You are on page 1of 14

Construction and Building Materials 30 (2012) 125–138

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Review

Durability of steel reinforced concrete in chloride environments: An overview


Xianming Shi a,b,⇑, Ning Xie c, Keith Fortune a, Jing Gong d
a
Corrosion and Sustainable Infrastructure Laboratory, Western Transportation Institute, P.O. Box 174250, College of Engineering, Montana State University,
Bozeman, MT 59717-4250, USA
b
Civil Engineering Department, 205 Cobleigh Hall, Montana State University, Bozeman, MT 59717-2220, USA
c
School of Transportation Science and Engineering, Harbin Institute of Technology, Harbin 150090, China
d
School of Civil Engineering and Architecture, Wuhan Polytechnic University, Wuhan 430023, China

a r t i c l e i n f o a b s t r a c t

Article history: Concrete is a unique composite material that is porous and highly heterogeneous. The durability of steel
Received 14 September 2011 reinforced concrete in chloride environments is of great interest to design engineers, infrastructure own-
Received in revised form 19 November 2011 ers and maintainers, and researchers. This review reports recent advances in the knowledge base relevant
Accepted 4 December 2011
to the durability of steel reinforced concrete in chloride environments, including: the role of mineral
Available online 29 December 2011
admixtures in concrete durability, the methods of measuring the chloride ingress into concrete, the chal-
lenges in assessing concrete durability from its chloride diffusivity, and the service life modeling of rein-
Keywords:
forced concrete in chloride-laden environments. It concludes with a look to the future, including research
Reinforced concrete
Chloride-induced corrosion
needs to be addressed.
Mineral admixtures Ó 2011 Elsevier Ltd. All rights reserved.
Chloride ingress into concrete
Chloride diffusivity
Service life modeling

Contents

1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
2. Role of mineral admixtures in concrete durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
2.1. FA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
2.2. UFFA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2.3. SF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2.4. GGBFS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
2.5. MK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3. Measuring the chloride ingress into concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
4. Challenges in assessing the durability of concrete from its chloride diffusivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
4.1. Chloride threshold. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
4.2. Chloride binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5. Service life modeling of steel reinforced concrete in chloride environments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

1. Background Concrete has also become the material of choice for the construc-
tion of structures exposed to extreme conditions [1]. Furthermore,
Concrete is the most widely used man-made building material sustainability has become an increasingly important characteristic
in the world, owing to its versatility and relatively low cost. for concrete infrastructure, as the production of Portland cement
(the most common binder in concrete) is an energy-intensive pro-
cess that accounts for a significant portion of global carbon dioxide
⇑ Corresponding author at: Corrosion and Sustainable Infrastructure Laboratory,
emissions and other greenhouse gases [2,3]. As such, even slight
Western Transportation Institute, P.O. Box 174250, College of Engineering, Montana
State University, Bozeman, MT 59717-4250, USA. Tel.: +1 406 994 6486; fax: +1 406 improvements in the design, production, construction, mainte-
994 1697. nance, and materials performance of concrete can have enormous
E-mail address: xianming_s@coe.montana.edu (X. Shi). social, economic and environmental impacts.

0950-0618/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.conbuildmat.2011.12.038
126 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

There are a variety of approaches to enhance the sustainability high tide: (1) adequate cover (4 in. for cast-in-place members
of concrete and reduce its environmental footprint. One attractive and 3 in. for prestressed components), and (2) low water-to-
approach is to use fly ash and other industrial byproducts as a cementitious-material (w/cm) ratio concrete with pozzolanic (fly
replacement for the Portland cement in concrete. Another ap- ash or silica fume) or corrosion inhibiting admixtures [6]. In light
proach is to enhance the durability of concrete infrastructure, since of advances in concrete technology and requirements of the AASH-
durability is a key cornerstone for sustainability. According to the TO Load and Resistance Factor Design (LRFD) for a 75-year design
ASCE 2009 Report Card for America’s Infrastructure, $2.2 trillion life, the California Department of Transportation (Caltrans) made
needs to be invested over 5 years to ‘‘bring the nation’s infrastruc- substantial changes to its Bridge Design Specifications (BDSs) Arti-
ture to a good condition’’ [4], which highlights the urgent need for cle 8.22 in 2000 and adopted the approach of using the chloride
research devoted to longer-lasting and ‘‘maintenance-free’’ con- diffusivity through concrete to determine the concrete cover
crete materials. requirements for structures subjected to chloride-bearing environ-
Chloride-induced rebar corrosion is one of the major forms of ments. The current BDS Article 8.22 provides guidance to the
environmental attack to reinforced concrete [5], which may lead design engineer in determining the required cement type, mini-
to reduction in the strength, serviceability, and esthetics of the mum required concrete cover, etc. for corrosion protection of var-
structure. The accumulation of corrosion products (oxides/hydrox- ious bridge members [11]. For instance, for bridge members
ides) in the concrete pore space near the steel rebar can build up exposed to corrosive soil or water (containing more than
hoop stresses around the rebar and result in cracking or spalling 500 ppm of chlorides), the maximum w/cm ratio shall not exceed
of the concrete, which in turn facilitates the ingress of moisture, 0.40. Mineral admixtures conforming to ASTM Designation C 618
oxygen, and chlorides to the embedded rebar and accelerates the Type F or N (e.g., fly ash) are required for all exposure conditions,
corrosion of steel [6]. except for ‘non-corrosive’ exposure conditions. For such bridge
Transportation agencies spend billions of dollars each year on members as precast piles and pile extensions exposed to corrosive
the construction and maintenance of concrete bridges, and the conditions, mineral admixtures conforming to ASTM Designation C
US national bridge inventory includes numerous reinforced con- 1240 (e.g., silica fume) may be required. The minimum concrete
crete structures with exposure to corrosive salt conditions in mar- cover required for bridge members ranges from 1 to 5 in., depen-
ine environments and in cold-climate regions. With extended dent on the bridge member type and exposure condition [11].
service life and reduced need for costly and difficult repair and Mineral admixtures, generally pozzolanic materials, are mainly
rehabilitation of concrete structures, the implementation of better glassy siliceous materials that may contain aluminous compounds
design practices for corrosion management will have immediate [10]. The reaction of such materials with Portlandite (calcium
positive impacts on the highway system, including cost savings, hydroxide) and water generates hydration products similar to
enhanced traveler safety, reduced traveler delays, and minimized those of Portland cement, i.e., calcium silicate hydrates (C–S–H),
environmental impacts. a rigid gel composed of extremely small particles with a layer
In this context, this review reports recent advances in the structure:
knowledge base relevant to the durability of steel reinforced con-
pozzolan þ water þ CaðOHÞ2 ! C—S—H
crete in chloride environments. Due to space limitations, this re-
view will focus on: the role of mineral admixtures in concrete This reaction can also be generally represented as:
durability, the methods of measuring the chloride ingress into con-
2
crete, the challenges in assessing concrete durability from its chlo- H4 SiO4 þ CaðOHÞ2 ! H2 SiO4 þ Ca2þ þ 2H2 O
ride diffusivity, and the service life modeling of reinforced concrete ! CaH2 SiO4  2H2 O
in chloride-laden environments. It concludes with a look to the fu-
ture, including research needs to be addressed. Other topics, while Note that the actual stoichiometry of the reaction may vary as a
related, will be omitted. These may include: the application of function of the Ca/Si ratio, available water molecules, etc., resulting
nanotechnology to reinforced concrete materials and structures in various C–S–H that may deviate from the general formula
[7–9], the electrochemical techniques to protect or rehabilitate (CaH2SiO42H2O).
chloride-contaminated concrete, the use of advanced materials In general, the use of mineral admixtures such as fly ash, silica
(e.g., fiber-reinforced polymers and cementitious composites) for fume, slag and metakaolin has been shown to enhance concrete
repair of reinforced concrete, and the emerging techniques (e.g., durability [12–14], by increasing chloride binding [15], decreasing
optical fiber sensors) for structural sensing and monitoring. chloride permeability [14,16], elevating threshold chloride content
[17], and/or improving the distribution of pore size and shape of
concrete matrix [18,19]. Mineral admixtures have been incorpo-
2. Role of mineral admixtures in concrete durability rated into binary, ternary and quaternary concrete mixes. Since
some of these materials are cheaper than Portland cement, there
There is general agreement that the most effective improve- is an economic advantage to wider use, in addition to potential
ment in concrete durability can be achieved at the design and benefits to the environment. Dhir and Jones [20] used the low-lime
materials selection stage of a project by using adequate concrete fly ash to develop concrete mixes with improved chloride resis-
cover and high quality concrete, as long as the concrete is properly tance, by improving the pore structure and binding capacity of
placed and cured. Usually, an increase in the thickness of the con- the concrete. They found that ternary blends (cement and fly ash
crete cover leads to beneficial effects, because it increases the bar- blended with silica fume or metakaolin) showed the highest chlo-
rier to the various aggressive species moving towards the ride resistance. Hossain et al. [21] found that the incorporation of
reinforcement and increases the time for corrosion to initiate. In ultra-fine fly ash (UFFA) improved the strength and chloride pene-
reality, however, the cover thickness cannot exceed certain limits, tration resistance of concrete, and the incorporation of silica fume
for mechanical and practical reasons [10]. To achieve desirable had even more pronounced benefits. They also found that the silica
strength, durability, and sustainability of concrete, recent innova- fume addition led to low slump and high early-age shrinkage
tion has focused on the use of chemical and mineral admixtures whereas the UFFA addition mitigated these two issues. As such, a
in concrete. The Florida Department of Transportation (DOT) ternary blend (with cement, silica fume and UFFA) was developed
adheres to the following specifications for concrete bridge sub- to feature high early-age strength, improved durability, low slump
structures within the 0–12 foot elevation range relative to mean and low free shrinkage. Thomas et al. [22] investigated the synergy
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 127

