Corrosion Science: Mohammad M. Kashani, Adam J. Crewe, Nicholas A. Alexander

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Corrosion Science 73 (2013) 208–221

Contents lists available at SciVerse ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Use of a 3D optical measurement technique for stochastic corrosion


pattern analysis of reinforcing bars subjected to accelerated corrosion
Mohammad M. Kashani ⇑, Adam J. Crewe, Nicholas A. Alexander
Dept. of Civil Engineering, University of Bristol, Bristol BS8 1TR, United Kingdom

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

Article history: The 3D corrosion patterns of 23 reinforcing bars subjected to accelerated corrosion are characterised
Received 1 December 2012 using an optical surface measurement technique. A stochastic signal processing methodology is
Accepted 28 March 2013 employed for corrosion pattern analysis of the measured data. The statistical analysis of corrosion pattern
Available online 18 April 2013
data shows that a lognormal distribution model can represent the non-uniform distribution of pitted sec-
tions along the corroded bars. It was observed that the frequency of corrosion is independent from the
Keywords: mass loss ratio and the length of the bars. Finally, a set of probabilistic distribution models for the geo-
A. Steel reinforced concrete
metrical properties of corroded bars is developed.
B. Modelling studies
C. Pitting corrosion
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction effects of inelastic buckling and low-cycle high amplitude fatigue.


Kashani et al. concluded that the most important parameter con-
The corrosion of reinforcing steel in reinforced concrete (RC) trolling the buckling mechanism is the non-uniform distribution
structures is a major issue facing bridge owners and managers. of pitting along the length of corroded bars. This conclusion is
This is one of the most important sources of engineering and eco- somewhat different from that which other researchers have drawn
nomic problems in developed countries. In the UK, the Department based just on tension tests. Failure and necking of a corroded bar in
of Transport (DoT) estimated that salt-induced corrosion damage tension normally occurs at the location of the maximum pit (min-
on motorway and trunk road bridges totalled £616.5 million in imum cross section area). However, when corroded bars are sub-
England and Wales alone [1] and these bridges represent only ject to compression, where inelastic buckling is critical, the
about 10% of the total bridge inventory in the country. By one buckling mechanism is significantly affected by the distribution
American estimate there is $150 billion worth of corrosion damage of pits along the bar. This is because the distribution of pitted areas
on their interstate highway bridges caused by de-icing and sea salt along a corroded bar creates major and minor axis as well as load
induced corrosion [2]. Indeed, severe corrosion of reinforcement eccentricity which are challenging to quantify. Moreover, for
has caused the complete collapse of one structure under live load- inelastic buckling analysis other cross sectional properties of the
ing as reported in [3]. However, many of these deteriorated struc- corroded bars such as the minimum second moment of area and
tures are located in earthquake prone regions where ductility of maximum load eccentricity are also required.
structural members is an important factor for dissipating energy Previously, several researchers have investigated, both experi-
in big earthquake events. Based on recent experimental studies mentally and analytically, the time and spatial variability of the
on corroded RC structural members, corrosion of reinforcement maximum pitting depth of corroded reinforcing bars, prestressing
will result in a significant reduction in ductility and in premature wires and steel pipelines [14–18]. The outcomes of these studies
buckling of longitudinal reinforcement in plastic hinge regions are a set of time-variant probabilistic distribution models of the
[4,5]. maximum pitting depth. These models have made a major contri-
In the past decade, several researchers have investigated the bution towards the structural safety and reliability analysis of RC
influence of corrosion on stress–strain behaviour of corroded rein- beams and slabs [19–23]. However, for capacity loss estimation
forcement in tension [6–11]. More recently Kashani et al. [12,13] of corrosion damaged columns subject to earthquakes i.e. bridge
investigated the influence of corrosion on the nonlinear stress– piers; buckling of vertical bars is one of the most important perfor-
strain behaviour of corroded reinforcement subjected to mono- mance parameters that limit the plastic rotation capacity of col-
tonic compression and to cyclic loading taking onto account the umns. Therefore, there is a need for probabilistic models that can
capture the time-dependent distribution of pitted areas along the
⇑ Corresponding author. Tel.: +44 7896368691. length of corroded bars. This has a significant influence on the
E-mail address: mehdi.kashani@bristol.ac.uk (M.M. Kashani). seismic reliability analysis of corroded RC structures. The research

0010-938X/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.03.037
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 209

presented in this paper addresses these issues using a novel corro- Power Supply
sion pattern analysis of corroded reinforcing bars. The corroded bar
specimens were produced using an accelerated corrosion tech-
nique in the laboratory. Using an advance 3D optical scanning
technique the surface corrosion pattern of bars was then mea-
sured. Based on the statistical analysis of corrosion pattern data,
a set of time-variant probabilistic models have been developed Stainless Steel Plate
for the geometrical properties of corroded bars. These models are (Cathode)
the input parameters of a probabilistic platform for future research
in spatial-time-dependent seismic reliability analysis of deteriorat-
ing systems.
RC Specimen
Immersed in NaCl
Solution
2. Experimental programme

2.1. Test specimens Reinforcement


(Anode)
A total of eight reinforced concrete specimens were cast as
shown in Fig. 1. Each specimen was dimensioned
250  250  700 mm and incorporated 8 No. 12 mm diameter Fig. 2. Schematic illustration of accelerated corrosion procedure.
reinforcing bars for use in a variety of buckling and cyclic tests.
The concrete mix was designed to have a mean compressive Assuming a uniform mass loss, the mean reduced diameter of
strength of 30 MPa at 28 days with a maximum aggregate size of the reinforcement can be estimated using Eq. (1) which gives an
12 mm. The specimens were cast with nominal cover of 25 mm. average residual diameter of reinforcement (DCorr) relative to the
mass loss:
pffiffiffiffiffiffiffiffiffiffiffiffi
2.2. Accelerated corrosion procedure DCorr ¼ D0 1c ð1Þ
where D0 is the initial diameter of the uncorroded bar and c is the
The corrosion process in the natural environment is very slow
mass loss ratio based on Eq. (2):
and time consuming. Even in a very severe environment it takes
 
