You are on page 1of 11

ACI MATERIALS JOURNAL

TECHNICAL PAPER

Title No. 113-M01

Relationship between Reinforcing Bar Corrosion and


Concrete Cracking
by Aimin Xu and Ahmad Shayan
This paper presents results of an investigation on the relationship
between reinforcement corrosion rust growth at the concrete-steel
interface and cracking of concrete cover, based on study of laboratory specimens containing various amounts of chloride ions.
The corrosion depth of steel bars to cause cover cracking (critical
depth) was found to vary from 5 m (0.2 mil) for 24 mm (0.9 in.)
bar under 25 mm (1 in.) cover, to 150 m (5.9 mil) for 6 mm (0.2 in.)
bar under 73 mm (2.9 in.) cover. The steel corrosion rate was found
to be closely related to the amount of chloride in concrete.
The applicability of existing equations on cover cracking is
assessed and a new mathematical equation is proposed. The
predicted critical corrosion depth by the equation matched with
the experimental data reasonably well. A mathematical equation
has been proposed for calculation of the time-to-cracking, which
uses this model in combination with the law of chloride diffusion
in concrete.
Keywords: chloride; concrete cover; corrosion; reinforcing steel;
time-to-cracking.

RESEARCH SIGNIFICANCE
In most studies of chloride-induced corrosion damage to
reinforced concrete (RC) structures, as well as the prediction
of service life, the prediction is often based on the initial
chloride diffusion into concrete until the threshold chloride level for the corrosion initiation is reached at the depth
of steel reinforcement. This prediction is obviously very
conservative and grossly underestimates the service life of
RC structures. The present research has enabled prediction
of the first cracking of cover concrete based on the depth of
corrosion on the reinforcing steel. This enables a more accurate prediction of service life of RC structures and remedial
intervention time.
INTRODUCTION
Concrete durability is of great interest to the owners and
operators of concrete structures, which have a replacement
value totaling billions of dollars. Of particular interest are
concrete structures located in aggressive environments with
exposure to seawater or saline groundwater, which are at
greater risk of chloride-induced corrosion. In the management of concrete structures, it is important to be able to
predict the progress of steel corrosion and time-to-cracking
so that maintenance intervention time can be determined and
budgeted for.
Existing models of service life prediction rely on assessing
the chloride ingress into concrete based on Ficks diffusion
equation and the premise that the service life of RC structure
is over when the chloride threshold is reached at reinforceACI Materials Journal/January-February 2016

ment depth.1-3 In reality, this is only the start of the second


stage of the deterioration, which combines the initiation and
development of steel corrosion and damage to the reinforced
concrete. Prediction of concrete cracking and steel section
loss should be the focus for this stage in the prediction of
service life of RC structures.
There is little quantitative information on the behavior of
reinforced concrete after the threshold level of chloride ions
has been reached at the reinforcement depth, and the effect
of progression of corrosion on the cracking of concrete cover
is not fully understood. In previous work, visible deterioration has been used as a criterion for time of intervention.
For example, Life-365 prediction software,3 based on the
result of Weyers,4 assumes a propagation period (defined
as the period between the corrosion initiation and end of
service life) to be 6 years. This assumption may not be valid
under all exposure conditions.
It is known that corrosion of reinforcement produces rust,
which has a greater volume than that of the steel lost due to
corrosion. This volume change at the steel-concrete interfacial zone results in building up of internal stresses in the
concrete surrounding the corroding steel. When the stress
exceeds the tensile strength of the concrete, cracks form in
the vicinity of the steel bars, which eventually extend to the
concrete surface. The volume change depends on the type of
rust formed, which, in turn, depends on the available moisture and oxygen. However, under certain conditions it would
not greatly vary and the amount of rust can be considered
proportional to that of corrosion depth.
This paper presents the result of an investigation on
the prediction of time-to-cracking of reinforced concrete,
induced by the corrosion of the reinforcing bars. The main
objectives of this work were:
Developing electrochemical data, including half-cell
potential and corrosion current density of the reinforcing steel for concrete specimens incorporating
various amounts of chloride.
Measuring corrosion-induced expansion in cover
concrete, across the corroding steel bars, and observing
the time-to-cracking.
Critically reviewing existing models in predicting the
time to crack formation.
ACI Materials Journal, V. 113, No. 1, January-February 2016.
MS No. M-2014-039.R6, doi: 10.14359/51688460, received February 19, 2015, and
reviewed under Institute publication policies. Copyright 2016, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including authors
closure, if any, will be published ten months from this journals date if the discussion
is received within four months of the papers print publication.

Amending the existing models or proposing new models


to predict the time-to-cracking, using the results of this
investigation.
The present paper is a follow-up to an earlier paper5
with more complete data on the cracking of concrete slabs
comprising three concrete cover thicknesses and three sizes
of reinforcing bars. The model proposed in the earlier paper
is further discussed in this paper.
BACKGROUND
The prediction of time-to-cracking has an important place
in the prediction of service life of RC structures, as it is the
beginning of more rapid deterioration. Crack formation is
greatly influenced by the rate of corrosion.
A few techniques are available for determining the rate of
corrosion, but accurate determination of time-to-cracking
remains a problem. Based on laboratory and field investigations, Broomfield et al.6 developed some broad criteria for
corrosion activity in RC, such as: corrosion current density
icorr < 0.1 A/cm2 (0.09 mA/ft2) for passive condition low to
moderate corrosion; icorr= 0.1 to 0.5 A/cm2 (0.09 to 0.46 mA/ft2);
icorr = 0.5 to 1 A/cm2 (0.46 to 0.93 mA/ft2) for moderate to high
corrosion; and icorr > 1 A/cm2(0.93 mA/ft2) for high corrosion
rate. Broomfield7 summarized corrosion potential criteria as Ecorr
of more negative than 300 mV, 500 mV, and 550 mV CSE,
corresponding to the above corrosion rate categories.
However, these categories cannot be readily converted to
data on crack development. Broomfield7 stated that approximately 10 m (0.4 mil) section loss or 30 m (1.2 mil)
rust growth was sufficient to cause cracking. Gonzalez et
al.8 concluded that a corrosion current density of 1 A/cm2
(0.93 mA/ft2) could result in damage in a period of 5 to
10years (equivalent to 58 m [2.3 mil] at 5 years, according
to Faradays Law).
Even larger variations can be found in published data as
reviewed by Liu9 and Al-Harthy et al.10 For example, less than
25 m (1 mil) steel corrosion would crack 22 mm (7/8 in.) thick
cover and that 734 m (29 mil) corrosion would be more
than enough to cause concrete spalling, concluded from
results of lab test and field investigation, respectively, by
Spellman and Stratfull11; 7.87 m (0.3 mil) by Baant12;
94m (3.7mil) by Clear13; 16 to 32 m (0.6 to 1.2 mil) by
Hladky14; 2.54 m (0.1 mil) for 20 mm (0.8 in.) bar under
40 mm (1.6 in.) cover by Suda et al.15; and 19.5 m (0.8 mil)
for 20 mm (0.8 in.) bar to crack 25 mm (1 in.) cover by Zhao
et al.16
The time-to-crackingthat is, the time required to
produce the quantity of rust that would create cracks in
concrete coveris a function of several factors, mainly: the
rate of steel corrosion, type of rust; the size of steel bar; the
thickness of cover concrete; and the strength and effective
modulus of elasticity of concrete. The microstructure of
concrete plays an important role in terms of allowing ingress
of chloride and transport of oxygen as well as distribution of
corrosion products.
A number of mathematical formulas incorporating these
factors have been proposed. The equation by Baant17 and
that by Liu9 are based on analysis of model elements of
concrete, which include an embedded steel bar. However,
4