between silica fume and fly ash, as silica fume compensates for the high-lime fly ashes) improved the resistance of concrete to carbon-
low early-strength pertinent to fly ash addition while fly ash com- ation whereas replacing cement with SCMs increased the carbon-
pensates for the workability issues pertinent to silica fume addi- ation depth. In both cases, however, the incorporation of SCMs
tion. The combinations of low dosage (3–6%) of silica fume and significantly lowered the total chloride content in concrete at all
moderate dosage (20–30%) of fly ash (despite its high lime content) depths other than the very external surface layer, for unknown rea-
were found very effective in reducing expansion due to alkali-silica sons. Gonen and Yazicioglu [34] found that the partial replacement
reaction (ASR) and in mitigating sulfate attack. The ternary blends of cement by silica fume (10% by mass of binder) or by both silica
showed great improvements in reducing chloride penetration and fume and fly ash (10% and 20% respectively) significantly improved
such reduction in diffusivity continued to increase with age. Bene- the concrete’s resistance to wetting–drying, capillary absorption
fits of ternary mixes (with 30–50% class C fly ash and 6–10% silica and carbonation. Shi et al. [27] found that in the presence of di-
fume in place of cement) in improving the compressive strength luted chloride solutions, the partial replacement of cement by
and reducing the chloride permeability of concrete were also re- 50% GGBFS (followed by the 10% ultrafine fly ash or 10% metakao-
ported by Hariharan et al. [23]. Khan [18] reported that a ternary lin replacement) was most beneficial to the freeze–thaw resistance
mix (with 20% class F fly ash and 10% silica fume) reduced the vol- of mortar specimens whereas the use of ordinary fly ash and silica
ume of large pores in mortar by approximately 50%, relative to the fume exacerbated their freeze thaw damage, all of which may be
control mix. Uysal et al. [24] found that the partial replacement of linked to the physical microstructure of the hardened concrete.
cement by FA or GGBFS in self-compacting concrete (SCC) mark- Shashiprakash and Thomas [35] found that mortars containing suf-
edly improved the workability of fresh concrete and the water ficient low-calcium fly ashes (e.g., 25% fly ash + 75% cement) or rel-
impermeability of hardened concrete. Such replacements also led atively low levels of ultrafine fly ash (e.g., 8–16% by mass of binder)
to better resistance to chloride penetration, with the 60% ground can improve the sulfate resistance of mortars made with high tri-
granulated blast-furnace slag (GGBFS) replacement of cement calcium aluminate [C3A] Portland cement. If the mortars contained
being the best. Thomas and Bamforth [16] modeled the chloride moderate to high levels of high-calcium fly ash (CaO > 20%), then a
diffusion in concrete using data from long-term field and labora- low-C3A cement or ternary cement blends with relatively low lev-
tory studies and showed that the incorporation of fly ash and slag els of silica fume (e.g., 3–6% by mass of binder) was needed to en-
may have little influence on the early-age chloride resistance but sure high sulfate resistance. All of these should be attributable to
dramatic benefits after a few years of exposure. Mangat et al. the changes in the chemical composition of cement hydrates in-
[25] investigated the partial replacement of cement by pulverized duced by the presence of fly ash or silica fume. Mineral admixtures
fuel ash, slag, and microsilica and showed microsilica to be the may slow the rate of strength gain in concrete, but do not adversely
most effective in enhancing the corrosion resistance of rebar in affect the long-term concrete strength [36] or even improve its
concrete. The microsilica addition was found to greatly increase strength properties [25,37,38]. Concrete mixes with high-volume
the pore volume in cement paste yet greatly decrease the chloride fly ash or high-volume fly ash and ground slag demonstrate good
penetration. Güneyisi et al. [26] investigated the rebar corrosion in workability, high compressive strength, and excellent durability
concrete made of blended cements which contained various pro- (negligible carbonation and very low chloride penetration) [39].
portions of Portland cement clinker, blast furnace slag, natural
pozzolans, and limestone powder. Relative to the plain cement
concrete, the specimens with blended cements showed superior 2.1. FA
corrosion performance and generally longer time to corrosion
cracking, which correlated very well with the splitting tensile Fly ash (FA) is a byproduct of coal combustion in the generation
strength data. of electricity, i.e., a finely segregated residue captured from the flue
Nonetheless, Shi et al. [27] demonstrated that laboratory-fabri- gas at coal-fired power plants. Most FA particles are spherical and
cated high-quality concrete specimens featured unusually low amorphous, ranging in size between 10 and 100 lm. With increas-
chloride diffusivity (Ds values in the order of 1013 m2/s, vs. the ing energy costs and heightened concerns about the impact of con-
normally reported 1012 m2/s), as measured by an accelerated crete construction and maintenance activities on the environment,
chloride migration test. In such case, the partial replacement of ce- there has been an attendant increase in interest and research activ-
ment by mineral admixtures (class F or class N fly ash, ultrafine fly ity on the use of FA and other recycled materials in concrete,
ash, silica fume, metakaolin, or GGBFS) showed little to no benefits including those targeting ASR prevention [40–42]. The effective-
in decreasing their chloride permeability. This is likely linked to ness of FA in mitigating mortar bar expansion induced by potas-
the fact that the chloride diffusivity in such concrete was largely sium-acetate-based deicer solution was found to depend on the
determined by the coarse aggregates instead of the cementitious lime content of FA and its dosage level [43].
paste phase. For the same mixes, the mortar specimens showed The use of FA as a supplemental binder in concrete is common:
Ds values in the order of 1011 m2/s, and the vast majority of mixes 15 of the 72 million tons of fly ash produced in the US in 2006 were
with mineral admixtures (except the one with 20% class F fly ash used for this purpose [44] and many states have allowed the use of
and 5% silica fume) showed a significant reduction in chloride per- performance-specified (ASTM C 1157) cements that contain FA.
meability. Zhang et al. [28] reported that the partial replacement of The efficacy of a particular FA in this regard however is difficult
cement by slag (20–40% by weight of binder) markedly reduced to predict and no single index value or combination of values is
the chloride permeability of concrete, as measured by the ASTM an infallible predictor of its performance in concrete [45]. Follow-
C 1202-97 (rapid chloride permeability test) method. Such reduc- ing the provisions of ASTM C 618, fly ashes can be divided into two
tion increased with the slag content. However, the addition of fly primary classes, F and C, based on their chemical composition
ash into the ternary binder increased the chloride permeability of resulting from the type of the coal burned. Normally class F FA is
concrete specimens, implying a negative impact of the fly ash on produced from anthracite or bituminous coals, whereas class C
the microstructure of hardened concrete. FA is produced from lignite or sub-bituminous coals [46]. ASTM
There is existing research demonstrating the use of mineral C 618 also specifies another class, N, typically for natural pozzo-
admixtures to affect other aspects of concrete durability, such as lans. This classification system, based on the silica, alumina, and
effectively mitigating the ASR-induced damage in concrete [29– iron oxide content of the ash, only indirectly indicates how the
32]. Papadakis [33] found that replacing aggregates with supple- ash will behave as an ingredient in concrete. Additional character-
mentary cementitious materials (SCMs, e.g., silica fume, low- and istics of importance include the calcium oxide content, fineness,
128 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

crystalline structure, and loss on ignition (LOI, an indicator of car- concrete was found to provide better corrosion protection to steel
bon content) of the ash. rebar due to of its higher resistance to chloride penetration. Oh
There are numerous studies on the effect of FA addition on the et al. [64] also reported lower chloride threshold values with
durability of concrete. Hedegaard and Hansen [47] argued that increasing addition of FA, whereas Schiessl and Breit [65] and Alon-
replacing cement with FA is likely degrade the water-tightness of so et al. [66] reported higher or similar threshold values respec-
concrete, as they found that 1 kg of cement would have to be re- tively when replacing cement with FA. For a concrete mix with
placed by 3 kg of FA in order to maintain the same level of water w/cm ratio of 0.37, the addition of FA (35% cement replacement)
permeability of hardened concrete (at 28 days and 56 days). Wong and silica fume (27% cement replacement) reduced the chloride
et al. [48] tested notched mortar specimens and concluded that a diffusion coefficient from 3.48  1012 m2/s to 7.35  1013 m2/s
15% cement replacement by class F FA enhanced the bond strength and 1.01  1012 m2/s, but also reduced the pore water pH from
of mortar-aggregate interface and fracture toughness. At high 13.84 to 13.39 and 13.47, respectively [67]. Other researchers
replacement levels (45% and 55%) the FA addition reduced the [68,69] also reported the reduction of pH in the pore solution as
interfacial bond strength and fracture toughness at 28 days but a result of FA addition. The reduction of pore water pH may explain
such reductions were recovered at 90 days. The FA replacement the research finding that the chloride threshold decreased with
at all levels was found to increase the interfacial fracture energy. increasing FA content in concrete, whereas the improved resis-
Gebler and Klieger [49] found that the use of certain fly ashes de- tance to chloride diffusion may explain the enhanced protection
graded the freeze–thaw resistance of air-entrained concrete when of embedded steel by the FA-admixed concrete [63]. Saraswathy
cured at low temperature and showed no noteworthy influence for and Song [70] investigated the effect of admixing activated FA on
other conditions. The incorporation of FA in concrete generally re- the corrosion resistance of concrete and found that the FA addition
duced the resistance of air-entrained concrete to deicing scaling significantly improved the corrosion performance of concrete up to
and showed little benefit on its resistance to chloride penetration. a critical moderate replacement level (20–30%) and the chemical
This is attributable to the absorption of air-entraining agent by FA, activation of FA worked the best.
which led to undesirable changes in the physical microstructure of
hardened concrete. Parande et al. [50] reported that concrete made 2.2. UFFA
with a FA-based Pozzolana Portland cement featured higher 150-
day compressive strength, less rebar corrosion in normal water Ultra-fine fly ash (UFFA) a relatively new pozzolanic admixture
and in domestic sewage water, and reduced chloride permeability, and there are a limited number of studies on the effect of its addi-
relative to the ordinary Portland cement concrete (PCC). Bouzou- tion on the durability of concrete. It is processed from ordinary FA
baa et al. [51] reported that high-volume FA concretes (with the to obtain finer particles. UFFA has been shown to feature higher
FA ground and blended in cement or unground and separately pozzolanic activity than ordinary fly ashes, to greatly reduce the
added at the mixer) exhibited comparable or superior mechanical water demand and air content of concrete, and to produce concrete
properties and durability characteristics (except deicing salt scal- of higher strength and lower porosity [71,72]. Hossain et al. [73]
ing resistance), relative to the ordinary PCC made with ASTM Type found that the restrained mortars containing UFFA or ordinary
I cement. Pacheco-Torgal and Jalali [52] reported that mortar spec- class F fly ash had lower residual stress levels, less free shrinkage,
imens made with cement containing 20% fly ash exhibited much increased cracking age, and decreased creep effect, relative to the
higher compressive strength both before and after exposure to sul- control. The UFFA addition led to more pronounced delay in the
furic acid solution (pH = 0.7), relative to those made with sulfate- age of cracking and in the reduction in creep effect, relative to
resistant cements. the ordinary fly ash. Subramaniam et al. [74] observed ‘‘a signifi-
In general, FA addition in concrete is considered an effective cant reduction in the autogenous shrinkage and an increase in
measure to mitigate chloride-induced corrosion of rebar in con- the age of restrained shrinkage cracking’’ in the concrete admixed
crete. For instance, using FA blended cement is known to reduce with UFFA, relative to the control and the concrete admixed with
chloride permeability and improve sulfate resistance of concrete silica fume. An ‘‘increase in the age of restrained shrinkage crack-
[53]. Dhir et al. [54] used the equilibrium method and found that ing and a significant increase in the compressive strength’’ were re-
the chloride-binding capacity of cement paste increased with the ported with increasing UFFA addition or decreasing w/cm ratio.
increase in FA replacement level up to 50% and then declined at
67%. In the case of admixed chloride, the increase in chloride bind- 2.3. SF
ing due to the replacement of FA was also found [55–58]. An in-
crease in chloride binding may be mainly ascribed to the high Silica fume (SF) is typically a byproduct of manufacturing sili-
alumina content in FA [20,56], which results in the formation of con and ferrosilicon alloys, i.e., a finely segregated residue captured
more Friedel’s salt [59]. The increase in chloride binding could also from the oxidized vapor on top of the electric arc furnaces. Most SF
be ascribed to the production of more gel during hydration, which particles are amorphous and ultra-fine in size, averaging from 0.1
results in better physical adsorption of chloride [60]. Other to 0.5 lm, or approximately one hundredth the size of the average
researchers [55,56,59] also found that partial replacement of ce- cement particle. Owing to its extreme fineness, large surface area
ment with FA has a positive effect on the chloride binding when and high silica content, SF is a chemically reactive pozzolan and
the cement paste was exposed to a chloride environment. How- its use in cementitious systems has been specified by ASTM C 1240.
ever, Nagataki et al. [61] found that the 30% replacement of cement Partial replacement of cement by SF up to 10% did not reduce
with FA reduced the chloride-binding capacity of cementitious the workability of fresh concrete, but slump loss with time was ob-
material in the case of external chlorides. Ampadu et al. [62] found served to increase with replacement level at low w/cm ratios [75].
the partial replacement of cement by FA only showed noteworthy As such, the SF addition is often accompanied by the use of a
benefits in reducing the chloride diffusivity in cement paste at later superplasticizer, i.e., high-range water reducer. Cong et al. [76] re-
ages of curing and a 40% replacement level was the best. Thomas ported the partial replacement of cement by SF coupled with the
[63] reported that chloride threshold decreased with increasing superplasticizer addition to increase the compressive strength of
of FA content in marine exposure. These threshold values obtained concrete, which was largely attributed to the improved strength
were 0.7%, 0.65%, 0.5% and 0.2% acid soluble chloride (by mass of of its cement paste.
cementitious material) for concrete with 0%, 15%, 30% and 50% SF is known to considerably reduce the permeability of concrete
FA, respectively. Despite of these lower threshold values, FA by refining its microstructure via both chemical and physical
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 129