years for a real structure to reach to a severe corrosion damage le- m0  m
vel. Therefore, this procedure needs to be accelerated in a labora-
c¼ ð2Þ
m0
tory environment where there are time limitations for
experiments. An accelerated corrosion technique (known as anodic where m0 is the mass per unit length of the original steel bar, m the
corrosion) was employed in this research. The concept of anodic final mass per unit length of the steel bar after removal of the cor-
corrosion is simple and consists of forming an electrochemical cir- rosion products. A more detailed discussion of the accelerated cor-
cuit using an external power supply. The reinforcing bars act as an rosion procedure can be found in [12,13].
anode in the cell and an external material acts as the cathode as
shown in Fig. 2. The external material used for the cathode in this 2.3. Summary of the experimental observations of mechanical testing
experiment was stainless steel plate.
The time required to get the desired corrosion level was esti- The experimental programme for this research consisted of
mated using Faraday’s 2nd Law of Electrolysis [12]. The corrosion three parts: (a) investigation of the influence of non-uniform corro-
current density used in these tests was 1.1 mA/cm2. This is about sion on the inelastic buckling behaviour of corroded bars under
100 times higher than the average corrosion current density re- monotonic compression loading, (b) investigation of the influence
ported under natural conditions. After the accelerated corrosion of non-uniform corrosion on the nonlinear cyclic response of cor-
procedure, the concrete specimens (shown in Fig. 1) were broken roded bars accounting for the effect of inelastic buckling and (c)
open and the corroded bars were carefully removed from the con- statistical analysis of the corrosion patterns of corroded bars. A
crete. To ensure that the concrete was completely removed from comprehensive description of the experimental procedure and de-
the corroded bars, a mechanical cleaning process using a bristle tailed discussion of the results of mechanical testing are reported
brush was used, in accordance with ASTM G1-03 [24]. The cor- elsewhere [12,13]. In this section only a summary of the important
roded bars were then washed with tap water and dried. The brush- observations that are related to this paper are reported.
ing and washing process was then repeated a second time. It The slenderness ratios of the corroded bars for the mechanical
should be noted that the same brushing process was applied to testing were chosen based on the ratios of horizontal tie spacing
the uncorroded control specimens and it was found that the effect in the common types of construction of RC columns (L) to bar
of brushing on the mass loss of base material was negligible. diameter (D) known as the L/D ratio. The L/D ratios tested were
5, 8, 10, 15, and 20 for the monotonic buckling tests and 5, 10
and 15 for the cyclic tests. Three control (uncorroded) specimens
8B12 were also tested for each slenderness ratio. It should be noted that
the measured mass loss of the bars was calculated over the buck-
ling length considered in each test.
A total of 120 corroded bars were prepared, 57 were used for
250 monotonic buckling tests and 40 for cyclic tests. A further 23 cor-
roded bars were randomly selected from the corroded bar speci-
mens (1 in 5) with different mass loss ratios for statistical
750 250 analysis of the corrosion patterns.
The results of the buckling tests showed that corrosion has a
Fig. 1. Test specimens subject to accelerated corrosion. significant influence on the buckling capacity of corroded bars.
210 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

The results indicated that a 10% mass loss will result in about a 20%
reduction in the buckling capacity of corroded bars with L/D of 8 or
more. It was found that the most important parameter affecting
the buckling mechanism was the distribution of the pitted areas
along the corroded bar. The non-uniform distribution of pitted sec-
tions results in a change in the buckling mechanism of the cor-
roded bars. Fig. 3 shows a corroded bar cut into five slices after a
buckling test. After slicing the bar the minimum diameter of each
section was measured with digital Vernier Caliper and is shown
under each section in Fig. 3. As shown in Fig. 3 the shape of cross
sections is no longer circular and also varies along the bar. This
phenomenon creates strong and weak axes along the bar and intro-
duces a load eccentricity. The combined effect of load eccentricity
and a reduction in cross section induces local stress concentrations
at critical cross sections. As a result, plastic hinges will form at the
location of these critical sections.
Similar buckling behaviour was observed in the compression
part of the cyclic tests. The non-uniform distribution of pitted cross Fig. 4. Coordinate system of the 3D surface measurement.
sections significantly reduced the low-cycle high amplitude fatigue
life of the corroded bars. The stress concentrations at severely pit- examples of the 3D solid elements created using the optical mea-
ted sections also resulted in premature fracture of the corroded surement data.
bars in tension. The reduction in buckling capacity was also more The software associated with the scanner is capable of generat-
severe due to the influence of the strain history combined with ing 3D meshes and generating cross sections and profiles through
the effect of low-cyclic high amplitude fatigue. It was also observed the solid element. For processing the data a 3D polygon mesh was
that corrosion affected the geometrical slenderness ratio of the cor- generated for each solid element. In order to explore the corrosion
roded bars which resulted in a significant reduction in the energy patterns and distribution of pits, for each bar cross sections were
dissipation capacity of the corroded bars. Therefore the uncertainty taken through the 3D solid mesh at 0.5 mm intervals along the
associated with the pitting distribution in corroded bars needs to whole length of the solid elements. The cross sections were trans-
be investigated probabilistically. This is the main objective of this formed to data points (x, y, z coordinates) and each cross section
paper and is discussed in the following sections. included data points at 0.03 mm intervals around the perimeter.
The raw data points were then exported to MATLAB [25] for
3. Investigation of corrosion patterns using an advanced optical post-processing and data analysis. Fig. 6 shows an example of cross
measurement technique sections extracted from the 3D model.
With reference to Fig. 6, the cross sections taken from two dif-
3.1. 3D optical measurement procedure ferent locations along the same bar are different in shape and area.
As expected, the change in the cross section shape and geometrical
As mentioned in Section 2.3, 23 corroded reinforcing bars were properties of the bar are not captured using the average reduced
taken out of the total of 120 samples for more refined geometrical cross section based on Eq. (1).
surface analysis of corrosion patterns. The reinforcing bars varied It should be pointed out that there is one section on the plot of
in length (from 220 mm to 400 mm) and had a range of mass loss Cross Section 2 in Fig. 6 which is larger than the nominal diameter
ratios (8.93–55.94%). of the original bar. This is due to the presence of ribs on the bar.
The surface pitting pattern of the corroded bars was measured This will influence the cross sectional area and geometrical proper-
using a structured light scanner with 5.0 MP resolution. The accu- ties of corroded bars taken from the scanned data. Therefore, the
racy of the measurement was set to 20 lm which was the highest raw data needs some processing to filter out the influence of the
resolution for the instrument. Given the high accuracy of the mea- ribs. This is discussed in detail in Section 4 of this paper.
surement technique it was very sensitive to any vibration during To better understand the corrosion patterns, the cross section of
the scanning process. Therefore, the corroded bars were securely the bars can be unwrapped (p 6 h 6 p) and the data presented in
fixed to a turntable to facilitate the scanning process. The direction the format of r and h in polar coordinates (Fig. 4). This also helps
and coordinate system of the measurement is shown in Fig. 4. This with investigating the influence of ribs on the corrosion patterns.
coordinate system has been used throughout this paper for the
analysis of the scanned data.
The outcome of the scanning process was a set of 3D solid mod-
els of the corroded bars that were identical to the real bars includ-
ing the very fine details of pitting pattern. Fig. 5a–c shows