Fig. 1Element of reinforced concrete illustrating corrosioninduced volume change and schematic expression of pressures on concrete.
other formulas based on similar model elements but using
different approaches, such as damage mechanics16 or largely
empirical formulas,18-20 resulted in quite a different relationship between cover cracking and these factors.
The commonly used model element (Fig. 1) is a unit
length of cylindrical concrete element containing a steel bar
at its center, which is subjected to an external pressure po and
the corrosion-induced internal pressure pi. The dimensions
include a = /2 the inner radius and b = C + a radius of the
element, where is the diameter of the bar and C is the cover
thickness.
The amount of rust that would result in cover cracking
proposed by Liu and Weyers21 was
Wcrit = rust [ (d s + d 0 ) + Wst /st ] , where d s =

Cf t a 2 + b 2
+ (1)
Eeff b 2 a 2

where rust is the density of rust; ft is the concrete tensile


strength; d0 represents a porous band surrounding the bar;
Eeff = E/(1 + ) is the effective modulus of elasticity of
concrete (where E and are the modulus of elasticity and
creep coefficient of concrete, respectively); is the Poissons ratio of concrete; Wst and st are the mass and density
of steel lost due to corrosion, respectively; and ds is the
radial displacement of concrete due to rust formation. This
equation includes the parameters necessary to be considered
in the corrosion-induced concrete cracking. However, the
present authors found ds in Eq. (1) did not confirm the test
results5 and proposed an equation for the critical depth of
steel corrosion that would result in cover cracking crit as

crit =

f t C (C + )(1 + )
(2)
Eeff

where is the relative volume change of steel due to corrosion, and the other symbols used are the same as for Eq. (1).
Both Eq. (1) and Eq. (2) will be further discussed in this paper.
EXPERIMENTAL INVESTIGATION
Details of specimens
Reinforced concrete slabs measuring 380 x 300 x 180 mm
(15 x 12 x 7 in.) were used in this study to develop data
on corrosion rate of reinforcing steel in concrete. Two types
of concrete of relatively poor quality (30 MPa [4351 psi])
and good quality (50 MPa [7252 psi]), three different sizes
(diameter) of steel bars and three different levels of chloride contamination (by dissolving NaCl in concrete mixing
ACI Materials Journal/January-February 2016

Table 1Variables employed in the experimental slabs


Concrete strength grade

30 MPa (4351 psi)

Cement content

420 kg/m3 (708 lb/yd3)

0.63

0.42

w/c
Chloride content (% by cement mass)

0.25, 0.98, 1.97

0.25, 0.75, 1.50

0.8 kg/m3 (1.348 lb/yd3),


3.15 kg/m3 (5.309 lb/yd3),
6.3 kg/m3 (10.619 lb/yd3)

1.05 kg/m3 (1.770 lb/yd3),


3.15 kg/m3 (5.309 lb/yd3),
6.3 kg/m3 (10.619 lb/yd3)

6 (0.24), 12 (0.47), 24 (0.94)

6 (0.24), 12 (0.47), 24 (0.94)

25 (1), 50 (2), 70 (2.8)

25 (1), 50 (2), 70 (2.8)

Chloride content (mass of Cl per unit volume of concrete)


Reinforcing steel bar diameter, mm (in.)

50 MPa (7252 psi)

320 kg/m (539 lb/yd )


3

Nominal concrete cover thickness, mm (in.)

Table 2Mixture proportions, density, and strength at 28 days


Mixture ID

Cement, kg/m3 (lb/yd3)

Mixture proportion
C:W:A:S

Cl/cement
(% by mass)

Cl/vol.,
kg/m3 (lb/yd3)

Density, kg/m3
(lb/yd3)

fc, MPa (psi)

n600

420 (708)

1:0.42:2.7:1.7

0.050

0.21 (0.35)

2495 (4205)

62.5 (9065)

n601

420 (708)

1:0.42:2.7:1.7

0.250

1.05 (1.77)

2500 (4215)

60.5 (8775)

n602

320 (539)

1:0.625:3.4:2.5

0.250

0.80 (1.35)

2480 (4180)

35.2 (5105)

n603

420 (708)

1:0.42:2.7:1.7

0.750

3.15 (5.31)

2505 (4225)

63.3 (9181)

n611

320 (539)

1:0.625:3.4:2.5

0.984

3.15 (5.31)

2440 (4110)

32.4 (4699)

n612

420 (708)

1:0.42:2.7:1.7

1.500

6.30 (10.62)

2485 (4195)

52.9 (7673)

n613

320 (539)

1:0.625:3.4:2.5

1.969

6.30 (10.62)

2440 (4110)

30.8 (4467)

water), and three concrete cover thicknesses were the main


variables employed in this work (Table 1).
The three levels of chloride were intended to generate
different levels of corrosion activity in steel bars. The
choice of the high chloride level was based on the results
by Liu9 who showed that, under external exposure conditions, concrete made with cement content of 635 lb/yd3
(376 kg/m3) and a water-cement ratio (w/c) of 0.45 having
an admixed Cl content of 12 lb/yd3 (7.1 kg/m3), cracks
appeared on the specimens with 16 mm (5/8 in.) bar under
25 mm (1 in.) cover at 0.72 year. It was anticipated that the
30 MPa (4351 psi) grade concrete with high level of chloride
would produce similar results.
The different concrete grades and cover thicknesses were
used to provide different levels of resistance to concrete
cracking. Figure 2 illustrates the main design feature of the
experimental slabs. The center-to-center distance between
steel bars was 100 mm (4 in.). All bars in a slab had the
same cover thickness, so that three slabs were made from
each concrete mixture at each chloride level. The initial
weight of each steel bar was recorded for later comparisons. The compressive strength of concrete at 28 days was
30 to 35MPa (4351 to 5076 psi), and 53 to 63 MPa (7687
to 9137 psi) for the 30 and 50 MPa (4351 and 7252 psi)
grade concrete, respectively; the strength decreased with the
increase in the amount of NaCl in concrete (Table 2).
A 50 MPa (7252 psi) grade concrete with 0.05% chloride per cement was originally cast as a trial mixture. After
2 weeks of initial moist curing, the three slabs made from
this mixture were placed outdoors. This group of slabs
were to provide data for the low corrosion rate measured by
nondestructive method.