pathways, and thus greatly reduce the risk of rebar corrosion in respectively when replacing cement with slag. Cheng et al. [89]
concrete. In light of available literature, Song et al. [77] developed investigated the corrosion behavior of reinforced concrete pris-
and verified a microstructure model for SF cement concrete, con- matic beams subjected to sustained loadings (37% and 75% of the
sidering the role of SF in ‘‘pore size refinement and matrix densifi- ultimate load) and 3.5% NaCl solution and found the slag concrete
cation, reduction in content of Ca(OH)2 and cement paste- to exhibit lower corrosion rate for a given reduction percentage in
aggregate interfacial refinement’’. The model suggests that the flexural rigidity (relative to the control). The partial replacement of
fineness of SF notably affects the permeability of SF cement con- cement by GGBFS reduced the electrical charge passing through
crete and the desirable range of SF replacement is 8–12% by mass the concrete (during the rapid chloride penetration test) and the
of binder. Selvaraj et al. [78] recently reviewed the influence of sil- water permeability of concrete. This was the result of GGBFS react-
ica fume on the factors relevant to the corrosion of reinforcement ing with water and Portlandite to form extra C–S–H gel and more
in concrete, including chloride diffusion, carbonation, oxygen dif- refined microstructure. A study of slag concrete after 25 years of
fusion, pore solution pH, and electrical resistivity of concrete. The exposure in a marine tidal zone [92] confirmed the beneficial role
partial replacement of cement by silica fume has been reported of slag in dramatically reducing the chloride ion penetration, espe-
to reduce the alkalinity of the pore solution and the chloride-bind- cially at relatively high replacement levels (45–65%) and low w/cm
ing capacity of hardened cement paste [79]. The reduction in pore ratio (0.40).
solution pH is mainly due to the pozzolanic reaction between sili-
con dioxide and the Portlandite [80]. The reduction in chloride-
2.5. MK
binding capacity by silica fume addition has been reported by
other researchers [56,79,81], as silica fume reduces the amount
Metakaolin (MK) is a material obtained by calcining clay min-
of aluminate phases in concrete that are able to chemically bind
eral kaolinite between 500–800 °C in an externally fired rotary kiln
chlorides and produces C–S–H that seem to have lower chloride
so that it loses water through dehydroxilization (i.e., removal of
sorption than C–S–H generated from cement hydration [81]. These
chemically bonded hydroxyl ions). MK generally has particle size
mechanisms can lead to dramatic increase in the Cl/OH ratio in
finer than cement but not as fine as silica fume and features a 2D
the pore solution and may be responsible for the reduction in the
order in crystal structure. MK is a highly reactive aluminosilicate
chloride threshold value of steel in concrete [17,82]. Dotto et al.
pozzolan and its use in cementitious systems has been specified
[83] observed that the silica fume addition led to significant
by ASTM C 618.
improvements in the corrosion performance of the concrete as well
The MK addition was found to increase early-age (1–3 days)
as in the compressive strength of concrete, whereas Page and Hav-
flexural strength of concrete by as much as 60%; and the finer
dahl [84] observed slightly higher corrosion rates of steel in ce-
MK (with surface area 25.4 vs. 11.1 m2/g) showed higher reactivity
ment paste containing silica fume.
and led to greater strength especially for compressive strength of
concrete with low w/cm ratio (0.40). MK was found to consume
2.4. GGBFS
Portlandite via pozzolanic reactivity and produce more refined
pore structure in concrete [93]. The sulfate resistance of both
Ground granulated blast-furnace slag (GGBFS) is a byproduct of
non-air-entrained and air-entrained concrete was found to in-
making iron and steel, i.e., a fine powder grounded from the glassy,
crease with the MK replacement level (from 5% to 10% and 15%),
granular material that forms when molten iron blast furnace slag is
by measuring expansion of concrete prisms and compressive
air quenched with water or steam. GGBFS features fineness and
strength reduction of concrete tubes [94]. The replacement of ce-
specific surface area similar to cement particles [85] and contains
ment or sand by MK (10% or 20% by weight of cement) can greatly
very limited amount of crystals. GGBFS is highly cementitious in
reduce the chloride permeability, gas permeability and sorptivity
nature and its use in mortar and concrete has been specified by
of concrete, by decreasing the mean pore size and improving uni-
ASTM C 989. Partial replacement of cement by GGBFS up to 80%
formity of the pore size distribution [95]. Shekarchi et al. [96]
was observed to reduce the compressive strength of concrete dur-
found that a 15% MK replacement of cement led to 20% increase
ing the first 28 days while the later-age strength increased with the
in the compressive strength of hardened concrete and reductions
slag replacement up to 60% [86]. Hadj-sadok et al. [85] reported
in its water penetration, water absorption, gas permeability, elec-
that 50% cement replacement by a GGBFS with moderate hydraulic
trical conductivity, ionic diffusion, and 28-day ASR expansion by
activity reduced the compressive strength of hardened mortar,
50%, 28%, 37%, 450%, 47%, and 82%, respectively. Both steady-
even after 90 days of curing. Meanwhile, the chloride diffusivity
and non-steady-state chloride diffusion tests showed that the MK
in mortar at this replacement level was reduced, along with a finer
addition in Portland cement mortar tends to enhance the resis-
porosity and lower water absorption at old ages (90 and 360 days).
tance to chloride transport through both the hydrated cement ma-
Partial replacement of cement by GGBFS up to 80% has demon-
trix and the paste-aggregate interfacial transition zone [97]. The
strated to improve the corrosion performance of concrete and the
MK addition was also found to compensate for the low early-age
50% replacement level in concrete imparted the best corrosion
compressive strength of concrete containing GGBFS [86].
resistance, which featured the corrosion initiation time of steel re-
bar 3.2–3.8 times as long as the control (depending on the C3A con-
tent in cement) [87]. The effect of GGBFS addition on the sulfate 3. Measuring the chloride ingress into concrete
resistance of concrete was more complex, depending on the
replacement level and the cement composition [87]. Corrosive agents, liquid or gaseous, may penetrate the concrete
GGBFS was found to considerably improve the pore structure of through capillary absorption, hydrostatic pressure, or diffusion.
concrete, increase its chloride-binding capacity (by forming more The ingress of gases, water or ions in aqueous solutions into con-
Friedel’s salt), and reduce its chloride diffusivity [88,89]. While crete takes place through pore spaces in the cement paste matrix
the slag addition improves both chemical and physical binding of and paste-aggregate interfaces or microcracks. For the durability
chloride [56,87,90], it decreases the pH of the pore solution [89]. of concrete, permeability is believed to be the most important
The effect of partial cement replacement by slag on the chloride characteristic [98], which is related to its microstructural proper-
threshold value is still controversial, as Dhir et al. [91] reported ties, such as the size, distribution, and interconnection of pores
lower threshold values for slag concrete whereas Schiessl and Breit and microcracks [99]. For reinforced concrete structures exposed
[65] and Oh et al. [64] reported higher or similar threshold values to chloride-laden environments, the chloride permeability of
130 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

Table 1
Summary of chloride penetration test methods [109].

Test Method Considers chloride ion At a constant Unaffected by conductors in the Approximate duration of test procedure
movement temperature concrete
Long AASHTO T259 (salt Yes Yes Yes 90 day after curing and conditioning
term ponding)
Bulk diffusion Yes Yes Yes 40–120 days after curing and conditioning
(Nordtest)
Short RCPT (T277) No No No 6h
term Electrical Mi action Yes Yes No Depends on voltage and concrete
Rapid migration (CTH) Yes Yes No 8h
Resistivity No Yes No 30 min
Pressure penetration Yes Yes Yes Depends on pressure and concrete (but
potentially long)
Other Sorptivity -lab No Yes Yes 1 week incl. conditioning
Sorptivity-field No Yes Yes 30 min
Propan-2-ol counter- No Yes Yes 14 days with thin paste samples
diffusion
Gas diffusion No Yes Yes 2–3 h

concrete has been recognized as a critical intrinsic property of the and drives the chlorides into the concrete under both convection
concrete [100]. Recent research has shed some light on the rela- and diffusion [106]. Recently, there has been an increasing demand
tionship between chloride diffusivity in concrete and its pore for rapid, reliable methods for testing the chloride ion penetration
microstructure or chemical composition. For instance, Saeki et al. resistance in a particular type of concrete and for testing the corro-
[101] tested the chloride diffusion coefficient of mortars featuring sion risk of rebar in a particular environment.
various w/cm, dosages of FA and GGBFS, curing period, and carbon- In the last decades, electric field migration tests have become
ation level. The hydration products and pore microstructure of very popular as they can greatly accelerate the chloride ingress
mortars were also analyzed. With such laboratory data, they devel- into concrete. Rapid Migration Test (RMT) is a method to measure
oped universal equations to predict the chloride diffusivity as a the electrical migration of chloride from one compartment with a
function of capillary pore volume, the specific surface area of hard- chloride solution to the other that is chloride-free [107]. The aver-
ened cementitious material or the amount and Ca/Si ratio of age depth of chloride penetration is obtained by spraying a color-
C–S–H. imetric indicator on the sample, and the value is then divided by
The chloride ingress into concrete and other cementitious mate- the product of the applied voltage and migration time to rate the
rials is a complex phenomenon involving multiple mechanisms. As concrete permeability. Castellote et al. [104] developed a method
such, a wide array of tests have been developed and used to eval- to derive both Dns and Ds from the migration test by monitoring
uate the resistance of chloride ion penetration into concrete. In the conductivity of the solution in the anodic compartment (desti-
1997, Stanish et al. [102] conducted a literature review to synthesis nation solution that was initially distilled water). Nonetheless, this
the state of the art pertinent to testing the chloride penetration method may produce misleading results when used to test high-
resistance of concrete (Table 1). And since then, there have been quality concrete over a long time period, since the anolyte conduc-
new advances in improving the test methods. tivity is very sensitive to chemical changes induced by the electro-
There are two types of natural penetration experiments gener- chemical reactions at the anode and the leachates from the test
ally used to measure the chloride diffusion coefficients of concrete. specimen.
One is the steady-state diffusion tests, such as the Diffusion Cell Test Rapid Chloride Permeability Test (RCPT) is a method that re-
in which a concrete specimen is used to separate a chloride solu- cords the amount of charge passed through a concrete sample in
tion from a chloride-free solution and periodical measurements order to evaluate its permeability [108]. By introducing the con-
of the chloride ion content are conducted until a steady state con- cept of ion mobility, the similarity between diffusion and migra-
dition is achieved. The other is from non-steady state tests that in- tion enables the determination of an ion diffusion coefficient
volve the ponding or immersion of concrete specimen for a specific from the migration tests. For PCC and mortar with no or little min-
duration before measuring the chloride penetration depth or pro- erals admixed, it has been shown that the total charge passed is
file, such as the Salt Ponding Test [99]. The diffusion coefficients ob- strongly correlated with the integral chloride content of ponding
tained are known as effective (Deff) and steady-state (Ds), or test [109] and with the chloride diffusion coefficient obtained from
apparent (Dapp) and non-steady-state (Dns) diffusion coefficient, an accelerated chloride migration test [110,111]. The RCPT has
respectively [103,104]. The ponding test has been standardized been standardized as AASHTO T 277 and ASTM C 1202, involving
as AASHTO T 259 and ASTM C 1543, involving the laborious anal- the classification of chloride permeability of concrete based on
ysis of chloride content at various depths of the sample after the charge in the first 6 h, which again is not sufficient time to dif-
90 days of ponding, which apparently is not sufficient time for ferentiate high-quality concrete mixes. Furthermore, the electrical
high-quality concrete to produce meaningful chloride concentra- charge passed in the test is related to all ions in the pore solution,
tion profile. not just chloride ions [102]. In addition, RCPT is not suitable for
Natural penetration tests (based on ASTM standards) are very evaluating the chloride permeability of concrete with supplemen-
time-consuming, especially for measuring the chloride diffusivity tary cementing materials [100,112,113], since the results may be
in high-quality concrete mixes. The diffusion tests often take a biased due to the change in the chemical composition of the pore
minimum of 1–3 years of exposure in simulated weathering condi- solution [103,114].
tions before any service life modeling can be conducted [105]. One Accelerated Chloride Migration Test (ACMT) can be considered a
way to accelerate the ingress of chloride into concrete is to apply a modified version of RMT and RCPT, which periodically measures
pressure. There is little research on this method, which exposes the accumulative chloride ion concentration in the destination
one face of the concrete to the chloride solution under pressure compartment either by the potentiometric titration method [115]
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 131