Fig. 3. Change in buckling mechanism due to variation in cross section shape and Fig. 5. 3D solid models generated using 3D scans of corroded bars: (a) 10.37%
non-uniform distribution of pitted sections (dimensions in mm). average mass loss (b) 30.87% average mass loss and (c) 55.94% average mass loss.
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 211

Using the equations above all of the geometrical properties of


the corroded cross sections can be calculated. The correlation be-
tween the geometrical properties of pitted sections, original sec-
tion and mass loss ratio can then be investigated. Using the
geometrical properties of the corroded bars the pitted cross sec-
tions could be transformed into an idealised arbitrary section for
buckling analysis using a fibre technique with the geometrical
properties equivalent to the pitted section. Load eccentricity can
also be quantified which is very important in inelastic buckling
analysis.

4. Frequency analysis and signal processing of corrosion pattern


data

4.1. Power spectral density estimate

A Power Spectral Density (PSD) is normally used to characterise


the frequency content of a time-series random signal [26]. The pur-
pose of estimating the PSD is to determine if there is any periodic
pattern in the data by observing peaks at the frequencies corre-
sponding to these periodicities. The Fast Fourier Transform (FFT)
is one of the well-known methods for frequency analysis of time-
Fig. 6. Cross section of a corroded bar with 54.34% average mass loss taken from the series signals. This algorithm transforms the signal from time do-
3D solid model. main to frequency domain to evaluate the periodic pattern of the
signal and estimate the PSD. Welch [27] has developed an efficient
Fig. 7a shows a typical 3D surface plot of an unwrapped cross sec- algorithm based on the FFT for estimation of the PSD. This algo-
tion of a corroded bar and Fig. 7b shows the contour plot of the rithm involves sectioning the time series, taking modified periodo-
same bar. The contour plot is a topographical map of the corrosion grams of these sections, and averaging these modified
pattern over the entire length of the bar. A localised pitting corro- periodograms. This is computationally more efficient than other
sion pattern and the ribs on the bar can be seen in Fig. 7. methods and is very useful for testing and measuring nonstation-
From the values of r and h all the geometrical properties of each arity in random data, further details are available in [27]. In this re-
cross section can be calculated over the length of each bar. The search the Welch method is employed for PSD estimation of the
method that was used for calculation of geometrical properties of measured data. It should be noted that the data in this paper are
bars is discussed in the following section. represented in spatial domain series. Therefore, the unit of fre-
quency is mm1.
3.2. Calculation of the geometrical properties of corroded bars using The radius data of each corroded bar r, calculated from the mea-
the 3D optical measurement data sured data (shown in Figs. 6 and 7), were saved in a 2D matrix. The
rows of the matrix represented the angles (hi) of each ri around the
As shown in Fig. 6a the cross section shape of corroded bars is a perimeter of section and columns of matrix represented the posi-
simple polygon (a non-self intersecting polygon). The area of a sim- tion of each ri along the length of the bar. The PSD of all radii along
ple polygon with n vertices can be calculated using Eq. (3) which is the bar for each hi was estimated for each bar. Fig. 8 shows an
based on Green’s Theorem. example of the PSD calculated for two different bars at h = p/2.
As shown in Fig. 8 the frequency at 0.1292 mm1 is identified as
1Xn1
A¼ ðy zi1  yiþ1 zi Þ ð3Þ the frequency of ribs and the next peak at 0.2701 mm1 as the sec-
2 i¼0 i ond harmonic of that frequency. This was validated by testing the
coherence of several PSDs of different bars. In addition, as an extra
where y and z are the coordinate of the vertices.
check the frequency of ribs was calculated by physical measure-
Subsequently by taking moments about the centre (0,0 coordi-
ment of rib spacing on the bars. The physical measurement showed
nate) the centroid and second moment of area of the polygon about
that the ribs were at 7.85 mm intervals which gives a frequency of
y and z axis can be calculated using the following equation:
1/7.85 = 0.1274 mm1. This rib frequency can be considered as
1 X
n1
noise along, what would otherwise be a smooth circular bar. There-
Cy ¼ ai ðyi þ yiþ1 Þ ð4Þ fore, it needs to be filtered from the data to leave just the influence
6A i¼0
of corrosion on the bars.

1 X
n1
Cz ¼ ai ðzi þ ziþ1 Þ ð5Þ 4.2. Design of band-stop filter
6A i¼0

In order to better distinguish the locations of the corrosion


1 X
n1
along the bars the repeating rib pattern has been removed by filter-
Iy ¼ ai ðz2i þ zi ziþ1 þ z2iþ1 Þ ð6Þ
12 i¼0 ing the raw profile data. A pair of 2nd order band-stop butterworth
filters were employed, one centred at 0.13 mm1 with a width of
1 X
n1 0.03 mm1, the other centred at 0.27 mm1 with a width of
Iz ¼ ai ðy2i þ yi yiþ1 þ y2iþ1 Þ ð7Þ 0.03 mm1. Two filters were used to remove both the main repeat
12 i¼0
period of the ribs and also the 2nd harmonic. Use of band-stop fil-
ters meant that only the rib pattern was targeted by the filtering.
ai ¼ ðyi zi1  yiþ1 zi Þ ð8Þ
Therefore, both lower and higher frequency corrosion components
212 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

Fig. 7. Corrosion pattern of a corroded bar with 54.23% mass loss: (a) 3D surface plot of corrosion pattern and (b) contour plot of corrosion pattern.