ACI Materials Journal/January-February 2016

Fig. 2Schematic illustration of specimen section. (Note:


Dimensions are in mm [in.].)
According to the empirical formulas and reference data
given by Australian standard AS 3600,22 the tensile strength,
ft, and E modulus were estimated by ft = 0.36fc and
E = 1.5 0.043fc, where fc is the compressive strength
(MPa), and is the concrete density (kg/m3). Poissons ratio
= 0.20; = 2.0 and 3.4 for 50 and 30 MPa (7252 and
4351 psi) grade concrete, respectively. Based on test results
(Table 2), the estimated average tensile strength is 2.1 and
2.8 MPa (299 and 404 psi), and E modulus is 29.9 and 41.5
GPa (4336 and 6018 ksi) for 30 and 50 MPa (4351 and
7252 psi) concrete, respectively.
Treatment of specimens
Each slab was placed in a container with water filled to
approximately 25 mm (1 in.) below the reinforcing bars.
NaCl was added to the water to minimize the release of
chloride ions from concrete. The amount of NaCl added was
60% of the chloride concentration, which would result from
the dissolution of the added amount of NaCl to the mixing
water. The containers were covered by plastic sheeting to
minimize the evaporation of water. The containers were
originally stored at room temperature, but after approx5

imately 1 year, only one slab (largest bar, thinnest cover)


had shown cracking. The corrosion rate was considered to be
too slow, therefore, slabs of one concrete were moved into
a constant-temperature room kept at 40C (104F), where
the corrosion current density became higher. The other slabs
were then moved into the 40C (104F) condition to accelerate the corrosion process.
Methods of measurement
Nondestructive evaluation of corrosion rateA corrosion rate meter was used to measure the corrosion current.
It measures the corrosion current over a confined length of
steel bar, the half-cell potential (coppercopper sulfate) of
the steel, and the electrical resistance of concrete. This instrument uses the linear polarization technique and calculates
the polarization resistance Rp and obtains the corrosion
current according to Stern and Geary relationship23

I Corr =

ba bc
ba + bc

B
(3)
Rp

where B = 26 mV, which had been assigned in the device;


and ba and bc are the anodic and cathodic Tafel slopes,
respectively. The polarization time used was 100 seconds.
Using the surface area of the steel within the confined
length, the corrosion current density icorr is calculated. Note
that the cathodic reaction is controlled by the oxygen diffusion, and the bc tends to , then B |ba|, which ranges from
30 to 40mV at 20C (68F) in a condition of pH > 13.24 The
corrosion current density calculated by the instrument based
on a constant B may not represent the actual corrosion rate
and the data need to be verified by directly measured mass
of corroded steel.
The amount of corroded steelThe rust on corroded
steel bars was cleaned off by tapping to break the bulky
rust pieces (which were collected for density determination) and then by brushing in a solution of 1 L HCl (sp.
gr. 1.19) containing 20 g Sb2O3 and 50 g SnCl2, as recommended by ASTM G1.25 The procedure had to be repeated
several times to clean off most of the rust. Cleaned bars were
weighed and the mass per length was calculated. The mass
loss per bar surface area due to corrosion was calculated by
w/AS = (wcorr wo)/(), where wo and wcorr are the mass per
unit length of the bar before and after corrosion, respectively.
Recording of cracksThe surfaces of slabs were visually examined, which was aided by spraying some water
on concrete and allowing it to dry. This revealed even small
cracks (less than 0.1 mm [4 mil]), which were marked on
the concrete surface, noting the date of observation. Crack
inspection was carried out throughout the period of corrosion current measurement and routine checking of the
containers. Note that a crack could only be observed when
it was wide enough, but then the actual time of cracking had
already passed. Therefore, the results were to be verified by
monitoring the cover concrete across the steel bars.
Monitoring of concrete length changeStainless steel
studs were glued on surface of each slab at the age of approximately 4 months. Two pairs of studs were fixed across each
6

Fig. 3Corrosion current density of steel bars; each is the


average value of nine bars.
steel bar at distance of 100 mm (4 in.) and along the steel bar
at spacing of 200 mm (8 in.) (Fig. 2). The length between
the studs was measured by DEMEC gauges on a regular
basis. The length change of bulk concrete parallel to the
steel bars was also monitored, which represents the volume
change of concrete as influenced by factors such as moisture
and temperature fluctuations. The values of length change
across the bars were corrected using that of the bulk concrete
and then plotted against time to trace the initial time of
crack formation.
EXPERIMENTAL RESULTS AND DISCUSSION
Corrosion current density
At the beginning, the corrosion current density was
measured at intervals of approximately twice per month and
then once per month when the slabs were stored at 23C
(73F) room condition. It was observed that the corrosion
current density of many bars decreased with time. The slabs
were then moved into the 40C (104F) room to accelerate
the corrosion. The later stage measurements were made at
longer intervals.
The results of corrosion current density measurement are
presented in Fig. 3, where each point is the average of all nine
bars of the same concrete mixture. The range of variations
between the nine bars of the same mixtures is approximately
20% (maximum 40%) of the average value (at 90% confidence level). The variation may be related to the different
bar sizes and cover thicknesses among the bars.
At early stages (for example, 14 days), the corrosion
current for steel bars in slabs of high chloride content and
lower strength concrete was very high, and it was difficult
to confine the measuring area of the steel bars. This might
have been due to the high conductivity of the concrete at
this stage. The confinement improved with the concrete age.
The corrosion current density (which is proportional to
the corrosion rate) distinctly depended on the grade and the
chloride content of the concrete; for example, the average
corrosion current density icorr, measured during the period when
the slabs were at 23C (73F), was 0.44 A/cm2 (0.41 mA/ft2)
for steel bars of the 30 MPa (4351 psi) grade concrete, and
0.27 A/cm2 (0.25 mA/ft2) for the 50 MPa (7252 psi) Grade
concrete, both containing high levels of chloride. After
moving into the 40C (104F) condition, the icorr of bars in
the slabs made with 30 MPa (4351 psi) grade concrete and
6.3 kg Cl/m3 (10.62 lb/yd3) increased by twofold, and cracks
ACI Materials Journal/January-February 2016

