Sei sulla pagina 1di 11

Magazine of Concrete Research, 2010, 62, No. 2, February, 91101 doi: 10.1680/macr.2008.62.2.

91

Pitting corrosion model for reinforced concrete structures in a chloride environment


M. S. Darmawan
Institute Technology of Surabaya

Different types of corrosion may arise in steel reinforcement, such as general corrosion, pitting corrosion, hydrogen embrittlement, stress corrosion cracking and corrosion fatigue. However, corrosion of steel reinforcement in reinforced concrete (RC) structures can be classified into two categories: general (uniform) corrosion and pitting (localised) corrosion. In general, concrete researchers use uniform corrosion to study the effect of corrosion on RC structures. This approach is not accurate for concrete structures subjected to chloride attack, which usually experiences pitting corrosion. This paper describes an accelerated corrosion test used to obtain statistical parameters of maximum pit-depths distribution of corroded steel in a RC structure. Using probabilistic analysis, these statistical parameters are combined with statistical parameters of RC beams (i.e. beam dimensions, concrete strength, steel yield strength, cover thickness, workmanship quality, in situ strength factor, model error for flexure and shear and also corrosion rate) to determine the effect of corrosion on flexural and shear strength of RC beams. Using the proposed pitting corrosion model improves service life prediction of RC structures in a chloride environment.

Introduction
Reinforced concrete (RC) structures cannot escape totally from corrosion, especially when they are subjected to an aggressive environment, such as the chloride environment. For RC structures, corrosion normally leads to serviceability problems (staining, cracking and spalling of concrete). However, if this problem is not rectified by repair action, it can lead to failure. Failure in this case is defined as fail to perform their intended function. Figure 1 shows a corrosion related problem of RC structures due to salt attack from seaspray. Chloride contamination is considered to be the primary cause of corrosion of RC structures. This can occur either from the application of de-icing salts in cold regions or exposure to sea-spray in marine environments and also in some rare cases from chloride contaminated concrete admixtures. The reaction between the hardened cement pastes with carbon dioxide (carbonation) and exposure to pollutants (sulphur dioxide (SO2 ), nitrogen oxide (NO x )) is also known to
Institute Technology of Surabaya (ITS), Indonesia (MACR 800133) Paper received 2 August 2008; last revised 6 February 2009; accepted 12 March 2009

Figure 1. Corrosion of reinforced concrete structures

cause corrosion to a lesser degree, especially in urban and industrialised areas. The corrosion of RC structures can take place if the concrete quality is not adequate, the concrete cover is less than that specified in the design, or through poor detailing during design and construction. Considerable research has been conducted on concrete material performance subjected to chloride exposure in the last two decades (Andrade et al., 1993; 91

www.concrete-research.com 1751-763X (Online) 0024-9831 (Print) # 2010 Thomas Telford Ltd