4.3. Result of band-stop filtering process

Fig. 10a and b shows a comparison of the contour plot of the


profile data of a corroded bar with 8.93% mass loss before and after
the filtering process. As it is evident from Fig. 10 that the rib pat-
tern has been removed from the raw data and the pattern of pitting
corrosion with all the fine details is left unaffected.
Fig. 11a also shows a longitudinal profile of a slice through the
length of another corroded bar with a 54.23% mass loss at
h = 2.631 rad. Fig. 11b shows the PSD estimation of the same profile
before and after the filtering process. It should be noted that the 3D
surface and contour plot of this corroded bar are previously shown
in Fig. 7. This bar had one of the most complex corrosion patterns
but the band stop filtering successfully kept the corrosion pattern
while removing the rib frequency.
Fig. 8. Welch PSD estimate of the radius r at h = p/2 for two bars with different mass From the filtered data the geometrical properties of each cor-
loss ratios.
roded bar were calculated. Fig. 12a shows examples of the varia-
tion, along the length of the bars, of the residual cross section
areas (A0 ) of three corroded bars normalised to the original uncor-
were maintained. Once the rib pattern was removed from the data roded cross section area (A0). Fig. 12b and c show the variation of
the radius data was then processed to give the geometrical proper- residual second moment of areas (I0y and I0z ) normalised to the ori-
ties of the bars at each location down their length based on the ginal second moment of area (I0) and Fig. 12d shows an example
method explained in Section 3.2 of this paper. Fig. 9 visualises orbit of centroid movement of a corroded bar with 54.23% mass
the band-stop filter designed for this analysis. loss in relation to the original bar diameter. Fig. 12 shows that
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 213

non-uniform corrosion results in a significant variation in the geo-


metrical properties of corroded bars along their length.
Fig. 13 shows the relationship between the ratio of minimum
cross section area A0min and the average reduced cross section area
Aave (i.e. the pitting coefficient of the minimum area) and the per-
centage mass loss w (w = 100c).
As shown in Fig. 13 there is a linear relationship, with a negative
slope, between A0min =Aav e and the percentage mass loss. This shows
that the pitting coefficient is not constant but decreases with an in-
crease in the percentage mass loss. In other words, the area loss
due to pitting increases with an increasing mass loss. The slope
of the regression line (0.0083 as shown in Fig. 13) is the rate of
change of the pitting coefficient. Fig. 13 only shows the pitting
coefficient for the minimum area (maximum area loss). However,
the pitting coefficient also varies (randomly distributed) over the
length of a corroded bar with a known percentage mass loss.
Accordingly, a probabilistic platform is developed in this paper to
Fig. 9. Butterworth band-stop filter visualisation. model the uncertainties associated with pitting distribution. The
details of this methodology are explained in the following sections.

5. Probabilistic-based modelling of time-variant distribution of


pitting corrosion

5.1. Probability distribution functions for geometrical properties of


corroded bars

Hawn [28] and Sheikh et al. [29] investigated the distribution of


maximum pitting depth of corroded steel pipelines. They sug-
gested that a Gumbel distribution can be used to model the distri-
bution of maximum pits. Darwaman and Stewart [16] studied the
probabilistic distribution of maximum pitting depth of corroded
prestressing wires. Stewart and Al-Harthy [21] investigated the
probabilistic distribution of the maximum pitting depth of cor-
roded reinforcement. Darwaman and Stewart [16] and Stewart
and Al-Harthy [21] both adopted a Gumble distribution function
to model the statistical distribution of the maximum pit depth.
In this section a set of probabilistic models are developed that
account for: (a) the change of cross section area over the length
Fig. 10. Comparison of the contour plot of a corroded bar with 8.93% mass loss of corroded bars, (b) the change in second moment of area of cross
before and after filtering the ribs pattern: (a) before filtering and (b) after filtering. sections over the length of corroded bars and (c) the influence of

Fig. 11. Pitting corrosion pattern of a corroded bar with 54.23% before and after filtering process: (a) longitudinal profile through the length at h = 2.631 rad and (b) PSD of the
profile shown in (a).
214 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

Fig. 12. Geometrical properties of corroded bars at section level: (a) normalised A0 (b) normalised I0m (c) normalised I0z and (d) orbit of centroid movement along the whole
length of a corroded bar with 54.23% mass loss.

corrosion on load eccentricity. These parameters are defined


below:

(a) Area pitting coefficient:

A0 ðxÞ
bðxÞ ¼ ð9Þ
Aa v e
where A0 ðxÞ is the cross section area including the pitting effect as a
function of length x, Aave is the average reduced cross section area
calculated using Eq. (1) (assuming a uniform volumetric mass loss)
and bðxÞ is the area pitting coefficient as a function of length x.

(b) Residual second moment of area coefficient:

In the calculation of the residual second moment of area consid-


eration needs to be given to the axis rotation. This is due the irreg-
ular shape of the corroded cross section and the movement of the
Fig. 13. Regression analysis of the pitting coefficient of the minimum cross section section centroid. As a result each cross section has three compo-
area. nent second moment of areas i.e., the second moment of areas at
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 215

the origin (0,0), the second moment of areas at the centroid, and coefficient of the residual minimum second moment of area is then
the second moment of areas about the principal axis. Given that defined in the following equation:
the minimum second moment of area is critical, it was decided
to investigate the probabilistic distribution of this variable. To ac-
I0min ðxÞ
count for the axis rotation, the minimum second moment of area gðxÞ ¼ ð10Þ
I0
about the principal axes was considered in the calculations. The

Fig. 14. Probability and cumulative probability graphs of the random variables.
216 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

Table 1
Statistical dependence of the lognormal distribution model parameters and mass loss ratio.

Model parameter Pearson Kendall Spearman


q P-value s P-value qsp P-value
b
lb 0.7965 9.20  106 0.7316 1.26  107 0.8961 3.35  106
rb 0.496 0.0189 0.4545 0.0026 0.6002 0.0038
g
lg 0.9839 2.04  1016 0.9134 6.87  1014 0.9808 3.11  106
rg 0.8849 4.53  108 7.49  101 4.91  108 0.904 3.66  106
m
lm 0.6844 3.16  104 0.6037 7.89  105 0.823 1.43  106
rm 0.1448 0.51 0.0966 0.54 0.1347 0.54

lb and rb are the lognormal model parameters related to the area pitting coefficient b.
lg and rg are the lognormal model parameters related to the residual second moment of area coefficient g.
lm and rm are the lognormal model parameters related to the load eccentricity coefficient m.