developed over the 24 mm (1 in.) diameter bars with 25


mm (1 in.) cover thickness. The icorr values for all steel bars
increased under the 40C (104F) condition, except in the
50 MPa (7252 psi) grade concrete with low chloride level.
The magnitude of corrosion rate increase with increased
temperature can be explained by the Arrhenius equation,26
according to which the ratio of reaction rate at two temperatures, T1 and T2, is exp[Ea(1/T1 1/T2)/R], where Ea is the
activation energy and R is the gas constant. From the data
presented by Melchers,27 who tested the corrosion rate of
steel coupons immersed in seawater, the value of Ea can be
calculated to be Ea = 37 kJ/mol at the initial stage. Thus, the
rate at 40C (T1 = 313 K) would be 2.3 times of that at 23C
(T2 = 296 K).
The measured icorr decreased with concrete age in the first
year (Fig. 3), during which there were no obvious cracking in
most slabs. This may be partly due to the increased concrete
resistivity with maturity, which reduced the mobility of electrons and consequently reduced the corrosion activity. There
was a large increase in icorr of bars in the high and medium
chloride level mixtures between 1 and 3 years, which may be
associated with the formation and widening of cracks around
the heavily corroding bars, which facilitates the supply of
oxygen and moisture for the corrosion. Later, despite the
formation of cracks in most of the slabs, the corrosion rate
was more or less constant, or decreased in one concrete
(n612), where icorr was higher than 1A/cm2 (0.93 mA/ft2).
This may be due to the decrease in oxygen diffusion through
the rust, which becomes thicker with time.
The corrosion rate of steel bars increased with the increase in
the chloride content in concrete, and could be approximated by
linear trend lines for both grades of concrete (Fig. 4). Comparatively, the corrosion rate was lower for the steel bars in the
higher-grade concrete than that in the lower-grade concrete,
at the same chloride-to-cement ratio. This difference can
be largely attributed to the lower oxygen supply and higher
electrical resistivity arising from the lower porosity of the
higher-grade concrete.
Measurement of steel mass loss
At the age of 3 years, when a number of slabs had developed cracking, a part of some corroded steel bars was
extracted by coring and cleaned to measure the mass loss
(cut ends of the steel bars were coated). At 5 years, all the
slabs with heavily corroded bars were split and the bars
taken out to determine their mass loss due to corrosion.
The distribution of rust was generally uniform along each
bar. However, there was much more rust on the top side, which
was closer to the slab surface than the bottom side of the bars,
which was closer to the wet part of the slabs. This difference
was due to blockage of the air path in concrete by water to the
bottom side of bars. It was also observed that concrete near
the corroded bars was stained by corrosion product, which
indicated that some of corrosion product penetrated into the
concrete pores and did not contribute to the expansion arising
from rust accumulation on the steel bar.
The mass loss per area was calculated according to the
reduction in mass per unit length of steel bar, which was

ACI Materials Journal/January-February 2016

Fig. 4Relation between corrosion current density (room


temperature) and chloride in concrete.
converted to corrosion depth using the density of steel, and
the results are included in Table 3.
Corrosion rate determined by nondestructive
testing method
The amount of steel corrosion is proportional to the corrosion current density and time. According to the Faradays
Law, the depth of steel corrosion, corr, is related to icorr by

corr =

M Fe
tr icorr , r (4)
zF steel r =1

where MFe = 55.85 g/mole (0.123 lb/mole); F = 96,486 C/mole


(Faraday constant); steel = 7860 kg/m3 (13,248 lb/yd3); z =
2 for steel corrosion (Fe Fe2+ + 2e)26; and icorr,r is the
corrosion current density at the time interval tr. For the ideal
case that icorr is constant, by inserting the aforementioned
data into Eq. (4), one obtains the correlation between the
icorr and corrosion rate: 1 A/cm2 = 11.6 m/year (0.46 mil/
year). In practice, the calculation was performed on each
steel bar using the average value of data obtained for the bar
measured at adjacent times.
Results of total charge per area of bars (time corrosion
current density) when the bar was extracted from slabs are
presented in Table 3. The ratio of the mass loss determined
depth to that by the nondestructive testing (NDT) method
was four times or higher.
Half-cell potential
GECOR6 measures the half-cell potential, and presents it
as corrosion potential, Ecorr. In general, the more negative
the Ecorr, the higher the icorr, although large scattering of data
occurs (Fig. 5). Typically at an Ecorr of approximately 300
mV, icorr was approximately 0.1 A/cm2 (0.1 mA/ft2) which
was categorized by Broomfield7 as boundary value for the
low-to-moderate corrosion rate measured by using confined
current technique. ASTM C87628 (before the 1991 version)
states the potential of 350 mV as the boundary value for
the probable occurrence of active corrosion. The corrosion
activity of steel bars in the current investigation largely
confirmed the above categories (Fig. 5). Figure 5 shows an
approximately linear relationship between the logarithm of
corrosion current density and the corrosion potential. This
7

Fig. 5Corrosion current density versus corrosion potential, temperature 40C (104F).
can be explained by that theoretically the anodic and cathodic
processes obey Butler-Volmer equation27that is, Icorr =
AJ0exp((FE)/(4RT)), where A and J0 are the average area
and average current density at the corrosion sites, respectively; E is the electrode potentials difference. The slope of
the linear trend line shown in Fig. 5 is 0.0085, which is close
to the theoretical value (0.0093) for 40C (104F) (T = 313 K),
calculated by F/(4RT). However, in the range where Ecorr is
more positive than 100 mV, values of icorr are greater than
that predicted by the linear relationship, indicating GECOR6
may have overestimated icorr.
Corrosion-induced expansion and time to cover
concrete cracking
After cover concrete across a steel bar expanded to more
than 10 to 20 m (0.4 to 0.8 mil) (100 to 200 microstrain),
further expansion would be definite and hairline cracks
appeared on concrete surface. Thus, the corrosion time
corresponding to 150 microstrain was taken as the time-tocracking. Two examples, over a 24 mm (0.94 in.) bar and the
other over a 12 mm (0.47 in.) bar, both under cover of 25 mm
(1 in.) are shown in Fig. 6.
The time-to-cracking observed in the slabs ranged from
145 days to 3 years, depending on the bar size, and the
cracking occurred when the steel corrosion depth reached
5 to 35 m (0.2 to 1.4 mil), depending on the cover thickness for the 24 mm (1 in.) diameter bars. This is in close
agreement with the literature.9,20 For 12 mm (0.5 in.) bars,
cracking occurred at a corrosion depth 8 to 78 m (0.3 to
3 mil), depending on the cover, which, however, could not
be verified by Eq. (1) proposed by Liu and Weyers21; for
example, the value of ds, the critical radial displacement of
concrete (Eq. (1)), computed at = 0.2 and d0 = 0, gives 6.7
to 26.9 m (0.27 to 1.06 mil) for 12 mm (0.5 in.) bars (for
the three cover thicknesses), which is less than that for larger
bars (ds for 24 mm [1 in.] bars is 7.3 to 27.4 m [0.29 to
1.08 mil]). This is opposite to the fact that smaller bars need
to corrode more to cause cracking. This was also noted by
some researchers.10

