. Build.
Mater., 25(8), 3345–3355.
Saito, M., and Ishimori, H. (1995). “Chloride permeability of concrete
under static and repeated compressive loading.” Cem. Concr. Res.,
25(4), 803–808.
Samaha, H. R., and Hover, K. C. (1992). “Influence of microcracking on the
mass transport properties of concrete.” ACI Mater. J., 89(4), 416–424.
Sharif, A., Loughlin, K. F., Azad, A. K., and Navaz, C. M. (1997).
“Determination of the effective chloride diffusion coefficient in concrete via a gas diffusion technique.” ACI Mater. J., 94(3), 227–233.
Shayanfar, M. A., Ghalehnovi, M., and Safiey, A. (2007). “Corrosion
effects on tension stiffening behavior of reinforced concrete.” Comput.
Concr., 4(5), 403–424.
Shi, X., et al. (2010). “Effect of chloride-based deicers on reinforced
concrete structures.” WA-RD 741.1, Corrosion & Sustainable Infrastructure Laboratory, Western Transportation Institute, Montana State Univ.,
Bozeman, MT.
Shigeru, A., Kanai, M., Takeda, N., and Ishizeki, Y. (2012). Seawatermixed concrete, concrete structure constructed with the same, and
design method of concrete structure constructed with seawater-mixed
concrete, EP 2 465 837 A1, Eueopean Patent Office.
Sindelar, R. L., Duncan, A. J., Dupont, M. E., Lam, P.-S., Louthan, M. R.,
and Skidmore, J. T. E. (2011). “Materials aging issues and aging management for extended storage and transportation of spent nuclear fuel.”
NUREG/CR-7116, SRNL-STI-2011-00005, Savannah River National
Laboratory, Aiken, SC.
Snyder, K. A. (2001). “The relationship between the formation factor and
the diffusion coefficient of porous materials saturated with concentrated
electrolytes: Theoretical and experimental considerations.” Concr. Sci.
Eng., 3(12), 216–224.
Song, H.-W., Pack, S.-W., and Moon, J.-S. (2006). “Durability evaluation
of concrete structures exposed to marine environment focusing on a
chloride build-up on concrete surface.” Int. Workshop on Life Cycle
Management of Coastal Concrete Structures, Port and Airport Research
Institute, Japan.
Spragg, R., Villani, C., Snyder, K., Bentz, D., Bullard, J., and Weiss, J.
(2013). “Factors that influence electrical resistivity measurements in
cementitious systems.” Transportation Research Record 2342, Transportation Research Board, Washington, DC, 90–98.
Stanish, K. (1997). “Corrosion effects on bond strength in reinforced
concrete.” M.A.S. thesis, Dept. of Civil Engineering, Univ. of Toronto,
Toronto.
Stanish, K. D., Hooton, R. D., and Thomas, M. D. A. (1997). “Testing
the chloride penetration resistance of concrete: A literature review.”
DTFH61-97-R-00022, Univ. of Toronto, Toronto, ON, Canada.
Stewart, M. G., and Al-Harthy, A. (2008). “Pitting corrosion and structural
reliability of corroding RC structures: Experimental data and probabilistic analysis.” Reliab. Eng. Syst. Safety, 93(3), 373–382.
Streicher, P. E., and Alexander, M. G. (1995). “A chloride conduction test
for concrete.” Cem. Concr. Res., 25(6), 1284–1294.
Sutter, L., Peterson, K., Julio-Betancourt, G., Hooton, D., Van Dam, T., and
Smith, K. (2008). “The deleterious chemical effecs of concenrtrated
deicing solutions on portland cement concrete.” Rep. SD2002-01-X,
Michigan Tech Transportation Institute, Houghton, MI.
Tang, L., and Nilsson, L.-O. (1991). “Chloride diffusivity in high strength
concrete.” Nordic Concr. Res., 11(1), 162–170.
Thomas, M. D. A., and Jones, M. R. (1996). “A critical review of service
life modelling of concretes exposed to chlorides.” Concr. Serv. Mankind
Radical Concr. Technol., R. K. Dhir and P. C. Hewlett, eds., E. & F. N.
Spon, London, 723–736.
Torres-Acosta, A., and Sagues, A. (2004). “Concrete cracking by localized
steel corrosion-geometric effects.” ACI Mater. J., 101(6), 501–507.
© ASCE
Transportation Research Board. (1991). “Highway deicing: Comparing salt
and calcium manesium acetate.” Special Rep. 235, National Research
Council, Washington, DC.