where I0min ðxÞ is the minimum residual second moment of area of the length of corroded bars. However, it should be noted that the
corroded section including the axis rotation and pitting effect as a values of the lognormal parameters are different in bars with dif-
function of length x, I0 is the second moment of area of the original ferent mass loss ratios. Therefore, it is important to find the corre-
uncorroded cross section and g(x) is the coefficient of the residual lation between the model parameters and mass loss ratio.
minimum second moment of area as a function of length x. The general form of the lognormal distribution model is defined
in the following equation:
(c) Load eccentricity coefficient: !
1 ðln g  lg Þ2
The load eccentricity is considered as the ratio of the product of f ðgjlg ; rg Þ ¼ pffiffiffiffiffiffiffi exp g>0 ð13Þ
g r g 2p 2r2g
the centroid in y and z axes to the original bar diameter of uncor-
roded bar as the following equation: where g is the random variable (either of b, g and m), and lg and rg
eðxÞ are the corresponding lognormal model parameters. The mean (Mg)
tðxÞ ¼ ð11Þ and standard deviation (Vg) of each random variable are functions of
r0
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi lg and rg as defined in the following equation:
!
eðxÞ ¼ cy ðxÞ2 þ cz ðxÞ2 ð12Þ r2g
Mg ¼ exp lg þ ð14Þ
where e(x) is the coefficient of load eccentricity ratio as a function of 2
length x, r0 is the original bar radius, cy(x) and cz(x) are the section
centroid in the y and z axes respectively. V g ¼ ðexpð2lg þ rg 2ÞÞðexpðr2g Þ  1Þ ð15Þ
The probability distribution is an important component in
uncertainty modelling and selection of a particular distribution The correlation of the random variables and the mass loss ratio
may significantly affect the characteristic values based on the se- was investigated using Pearson’s linear correlation coefficient (q)
lected distribution model. The modelling strategy that Sriramula and nonparametric hypothesis tests such as Kendall’s rank correla-
and Chryssanthopoulos [30] used for uncertainly modelling of tion coefficient (s) and Spearman’s rank correlation coefficient
the mechanical properties of GFRP panels is employed here for (qsp) [32]. The calculated correlation coefficients together with P-
uncertainty modelling of the random variables (b, g and m). values for each method at 0.05 significance are shown in Table 1.
Several distribution models have been fitted to the observed The results of the correlation analysis show that there is very
random variables (b, g and m). The goodness of fit was investigated strong correlation between most of the model parameters and
using Chi-square (C-S) and Kolmogorov–Smirnov (K–S) hypothesis the mass loss ratio. However it was found that there is only a very
tests [31]. It was found that the two parameter lognormal distribu- weak correlation between the model parameter rm and the mass
tion is supported by both tests at a significance level of 0.05 (95% loss ratio. This is clear from the corresponding P-values of the ran-
confidence) for all three random variables (b, g and m). In order dom variables b, g and lm which are all less than the considered
to visualise the goodness of fit for demonstrative purposes, a com- significance level (0.05). This shows that the dependence of the
parison between representative probability graphs of the random model parameters of random variables b, g and lm to the mass loss
variables and a Weibull, normal and lognormal fit is presented in ratio is statistically significant. However, the calculated P-values of
Fig. 14a–e. Fig. 14b–f shows example cumulative probability the model parameter rm for the random variable m are all greater
graphs of the random variables presented in Fig. 14a–e and pre- than the considered significance level. This shows that there is
dicted values using the lognormal fit together with 95% prediction no correlation between the model parameter rm and the mass loss
bounds. ratio. Therefore, the dependence of the model parameter rm and
From Fig. 14 it is evident that the predicted values based on a the mass loss ratio is statically insignificant. It should be noted that
lognormal fit are generally within the 95% prediction bounds. other distribution models i.e. Weibull distribution; show the same
Therefore, based on null hypothesis tests it can be concluded that result for the random variable m.
a lognormal distribution model can represent the distribution of In summary, the results of the correlation analysis show that as
the random variables (b, g and m). the level of corrosion increases the mean values of random vari-
ables b and g decrease (indicating an increase in the reduction of
5.2. Correlation between the parameters of the probabilistic models area and second moment of area) and the standard deviations in-
and the mass loss ratio crease (indicating that the variation of the data is increasing).
However, the mean value of random variable m increases as the le-
As shown in the previous section the lognormal distribution vel of corrosion increases (indicating that the load eccentricity is
model can represent the distribution of the corrosion pattern over increasing) but, given that the variation of the centroid is
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 217

numerically very small, the change in mass loss ratio does not have where lg and rg are the model parameters of random variables (b, g
a significant influence on the standard deviation of variable m. and m), a, b, c and d are the regression coefficients that are defined in
The interrelationship between the model parameters and the Table 2 and w is the percentage mass loss (w = 100c).
mass loss ratio is modelled using regression analysis of the data. It should be noted that, given that the model parameter rm is
The results of the regression analysis are shown in Fig. 15a–e. statically independent from the mass loss ratio, the average value
The relationships between the model parameters and the mass loss of rm = 0.685 is considered in the probabilistic models.
ratio are defined by empirical Eqs. (16) and (17) which are based Using empirical Eqs. (16) and (17) it is possible to generate a set
on the regression analysis. of probabilistic distribution models for the random variables (b, g
and m) for any given mass loss ratio. In other words, these models
lg ¼ awb ð16Þ can represent the time-variant distribution of the geometrical
properties of corroded bars. The results of these models with a
rg ¼ cwd ð17Þ numerical example are discussed in the Section 5.3.

Fig. 15. Regression analysis of the probabilistic model parameters.


218 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

Table 2
The proposed probabilistic model parameters as a function of the mass loss ratio.

Model parameter a b c d
b
lb 0.000052 1.825
rb 0.0006491 1.526
g
lg 0.008811 1.354
rg 0.001768 1.495
m
lm 8.473 0.304

lb and rb are the lognormal model parameters related to the area pitting coefficient
b.
lg and rg are the lognormal model parameters related to the residual second
moment of area coefficient g.
lm and rm are the lognormal model parameters related to the load eccentricity
coefficient m.

Fig. 17. CCF plot of two pairs of corroded bars with different mass loss ratios and
lengths: (a) 200 mm length with 21. 42% and 36.40% mass loss and (b) 300 mm
length with 19.93% and 30.32% mass loss.