Fig. 6Development of corrosion and expansion. Vertical


axis maximum value: 50m (2.0 mil).
MODELING OF COVER CRACKING
Expansion of steel due to corrosion
The relative volume change (V/Vsteel) due to conversion
of steel to rust, is defined as = V/Vsteel =steel/(rust) 1,
where is the density, and = msteel/mrust is the mass fraction of iron (Fe) in the rust. For typical rusts Fe3O4, Fe2O3,
Fe(OH)2, and Fe(OH)3, = 0.7236, 0.6994, 0.6215, and
0.5226, respectively.
The rust formed on steel bars was collected and its density
was determined by the water replacement method. The
average density of dry rust was 3460 kg/m3 (5832 lb/yd3).
Comparing with the chart given in ACI 222R-01,29 the rust
was likely to be in form of Fe(OH)2, and = 2.66.
Crack formationanother look at theories
The effect of corrosion on stress buildup in concrete is
analyzed using the cylindrical element shown in Fig. 1.
To simplify the model, d0 proposed by Liu9 is ignored.
Timoshenko and Goodier30 used polar coordinates to analyze
the stress and strain in such an element. For a cylinder
subjected to internal pressure pi and external pressure po,
stresses as functions of r (the distance from the center) are

r =

a 2 b 2 ( po pi ) pi a 2 po b 2
+
, and
r 2 (b 2 a 2 )
b2 a 2

a 2 b 2 ( po pi ) pi a 2 po b 2
+
r 2 (b 2 a 2 )
b2 a 2

(5)

where r is the normal stress in the radial direction (compressive); is the normal stress in the circumferential direction
(tensile); and a and b are as defined previously. The strains are
8

ACI Materials Journal/January-February 2016

r =

1
1
( r ), and = ( r ) (6)
E
E

If there is no external pressure (po = 0), both stresses reach


the maximum values at r = a
( r ) max = pi , and ( ) max = pi

a 2 + b2
(7)
b2 a 2

In derivation of Eq. (1), Liu9 cited an expression to


correlate the rust-caused displacement with the internal
pressure as
pi =

2 Ed s
(8)
a 2 + b2

( + 2 d 0 ) 2

b a2

The pressure described by this equation is that as if to insert


a bar into a hole that is smaller in diameter. Thus, when ds is
the same, the smaller the hole size, the larger the generated
pressure. However, it is incorrect to apply this equation for
the corrosion-induced pressure in concrete cover.
Crack formation due to expansion of bar
For the element of a unit length (l), the relative volume
increase due to bar corrosion is V/V, which equals
lA/(lA) = A/A; A/A is the relative increase in cross-sectional area. Let be the thickness of rust less the corroded
steel, for / << 1, the corrosion resulted strain is
=

steel
A

(9)

= 2
=
2
2
A (C + / 2) ( / 2)
C + C C (C + )

where steel is the corrosion depth.


Let the maximum tensile stress equal the concrete tensile
strength, that is, ft =,max, and insert it into Eq. (7), then one
obtains ft = pi[(a2 + b2)/(b2 a2)]. The maximum radial strain
(by Eq. (6)) is
r

max

pi
p (a 2 + b 2 ) pi
i 2
=
E
E
E (b a 2 )


a 2 + b2 ft b2 a 2
1 + b 2 a 2 = E b 2 + a 2 +

(10)
It may be noted that a number of researchers used incorrect assumptions to correlate the stress or strain to concrete
strength. These include using the thin-wall model to correlate
the pressure to concrete tensile strength, that is, pi =
2C ft,9,31-33 or reasoning that cracking starts when the loop
strain at steel-concrete boundary reaches concrete ultimate
strain, that is, ,max = t = ft/E.16,34
Letting Eq. (9) equal Eq. (10), one obtains the critical
corrosion depth for the first crack to occur
crit =

f t C (C + ) 1 (a / b) 2

f t C (C + ) b 2 a 2
+
+

=
2
2
2

Eeff b + a
Eeff 1+ (a / b)

(11)

ACI Materials Journal/January-February 2016

Fig. 7Critical corrosion depth for 50 MPa (7252 psi)


concrete.
For reinforcing bar in concrete structure, (a/b)2 is a small
value. The largest one for specimens in this research is 0.105
(C = 25, = 24), which gives the ratio (1 (a/b)2)/(1 +
(a/b)2) = 0.81; for the other combinations, this ratio varies
from 0.93 to 1. Equation (11) can be simplified to Eq. (2) for
most of cover-bar combinations.
Comparison of model calculation and test results
The tensile strength, E modulus, Poissons ratio, and creep
coefficient of the concrete were estimated according to
Australian standard AS 3600: ft = 0.36fc, E = 1.5 0.043fc,
Eeff = E/(1 + ), Poisson ratio = 0.20; = 2.0 and 3.4 for
50 and 30 MPa (7252 and 4351 psi) grade concrete, respectively.22 Using the compressive strength and density (Table 2),
the average estimated tensile strength is 2.1 MPa (299 psi)
and 2.8 MPa (404 psi); E modulus is 29.9 GPa (4336 kpsi)
and 41.5 GPa (6018 kpsi) for 30 MPa and 50 MPa concrete,
respectively.
The critical corrosion depth calculated by Eq. (11)
decreases with the bar sizes and increases with the bar
covers as depicted in Fig. 7. It is noticed that there is no
significant difference in the calculated crit for the two grades
of concrete because the ratio ft/Eeff is similar.
On the assumption that the actual corrosion current
density is four times that determined by GECOR6, the value
of corrosion depth for each bar is calculated using Eq. (4) for
the first crack (the time is traced by the length change). The
test results and the predicted critical depth, crit, are shown in
Table 3. crit ranges from 5 to 150 m (0.2 to 5.9 mil) for the
specimens of this research.
The crack formation prediction was reasonably well verified by most of the data, and some cracks were found based
on the prediction. The exceptions are: in some slabs, lateral
cracks between bars occurred before the cover cracking;
there were a few cases where cracks in cover occurred
before corrosion reached the critical value. The latter may
be due to the fact that actual corrosion rate was greater than
that estimated based on the NDT result. Other factors, such
as the estimated concrete tensile strength and modulus of
elasticity, assumed Poissons ratio, the creep coefficient, and
presence of pores (pore near steel bar would accommodate
some corrosion product35), all could contribute to the differ9

Table 3Cover crack formation and depth of corrosion in 5 years


ID

Bar nominal,
mm (in.)

Cover, actual,
mm (in.)