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Darmawan Cady and Weyers, 1983; Gonzales et al., 1995; Liu and Weyers, 1996; Mehta, 1991; Torres-Acosta and Martinez-Madrid, 2003; Tuutti, 1982). As natural corrosion takes many years to observe, many researchers use an accelerated corrosion test to study the effect of corrosion on concrete structures (Alonso et al., 1998; Andrade et al., 1993; Darmawan and Stewart, 2006; Vu et al., 2005). This approach can simulate significant corrosion in a relatively short time. The two most common types of corrosion phenomena that affect reinforcing steel are general and pitting (localised) corrosion. General corrosion is found when there is a relatively uniform surface attack to the steel. This rarely happens in steel embedded in concrete exposed to an aggressive chloride environment, but is more likely a result of carbonation. On the other hand, pitting corrosion is the most likely corrosion phenomena to occur in RC structures in a chloride environment. This corrosion arises especially at places where the metal protective layer (passivation) has been damaged by aggressive ions (e.g. chlorides). Chlorides may come from different sources, such as salt spray in marine environments, from the application of de-icing salts and also in some cases from concrete admixtures. Corrosion can only start when the passive layer is destroyed and there is enough oxygen and water to maintain the corrosion reaction (Hausmann, 1967; Nakamura and Kanako, 1975). The passive layer may be destroyed in the presence of chloride ions (depassivation). The depassivation of the steel takes place when the level of chloride ions at the steel surface exceeds a certain threshold limit. The time required for the chloride concentration at the steel surface to reach the threshold chloride concentration needed to destroy the passive layer of the steel is defined as corrosion initiation, and this represents the first stage of deterioration (see Figure 2). The second stage is called corrosion propagation, where reinforcing steel corrodes causing loss of area (metal loss) and reduces structural capacity. Different approaches have been made to model chloride penetration in concrete; these range from analytical models based on physical mechanisms such as diffusion, absorption, electrostatic fields (Bazant, 1979; Bentz et al. 1999; Nagesh and Bhattacharjee, 1998; Robert et al., 2000), to empirical models based on experiments (Hong and Hooton, 1999; Li, 2002), and/ or field data (Bamforth, 1999; Liam et al., 1992). The
Structural capacity b

different approaches proposed clearly underline that the actual chloride penetration process is very complicated, and may involve a combination of processes contributing towards the overall ingress of chlorides (Papadakis et al., 1996). Of all the available models, it is widely assumed that the model based on diffusion theory represents the chloride ingress in concrete. This view is still widely held, despite the fact that some of the assumptions used are not in agreement with the actual service conditions, but, as an empirical approach has been shown to be relatively robust. Hence, for this study only corrosion propagation will be discussed. Corrosion propagation is mostly modelled by assuming a relatively uniform loss of material thickness (see Figure 3), such as used by Thoft-Christensen and Hansen (1993) and also Vu and Stewart (2000). However, this approach is not accurate for concrete structures subjected to chloride attack, which usually experiences pitting corrosion. Val and Melchers (1997) proposed an alternative model for corrosion of RC structures, by assuming non-uniform loss of material thickness of steel (pitting corrosion), see Figure 4 for pit-configuration. This model was developed based on accelerated corrosion tests performed by Gonzales et al. (1995) However, the specimens used in this test are relatively small (150 mm s/d 300 mm), which do not represent field condition and also the number of specimens is very limited. Hence, the statistical parameter of maximum pit-depths distribution is not known from this test. The maximum pit-depth of corroded reinforcing steel is an important parameter as it is the likely place of critical (minimum) section of corroded reinforcing steel. Darmawan and Stewart (2006) have developed a pitting corrosion model for prestressed concrete (PC) structures subjected to chloride attack. This model was developed from an accelerated corrosion test using four slabs, each of dimensions 1500 mm 3 1000 mm 3

D0

Figure 3. General corrosion

Pit-depth (a)

Corrosion initiation

Corrosion propagation

D0

Figure 2. Deterioration model

Figure 4. Pit configuration (Val and Melchers, 1997) Magazine of Concrete Research, 2010, 62, No. 2

92

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Pitting corrosion model for reinforced concrete structures in a chloride environment 250 mm with wires/strands. From this test, Darmawan and Stewart (2006) have determined that the distribution of maximum pit-depths for prestressing wire is best represented by the Gumbel (EV-type I). The predicted Gumbel distribution of maximum pitdepth (a) at any time of exposure T (years), corrosion rate icorr (A/cm2 ) and wire length L is     a a : 0:54 e 0 54 f a T , icorr , L 0:54 e e (1) where complete experimental set-up used is shown in Figure 5. The accelerated corrosion process was introduced to the rebar using an electric current, which was induced from a power supply through a current regulator. This current regulator can maintain a constant current during the test. At the completion of each corrosion test the specimen was broken up and the steel then cleaned, dried and weighed using the method as specified by Standard practice for preparing, cleaning, and evaluating corrosion test specimens (ASTM G1-03 (ASTM, 1999)). The pit-depth in the corroded steel was then measured using a micrometer gauge. However, it must be mentioned here that measuring pit-depth using a &