or using a chloride sensor [110,111]. The test lasts until significant cations) to reach Clth at the rebar depth is defined as time-to-
chloride ion concentration is detected in the destination compart- corrosion, Ti. The Clth is an important parameter in modeling and
ment, which could be hours, days, or weeks depending on the predicting the Ti and subsequently in assessing the service life of
thickness and quality of the test specimen and the applied voltage. reinforced concrete in chloride-laden environments [90,123].
Cho and Chiang [115] investigated the chloride diffusivity in con- While the chloride ion (Cl) has only a small influence on pore
crete specimens with various w/cm ratio (0.35–0.65) and slag con- water pH, concentrations as low as 0.6 kg/m3 by weight of concrete
tent (0–70%). They found good and very poor correlation between have been projected to compromise steel passivity [124]. For stable
the charge passed and the non-steady-state diffusion coefficient pit growth to be sustained, the relative concentrations of aggres-
(Dns) obtained from the ponding test, for concrete with and with- sive Cl and inhibitive OH should be above a certain ratio, other-
out slag respectively. For both types of concrete, there was a linear wise repassivation will occur [67].
correlation between the steady-state diffusion coefficient (Ds) ob- The Clth data in published literature scatter over a wide range of
tained from the ACMT and the Dns obtained from the ponding test, values [125–127]. One main reason is the inherently heteroge-
suggesting the ACMT to be a reliable accelerated test method. If the neous microstructure of the multi-phase concrete matrix. Another
applied voltage is too high (e.g., 60 V), the Joule effect may lead to a possible reason is that the chloride threshold has different
higher value of electrical charge passed during RCPT or a higher Ds definitions and measurement methods [128–130]. The chloride-
value during ACMT, i.e., the temperature increase of the test solu- to-hydroxyl ionic concentration ratio ([Cl]/[OH]) has been
tions [92], which can be mitigated by substantially increasing the traditionally considered to be a more reliable indicator than the
volume of test solutions (e.g., from 250 mL to 4.5 L) [115]. Further- chloride concentration (often expressed as total chloride content
more, the geometric shape of the test cell and the resistivity of the by the weight of cement or concrete or free chloride concentration
concrete specimen could affect the test results [116,117]. Vivas in concrete pore solution), considering that the competition be-
et al. [118] investigated the chloride diffusivity in 19 concrete tween aggressive Cl and inhibitive OH governs the pitting and
mixes prepared with materials typically used in construction in repassivation of steel. Research in aqueous solutions has indicated
the state of Florida. They found that the RMT test results had sim- that for chloride-contaminated concrete the pitting corrosion
ilar or better correlation with the 364-day bulk diffusion test than occurs only above a critical [Cl]/[OH] ratio [130]. Through a
those from the RCPT or the surface resistivity test and were less af- probability simulation model, the threshold [Cl]/[OH] for corro-
fected by the presence of SCMs in concrete. sion of bare steel rods in high pH solutions was once predicted to
be 0.66 in the presence of oxygen bubbles attached to the steel and
1.4 in the case of air. Such result agreed favorably with experimen-
4. Challenges in assessing the durability of concrete from its
tal data. In the same model, it was concluded that the threshold ra-
chloride diffusivity
tio should be about 1.4 for typical reinforced concrete and in excess
of 3 for high quality concrete with minimal air voids [121]. A num-
The length of the corrosion propagation stage in concrete is usu-
ber of studies [52,90,131–134] exposed reinforcing steel bars to
ally found to be relatively short, typically a few years. As a result,
simulated concrete pore solutions and revealed that the threshold
much of the emphasis on achieving concrete durability of 75 years
[Cl]/[OH] ratio increased with increasing pH. Yu et al. [135]
or longer is put on achieving a long corrosion initiation stage [119],
tested the corrosion behavior of two types of steel rebars (sand-
which is a function of the chloride transport properties of concrete
blasted or pre-rusted) in simulated concrete pore solutions and
(usually the diffusion coefficient), the surface chloride content dic-
concluded that the Clth values derived from open circuit potential
tated by the environment, the concrete cover thickness, and the
(OCP) and linear polarization resistance (LPR) observations agreed
chloride threshold concentration determining the onset of active
well with each other, and can be expressed in either [Cl] or [Cl]/
corrosion. Note that Pettersson [120] argued that the propagation
[OH]. Note that the threshold Cl/OH ratio in mortars has reported
period could be as long as 50 years or more for high performance
higher results (1.17–3.98) than that found in synthetic pore solu-
concrete featuring high electric resistivity and very limited oxygen
tion (0.25–0.8) [126]. Ann and Song [129] argued that the ratio
availability.
of total chloride content to acid neutralization capacity, [Cl]/
Both the chloride ingress into concrete and the subsequent corro-
[H+], best presents the chloride threshold level since it takes into
sion initiation of rebar in concrete are complex processes, which are
account ‘‘all potentially important inhibitive (cement hydration
influenced by numerous factors (e.g., temperature cycles and wet–
products) and aggressive (total chloride) factors’’. The different
dry cycles experienced by field concrete). Costa and Appleton
methods used to assess the chloride content or its profile in con-
[121] exposed 54 concrete panels to marine environments for
crete contributed to the variability in reported Clth values. Tradi-
3–5 years, and the study examined three concrete mixes and five
tionally, the coring method is commonly used, which involves
exposure conditions (from the tidal to the atmospheric zone). It con-
acquisition of one or more cores from sound concrete between
cluded that both the chloride diffusion coefficient and the surface
reinforcements at the time of active corrosion initiation. The cores
chloride concentration were time-dependent, which has consider-
are sliced and analyzed for their chloride content, and the chloride
able implications in predicting the chloride ingress into concrete
content in the slice near the rebar depth is defined as Clth. Recently,
and the risk of rebar corrosion in concrete. Therefore, challenges
both experimental [136,137] and modeling [138–141] studies
are inherent in assessing concrete durability from its chloride diffu-
unraveled that chloride content at the top of the rebar trace was
sivity. Despite these concerns, the following sections provide a syn-
higher than that at the same depth away from the rebar, owing
thesis of information pertinent to the determination of chloride
to the relatively low content of coarse aggregates in the vicinity
threshold and the quantification of chloride binding effect.
of the rebar [142] and the rebar serving as a physical barrier to
chloride migration. Thus, it is more reasonable to express Clth with
4.1. Chloride threshold the chloride content on the top of rebar trace than that acquired
from the core sample. Yu and Hartt [143] reported that the pres-
Chloride threshold of rebar in concrete, Clth, can be defined as ence of rebar affects the measured Clth and Ti and developed a
the content of chloride at the depth of the rebar that is necessary three-shell chloride diffusion model to account for such effects in
to sustain localized breakdown of its passive film and hence initi- the rebar vicinity zone. They found that rebar ‘‘with larger size,
ate its active corrosion [122]. The time it takes for chloride ions better corrosion resistance, less cover thickness, or a more flat
from external sources (marine environments or deicing salt appli- leading edge causes more pronounced Ti reduction’’.
132 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

The lack of universally accepted chloride threshold value is also found to be near three times the free chloride diffusivity [15]. In
attributable to the numerous factors that affect steel corrosion in a Florida DOT study, the relationship between bound and free chlo-
concrete, such as: the pH of concrete pore solution, the electro- rides was found to follow the Langmuir adsorption isotherm [67].
chemical potential of the steel, and the physical condition of the While previous studies [60,157] suggest that only free chloride
steel/concrete interface. Shi et al. [144] conducted a set of labora- ions in the pore solution are responsible for initiating corrosion
tory tests to assess the corrosion potential (Ecorr) and pitting poten- of the steel, Glass et al. [125,158,159] indicated that bound chlo-
tial (Epit) of steel coupons in simulated concrete pore solutions. The rides may also present a significant risk to steel. One possible rea-
data were then used to establish a phenomenological model corre- son is that a large part of bound chlorides are released as soon as
lating the influential factors with the pitting risk (characterized by the pH drops to values below 12 [6].
Ecorr–Epit). The model predictions indicate that the threshold [Cl]/ Chloride binding further complicates the determination of the
[OH] of steel rebar in simulated concrete pore solutions is a func- threshold [Cl]/[OH] to initiate corrosion of steel in concrete
tion of dissolved oxygen concentration, pH and chloride binding, [160], and the chloride binding and pH of pore solution are two in-
instead of a unique value. The pH of concrete pore solution mainly ter-related parameters. It has been observed that the pH of NaCl-
depends on the type of cement and additions and the carbonation containing alkaline solution increases as the chloride binding in-
level of concrete [145–148]. The potential of the steel is not only creases [154]. Reducing the pH in concrete may destabilize the
related to the steel type and surface condition (e.g., roughness) chloroaluminate and thus reduce the percent of bound chloride
but also the oxygen availability at the steel surface; the latter is af- [10,129], and carbonation of concrete can reduce the chloride-
fected by the moisture content of concrete [66,133,138]. The phys- binding capacity [130,161] and facilitate the chloride intrusion
ical condition of the steel/concrete interface (especially entrapped [162]. Nonetheless, above pH 12.6 chloride binding evidently de-
air void content) was found to be more dominant in the Clth than creases with increasing OH and an increase in pH can thus result
chloride binding or buffering capacity of cement matrix or binders in higher [Cl]/[OH] [163].
[129]. Voids that can be normally found in real structures due to The chloride-binding capacity is affected by numerous factors,
incomplete compaction may weaken the layer of cement hydration such as the C3A and alkali contents of cement [164], use of mineral
products deposited at the steel/concrete interface and thus may fa- admixtures [15,54–61,79,81,88,89], cation of the chloride salt
vor local acidification required for sustained propagation of pits. [165], temperature, and degree of hydration [156]. A popular
The presence of air voids, as well as crevices and microcracks, method of measuring chloride-binding capacity was introduced
may decrease the chloride threshold [149,150,142,151]. In addi- by Tang and Nilsson [166], which uses crushed specimens for test-
tion, the presence of sulfate ions, the temperature and the concrete ing and then calculates the chloride-binding capacity of a cementi-
mix proportions and quality may affect the chloride threshold tious material to the amount of cement in it [167]. However,
[127,128,134,151]. Li and Sagüés [134] listed a wide array of inter- Jiricková and Cerný [167] demonstrated that this method provides
nal and external factors defining the Clth, such as: the composition, a upper limit to the actual chloride-binding capacity, since the
surface condition, and configuration of rebar; the concrete chemis- measurement is affected by the size of the porous specimen and
try (type and amount of cement and admixtures, type and porosity a real concrete specimen would have bigger pores and lower spe-
of aggregates, w/cm ratio, degree of hydration, etc.); the type and cific surface than a crushed specimen.
source of chloride; and the service environment (humidity, tem-
perature, cracking of concrete, etc.). Angst et al. [128] summarized 5. Service life modeling of steel reinforced concrete in chloride
the state of the art in the Clth research in a recent review. environments
Furthermore, it has been argued that the Clth (and Ti) should be
treated as a distributed parameter represented by a probability Through the use of concrete deterioration models, cost-effective
function, in light of the statistical nature of the processes involved decisions can be made concerning the appropriate time to repair or
(e.g., chloride ingress and pitting initiation) and the inherent heter- replace existing structures, and the most effective corrosion con-
ogeneities of the concrete matrix [134,142,151]. Li and Sagüés trol strategies. There are software packages available to aid in con-
[134] suggested the incorporation of a Clth variability term in the crete modeling and service life forecasting, such as 4SIGHT and
service life prediction procedures and durability models. Hartt CONLIFE available from the National Institute of Standards and
and Nam [145] reported a range of values for Clth and Ti with seem- Technology [169], Life-365™ (www.life-365.org/), as well as con-
ingly identical slabs and the same exposure condition, as variabil- ventional Finite Difference Method (FDM) applications. Khatri
ity was introduced by microstructural factors such as the size and and Sirivivatnanon [170] proposed a model for predicting the ser-
distribution of entrapped air voids, and the arrangement of aggre- vice life of reinforced concrete structures, where the acceptable le-
gates which can markedly affect the tortuosity of chloride ingress vel of deterioration is related to the presence of chloride ions on
path. Such effect of non-diffusive coarse aggregates on the reported the rebar surface. As such, the service life was defined as the time
variability in Clth and Ti was further examined by Yu and Hartt required for transport processes to raise the chloride content at the
[152], using a modeling approach. Currently there is limited re- depth of the rebar to the threshold level for pitting corrosion. It
search on the probabilistic nature of how mix design and other fac- should be mentioned that the service life is not a fixed value as cal-
tors impact corrosion initiation (as indicated by Clth and Ti) culated by a deterministic model, but instead it is a range of values
[145,138,139]. determined by material characteristics, cover depth, and severity
of service environments.
4.2. Chloride binding The diffusion of chloride ions in water-saturated cementitious
materials is a complex process involving various physical and
In concrete, chlorides can exist either in the pore solution, chemical interactions. The chloride ions can be bound either phys-
chemically bound with concrete C3A (tricalcium aluminate) or ically or chemically by cement paste, thereby lowering the fraction
C4AF phases (e.g., Friedel’s salt, 3CaOAl2O3CaCl210H2O), or phys- of free chlorides that can diffuse freely in the pore solutions. In
ically held to the surface of hydration products (e.g., adsorption on addition, the internal electric field formed by the cations and anions
C–S–H) [153–155]. Chloride binding removes chloride ions from will speed up the ions that have low diffusion coefficients and slow
the pore solution, and slows down the rate of penetration [156]. down the ions that have high diffusion coefficients to maintain the
With water-soluble and acid-soluble chlorides referred to as free electro-neutrality condition. If we ignore such complex internal
and total chlorides respectively, the total chloride diffusivity was processes and treat the diffusion problem phenomenologically,
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 133