Age, days

corr, m (mil)

crit, m (mil)

tr icorr, r, yrA/cm2*

Depth, m (mil)*

n601-1-06

6 (1/4)

26.5 (1.0)

NA

12.8 (0.5)

0.25

15.6 (0.6)

n601-1-12

12 (1/2)

28.2 (1.1)

NA

8.1 (0.3)

0.26

50.4 (2.0)

lateral

n601-1-24

24 (1)

25.6 (1.0)

1014

5.9 (0.2)

4.1 (0.2)

0.27+

32.1 (1.3)+

n601-2-06

6 (1/4)

53.4 (2.1)

NA

47.7 (1.9)

0.21

24.5 (1.0)

n601-2-12

12 (1/2)

54.0 (2.1)

NA

26.5 (1.0)

0.29

42.4 (1.7)

n601-2-24

24 (1)

53.4 (2.1)

1457

9.8 (0.4)

15.0 (0.6)

0.30+

50.1 (2.0)+

n601-3-06

6 (1/4)

72.2 (2.8)

NA

85.1 (3.4)

0.34

28.2 (1.1)

n601-3-12

12 (1/2)

71.6 (2.8)

NA

44.8 (1.8)

0.35+

17.8 (0.7)

n601-3-24

24 (1)

72.6 (2.9)

1824

18.5 (0.7)

26.0 (1.0)

0.41

67.2 (2.6)+

n602-1-06

6 (1/4)

27.0 (1.1)

NA

19.7 (0.8)

1.52

28.7 (1.1)

lateral

n602-1-12

12 (1/2)

27.4 (1.1)

<990

27.7 (1.1)

11.5 (0.5)

4.50+

117 (4.6)+

n602-1-24

24 (1)

27.4 (1.1)

540

7.9 (0.3)

6.9 (0.3)

3.85+

276 (10.9)+

n602-2-06

6 (1/4)

53.5 (2.1)

NA

71.2 (2.8)

1.72

43.0 (1.7)

n602-2-12

12 (1/2)

53.3 (2.1)

1308

37.1 (1.5)

38.4 (1.5)

2.61

171 (6.7)

n602-2-24

24 (1)

54.7 (2.2)

805

17.6 (0.7)

23.2 (0.9)

3.09+

214 (8.4)+

n602-3-06

6 (1/4)

72.5 (2.9)

NA

127.5 (5.0)

1.91

42.4 (1.7)

n602-3-12

12 (1/2)

71.8 (2.8)

<1639

91.2 (3.6)

66.9 (2.6)

2.64

76.6 (3.0)

n602-3-24

24 (1)

73.2 (2.9)

<1458

42.9 (1.7)

39.2 (1.5)

1.85+

45.5 (1.8)+

n603-1-06

6 (1/4)

26.5 (1.0)

<1131

16.4 (0.6)

12.7 (0.5)

0.63

24.6 (1.0)

n603-1-12

12 (1/2)

26.3 (1.0)

<803

12.3 (0.5)

7.2 (0.3)

0.77

84.5 (3.3)

n603-1-24

24 (1)

26.3 (1.0)

<606

8.7 (0.3)

4.3 (0.2)

1.22+

280 (11)+

n603-2-06

6 (1/4)

54.6 (2.2)

NA

49.6 (2.0)

0.63

28.2 (1.1)

n603-2-12

12 (1/2)

53.2 (2.1)

1131

36.0 (1.4)

25.7 (1.0)

1.24+

51.9 (2.0)+

n603-2-24

24 (1)

54.2 (2.1)

1272

27.3 (1.1)

15.3 (0.6)

0.80+

68.1 (2.7)+

n603-3-06

6 (1/4)

72.5 (2.9)

NA

85.5 (3.4)

0.83

25.4 (1.0)

n603-3-12

12 (1/2)

72.0 (2.8)

NA

45.1 (1.8)

1.17

55.5 (2.2)

n603-3-24

24 (1)

73.2 (2.9)

965

17.5 (0.7)

26.3 (1.0)

0.68

61.2 (2.4)+

n611-1-06

6 (1/4)

25.7 (1.0)

<797

39.3 (1.5)

18.4 (0.7)

8.26

NA

n611-1-12

12 (1/2)

26.0 (1.0)

732

51.3 (2.0)

10.7 (0.4)

12.47+

698 (27.5)+

y, lateral

n611-1-24

24 (1)

25.7 (1.0)

244

7.1 (0.3)

6.3 (0.2)

10.78+

982 (38.7)+

n611-2-06

6 (1/4)

53.5 (2.1)

NA

73.0 (2.9)

13.12

NA

lateral

n611-2-12

12 (1/2)

53.7 (2.1)

<595

39.8 (1.6)

39.9 (1.6)

10.98

NA

n611-2-24

24 (1)

54.7 (2.2)

492

15.0 (0.6)

23.8 (0.9)

6.31

NA

n611-3-06

6 (1/4)

73.0 (2.9)

NA

132.5(5.2)

8.72

NA

lateral

n611-3-12

12 (1/2)

71.8 (2.8)

797

53.4 (2.1)

68.6 (2.7)

11.41+

704 (27.7)+

n611-3-24

24 (1)

72.0 (2.8)

541

17.1 (0.7)

39.0 (1.5)

7.57

667 (26.3)+

n612-1-06

6 (1/4)

26.6 (1.1)

NA

13.0 (0.5)

3.22

NA

n, lateral

n612-1-12

12 (1/2)

28.2 (1.1)

<1500

128 (5.0)

8.2 (0.3)

4.67

NA

y, lateral

n612-1-24

24 (1)

27.2 (1.1)

<450

17.8 (0.7)

4.6 (0.2)

5.51

NA

n612-2-06

6 (1/4)

48.9 (1.9)

NA

40.7 (1.6)

5.82

NA

n, lateral

n612-2-12

12 (1/2)

53.7 (2.1)

NA

26.4 (1.0)

6.43

NA

lateral

n612-2-24

24 (1)

53.1 (2.1)

<562

30.9 (1.2)

14.9 (0.6)

7.82

NA

n612-3-06

6 (1/4)

71.0 (2.8)

NA

83.0 (3.3)

2.32

55.6 (2.2)

n612-3-12

12 (1/2)

71.0 (2.8)

795

35.5 (1.4)

44.4 (1.7)

3.15

166 (6.5)

y, lateral

n612-3-24

24 (1)

73.5 (2.9)

860

33.3 (1.3)

26.8 (1.1)

2.43

241 (9.5)

n613-1-06

6 (1/4)

26.9 (1.1)

NA

20.0 (0.8)

23.07

NA

y, lateral

n613-1-12

12 (1/2)

29.0 (1.1)

<307

19.8 (0.8)

13.0 (0.5)

25.66

NA

y, lateral

n613-1-24

24 (1)

26.0 (1.0)

145

7.9 (0.3)

6.4 (0.3)

7.07

NA

n613-2-06

6 (1/4)

53.8 (2.1)

NA

73.7 (2.9)

16.51

NA

y, lateral

n613-2-12

12 (1/2)