'2 # k 1 T D2 Do 0:0232 3 icorr (1) 1 1 o 1 # "  & '2 k 1 T D2 Do 0:0232 3 icorr (1) 1 1 o 1 o 

"

(2)

and To is given by 2 1 T o exp ln 1 o exp 6 4

0 B @

1 icorr exp 3 T o exp k 1 icorr (1) k 3 icorr (1)  ln L Lo  o exp T > 1year

13 C7 A5 (3) (4) (5)

1 o exp

icorr (T ) icorr (1) 3 kT

and where oexp and o-exp are the parameters of the Gumbel distribution as obtained from the accelerated corrosion tests, Do is the initial diameter of the wire, Lo is the length of the wire used to record the maximum pit-depths, To-exp is the period of the experiment, icorr-exp is the corrosion rate used in the experiment, icorr (1) is the initial corrosion rate and k and are empirical factors. Using field data of RC slabs exposed to the environment reported by Liu and Weyers (1996), Vu and Stewart (2000) proposed that k 0.85 and 0.29. If k 1 and 0 then icorr (T) becomes constant with time (i.e. time-invariant corrosion rate). See the Appendix for full details of the model development.

micrometer gauge was not easy as different shapes of pit may arise. Hence, at least five measurements of pitdepth for each pit have to be carried out and the maximum value of measurement was recorded as maximum pit-depth. The soffit of the specimen was immersed in a 5% sodium chloride (NaCl) solution. The reinforcing steel acts as an anode while the stainless steel plate immersed in the sodium chloride solution acts as the cathode and the pore fluid in the concrete is the electrolyte. Type GP cement (AS 3972-1997 (SAA, 1997)), also known as type I cement (ASTM C150-02 (ASTM, 2007)), was used for both test series and 3% of calcium chloride (CaCl2 ) by weight of cement was added to the concrete mix to induce corrosion. All concrete mixes had the same watercement ratio and mix-design (w/c 0.5; f c 40 MPa). All specimens were moist9 cured for 28 days before testing. The test used two concrete slabs, each of dimensions 1500 mm 3 1000 mm 3 250 mm with reinforcing steel. Nine 16 mm diameter bars of 100 mm equal spacing were placed in each slab, each with a 70 mm concrete cover (see Figure 6). The nominal corrosion rate introduced in the corrosion test is 100 A/cm2 for a period of 28 days.
Power supply and current regulator Rebar 5% sodium chloride solution

Experimental methods
As mentioned previously, Darmawan and Stewart (2006) has successfully developed pitting corrosion model for PC structures subjected to chloride attack. In this part, the pitting corrosion model for RC structures will be developed using a similar approach. Hence, the accelerated pitting corrosion test for RC structures is used to obtain statistical parameters of maximum pitdepths distribution of corroded reinforcing steel. The
Magazine of Concrete Research, 2010, 62, No. 2