Treadaway, K. W. J., Cox, R. N., and Brown, B. L. (1989). “Durability
of corrosion resisting steels in concrete.” Proc. Inst. Civ. Eng., 86(2),
305–331.
Trejo, D., and Pillai, R. G. (2004). “Accelerated chloride threshold
testing—Part II: Corrosion resistant reinforcement.” ACI Mater. J.,
101(1), 57–64.
Tumidajski, P. J., and Schumacher, A. S. (1996). “On the relationship
between the formation factor and propan-2-ol diffusivity in mortars.”
Cem. Concr. Res., 26(9), 1301–1306.
Turk, K., Karatas, M., and Gonen, T. (2013). “Effect of fly ash and silica
fume on compressive strength, sorptivity and carbonation of SCC.”
KSCE J. Civ. Eng., 17(1), 202–209.
Val, D. V., and Chernin, L. (2012). “Cover cracking in reinforced
concrete elements due to corrosion.” Struct. Infrastruct. Eng., 8(6),
569–581.
Val, D. V., Stewart, M. G., and Melchers, R. E. (1998). “Effect of reinforcement corrosion on reliability of highway bridges.” Eng. Struct., 20(11),
1010–1019.
Valcuende, M., and Parra, C. (2010). “Natural carbonation of selfcompacting concretes.” Constr. Build. Mater., 24(5), 848–853.
Valenta, O. (1969). “Kinetics of water penetration into concrete as an
important factor of its deterioration and of reinforcement corrosion.”
RILEM Int. Symp.: Durability of Concrete—Part 1, Academia, Prague,
A177–A189.
Vidal, T., Castel, A., and Francois, R. (2004). “Analyzing crack width
to predict corrosion in reinforced concrete.” Cem. Concr. Res., 34(1),
165–174.
Villagrán-Zaccardi, Y. A., Zega, C. J., and Di Maio, Á. A. (2008). “Chloride penetration and binding in recycled concrete.” J. Mater. Civ. Eng.,
10.1061/(ASCE)0899-1561(2008)20:6(449), 449–455.
Vu, K. A. T., and Stewart, M. G. (2000). “Structural reliability of concrete
bridges including improved chloride-induced corrosion models.” Struct.
Safety, 22(4), 313-333.
Wang, K., Nelsen, D. E., and Nixon, W. A. (2006). “Damaging effects of
deicing chemicals on concrete materials.” Cem. Concr. Compos., 28(2),
173–188.
Wang, X.-M., and Zhao, H.-Y. (1993). The residual service life prediction
of R.C. structures, E & FN Spon, Omiya, Japan.
Webster, P. M. (2000). “The assessment of corrosion-damaged concrete structures.” Ph.D. thesis, School of Civil Engineering, Univ. of
Birmingham, Birmingham, U.K.
Wegian, F. M. (2010). “Effect of seawater for mixing and curing on structural concrete.” IES J. Part A Civ. Struct. Eng., 3(4), 235–243.
Whiting, D. (1981). “Rapid measurement of the chloride permeability of
concrete.” Public Roads, 45(3), 101–112.
Xi, Y., and Bazant, Z. P. (1999). “Modeling chloride penetration in saturated concrete.” J. Mater. Civ. Eng., 10.1061/(ASCE)0899-1561(1999)
11:1(58), 58–65.
Yoon, I.-S., Copuroglu, O., and Park, K.-B. (2007). “Effect of global
climatic change on carbonation progress of concrete.” Atmos. Environ.,
41(34), 7274–7285.
Yu, H., and Hartt, W. H. (2007). “Effects of reinforcement and coarse
aggregates on chloride ingress into concrete and time-to-corrosion:
Part 1—Spatial chloride distribution and implications.” Corrosion,
63(9), 843–849.
Zhang, M. H., and Gjorv, O. E. (1991). “Permeability of high strength lightweight concrete.” ACI Mater. J., 88(5), 463–469.
Zhang, S.-F., Lu, C.-H., and Liu, R.-G. (2011). “Experimental determination of chloride penetration in cracked concrete beams.” Procedia Eng.,
24, 380–384.
Zhang, W., Shang, D., and Gu, X. (2006). “Stress-strain relationship of
corroded steel bars.” J. Tongji Univ., 34(5), 586–592 (in Chinese).
Zhang, W.-M., and Ba, H.-J. (2011). “Accelerated life test of concrete
in chloride environment.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT
.1943-5533.0000182, 330–334.
04014245-17
J. Mater. Civ. Eng.
J. Mater. Civ. Eng.