53.2 (2.1)

<483

43.9 (1.7)

39.3 (1.5)

19.90

NA

y, lateral

n613-2-24

24 (1)

53.8 (2.1)

329

16.6 (0.7)

23.1 (0.9)

17.43

NA

n613-3-06

6 (1/4)

72.7 (2.9)

NA

131.4 (5.2)

19.00

NA

y, lateral

n613-3-12

12 (1/2)

72.6 (2.9)

526

53.6 (2.1)

70.0 (2.8)

11.34

751 (30)

n613-3-24

24 (1)

72.3 (2.8)

473

26.6 (1.0)

39.3 (1.5)

13.31+

904 (36)+

Cover crack

*Mass loss test determined corrosion depth at 5 years, except those with +, which were done at 3 years; tricorr,r is time GECOR6 reading (1 A/cm2 = 0.929 mA/ft2).

y is crack observed; lateral is crack between bars occurred before cover cracking; and n is crack was not formed at 5 years. Slabs of n600 did not crack and the results are not
presented herein.

10

ACI Materials Journal/January-February 2016

ence between the estimated critical corrosion depth and that


of actual corrosion depth.
Time to cover concrete cracking
Denoting t1 and t2 as the times to corrosion initiation and
cracking, respectively, calculation of the time-to-cracking is
to find time t2 when (t2) = crit. For the simplest case that
is, the corrosion rate is proportional to the chloride content
in concrete, and chloride distribution obeys the Ficks diffusion lawthe corrosion depth is
t2

t2

(t2 ) = t kC (t , X C )dt = k t [Ci + (CS Ci )(1 erf ( z ))] dt


z23t2 z13t1 z25t2 z15t1 z27 t2 z17 t1
4

= k CS (t2 t1 ) (CS Ci )

+
...
z2 t2 z1t1 +
3
30
210

(12)
where k is the ratio of corrosion rate to the chloride content
(unit: depth/time/chloride content; at t <t1, k = 0), XC is the
cover depth and C(t, XC) is the chloride distribution; CS
and Ci are the chloride content at concrete surface and that
initially present in concrete, respectively; z1 = XC/(4Dt1)
and z2 = XC/(4Dt2), where D is the diffusion coefficient.
When both z1 and z2 < 1, the residual of calculation using
terms shown in Eq. (12) is less than 1%. The Taylor expansion for the error function
erf ( z ) =

2
1 z3 1 z5
1
z 2 n 1
+
+ (1) n 1
+
z
(n 1)! 2n 1
1! 3 2 ! 5

is used for the derivation of Eq. (12). If the steel corrosion


rate is nonlinearly dependent on the chloride content, then it
is necessary to divide the range into more sections where in
each section the dependency is linear.
SUMMARY AND CONCLUSIONS
This paper presents corrosion data of reinforcing steel
bars embedded in concrete premixed with sodium chloride,
and the occurrence of cracks in concrete. A theoretical relation between the critical depth of corrosion and influencing
factors has been established. The main findings are:
The relationship between the corrosion current density
and corrosion potential can be explained by the ButlerVolmer equation.
The corrosion current density measured on steel bars at
room temperature approximately linearly depends on
the amount of chloride in concrete.
The critical corrosion depth crit, the corrosion needed
to cause cover cracking, is larger for smaller-sized bars
and thicker cover.
A higher concrete strength may not mean a larger critical depth because ft/[E/(1 + )] may not be higher.
Time-to-cracking because the corrosion initiation can
be predicted by computing the time during which the
steel corrosion depth reaches crit. An equation for the
prediction is derived and presented in this paper.

ACI Materials Journal/January-February 2016

AUTHOR BIOS

Aimin Xu is a Senior Research Engineer at ARRB Group Ltd, Melbourne,


Australia. He received his BS in civil engineering from Southeast University, Nanjing, China; and Licentiate and PhD in civil engineering from
Chalmers University of Technology, Gthenburg, Sweden. His research
interests include reinforced concrete structure durability and various
aspects of concrete technology.
Ahmad Shayan is a Chief Scientist at ARRB Group Ltd and Adjunct
Professor in the Department of Civil Engineering, Swinburne University
of Technology, Melbourne, Australia. He received his PhD from Sydney
University, Sydney, Australia. His research interests include durability of
concrete structures.

ACKNOWLEDGMENTS

The authors wish to express their gratitude and sincere appreciation to the
Austroads for financing this research work.

REFERENCES

1. The European UnionBrite EuRam III, DuraCreteProbabilistic


Performance Based Durability Design of Concrete Structures, Document
BE95-1347/R0, Gouda, the Netherlands, 1999, 86 pp.
2. Ferreira, M.; rskog, V.; Jalali, S.; and Gjrv, O. E., Software for
Probability-Based Durability Analysis of Concrete Structures, Proceedings of the Fourth International Conference on Concrete Under Severe
Conditions: Environment and Loading, B. H. Oh, K. Sakai, O. E. Gjrv, and
N. Banthia, eds., Seoul National University and Korea Concrete Institute,
Seoul, Korea, 2004, pp. 1015-1024.
3. Silica Fume Association, Life-365: Service Life Prediction Model
and Computer Program for Predicting the Service life and Life-Cycle Costs
of Reinforced Concrete Exposed to Chlorides, V 2.6, 2008, 67 pp. http://
www.nrmca.org/research/life365v2usersmanual.pdf
4. Weyers, R. E., Service Life Model for Concrete Structures in Chloride Laden Environments, ACI Materials Journal, V. 95, No. 4, July-Aug.
1998, pp. 445-453.
5. Shayan, A., and Xu, A., Relationship between Rust Growth, Cover
Thickness and Concrete Cracking, Proceedings of the Australasian
Corrosion Association: Corrosion Control & NDT, V. 2, Paper No. 96,
Melbourne, Australia, 2003, 13 pp.
6. Broomfield, J. P.; Rodriguez, J.; Ortega, L. M.; and Garcia, A. M.,
Corrosion Rate Measurements in Reinforced Concrete Structures by a
Linear Polarization Device, Concrete Bridges in Aggressive Environments,
SP-151, R. E. Weyers, ed., American Concrete Institute, Farmington Hills,
MI, 1994, pp. 163-181.
7. Broomfield, J. P., Corrosion of Steel in Concrete: Understanding,
Investigation and Repair, E&FN Spon, London, UK, 1992, 240 pp.
8. Gonzalez, J. A.; Feliu, S.; and Rodriguez, P., Threshold Steel Corrosion Rates for Durability Problems in Reinforced Structures, Corrosion
Engineering, V. 53, No. 1, Jan. 1997, pp. 65-71.
9. Liu, Y., Modelling the Time-to-Corrosion Cracking of the Cover
Concrete in Chloride Contaminated Reinforced Concrete Structures,
doctoral dissertation, Virginia Polytechnic Institute and State University,
Blacksburg, VA, Oct. 1996, 117 pp.
10. Al-Harthy, A. S.; Stewart, M. G.; and Mullard, J., Concrete Cover
Cracking Caused by Steel Reinforcement Corrosion, Magazine of Concrete
Research, V. 63, No. 9, 2011, pp. 655-667. doi: 10.1680/macr.2011.63.9.655
11. Spellman, D. L., and Stratfull, R. F., Chlorides and Bridge Deck
Deterioration, State of California Department of Public Works, Division of
Highways, Materials and Research Department, 1969, 30 pp.
12. Baant, Z. P., Physical Model for Steel Corrosion in Sea StructuresApplications, Journal of the Structural Division, ASCE, V. 105,
No. 6, June 1979, pp. 1155-1166.
13. Clear, K. C., Measuring Rate of Corrosion of Steel in Field Concrete
Structures, Transportation Research Record No. 1211, Transportation
Research Board, National Research Council, Washington, DC, 1989, pp. 28-37.
14. Hladky, K., Development in Rate of Corrosion Measurements for
Reinforced Concrete Structures, CORROSION 89/169, NACE, Houston,
TX, 1989.
15. Suda, K.; Misra, S.; and Motohashi, K., Corrosion Products of Reinforcing Bars Embedded in Concrete, Corrosion Science, V. 35, No. 5-8,
1993, pp. 1543-1549. doi: 10.1016/0010-938X(93)90382-Q
16. Zhao, Y.; Yu, J.; and Jin, W., Damage Analysis and Cracking Model of
Reinforced Concrete Structures with Rebar Corrosion, Corrosion Science,
V. 53, No. 10, 2011, pp. 3388-3397. doi: 10.1016/j.corsci.2011.06.018
17. Baant, Z. P., Physical Model for Steel Corrosion in Sea StructuresTheory, Journal of the Structural Division, ASCE, V. 105, No. 6,
June 1979, pp. 1137-1153.