Stainless steel plate

Concrete beam or slab

Figure 5. Experimental set-up

93

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Darmawan
1000 mm

predicted maximum pit-depths: mm

35 30 25 20 15

icorr-exp

150 A/cm2; To-exp

28 days; Lo

325 mm; n

72

1500 mm

100 mm

Normal Log-normal Gamma Gumbel Weibull Perfect fit

1)]
1

[i /(n
Reinforcing steel bar ( 16 mm) 250 mm 70 mm

10 05 05 10 15 20 25

CDF

30

35

Maximum pit-depths/a: mm

Figure 6. Specimen for accelerated corrosion test

Figure 7. Inverse CDF (CDF 1 ) plots for maximum pitdepths in a reinforcing bar

Results
Statistical parameters of maximum pit-depths distribution The maximum pit-depth along the 1500 mm length of rebar was measured for each 325 mm rebar length, except the 100 mm length of rebar at the two ends of the slab. This was done to avoid edge effects, since pit-depths in this area were observed to be greater than at other areas. This observation is consistent with that reported in the literature (Bond, 1972). Normal, lognormal, Gamma, Gumbel and Weibull distributions were then fitted to the measured pit depth data, see Figure 7. All these distributions passed the KolmogorovSmirnov goodness-of-fit test at the 5% significant level. However, inverse cumulative distribution function (CDF 1 ) plots of the distribution show that the distribution of maximum pit-depths in the reinforcing steel can be best represented using the Gumbel (EV-type I) distribution because this distribution gives a better fit to the upper tail of the histogram of the measured maximum pit-depths. Hence, the observation in these tests that the Gumbel distribution (EV-type I) provides the best fit to maximum pit-depth data is consistent with pitting phenomena for prestressing wires/strand (Darmawan and Stewart, 2006) and also for aluminium alloys and steel plates and pipes (Aziz, 1956; Eldredge, 1957; Finley and Toncre, 1964; Hawn, 1997; Sheikh et al., 1990; Vajo et al., 2003).

Table 1 gives the summary of the statistical parameters obtained from the test data. Note that the pitting corrosion statistical parameters oexp and o-exp are indicative only and increased confidence in predictions will be obtained if these parameters are based on tests which more closely represent actual field conditions (i.e. natural corrosion). However, this kind of test will take many years to complete. Table 1 shows that the measured corrosion rates are higher than the nominal corrosion rates. Previous experimental tests showed a similar trend (Alonso et al., 1998; Vu, 2003). Higher measured corrosion rates are generally caused by acidification owing to a hydrolysis reaction that occurs during the corrosion process and also, instead of dissolving electrolytically, there are some parts of the metal that spall out of the metal surface when the surrounding material is oxidizing (Alonso et al., 1998). The effect of length of reinforcing bar Figure 8 shows the effect of length of the reinforcing bar on the distribution of maximum pit-depths. As expected, as the length of the rebar increases the probability of finding larger pit-depths also increases. As seen in Figure 8, increasing the length from 1 m to 15 m will increase the mean maximum pit-depths from 4.4 mm to 6.4 mm. However, it should be realised that the use of extreme value statistics to extrapolate distributions of maximum pit-depths for long lengths L of

Table 1. Statistical parameters for the maximum pit-depths


To-exp : years icorr-exp : A/cm2 Length Lo : mm Mean 0.076712 150 325 1.86 a: mm COV 0.22 1.68 2.99 oexp oexp

o-exp, o-exp are Gumbel parameters; COV coefficient of variation

94

Magazine of Concrete Research, 2010, 62, No. 2

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Pitting corrosion model for reinforced concrete structures in a chloride environment


06 icorr 05 50 A/cm2 T 10 years 100 T 10 years 120 icorr 50 A/cm2

Probability density

04 03 02 01 00 00 20 40 60 80 Maximum pit-depths: mm 100 120

Probability density

80 T 60 T 40 20 00 00 02 04 Apit /Ao 06 08 10 30 years T 40 years T 50 years 20 years

L L L L

1m 5m 10 m 15 m

Figure 8. The effect of length on the distribution of maximum pit-depths for a reinforcing bar

Figure 10. Distribution of reduction of cross-sectional area of a reinforcing bar with time

reinforcing bar (e.g. 30 m) may not be totally accurate as there is likely to be some upper physical limit to maximum pit-depths. Therefore, extreme value statistics should be used cautiously. The effect of time since corrosion initiation Figure 9 illustrates the effect of time since corrosion initiation T on the distribution of maximum pit-depths of a reinforcing bar for corrosion rate (icorr ) 5.0 A/ cm2 . As the time T increases from 10 years to 50 years, the mean maximum pit depths increase from 3.5 mm to 7.2 mm. Similar to the effect of length L, the extrapolation using this model to determine the distribution of maximum pit-depths of any time T beyond the time used in the corrosion tests also needs further validation. Note that the value of To for the specimen used in the accelerated corrosion test is 3 years for corrosion rate of 5.0 A/cm2 . Figure 10 illustrates the effect of time since corrosion initiation on the reduction of cross-sectional area at time T [(Apit (T)/Ao )]. Clearly, pitting corrosion leads to a significant reduction of cross-section and accordingly reduces the strength of the reinforcing bar. Figure
06 icorr 05 50 A/cm2