the temporal and spatial evolution of chloride-ion concentration the chloride ingress into concrete and the service life model
can be calculated based on Fick’s second law in Eq. (1) [171,172], developed for a wet–dry environment was then verified with data
which has been used extensively by many researchers to calculate obtained from a concrete pier exposed in the marine environment
the chloride concentrations for various concrete cover depths at dif- for 22 years. Hong and Hooton [178] exposed concrete specimens
ferent exposure time intervals: containing slag and/or SF to various wetting–drying regimes and
found that ‘‘longer drying times increase the rate of chloride in-
@C @2C gress’’. Boddy et al. [179] developed a multi-mechanistic transport
¼ Dc 2 ð1Þ
@t @x model for chloride penetration which considers diffusion, perme-
where C is the concentration (mol m3); t is the time (s); D is the ability, chloride binding and wicking as well as the time-depen-
chloride diffusion coefficient (m2 s1); v is the position (m). In most dent nature of concrete properties (e.g., diffusivity). Marchand
work, the service life or the length of the corrosion initiation stage is and Samson [180] argued that ‘‘using Fick’s law of diffusion to ana-
approximated with the following simplified assumptions: lyze chloride ingress in concrete structure is inappropriate’’ and
emphasized ‘‘the dependence of the chloride apparent diffusion
 The concrete is initially chloride-free, and the concrete acts as a coefficient on the external environment’’. They also discussed the
physical barrier to protect the rebar. The rebar corrosion is trig- limitations of using simplified models for the design of new con-
gered only when the concrete in contact with the steel becomes struction and rehabilitation of existing structures.
contaminated with chloride ions exceeding a threshold concen- Concrete is a multiphase porous composite consisting of cement
tration value. paste and aggregates, in both of which phases pores exist. Chloride
 Chloride ions progress inward from the external surface of the ions can only diffuse in the pore solutions. Effective diffusion coef-
concrete, which is covered by aqueous solutions of chlorides. ficients of chloride, characterizing the resistance of concrete to dif-
Therefore, the concrete immediately below the surface acquires fusive ingress of chlorides, are therefore considered to be a
a surface chloride concentration that remains unchanged in the function of the characteristics of the cement paste and the aggre-
simulation. gates. The rate at which chloride ions penetrate into water-satu-
 Chloride ions progress inward by simple diffusion, driven by the rated concrete depends on the diffusion coefficients in cement
gradient of the concentration of chloride ions in the concrete. paste and aggregates as well as the aggregates fraction. Addition-
The effective diffusion coefficient is constant with time and ally, the rate of ingress is also influenced by the cement paste/
space, and is a property of the concrete between the concrete aggregate interfacial (ITZ) zones and internal cracks. Prediction of
surface and the steel rebar. effective chloride diffusion coefficients in concrete based on its
mixture proportions is needed for service life evaluation. If the
Based on these assumptions, an analytical solution exists to pre- roles of ITZ and internal cracks on chloride ingress can be ignored,
dict the spatial and temporal evolution of chloride concentration the effective diffusion coefficient of chloride in water-saturated
profiles in concrete, which is given by [170,171,173]: concrete can be calculated by the following equation:
   ½ðDa  Dp ÞV a þ ðDp þ Da ÞDp
x Dc ¼ ð4Þ
C t ¼ C s 1  erf pffiffiffiffiffiffiffi ð2Þ ðDp þ Da Þ þ ðDp  Da ÞV a
2 Dc t
where x is the concrete cover depth; Ct is the chloride concentration where Dp and Da are the chloride ion diffusion coefficients in the ce-
at cover depth; Cs is the surface chloride concentration; t is time; Dc ment paste and aggregates, respectively, and Va is the aggregate vol-
is the effective diffusion coefficient in concrete; erf is the Gaussian ume fraction.
error function as below: If the aggregates have a lower diffusion coefficient than the ce-
Z ment paste, the concrete will have a lower diffusion coefficient
z
2 2 than the cement paste. Therefore, chloride ion ingress will decrease
erf ðzÞ ¼ pffiffiffiffi et dt ð3Þ
p 0 with the increase in the aggregate volume, and controls are neces-
sary on the quantity of cement paste and aggregates to achieve a
Eq. (2) can be used to calculate the service life of a reinforced con- specified service life. The diffusion coefficient of chloride ions in
crete structure, provided that Ct, Cs, Dc and x are known. aqueous solutions [181], cement paste [182] and marble [183]
The application of Fick’s law and a constant diffusivity of the has the order of 109 m2/s, 1012 m2/s and 1015 m2/s, respec-
concrete assume the concrete to be fully saturated and free of tively. Accordingly, it can be assumed that diffusion of chloride
time-dependent cover cracking. In the field environment, however, ions in aggregates of concrete is negligible, leading to:
the reinforced concrete generally experiences wetting and drying
cycles and diffusion may not be the main mechanism of chloride ð1  V a Þ
Dc ¼ Dp ð5Þ
ingress. Spragg et al. [174] revealed that wetting and drying behav- ð1 þ V a Þ
iors of concrete exposed to deicing salt solutions differed from
those exposed to water. Castro et al. [175] studied chloride concen- For instance, at Va = 0.7, the effective diffusion coefficient predicted
tration profiles from field exposed concrete and revealed that ‘‘the from Eq. (5) is:
environmental conditions of the tropical marine climate in cylin- Dc ¼ 0:1765Dp ð6Þ
drical concrete specimens promoted the formation of two zones:
one internal that is always dampened and one external that is Ti and Clth are important service life determinants for reinforced
always wetting and drying’’. Meijers and Schlangen [176] pre- concrete structures in chloride-laden environments. Numerous lit-
sented a 2D chloride ingress model that couples the chloride trans- eratures have discussed, through experimental or modeling ap-
port with heat and moisture transport. The model simulates the proaches, these two determinants in conventional PCC and their
concrete exposed to the splash and tidal zone where wetting and relationships with the type of cement and reinforcement, mix de-
drying occur concurrently and found that ‘‘the drying above the sign, exposure conditions, and other factors. It should be cautioned,
water-level enhances the chloride ingress below the water-level, however, that these relationships feature a probabilistic nature
and that accumulation of chloride ions occurs just above the [151,184]. Kirpatrick et al. [173] incorporated the probabilistic con-
water-level behind the drying face’’. Guimaraes and Helene [177] siderations of chloride-induced rebar corrosion into a service life
confirmed that the saturation degree of concrete notably affects model. The model predicts the time to first repair and subsequent
134 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

rehabilitation as a function of surface chloride concentration, crack initiation and propagation. Chen [196] also suggested future
apparent chloride diffusion coefficient, and clear cover depth, using improvements for the model could consider additional deteriora-
data collected from 10 Virginia concrete bridge decks exposed to tion processes such as freeze–thaw cycling, sulfate attack, alkali-
chloride deicers. aggregate reactions, and leaching of concrete constituents.
Williamson et al. [185] validated the probabilistic model that
predicts the time to first repair (2% deterioration) of bridge decks
under various chloride exposure conditions, using data gathered 6. Concluding remarks
from 10 bare steel bridge decks constructed in Virginia between
1965 and 1968. Sagüés and Kranc [186,187] developed service life Chloride ingress into concrete is a complex process, which in
models for marine substructures that incorporate statistical stoch- many environments is further complicated by the temperature cy-
astics and take preexisting cracks, strong localized corrosion spots, cles and wet–dry cycles experienced by the reinforced concrete
and/or structure geometry into account. Liu and Shi [168] devel- structures. While there are numerous existing experimental or
oped a 2D finite element method (FEM) model that considers the modeling studies related to the transport of chlorides in concrete,
stochastic nature of inputs (surface chloride concentration, con- the measurement of chloride ingress into concrete is technically
crete cover depth, chloride diffusion coefficient, and chloride challenging. Furthermore, it is also difficult to assess the durability
threshold) as well as the interactions of multiple ionic species dur- of reinforced concrete from its chloride diffusivity, partially owing
ing their transport in concrete. The model predicts the service life to the heterogeneous nature of the concrete matrix, the difficulty
of steel reinforced concrete structures as a function of mix design, in fabricating concrete in a reproducible manner, and the inher-
surface chloride concentration, cracking level, amount of coarse ently probabilistic nature of species transport in concrete and of
aggregates, and concrete cover depth. A recent study confirmed chloride-induced corrosion of rebar in concrete. The processes
and analyzed the probabilistic distribution of Ti and Ct for both reg- and procedures used in the new construction should be closely
ular concrete and SCC slabs, and the air voids at the rebar-concrete supervised under a systematic quality assurance program, in order
interface were also found to contribute the uncertainty inherent in to achieve the great potential of reinforced concrete as a construc-
the service life of reinforced concrete in chloride-laden environ- tion material and to manage corrosion risks pro-actively. The
ments [151]. importance of good construction and curing practices cannot be
Note that the models discussed above assume rapid propaga- overemphasized, as they greatly reduce the risk of cracking and re-
tion of rebar corrosion once it is initiated in the concrete and thus bar corrosion in concrete.
focus only on the initiation period for chloride-induced corrosion. Modeling is a useful tool to provide quantitative understanding
Pettersson [188] argued that the propagation period could be as of key processes and their interactions that define the service life
long as 50 years or more for high performance concrete featuring of reinforced concrete in chloride-laden environments. More
high electric resistivity and very limited oxygen availability. There improvements could be made to the existing service life models
are also service life models that consider both the initiation and for reinforced concrete so that they can better represent or simu-
propagation of rebar corrosion, such as the 1D stochastic corrosion late the field behavior of reinforced concrete structures. This could
damage progression models for black bars and for corrosion-resis- be accomplished by taking into account the time-dependency of
tant bars presented by Hartt [189,190], the 3D chemo-thermo- transport properties of concrete [16,197], the repair or replace-
hygro-mechanical model presented by Ožbolt et al. [191] and the ment of concrete cover [198], the corrosion propagation [120],
time-to-cracking and overall deterioration model presented by the chloride penetration mechanisms other than diffusion (e.g.,
Broomfield [192]. A recent review on the prediction of corrosion wicking) [179], the structure geometry [186], the environmental
rate in reinforced concrete structures was provided by Otieno humidity and temperature fluctuations and the decay of structures
et al. [193]. under coupled physical, chemical, and mechanical deterioration
The service life of reinforced concrete in the field environment processes [195,196], etc. With continued improvements on such
can be very difficult to predict or model, as it depends on not only service life models, they could also be used for life cycle costing
the chloride-related durability but also on mechanical properties and for the timing of repair or rehabilitation strategies.
of the concrete. For instance, the cracking of concrete (due to envi- The use of mineral admixtures is promising in light of their
ronmental or mechanical stresses) can greatly affect the ingress of potential benefits on the durability and sustainability of reinforced
chlorides and other deleterious species from the surrounding envi- concrete. When used to partially replace cement, such materials as
ronment and the durability and serviceability of the concrete. fly ash, silica fume, metakaolin and slag have shown to generally
Francois et al. [194] conducted a corrosion study by exposing loaded improve the resistance of concrete to chloride penetration. The
reinforced concrete beams to salt fog for 12 years. They found that use of fly ash, slag, and other industrial byproducts may translate
the tensile micro-cracking in concrete as a result of service loading to cost savings and reduced energy use, greenhouse gas emissions
was responsible for accelerating chloride penetration whereas the and landfill waste, without sacrificing quality and long-term per-
existence and width of microcracks (with width less than 0.5 mm) formance of the concrete. Additional research, however, is needed
did not influence the development of the rebar corrosion. to validate the use of such modified mixes for corrosion mitigation
Lin et al. [195] presented a deterministic service life FEM model in chloride environments.
of reinforce concrete structures, which incorporates chloride bind- Currently, there is a considerable amount of variability in deter-
ing, diffusion and convection, environment humidity and tempera- mining chloride diffusion coefficients as an indicator of the
ture fluctuations, and the decay of structural performance. As such, durability of both conventional and unconventional concrete struc-
the model considers the interactions between the transport of tures. In particular, the chloride diffusivity depends on the con-
moisture and chloride and the decay of structures under coupled crete pore structure and on all the factors that determine it, such
physical, chemical, and mechanical deterioration processes. Chen as: mix design parameters (w/cm ratio, type and proportion of
[196] presented a probabilistic service life FEM model that also con- mineral admixtures and cement, compaction, curing, etc.) and
siders the coupled deterioration processes, including heat transfer presence of cracks. The chloride diffusion coefficient is also a func-
and associated expansion and contraction, moisture transport and tion of chloride exposure condition (submerged, splash, atmo-
associated wetting expansion and drying shrinkage, carbon dioxide sphere, etc.) and the length of exposure, partly due to hydration
transport and associated carbonation, chloride penetration and of slowly reacting cementitious constituents such as blast furnace
associated rebar corrosion and rust expansion, and subsequent slag or fly ash. More testing and research are warranted in order to
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 135