11

18. Andrade, C.; Alonso, C.; Rodriguez, J.; and Ortega, L. M., On-Site
Corrosion Rate Monitoring and Its Use to Assess Structural Condition,
Rehabilitation of Structures, Proceedings of the 2nd International RILEM/
CSIRO/ACRA Conference, Melbourne, Australia, 1998, pp. 144-153.
19. Vu, K.; Stewart, M. G.; and Clark, L. A., Pressure Required to Cause
Cover Cracking of Concrete due to Reinforcement Corrosion, ACI Structural Journal, V. 102, No. 5, Sept.-Oct. 2005, pp. 719-726.
20. Lu, C.; Jin, W.; and Liu, R., Reinforcement Corrosion-Induced
Cover Cracking and Its Time Prediction for Reinforced Concrete Structures, Corrosion Science, V. 53, No. 4, 2011, pp. 1337-1347. doi: 10.1016/j.
corsci.2010.12.026
21. Liu, Y., and Weyers, R. E., Modeling the Time-to-Corrosion
Cracking of the Cover Concrete in Chloride Contaminated Reinforced
Concrete Structures, ACI Materials Journal, V. 95, No. 6, Nov.-Dec. 1998,
pp. 675-681.
22. AS 3600-2009, Australian Standard: Concrete Structures, Standards Australia, Sydney, Australia, 2009, pp. 37-38.
23. Gonzlez, J. A.; Molina, A.; Escudero, M. L.; and Andrade, C.,
Errors in Electrochemical Evaluation of Very Small Corrosion Rates
I. Polarization Resistance Method Applied to Corrosion of Steel in
Concrete, Corrosion Science, V. 25, No. 10, 1985, pp. 917-930. doi:
10.1016/0010-938X(85)90021-6
24. Song, G., and Shayan, A., Determination of Corrosion Rate of Reinforcement in Concrete Based on Polarisation Resistance, Transport 98
Proceedings of the 19th ARRB Conference, Sydney, Australia, Dec. 1998,
pp. 263-278.
25. ASTM G1-90(99), Standard Practice for Preparing, Cleaning, and
Evaluating Corrosion Test Specimens, ASTM International, West Conshohocken, PA, 1999, 8 pp.
26. Atkins, P. W., Physical Chemistry, third edition, Oxford University
Press, Oxford, UK, 1986, pp. 699, 804-805.

12

27. Melchers, R. E., Recent Progress in the Modeling of Corrosion of Structural Steel Immersed in Seawaters, Journal of Infrastructure Systems, ASCE, V. 12, No. 3, 2006, pp. 154-162. doi: 10.1061/
(ASCE)1076-0342(2006)12:3(154)
28. ASTM C876-91, Standard Test Method for Half-Cell Potentials
of Uncoated Reinforcing Steel in Concrete, ASTM International, West
Conshohocken, PA, 1991, 6 pp.
29. ACI Committee 222, Protection of Metals in Concrete against
Corrosion (ACI 222R-01), American Concrete Institute, Farmington Hills,
MI, 2001, 41 pp.
30. Timoshenko, S., and Goodier, J. N., Theory of Elasticity, third
edition, McGraw-Hill Book Co. Inc., New York, 1970, pp. 65-77.
31. Li, C. Q.; Lawanwisut, W.; Zheng, J. J.; and Kijawatworawet, W.,
Crack Width Due to Corroded Bar in Reinforced Concrete Structures,
International Journal of Materials & Structural Reliability, V. 3, No. 2,
2005, pp. 87-94.
32. Bhargava, K.; Ghosh, A. K.; Mori, Y.; and Ramanujam, S., Modelling of Time to Corrosion-Induced Cover Cracking in Reinforced Concrete
Structures, Cement and Concrete Research, V. 35, No. 11, 2005, pp. 22032218. doi: 10.1016/j.cemconres.2005.06.007
33. Cusson, D.; Lounis, Z.; and Daigle, L., Durability Monitoring for
Improved Service Life Predictions of Concrete Bridge Decks in Corrosive Environments, NRCC-52708, National Research Council of Canada,
Ottawa, ON, Canada, 2011, 40 pp.
34. Cao, C.; Cheung, M. M. S.; and Chan, B. Y. B., Modeling of Interaction between Corrosion-Induced Concrete Cover Crack and Steel Corrosion Rate, Corrosion Science, V. 69, 2013, pp. 97-109. doi: 10.1016/j.
corsci.2012.11.028
35. Kim, K. H.; Jiang, S. Y.; Jiang, B. S.; and Oh, B. H., Modeling
Mechanical Behavior of Reinforced Concrete Due to Corrosion of Steel
Bar, ACI Materials Journal, V. 107, No. 2, Mar.-Apr. 2010, pp. 106-113.

ACI Materials Journal/January-February 2016

Reproduced with permission of the copyright owner. Further reproduction prohibited without
permission.

You might also like