Fig. 16a and b shows example ACF graphs for two corroded bars
with different lengths and mass loss ratios.
The length L0 shown in Fig. 16 indicates the length lag at which
there is no correlation between the cross section areas. In other
words the similarity between the cross section area at length l of
a corroded bar and the cross section area at length l + L0 is zero.
As shown in Fig. 16 the value of L0 is different for bars with differ-
ent lengths and mass loss ratios. Therefore, a comparison between
corroded bars with the same length but different mass loss ratios
was made using a CCF. Fig. 17a and b shows two example CCF
graphs each for a pair of corroded bars with the same length but
with different mass loss ratios.
Fig. 16. ACF for the cross section area of two corroded bars with different lengths
As shown in Figs. 16 and 17 the length L0 generally varied be-
and mass loss ratios: (a) 36.40% mass loss and (b) 19.93% mass loss.
tween 20 mm and 50 mm in all bars. The ACF and CCF analysis
showed that, in most cases, within a length (l) 10 mm 6 l 6 20 mm
So far we have discussed the probabilistic distribution of geo- the variation in correlation coefficient is low, while if l P 20 mm
metrical properties of corroded bars as function of the mass loss ra- the variation in correlation coefficient can be large and hence the
tio. However, an important parameter that needs to be considered variation between cross section areas is significant. It was found
in using the above probabilistic models is the length dependence of that within a critical length (Lc) equal or less than 10 mm the AFC
the model parameters. This is discussed in the following section. is always above 0.7 and CCF is always above 70% of the maximum
calculated CCF. In other words, the variation of cross section areas
within 10 mm of any given point along the bar is relatively small.
5.3. Autocorrelation function, cross-correlation function and spatial The value of Lc is important in spatial-variant probabilistic model-
variability of corrosion pattern ling of pitting distribution in corroded bars. For example, if a cor-
roded bar is 100 mm long it can be idealised by 10 mm segments.
The cross-correlation function (CCF) indicates the correlation, or The above probabilistic models can then be considered separately
similarity, of two time-series random variable as function of a for each 10 mm segment of the whole bar. It should be pointed
time-lag applied to one of them. This is widely used in signal pro- out that if a smaller Lc (Lc 6 10 mm) is considered, more accurate
cessing i.e. searching a long signal for a shorter signal and in pat- results can be obtained, however the analysis process will be
tern recognition [26]. The autocorrelation function (ACF) is a computationally more expensive. On the other hand, if a larger crit-
special case of the CCF being the CCF of the signal with itself. The ical length (Lc P 10 mm) is considered, there is always a possibility
ACF indicates the similarity/correlation between observed data as that we lose some of the variation in the corrosion pattern data.
a function of the time separation between points [26]. This is an Further analytical and experimental research is required to
efficient mathematical tool for finding repeating patterns, such as investigate the influence of Lc on the mechanical response predic-
the presence of a periodic signal which has been hidden by noise tion of corroded bars. There are other factors that might influence
in the time-domain. The ACF is often used in signal processing the length Lc such as corrosion rate, concrete mix (cement and
for analysing functions or series of values, such as time-domain aggregate types) and exposure type (chloride induced corrosion
signals. or carbonation induced corrosion). Given that all of the test speci-
In this research the CCF and ACF have been used to investigate mens in this research were corroded at the same corrosion rate and
the spatial variability of the cross section areas along the corroded were taken from concrete specimens with the same concrete mix,
bars as a function of length separation between the cross sections. these parameters could not be investigated in this experiment. This
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 219

Fig. 18. Results of time-variant probabilistic models: (a) and (b) variation of area pitting coefficient with time, (c) and (d) variation of residual minimum second moment of
area with time and (e) and (f) variation of load eccentricity coefficient of corroded reinforcement with time.

is an important area for further research and the methodology taken based on the correlation analysis of the various corrosion
developed in this work will allow researchers to analyse the corro- patterns found in this work. Therefore, the probabilistic models
sion patterns of any corroded reinforcement in the future. Never- developed in this research combined with the suggested Lc can
theless, a suggested conservative value of Lc = 10 mm can be safely be used in uncertainty modelling of corrosion pattern.
220 M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221