address some of the existing knowledge gaps and to shed light on [21] Hossain AB, Shrestha S, Summers J. Properties of concrete incorporating
ultrafine fly ash and silica fume. Transp Res Rec: J Transp Res Brd
the fundamental relationships.
2009;2113:41–6.
Existing chloride permeability tests are either very time-con- [22] Thomas MDA, Shehata MH, Shashiprakash SG, Hopkins DS, Cail K. Use of
suming for high quality concrete mixes or too biased to provide ternary cementitious systems containing silica fume and fly ash in concrete.
reliable chloride diffusion coefficients. Additional research is thus Cem Concr Res 1999;29:1207–14.
[23] Hariharan AR, Santhi AS, Mohan Ganesh G. Effect of ternary cementitious
needed to establish standard, reliable, and rapid test methods for system on compressive strength and resistance to chloride ion penetration.
determining chloride diffusion coefficients and chloride thresholds Int J Civil Struct Eng 2011;1(4):696–706.
in unconventional concrete. Such methods are anticipated to help [24] Uysal M, Yilmaz K, Ipek M. The effect of mineral admixtures on mechanical
properties, chloride ion permeability and impermeability of self-compacting
the transition from prescriptive specifications of concrete mixes concrete. Constr Build Mater 2012;27(1):263–70.
to more performance-based specifications, which then would al- [25] Mangat PS, Khatib JM, Molloy BT. Microstructure, chloride diffusion and
low more innovation and flexibility in the materials selection of reinforcement corrosion in blended cement paste and concrete. Cem Concr
Compos 1994;16(2):73–81.
concrete and likely facilitate the paradigm shift from conventional [26] Güneyisi E, Özturan T, Gesolu M. Performance of plain and blended cement
PCC to environmentally friendly concretes. concretes against corrosion cracking. In: Konsta-Gdoutos MS, editor.
Measuring, monitoring and modeling concrete properties. Netherlands:
Springer; 2006. p. 189–98.
Acknowledgements [27] Shi X, Yang Z, Liu Y, Cross D. Strength and corrosion properties of Portland
cement mortar and concrete with mineral admixtures. Constr Build Mater
The first author acknowledges the financial support provided by 2011;25(8):3245–56.
[28] Zhang D, Li X, Ma X, Wang Z. Effects of mineral admixtures on the chloride
the California Department of Transportation and the Research & permeability of hydraulic concrete. Adv Mater Res 2011;8:168–70.
Innovative Technology Administration at the US Department of [29] Xu GJZ, Watt DF, Hudec PP. Effectiveness of mineral admixtures in reducing
Transportation for this work. Gratitude is also due to Dr. Yajun ASR expansion. Cem Concr Res 1995;25(6):1225–36.
[30] Malvar LJ, Cline GD, Burke DF, Rollings R, Sherman TW, Greene JL. Alkali-silica
Liu for his assistance with the modeling section. reaction mitigation: state of the art and recommendations. ACI Mater J
2002;99:480–9.
[31] Taha B, Nounu G. Using lithium nitrate and pozzolanic glass powder in
References concrete as ASR suppressors. Cem Concr Compos 2008;30:497–505.
[32] Shehata MH, Thomas MDA. The role of alkali content of Portland cement on
[1] Lomborg B. The skeptical environmentalist: measuring the real state of the the expansion of concrete prisms containing reactive aggregates and
world. Cambridge, United Kingdom: Cambridge University Press; 2001. p. supplementary cementing materials. Cem Concr Res 2010;40(4):
512–40. 569–74.
[2] Mehta PK. Reducing the environmental impact of concrete. Concr Int [33] Papadakis VG. Effect of supplementary cementing materials on concrete
2001;23(10):61–6. resistance against carbonation and chloride ingress. Cem Concr Res
[3] Van Dam TJ, Smartz BW. Use of performance specified (ASTM C1157) cements 2000;30(2):291–9.
in Colorado transportation projects: case studies. In: TRB 89th annual [34] Gonen T, Yazicioglu S. The influence of mineral admixtures on the short and
meeting compendium of papers DVD, Transportation Research Board, long-term performance of concrete. Build Environ 2007;42:3080–5.
Washington DC, 2010. [35] Shashiprakash SG, Thomas MDA. Sulfate resistance of mortars containing
[4] ASCE. Report card for America’s infrastructure. American Society of Civil high-calcium fly ashes and combinations of highly reactive Pozzolans and fly
Engineers. <http://www.infrastructurereportcard.org/> [accessed 01.06.10]. ash. ACI Special Publication, vol. 199; 2001. p. 221–38.
[5] Samples L, Ramirez J. Methods of corrosion protection and durability of [36] California Department of Transportation. Memo to designers 1–5: protection
concrete bridge decks reinforced with epoxy-coated bars, phase I, FHWA/IN/ of reinforcement against corrosion due to chlorides, acids and sulfates; 2002.
JTRP-98/15. Purdue University, IN, 1999. [37] Mehta PK. Pozzolanic and cementitious by-products in concrete – another
[6] Bertolini L, Elsener B, Pedeferri P, Polder R. Corrosion of steel in concrete: look. In: Malhotra VM, editor. Proceedings of the third CANMENT/ACI
prevention, diagnosis, repair. KgaA, Weinheim: Wiley-VCH, Verlag GmbH & international conference, American Concrete Institute, Michigan, ACI SP
Co.; 2004. 114, vol. 1; 1989. p. 1–43.
[7] Morefield SW, Hock VF, Weiss Jr CA, Malone PG. Application of electrokinetic [38] Mehta PK. Role of pozzolanic and cementitious material in sustainable
nanoparticle migration in the production of novel concrete-based composites, development of concrete industry. In: Malhotra VM, editor. Proceedings of
December 2008. <http://www.oai.dtic.mil/oai/oai?verb=getRecord&meta the sixth CANMET/ACI/JCI, international conference on fly ash, silica fume and
dataPrefix=html&identifier=ADA504199> [accessed 19.10.10]. natural Pozzolans in concrete, American Concrete Institute, Michigan, ACI SP-
[8] Cardenas HE, Syed F. Rapid electrokinetic nanoparticle transport in concrete. 178, vol. 1; 1998. p. 1–20.
In: Grantham M, Salomoni V, Majorana C, editors. Concrete solutions. CRC [39] Collepardi S, Corinaldesi V, Moriconi G, Bonora G, Collepardi M. Durability of
Press; 2009. doi: 10.1201/9780203864005.ch1. high-performance concretes with pozzolanic and composite cements, ACI
[9] Cardenas HE, Struble LJ. Electrokinetic nanoparticle treatment of hardened special report SP 192-10, American Concrete Institute; 2000.
cement paste for reduction of permeability. J Mater Civil Eng [40] Malvar LJ, Lenke LR. Efficiency of fly ash in mitigating alkali silica reaction
2006;18(4):554–60. based on chemical composition. ACI Mater J 2006;103(5):319–26.
[10] Hartt W, Charvin S, Lee S. Influence of permeability reducing and corrosion [41] Lenke LR, Malvar LJ. Alkali silica reaction criteria for accelerated mortar bar
inhibiting admixtures in concrete upon initiation of salt induced embedded tests based on field performance data. In: Proceedings of the world of coal
metal corrosion, prepared for the Florida Department of Transportation; ash, KY, May 4–7, 2009.
1999. [42] Malvar LJ. ASR prevention in DOD. In: TRB 89th annual meeting compendium
[11] California Department of Transportation, bridge design specifications, section of papers DVD, Transportation Research Board, Washington DC, 2010.
9 – reinforced concrete; 2003. [43] Rangaraju PR, Desai J. Effectiveness of fly ash and slag in mitigating alkali–
[12] Toutanji HA, Delatte D. Supplementary materials to enhance bridge deck silica reaction induced by deicing chemicals. J Mater Civil Eng
durability, prepared for University Transportation Center for Alabama; 2001. 2009;21:19–31.
[13] Gruber KA, Ramlochan T, Boddy A, Hooton RD, Thomas MDA. Increasing [44] ACAA. American Coal Ash Association; 2007. <http://www.acaa-usa.org/>.
concrete durability with high-reactivity metakaolin. Cem Concr Compos [45] Joshi RC, Lohtia RP. Fly ash in concrete: production, properties and uses.
2001;23(6):479–84. Advances in concrete technology, vol. 2. Amsterdam, Netherlands: Gordon
[14] Choi Y-S, Kim J-G, Lee K-M. Corrosion behavior of steel bar embedded in fly and Breach Science Publishers; 1997.
ash concrete. Corros Sci 2006;48(7):1733–45. [46] Scheetz BE, Olanrewaju J. Determination of the rate of formation of
[15] Lu X, Li C, Zhang H. Relationship between the free and total chloride hydroceramic waste forms made with INEEL calcined wastes, final report
diffusivity in concrete. Cem Concr Res 2002;32(2):323–6. prepared for the Department of Energy, 2001.
[16] Thomas MDA, Bamforth PB. Modelling chloride diffusion in concrete: effect of [47] Hedegaard BE, Hansen TC. Water permeability of fly ash concretes. Mater
fly ash and slag. Cem Concr Res 1999;29(4):487–95. Struct 1992;25:381–7.
[17] Manera M, Vennesland Ø, Bertolini L. Chloride threshold for rebar corrosion [48] Wong YL, Lam L, Poon CS, Zhou FP. Properties of fly ash-modified cement
in concrete with addition of silica fume. Corros Sci 2008;50(2):554–60. mortar-aggregate interfaces. Cem Concr Res 1999;29:1905–13.
[18] Khan MI. Nanostructure and microstructure of cement concrete incorporating [49] Gebler SH, Klieger P. ‘Effect of fly ash on the durability of air-entrained
multicementitious composites. Transp Res Rec: J Transp Res Brd concrete. In: Proceedings of the 2nd international conference on the use of fly
2010;2141:21–7. ash, silica fume, slag, and other mineral by-products in concrete, April 21–25,
[19] Yang CC, Cho SW. An electrochemical method for accelerated chloride 1986, Madrid, Spain. ACI Publication SP-91.
migration test of diffusion coefficient in cement-based materials. Mater Phys [50] Parande AK, Ramesh Babu B, Pandi K, Karthikeyan MS, Palaniswamy N.
Chem 2003;81(1):116–25. Environmental effects on concrete using ordinary and pozzolana Portland
[20] Dhir RK, Jones MR. Development of chloride-resisting concrete using fly ash. cement. Constr Build Mater 2011;25:288–97.
Fuel 1999;78(2):137–42.
136 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

[51] Bouzoubaa N, Zhang MH, Malhotra VM, Golden, DM. Mechanical properties [85] Hadj-sadok A, Kenai S, Courard L, Darimont A. Microstructure and durability
and durability of laboratory produced high-volume fly ash blended cements. of mortars modified with medium active blast furnace slag. Constr Build
ACI Special Publication, vol. 199; 2001. p. 55–82. Mater 2011;25:1018–25.
[52] Pacheco-Torgal F, Jalali S. Sulphuric acid resistance of plain, polymer modified, [86] Khatib JM, Hibbert JJ. Selected engineering properties of concrete
and fly ash cement concretes. Constr Build Mater 2009;23:3485–91. incorporating slag and metakaolin. Constr Build Mater 2005;19(6):
[53] Saraswathy V, Song H-W. Corrosion performance of fly ash blended cement 460–72.
concrete: a state-of-art review. Corros Rev 2006;24(1–2):87–122. [87] Al-Gahtani AS, Rasheeduzzafar, Al-Saadoun SS. Rebar corrosion and sulfate
[54] Dhir RK, El-Mohr MAK, Dyer TD. Developing chloride resisting concrete using resistance of blast-furnace slag cement. J Mater Civil Eng 1993;6(2):223–33.
PFA. Cem Concr Res 1997;27(11):1633–9. [88] Luo R, Cai Y, Wang C, Huang X. Study of chloride binding and diffusion in
[55] Byfors K. Chloride binding in cement paste. Nordic Concr Res 1986;5:27–38. GGBS concrete. Cem Concr Res 2003;33(1):1–7.
[56] Arya C, Buenfeld NR, Newman JB. Factors influencing chloride binding in [89] Cheng A, Huang R, Wu J-K, Chen C-H. Influence of GGBS on durability and
concrete. Cem Concr Res 1990;20(2):291–300. corrosion behavior of reinforced concrete. Mater Chem Phys
[57] Arya C, Xu Y. Effect of cement type on chloride binding and corrosion of steel 2005;93:404–11.
in concrete. Cem Concr Res 1995;25(4):893–902. [90] Gouda VK. Corrosion and corrosion inhibition of reinforced steel. I. Immersed
[58] Holden WR, Page CL, Short NR. Corrosion of reinforcement in concrete in alkaline solutions. Br Corros J 1970;5:198–203.
construction. In: Crane AP, editor. Chichester: Ellis Horwood Ltd.; 1983. p. [91] Dhir RK, El-Mohr MAK, Dyer TD. Chloride binding in GGBS concrete. Cem
143–50. Concr Res 2000;26(12):1767–73.
[59] Wiens U, Schiessl P. Chloride binding of cement paste containing fly ash. In: [92] Thomas MDA, Scott A, Bremner T, Bilodeau A, Day D. Performance of slag
Justnes H, editor. Proceedings of the 10th ICCC, Goteborg, Sweden, 1997. p. 4– concrete in marine environment. ACI Mater J 2008;105(6):628–38.
10. [93] Justice JM, Kurtis KE. Influence of metakaolin surface area on properties of
[60] Kayyali OA, Haque MN. The Cl:OH ratio in chloride-contaminated concrete cement-based materials. J Mater Civil Eng 2007;19(9):762–71.
– a most important criterion. Mag Concr Res 1995;47:235–42. [94] Al-Akhras NM. Durability of metakaolin concrete to sulfate attack. Cem Concr
[61] Nagataki S, Otsuki N, Wee TH, Nakashita K. Condensation of chloride ion in Res 2006;36(9):1727–34.
hardened cement matrix materials and on embedded steel bars. ACI Mater J [95] Badogiannis E, Tsivilis S. Exploitation of poor Greek Kaolins: durability of
1993;90(4):323–32. metakaolin concrete. Cem Concr Compos 2009;31(2):128–33.
[62] Ampadu KO, Torii K, Kawamura M. Beneficial effect of fly ash on chloride [96] Shekarchi M, Bonakdar A, Bakhshi M, Mirdamadi A, Mobasher B. Transport
diffusivity of hardened cement paste. Cem Concr Res 1999;29(4):585–90. properties in metakaolin blended concrete. Constr Build Mater 2010;24.
[63] Thomas M. Chloride threshold in marine concrete. Cem Concr Res [97] Asbridge AH, Chadbourn GA, Page CL. Effects of metakaolin and the interfacial
1996;26(4):513–9. transition zone on the diffusion of chloride ions through cement mortars.
[64] Oh BH, Yang SY, Shin YS. Experimental investigation of the threshold chloride Cem Concr Res 2001;31(11):1567–72.
concentration for corrosion initiation in reinforced concrete structures. Mag [98] Baykal M. Implementation of durability models for portland cement concrete
Concr Res 2003;55:117–24. into performance-based specifications. Austin, TX: The University of Texas at
[65] Schiessl P, Breit W. Local repair measures at concrete structures damaged by Austin; 2000.
reinforcement corrosion – aspects of durability. In: Proceedings of the 4th [99] Savas BZ. Effects of microstructure on durability of concrete. Raleigh,
international symposium on corrosion of reinforcement in concrete NC: North Carolina State University; 1999.
construction, The Royal Society of Chemistry, Cambridge, 1996. p. 525–34. [100] Wee T, Suryavanshi A, Tin S. Evaluation of rapid chloride permeability test
[66] Alonso C, Castellote M, Andrade C. Chloride threshold dependence of pitting (RCPT) results for concrete containing mineral admixture. ACI Mater J
potential of reinforcements. Electrochim Acta 2002;47:3469–81. 2000;97(2):221–32.
[67] Page CL, Short NR, Holden WR. A study of the quantitative relationship [101] Saeki T, Sasaki K, Shinada K. Estimation of chloride diffusion coefficient of
between strength and pore-size distribution of porous materials. Cem Concr concrete using mineral admixtures. J Adv Concr Technol 2006;4(3):
Res 1986;16(1):79–86. 385–94.
[68] Diamond S. Two Danish flyashes on alkali contents of pore solutions of [102] Stanish KD, Hooton RD, Thomas MDA. Testing the chloride penetration
cement-flyash pastes. Cem Concr Res 1981;11:383–94. resistance of concrete: a literature review. Toronto, ON: Canada, University of
[69] Kawamura M, Kayyali OA, Haque MN. Effect of flyash on pore solution Toronto, Toronto, ON, Canada; 1997.
composition in calcium and sodium chloride-bearing mortars. Cem Concr Res [103] Andrade C, Castellote M, Alonso C, González C. Non-steady-state chloride
1988;18:763–73. diffusion coefficients obtained from migration and natural diffusion tests.
[70] Saraswathy V, Song H-W. Effectiveness of fly ash activation on the corrosion Part I: Comparison between several methods of calculation. Mater Struct
performance of steel embedded in concrete. Mag Concr Res 2000;33(1):21–8.
2007;59(9):651–61. [104] Castellote M, Andrade C, Alonso C. Measurement of the steady and non-
[71] Obla KH, Hill RL, Thomas MDA, Shashiprakash SG, Perebatova O. Properties of steady-state chloride diffusion coefficients in a migration test by means of
concrete containing ultra-fine fly ash. ACI Mater J 2003;100(5):1–8. monitoring the conductivity in the Anolyte chamber. Comparison with
[72] Jeon S, Nam J-H, An J-H, Kwon S-A. Physical properties of rapid-setting Natural Diffusion Tests. Cem Concr Res 2001;31(10):1411–20.
concrete using ultra fine fly ash. in proceedings of selected papers from the [105] Husain A, Al-Bahar S, Salam SA, Al-Shamali O. Accelerated AC impedance
2009 GeoHunan international conference, new technologies in construction testing for prequalification of marine construction materials. Desalination
and rehabilitation of Portland cement concrete pavement and bridge deck 2004;165:377–84.
pavement, GSP 196. [106] Shi X, Liu Y, Mooney M, Berry M, Hubbard B, Fay L, et al. Effect of chloride-
[73] Hossain AB, Fonseka A, Bullock H. Early age stress development, relaxation, based deicers on reinforced concrete structures. Final report prepared for the
and cracking in restrained low W/B ultrafine fly ash mortars. J Adv Concr Washington State Department of Transportation, July 2010.
Technol 2008;6:261–71. [107] Stanish K, Hooton RD, Thomas MDA. A novel method for describing chloride
[74] Subramaniam KV, Gromotka R, Shah SP, Obla K, Hill R. Influence of ultrafine ion transport due to an electrical gradient in concrete: Part 1. Theoretical
fly ash on the early age response and the shrinkage cracking potential of description. Cem Concr Res 2004;34(1):43–9.
concrete. J Mater Civil Eng 2005;17(1):45–53. [108] Ahmed MS, Kayali O, Anderson W. Evaluation of binary and ternary blends of
[75] Duval R, Kadri EH. Influence of silica fume on the workability and the pozzolanic materials using the rapid chloride permeability test. J Mater Civil
compressive strength of high-performance concretes. Cem Concr Res Eng 2009;21(9):446–53.
1998;28(4):533–47. [109] Whiting D. Rapid measurement of the chloride permeability of concrete.
[76] Cong X, Gong S, Darwin D, McCabe SL. Role of silica fume in compressive Public Roads 1981;45(3):101–12.
strength of cement paste, mortar, and concrete. Mater J 1989;89(4):375–87. [110] Yang Z, Shi X, Creighton AT, Peterson MM. Effect of styrene–butadiene rubber
[77] Song H-W, Pack S-W, Nam S-H, Jang J-C, Saraswathy V. Estimation of the latex on the chloride permeability and microstructure of Portland cement
permeability of silica fume cement concrete. Constr Build Mater mortars. Constr Build Mater 2009;23(6):2283–90.
2010;24:315–21. [111] He X, Shi X. Chloride permeability and microstructure of Portland cement
[78] Selvaraj R, Muralidharan S, Srinivasan S. The influence of silica fume on the mortars incorporating nanomaterials. Transp Res Rec: J Transp Res Brd
factors affecting the corrosion of reinforcement in concrete – a review. Struct 2008;2070:13–21.
Concr J FIB 2003;4(1):19–24. [112] Shi C, Stegemann JA, Caldwell RJ. Effect of supplementary cementing
[79] Page CL, Vennesland Ø. Pore solution composition and chloride binding materials on the specific conductivity of pore solution and its implications
capacity of silica-fume cement pastes. Mater Struct 1983;16(1):19–25. on the rapid chloride permeability test (AASHTO T277 and ASTM C1202)
[80] Byfors K. Influence of silica fume and flyash on chloride diffusion and ph results. ACI Mater J 1998;95(4):389–94.
values in cement paste. Cem Concr Res 1987;17:115–30. [113] Feldman R, Prudencio LR, Chan G. Rapid chloride permeability test on blend
[81] Larsen CK. Chloride binding in concrete, doctoral thesis, report no. 1998: 101, cement and other concretes: correlations between charge, initial current and
Norwegian University of Science and Technology, NTNU, 1998. conductivity. Constr Build Mater 1999;13:149–54.
[82] Pettersson K. Chloride threshold value and the corrosion rate in reinforced [114] Wee TH, Suryavanshi AK, Tin SS. Evaluation of rapid chloride permeability
concrete. In: Proceedings of the Nordic Seminar, Lund, 1995. p. 257–66. test (RCPT) results for concrete containing mineral admixtures. ACI Mater J
[83] Dotto JMR, de Abreu AG, Dal Molin DCC, Müller IL. Influence of silica fume 2000;97(2):221–32.
addition on concretes physical properties and on corrosion behaviour of [115] Cho SW, Chiang SC. Using the chloride migration rate to predict the chloride
reinforcement bars. Cem Concr Compos 2004;26(1):31–9. penetration resistance of concrete. In: Konsta-Gdoutos MS, editor. Measuring,
[84] Page CL, Havdahl J. Electrochemical monitoring of corrosion of steel in monitoring and modeling concrete properties. Netherlands: Springer; 2006.
microsilica cement pastes. Mater Struct 1985;18:41–7. p. 575–81.
X. Shi et al. / Construction and Building Materials 30 (2012) 125–138 137