5.4. Results of probabilistic models ture. However, the models developed in this paper can predict the
cross section loss as well as other geometrical properties of the cor-
In order to demonstrate the utilisation of the probabilistic mod- roded bars based on a non-destructive predictive method. This will
els that were developed in Section 5.2, some of the results of these significantly improve the maintenance optimisation techniques for
models are presented in this section. the entire bridge network. Furthermore, if the bridge network is lo-
Consider a RC bridge pier that is subject to chloride attack due cated in an earthquake prone region where the response of struc-
to the splashing of de-icing salts from the adjacent road. Assume ture is severely affected by corrosion of reinforcement [4,5] the
the cover concrete C = 50 mm, the water cement ratio w/c = 0.4 time-dependent seismic performance of the network is very
and the diameter of main reinforcement is 20 mm. The deteriora- important. Therefore, in recent years, other researchers have inves-
tion mechanism has two main stages which are (a) the corrosion tigated the time-dependent seismic performance of corroded
initiation period and (b) the corrosion propagation period bridges using finite element analysis assuming a uniform corrosion
[33,34]. Several researchers have investigated the critical time to [42–45]. The result of this research provides a probabilistic plat-
corrosion initiation, caused by diffusion of chloride ions through form allowing researchers to create more realistic uncertainty
the cover concrete to the reinforcement, and as a result of their ef- modelling for seismic reliability analysis of corroded RC bridges
fort several probabilistic models have been developed [33,34]. and other structures.
Other researchers studied the corrosion propagation period i.e.
crack initiation and crack propagation due to the penetration of
corrosion products into concrete [35–39]. In this example we only 6. Conclusion
consider the corrosion propagation period to demonstrate the
application of the proposed models. Once corrosion is initiated, The non-uniform pitting corrosion pattern of reinforcement
using Eqs. (18)–(21) the average reduced area and the mass loss ra- subjected to accelerated corrosion has been measured using a 3D
tio can be calculated. optical scanning technique. The data generated by the optical mea-
The average reduced diameter of a corroded bar assuming a surement has been used in spatial-domain stochastic corrosion
uniform mass loss can be calculated using the following equation: pattern analysis and uncertainty modelling of the corroded sec-
tions. The main outcomes of the research presented in this paper
Dcorr ðtÞ ¼ D0  2Pav e ðtÞ ð18Þ
are summarised below:
where Dcorr (t) is the average reduced diameter of a corroded rein-
forcing bar at time t after corrosion initiation, D0 is the original (1) It was found that the frequency of the corrosion pattern
bar diameter and Pave (t) is the average corrosion penetration depth ranges from 0.0071 mm1 to 0.0055 mm1 and is indepen-
based on the uniform volumetric mass loss at time t after corrosion dent from the mass loss ratio and the length of the bars. This
initiation. Pave (t) can be calculated using the following equation: shows that the corrosion frequency is a time-invariant vari-
Z able. In other words the frequency of corrosion depends on
t
Pav e ðtÞ ¼ j icorr ðtÞdt ð19Þ the nature of corrosion not the duration of corrosion. How-
T cr ever, this conclusion is drawn based on the results of this
experimental programme. Further research is required for
where icorr(t) is the corrosion current density at time t after corro-
investigation of the influence of other parameters such as
sion initiation in lA/cm2 and j = 0.0116 is the conversion factor
corrosion rate, exposure type and concrete mix on the fre-
from lA/cm2 to mm/year and Tcr is the critical corrosion initiation
quency of corrosion pattern.
time.
(2) Based on the regression analysis of the maximum cross sec-
The time dependent corrosion current density can be estimated
tion loss, it was found that the pitting effect is function of
using the following equation:
mass loss ratio. Therefore, it is a time-variant phenomenon.
icorr ðtÞ ¼ 0:85 icorr0 ðt  T cr Þ0:3 ð20Þ (3) The statistical analysis based on null hypothesis tests
showed that the lognormal distribution model can represent
the non-uniform distribution of geometrical properties of
27ð1  w=cÞ1:64
icorr0 ¼ ð21Þ corroded bars including the pitting effect.
C (4) The correlations between the change in geometrical proper-
where icorr0 is the corrosion current density at corrosion initiation ties and the length of the bars were investigated by calcula-
time Tcr, w/c is the water cement ratio of concrete and C is the thick- tion of ACF for each corroded bar and CCF for pairs of
ness of cover concrete. Eq. (20) shows that corrosion current density corroded bars. It was found that the correlation coefficient
decreases after corrosion initiation time. This is due to the forma- is always more than 0.7 over a critical length Lc = 10 mm.
tion of corrosion products around the surface of reinforcement. This Therefore, this can be used in probabilistic modelling of cor-
reduces the diffusion of the iron ions away from the reinforcement roded bars using the developed probabilistic models in this
surface which will result in reduction of corrosion rate with time paper. Further analytical research (e.g. stochastic finite ele-
[33]. ment analysis) is required to investigate the sensitivity of
Subsequently, using the probabilistic models described in Sec- the response of corroded bars (under tension, compression
tion 5.2, the time-variant distribution models of the geometrical and cyclic load) and RC elements to the chosen value of Lc.
properties of corroded bar can be generated. Fig. 18a–f shows the (5) The probabilistic models developed in this paper are based
results of the time variant probabilistic models for the given on accelerated corrosion tests and there is a need for further
example. model calibration and comparison of these models with nat-
The results of probabilistic models shown in Fig. 18 can then be urally corroded bars. Nevertheless the results of previous
used in a structural safety and reliability analysis of the bridge pier experimental studies [46–49] showed that the response of
in this example. accelerated and naturally corroded bars with similar mass
The current standards for the assessment of corroded structures loss ratios is similar in tension tests.
such as BA 51/95 and BA 38/93 in the UK [40,41] provide simplified (6) Further research is required to investigate the influence of
approaches for estimation of the cross section loss of corroded corrosion pattern on the global response of corroded RC
reinforcement which requires destructive sampling from the struc- structures subject to earthquake loading. However, the
M.M. Kashani et al. / Corrosion Science 73 (2013) 208–221 221