[116] Ghosh P, Tikalsky PJ. Determination of diffusion coefficients and corrosion [147] Hassain SE, Al-Gahtani AS, Rasheeduzzafar. Chloride threshold for corrosion
initiation time for ternary cementitious mixtures. In: TRB 90th annual of reinforcement in concrete. ACI Mater J 1996;93(6):1–5.
meeting compendium of papers DVD, Transportation Research Board, [148] Moreno M, Morris W, Alvarez MG, Duffo9 GS. Corrosion of reinforcing steel in
Washington DC, 2011. simulated concrete pore solutions – effect of carbonation and chloride
[117] Diaz B, Freire L, Nóvoa XR, Puga B, Vivier V. Resistivity of cementitious content. Corros Sci 2004;46:2681–99.
materials measured in diaphragm migration cells: the effect of the [149] Monfore GE, Verbeck GJ. Corrosion of prestressed wire in concrete. ACI Mater
experimental set-up. Cem Concr Res 2010;40(10):1465–70. J 1960;57(5):491–7.
[118] Vivas E, Boyd A, Hamilton III HR. Permeability of concrete – comparison of [150] Söylev TA, Francois R. Corrosion of reinforcement in relation to presence of
conductivity and diffusion methods, final report prepared for the Florida defects at the interface between steel and concrete. J Mater Civil Eng
Department of Transportation, June 2007. 2005;17(4):447–55.
[119] Weyers RE. Corrosion service life model. In: Silva-Araya WP, de Rincon OT, [151] Yu H, Shi X, Hartt WH, Lu B. Laboratory investigation of reinforcement
O’Neill LP, editors. Repair and rehabilitation of reinforced concrete corrosion initiation and chloride threshold concentration for self-compacting
structures: the state of the art. Reston, VA: American Society of Civil concrete. Cem Concr Res 2010;40(10):1507–16.
Engineers; 1998. p. 105. [152] Yu H, Hartt WH. Modelling corrosion initiation of reinforcing steel in
[120] Pettersson K. Service life of concrete structure including the propagation concrete: effect of non-diffusive coarse aggregate. J Compos Mater
time. In: Concrete under severe conditions 2: environment and loading: 2011;45(2):153–69.
proceedings of the 2nd international conference on concrete under severe [153] Berman HA. Determination of chloride in hardened Portland cement paste,
conditions, CONSEC’98, Troms/, Norway, June 21–24, 1998. p. 489–98. mortar, and concrete. J Mater 1972;7(3):330–5.
[121] Costa A, Appleton J. Chloride penetration into concrete in marine [154] Harald J. A review of chloride binding in cementitious systems. Nordic Concr
environment – Part I: Main parameters affecting chloride penetration. Res 1998;21:1–6.
Mater Struct 1999;32:252–9. [155] Anik D, Jacques M, Jean-Pierre O, Simone J, Kati H. Chloride binding capacity
[122] Schiessl P, Raupach M. Influence of concrete composition and microclimate of various hydrated cement systems. Adv Cem Based Mater 1997;6:28–35.
on the critical chloride content in concrete. In: Page CL, Treadaway KWJ, [156] Buenfeld NR, Glass GK, Hassanein AM, Zhang J-Z. Chloride transport in
Bamforth PB, editors. Corrosion of reinforcement in concrete. London concrete subjected to an electric field. J Mater Civil Eng 1998;10(4):220–8.
(UK): Elsevier Applied Science; 1990. p. 49–58. [157] Suryavanshi AK, Scantlebury JD, Lyon SB. Corrosion of reinforcement steel
[123] Stratfull RF. The corrosion of steel in a reinforced concrete bridge. Corrosion embedded in high water–cement ratio concrete contaminated with chloride.
1956;13(3):173–8. Cem Concr Compos 1998;20(3):263–9.
[124] Hartt WH, Rodney GP, Virginie L, Lysogorski DK. A critical literature review of [158] Glass GK, Buenfeld NR. The influence of chloride binding on the chloride-
high-performance corrosion reinforcements in concrete bridge applications. induced corrosion risk in reinforced concrete. Corros Sci 2000;42(2):329–44.
FHWA-HRT-04-093, 2004. [159] Reddy B, Glass GK, Lim PJ, Buenfeld NR. On the corrosion risk presented by
[125] Glass GK, Buenfeld NR. The presentation of the chloride threshold level for chloride bound in concrete. Cem Concr Compos 2002;24(1):1–5.
corrosion of steel in concrete. Corros Sci 1997;39(5):1001–13. [160] Shi X, Nguyen TA, Kumar P, Liu Y. A phenomenological model for the chloride
[126] Alonso C, Andrade C, Castellote M, Castro P. Chloride threshold values to threshold of pitting corrosion of steel in simulated concrete pore solutions.
depassivate reinforcing bars embedded in a standardized OPC mortar. Cem Anti-Corros. Meth Mater 2011;58(4):179–89.
Concr Res 2000;30(7):1047–55. [161] Neville A. Chloride attack of reinforced concrete: an overview. Mater Struct
[127] Glass GK, Buenfeld NR. Chloride-induced corrosion of steel in concrete. Prog 1995;28:63–70.
Struct Eng Mater 2000;2(4):448–58. [162] Traetteberg A. The mechanism of chloride penetration in concrete, SINTEF
[128] Angst U, Elsener B, Larsen CK, Vennesland Ø. Critical chloride content in report STF65 A77070, 1977. 51 pp.
reinforced concrete – a review. Cem Concr Res 2009;39:1122–38. [163] Tritthart J. Chloride binding in cement II. The influence of the hydroxide
[129] Ann KY, Song H-W. Chloride threshold level for corrosion of steel in concrete. concentration in the pore solution of hardened cement paste on chloride
Corros Sci 2007;49:4113–33. binding. Cem Concr Res 1989;19(5):683–91.
[130] Kayyali OA, Hague MN. Chloride penetration and the ratio of Cl/OH in the [164] Rasheeduzzafar, Hussain SE, Al-Saadoun SS. Effect of cement composition on
pores of cement paste. Cem Concr Res 1988;18:895–900. chloride binding and corrosion of reinforcing steel in concrete. Cem Concr Res
[131] Yonezawa T, Ashworth V, Procter RPM. Pore solution composition and 1991;21(5):777–94.
chloride effects on the corrosion of steel in concrete. Corrosion [165] Andrade C, Page CL. Pore solution chemistry and corrosion in hydrated
1988;44(7):489–93. cement systems containing chloride salts: a study of cation specific effects. Br
[132] Breit W. Critical chloride content – investigations of steel in alkaline chloride Corros J 1986;21(1):49–53.
solutions. Mater Corros 1998;49(6):539–50. [166] Tang L, Nilsson LO. Chloride binding capacity and binding isotherms of OPC
[133] Li L, Sagüés AA. Chloride corrosion threshold of reinforcing steel in alkaline pastes and mortars. Cem Concr Res 1993;23:247–53.
solutions – open-circuit immersion tests. Corrosion 2001;57(1):19–28. [167] Jiricková M, Cerný R. Chloride binding in building materials. J Build Phys
[134] Li L, Sagüés AA. Metallurgical effects on chloride ion corrosion threshold of 2006;29(3):189–200.
steel in concrete, final report for the Florida Department of Transportation. [168] Liu Y, Shi X. Stochastically modeling of service life of concrete structures in
Tallahassee, FL, report no. WPI 0510806, November 2001. chloride-laden environments. J Mater Civil Eng 2012, in press.
[135] Yu H, Chiang K-TK, Yang L. Threshold chloride level and characteristics of [169] NIST, Publicly available computer models. http://ciks.cbt.nist.gov/~bentz/
reinforcement corrosion initiation in simulated concrete pore solutions. phpct/cmml.html [accessed 01.06.10].
Constr Build Mater 2012;26(1):723–9. [170] Khatri RP, Sirivivatnanon V. Characteristic service life for concrete exposed to
[136] Yu Y, Hartt WH. Effect of reinforcement and coarse aggregates on chloride marine environments. Cem Concr Res 2004;34(5):745–52.
ingress into concrete and time-to-corrosion: Part I – Spatial chloride [171] Erdoğdu S, Kondratova IL, Bremner TW. Determination of chloride diffusion
distribution and implications. Corrosion 2007;63(9):843–50. coefficient of concrete using open-circuit potential measurements. Cem
[137] Yu H, Hartt WH. Effects of reinforcement and coarse aggregates on chloride Concr Res 2004;34(4):603–9.
ingress into concrete and time-to-corrosion: Part 1 – Spatial chloride [172] Zhang J, Mcloughlin I, Buenfeld N. Modelling of chloride diffusion into
distribution and implications. Corrosion 2007;63(9):843–9. surface-treated concrete. Cem Concr Compos 1998;20(4):253–61.
[138] Cros P, Hartt WH, Yu H. Effects of reinforcement on chloride intrusion into [173] Kirkpatrick TJ, Weyers RE, Anderson-Cook CM, Sprinkel MM. Probabilistic
concrete and time-to-corrosion. In: Proceedings of the 16th international model for the chloride-induced corrosion service life of bridge decks. Cem
corrosion congress, September 22–26, 2005, Beijing, China. Concr Res 2002;32(12):1943–60.
[139] Hansen EJ, Saouma VE. Numerical simulation of reinforced concrete [174] Spragg RP, Castro J, Li W, Pour-Ghaz M, Huang P-T, Weiss J. Wetting and
deterioration: Part I: Chloride diffusion. ACI Mater J 1999;96(2):173–80. drying of concrete using aqueous solutions containing deicing salts. Cem
[140] Oh BH, Jang BS. Chloride diffusion analysis of concrete structures considering Concr Compos 2011;33:535–42.
effects of reinforcing. ACI Mater J 2003;100(2):143–9. [175] Castro P, De Rincon OT, Pazini EJ. Interpretation of chloride profiles from
[141] Kranc SC, Sagüés AA, Presuel-Moreno F. Decreased corrosion initiation time concrete exposed to tropical marine environments. Cem Concr Res
of steel in concrete due to reinforcing bar obstruction of diffusional flow. ACI 2001;31:529–37.
Mater J 2002;99(1):51–3. [176] Meijers SJH, Schlangen E. 2D-analysis of chloride ingress in the tidal zone of
[142] Yu H, Himiob RJ, Hartt WH. Effects of reinforcement and coarse aggregates on marine concrete structures. In: Weiss J, Kovler K, Marchand J, Mindess S,
chloride ingress into concrete and time-to-corrosion: Part 2: Spatial editors. Proceedings of the international RILEM symposium on concrete
distribution of coarse aggregates. Corrosion 2007;63(10):924–31. science and engineering: a tribute to Arnon Bentur, 2004. doi: 10.1617/
[143] Yu H, Hartt WH. Correction of chloride threshold concentration and time-to- 2912143926.045.
corrosion due to reinforcement presence. Mater Corros 2010;61(9999): [177] Guimaraes ATC, Helene PRL. Diffusion of Chloride Ions in Unsaturated
1–8. Concrete: Forecast of Service Life in a Wet-Dry Environment. ACI Special
[144] Shi X, Nguyen TA, Kumar P, Liu Y. A phenomenological model for the chloride Publication, vol. 229; 2005. p. 175–94.
threshold of pitting corrosion of steel in simulated concrete pore solutions. [178] Hong K, Hooton RD. Effects of cyclic chloride exposure on penetration of
Anti-Corros Methods Mater 2011;58(4):179–89. concrete cover. Cem Concr Res 1999;29:1379–86.
[145] Hartt WH, Nam J. Effect of cement alkalinity on chloride threshold and time- [179] Boddy A, Bentz E, Thomas MDA, Hooton RD. An overview and sensitivity
to-corrosion of reinforcing steel in concrete. Corrosion 2008;64(8):671–80. study of a multimechanistic chloride transport model. Cem Concr Res
[146] Dehwah HAF, Maslehuddin M, Austin SA. Effect of cement alkalinity on pore 1999;29(6):827–37.
solution chemistry and chloride-induced reinforcement corrosion. ACI Mater [180] Marchand J, Samson E. Predicting the service-life of concrete structures –
J 2002;99(3):227–35. limitations of simplified models. Cem Concr Compos 2009;31:515–21.
138 X. Shi et al. / Construction and Building Materials 30 (2012) 125–138

[181] Lee SH, Rasaiah JC. Molecular dynamics simulation of ion mobility. 2. Alkali [190] Hartt, WH. Corrosion initiation projection for reinforced concrete exposed to
metal and halide ions using the SPC/E model for water at 25 °C. J Phys Chem chlorides – Part II: Corrosion resistant bars. In: Proceedings of CORROSION
1996;100(4):1420–5. 2011 conference, NACE international, paper no. 11009, Houston, Texas. p. 11.
[182] Castellote M, Alonso C, Andrade C, Chadbourn GA, Page CL. Oxygen and [191] Ožbolt J, Balabanić G, Kušter M. 3D numerical modelling of steel corrosion in
chloride diffusion in cement pastes as a validation of chloride diffusion concrete structures. Corros Sci, in press. doi: 10.1016/j.corsci.2011.08.026.
coefficients obtained by steady-state migration tests. Cem Concr Res [192] Broomfield JP. Modelling the rate of deterioration of reinforced concrete
2001;31(4):621–5. structures. In: Proceedings of CORROSION 2011 conference, NACE
[183] Hobbs DW. Aggregate influence on chloride ion diffusion into concrete. Cem international, paper no. 11003, Houston, Texas. p. 13.
Concr Res 1999;29(12):1995–8. [193] Otieno M, Beushausen H, Alexander M. Prediction of corrosion rate in RC
[184] Engelund S, Sørensen JD. A probabilistic model for chloride-ingress and structures – a critical review. In: Andrade C, Mancini G, editors. Modelling of
initiation of corrosion in reinforced concrete structures. Struct Saf corroding concrete structures, RILEM, 2011. p. 15–37. doi: 10.1007/978-94-
1998;20(1):69–89. 007-0677-4_2.
[185] Williamson GS, Weyers RE, Brown MC, Ramniceanu A, Sprinkel MM. [194] Francois R, Arliguie G, Castel A. Influence of service cracking on service life of
Validation of probability-based chloride-induced corrosion service-life reinforced concrete. In: Concrete under severe conditions 2: environment
model. ACI Mater J 2008;105(4):375–80. and loading: proceedings of the 2nd international conference on concrete
[186] Kranc SC, Sagüés AA. Advanced analysis of chloride ion penetration profiles in under severe conditions, CONSEC’98, Troms/, Norway, June 21–24, 1998. p.
marine substructures, final report prepared for the Florida Department of 143–52.
Transportation, 2003. [195] Lin G, Liu Y, Xiang Z. Numerical modeling for predicting service life of
[187] Sagüés AA, Kranc SC, Presuel-Moreno F, Rey D, Torres-Acosta A, Yao L. reinforced concrete structures exposed to chloride environments. Cem Concr
Corrosion forecasting for 75-year durability design of reinforced concrete, Compos 2010;32(9):571–9.
final report prepared for the Florida Department of Transportation, 2001. [196] Chen D. Computational framework for durability assessment of reinforced
[188] Pettersson K. Service life of concrete structure including the propagation concrete structures under coupled deterioration processes. Doctoral thesis,
time. In: Concrete under severe conditions 2: environment and loading: PhD in civil engineering, Vanderbilt University, August 2006.
Proceedings of the 2nd international conference on concrete under severe [197] Nokken M, Boddy A, Hooton RD, Thomas MDA. Time dependent diffusion in
conditions, CONSEC’98, Troms (Norway, June 21–24, 1998, p. 489–98. concrete – three laboratory studies. Cem Concr Res 2006;36(1):200–7.
[189] Hartt WH. Corrosion initiation projection for reinforced concrete exposed to [198] Song H-W, Shim H-B, Petcherdchoo A, Park S-K. Service life prediction of
chlorides – Part I: Black bars. In: Proceedings of CORROSION 2011 conference, repaired concrete structures under chloride environment using finite
NACE international, paper no. 11006, Houston, Texas. p. 18. difference method. Cem Concr Compos 2009;31:120–7.

You might also like