probabilistic distribution models developed in this paper are [21] M.G. Stewart, A. Al-Harthy, Pitting corrosion and structural reliability of
corroding RC structures: experimental data and probabilistic analysis, Relib.
currently the only available spatial-time-variant probabilis-
Eng. Syst. Safe 93 (2008) 373–382.
tic models for distribution of pitted cross sections along [22] M.S. Darmawan, M.G. Stewart, Spatial time-dependent reliability analysis of
the corroded bars. corroding pretensioned prestressed concrete bridge girders, Struct. Saf. 29
(2007) 16–31.
[23] D.V. Val, R.E. Melchers, Reliability of deteriorating RC slab bridges, J. Struct.
Eng. 123 (12) (1997) 1638–1644.
Acknowledgements [24] ASTM G1-03 Standard Practice for preparing, cleaning, and evaluating
corrosion test specimens, ASTM Int’l 2011.
[25] MTLAB R2012b, The MathWorks lnc, 1994–2012 <http://
The experimental work is funded by Earthquake Engineering www.mathworks.com>.
Research Centre (EERC) at the University of Bristol. Any findings, [26] J.D. Cryer, K.S. Chan, Time Series Analysis with Applications in R, second ed.,
opinions and recommendations provided in this paper are only Springer, New York, 2008.
[27] P.D. Welch, The use of fast fourier transform for the estimation of power
based on the author’s view. spectra a method based on time averaging over short, modified periodograms,
IEEE Trans. Audio Electroacoust. 15 (2) (1967) 70–73.
References [28] D.E. Hawn, Extreme value prediction of maximum pits on pipelines, Mater.
Perform. 3 (29) (1977) 29–32.
[29] A.K. Sheikh, J.K. Boah, D.A. Hansen, Statistical modelling of pitting corrosion
[1] E.J. Wallbank, The performance of concrete in bridges: a survey of 200 highway
and pipeline reliability, Corrosion 46 (3) (1990) 190–197.
bridges, London, 1989.
[30] S. Sriramula, M.K. Chryssanthopoulos, Probabilistic models for spatially
[2] Transportation Research Board, Highway deicing: comparing salt and calcium
varying mechanical properties of in-service GFRP cladding panels, J. Compos.
magnesium acetate, special report 235, National Research Council,
Constr. 13 (2) (2009) 159–167.
Washington, DC, 1991.
[31] F.J. Massey, The Kolmogorov–Smirnov test for goodness of fit, J. Am. Stat.
[3] J.P. Broomfield, Corrosion of Steel in Concrete: Understanding, Investigation
Assoc. 46 (253) (1951) 68–78.
and Repair, second ed. Taylor & Francis, London, 2007.
[32] J.D. Gibbons, Nonparametric Statistical Inference, fourth ed., Basel, 2003.
[4] Y. Ou, L. Tsai, H. Chen, Cyclic performance of large-scale corroded reinforced
[33] K.A.T. Vu, M.G. Stewart, Structural reliability of concrete bridges including
concrete beams, Earthq. Eng. Struct. D 41 (2011) 592–603.
improved chloride-induced corrosion models, Struct. Saf. 22 (2000) 313–333.
[5] Y. Ma, Y. Che, J. Gong, Behavior of corrosion damaged circular reinforced
[34] M.M. Rafiq, M.K. Chryssanthopoulos, T. Onoufriou, Performance updating of
concrete columns under cyclic loading, Constr. Build. Mater. 29 (2012) 548–
concrete bridges using proactive health monitoring methods, Relib. Eng. Syst.
556.
Safe. 86 (2004) 247–256.
[6] A.A. Almusallam, Effect of degree of corrosion on the properties of reinforcing
[35] Y. Zhao, Y. Wu, W. Jin, Distribution of millscale on corroded steel bars and
steel bars, Constr. Build. Mater. 15 (2001) 361–368.
penetration of steel corrosion products in concrete, Corros. Sci. 66 (2013) 160–
[7] Y.G. Du, L.A. Clark, A.H.C. Chan, Residual capacity of corroded reinforcing bars,
168.
Mag. Concr. Res. 57 (3) (2005) 135–147.
[36] J. Ozbolt, G. Balabanic, M. Kušter, 3D Numerical modelling of steel corrosion in
[8] Y.G. Du, L.A. Clark, A.H.C. Chan, Effect of corrosion on ductility of reinforcing
concrete structures, Corros. Sci. 53 (2011) 4166–4177.
bars, Mag. Concr. Res. 57 (7) (2005) 407–419.
[37] Y. Zhao, J. Yu, B. Hu, W. Jin, Crack shape and rust distribution in corrosion-
[9] J. Cairns, G.A. Plizzari, Y.G. Du, D.W. Law, F. Chiara, Mechanical properties of
induced cracking concrete, Corros. Sci. 55 (2012) 385–393.
corrosion-damaged reinforcement, ACI Mater. J. 102 (4) (2005) 256–264.
[38] Y. Zhao, J. Yu, W. Jin, Damage analysis and cracking model of reinforced
[10] C.A. Apostolopoulos, M.P. Papadopoulos, S.G. Pantelakis, Tensile behavior of
concrete structures with rebar corrosion, Corros. Sci. 53 (2011) 3388–3397.
corroded reinforcing steel bars BSt 500s, Constr. Build. Mater. 20 (2006) 782–
[39] Y. Zhao, B. Hu, J. Yu, W. Jin, Non-uniform distribution of rust layer around steel
789.
bar in concrete, Corros. Sci. 53 (2011) 4300–4308.
[11] C.A. Apostolopoulos, Mechanical behavior of corroded reinforcing steel bars
[40] Design Manual for Roads and Bridges, The assessment of concrete structures
S500s tempcore under low cycle fatigue, Constr. Build. Mater. 21 (2007) 1447–
affected by steel corrosion, BA 51/95, UK.
1456.
[41] Design Manual for Roads and Bridges. Assessment of the fatigue life of
[12] M.M. Kashani, A.J. Crewe, N.A. Alexander, Nonlinear stress–strain behaviour of
corroded or damaged reinforcing bars, BA 38/93, UK.
corrosion-damaged reinforcing bars including inelastic buckling, Eng. Struct.
[42] D. Choe, P. Gardoni, D. Rosowsky, T. Haukaas, Probabilistic capacity models
48 (2013) 417–429.
and seismic fragility estimates for RC columns subject to corrosion, Reliab.
[13] M.M. Kashani, A.J. Crewe, N.A. Alexander, Nonlinear cyclic response of
Eng. Syst. Safe 93 (2008) 383–393.
corrosion-damaged reinforcing bar with the effect of buckling, Constr. Build.
[43] L. Berto, R. Vitaliani, A. Saetta, P. Simioni, Seismic assessment of existing RC
Mater. 41 (2013) 388–400.
structures affected by degradation phenomena, Struct. Saf. 31 (2009) 284–297.
[14] J.A. Gonzalez, C. Andrade, C. Alonso, S. Feliu, Comparison of rates of general
[44] J. Ghosh, J.E. Padgett, Aging considerations in the development of time-
corrosion and maximum pitting penetration on concrete embedded steel
dependent seismic fragility curves, J. Struct. Eng. 136 (12) (2010) 1497–1511.
reinforcement, Cement Concr. Res. 25 (2) (1995) 257–264.
[45] A. Alipour, B. Shafei, M. Shinozuka, Performance evaluation of deteriorating
[15] M.S. Darmawna, Pitting corrosion model for reinforced concrete structures in a
highway bridges located in high seismic areas, J. Bridge Eng. 6 (5) (2011) 597–
chloride environment, Mag. Concr. Res. 62 (2) (2010) 91–101.
611.
[16] M.S. Darmawna, M.G. Stewart, Effect of pitting corrosion on capacity of
[46] C.A. Apostolopoulos, V.G. Papadakis, Consequences of steel corrosion on the
prestressing wires, Mag. Concr. Res. 49 (2) (2007) 131–139.
ductility properties of reinforcement bar, Constr. Build. Mater. 22 (2008)
[17] F. Caleyo, J.C. Velázquez, A. Valor, J.M. Hallen, Probability distribution of pitting
2316–2324.
corrosion depth and rate in underground pipelines: a Monte Carlo study,
[47] R. Palssom, M.A. Mirza, Mechanical response of corroded steel reinforcement
Corros. Sci. 51 (2009) 1925–1934.
of abandoned concrete bridge, ACI Struct. J. 99 (2) (2002) 157–162.
[18] H. Kim, S. Tae, H. Lee, S. Lee, T. Noguchi, Evaluation of mechanical performance
[48] W. Zhang, X. Song, X. Gu, S. Li, Tensile and fatigue behavior of corroded rebars,
of corroded reinforcement considering the surface shape, ISIJ Int. 49 (9) (2009)
Constr. Build. Mater. 34 (2012) 409–417.
1392–1400.
[49] C.A. Apostolopoulos, S. Demis, V.G. Papadakis, Chloride-induced corrosion of
[19] M.G. Stewart, Q. Suo, Extent of spatially variable corrosion damage as an
steel reinforcement – mechanical performance and pit depth analysis, Constr.
indicator of strength and time-dependent reliability of RC beams, Eng. Struct.
Build. Mater. 38 (2013) 139–146.
31 (2009) 198–207.
[20] M.G. Steward, Mechanical behaviour of pitting corrosion of flexural and shear
reinforcement and its effect on structural reliability of corroding RC beams,
Struct. Saf. 31 (2009) 19–30.

You might also like