10 also indicates that after 50 years since corrosion initiation the mean reduction in cross-sectional area is about 30%, however, there is 2% chance that the reduction is higher than 80%. The presence of pitting with various depths and shapes may raise questions regarding mode of failure of corroded reinforcing steel and the possibility of stress concentration associated with pit geometry. However, Darmawan and Stewart (2006) have shown that for corroded prestressing steel ( fy 1700 MPa), the mode of failure is yielding and the use of the linearelastic fracture analysis (LEFA) to determine the ultimate load of corroded prestressing steel is not accurate. Therefore, for much lower strength steel such as reinforcing steel ( fy 460 MPa), the possibility of stress concentration associated with pit geometry is very unlikely. The effect of corrosion rate Figures 11 and 12 demonstrate the effect of different corrosion rates on the distribution of maximum pit10 icorr 08 05 A/cm2 icorr 06 icorr 04 30 A/cm2 10 A/cm2 T 25 years

Probability density

04 03 02 01 00 00 20 40 60 80 100 120

T T T T T

10 years 20 years 30 years 40 years 50 years

Probability density

02

icorr

50 A/cm2

00 00 140 160 20 40 60 80 Maximum pit-depths: mm 100 120

Maximum pit-depths: mm

Figure 9. Distribution of maximum pit-depths with time Magazine of Concrete Research, 2010, 62, No. 2

Figure 11. The effect of corrosion rates on distribution of maximum pit-depths of rebar for time-invariant corrosion rates

95

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Darmawan
35 icorr 30 25 20 icorr 15 10 05 00 00 20 40 60 80 Maximum pit-depths: mm 100 120 icorr 50 A/cm2 30 A/cm2 icorr 10 A/cm2 05 A/cm2 T 25 years

Figure 12. The effect of corrosion rates on distribution of maximum pit-depths of rebar for time-variant corrosion rates

depths for a reinforcing bar for time-invariant and time-variant corrosion rates, respectively. As can be seen, the two corrosion rate scenarios yield significantly different results. Since the time-variant corrosion rate assumes that corrosion rate decreases with time, the maximum pit-depths are expected to be lower as shown in Figure 11. For time-invariant corrosion rate, changes the corrosion rate from 0.5 A/cm2 to 5.0 A/ cm2 increases the mean maximum pit-depths from 2.0 mm to 6.0 mm, whereas for time-variant corrosion rate the mean maximum pit-depth changes from 0.8 mm to 2.5 mm.

Carlo simulation will be used to determine the distribution of flexural and shear strength of concrete beams. The statistical parameters of RC beams used in the probabilistic analysis are given in Table 2. Table 2 shows that the parameters influencing flexural and shear strength of RC beams have some uncertainty (random variables). The coefficients of variation (COVs) of these parameters range from 0.046 to 0.3. For pitting model parameter, the COV is also on the high side (e.g. 0.22), see Table 1. Therefore, flexural and shear strength determination of RC beams based on deterministic analysis is inaccurate. Note that all of these parameters were obtained in developed countries (e.g. USA). For developing countries, such as Indonesia, the COV of these parameters is expected to be higher than is shown in Table 2. A previous study by Darmawan and Stewart (2007) showed that considering spatial variation of pit-depth along the prestressing steel has increased the probability of strength failure of a prestressed concrete girder by 10%. As RC structures use much larger diameter than prestressing steel (e.g. 4 to 5 mm), the effect of considering spatial variation of pit-depth along the bars for a RC beam is expected to be lower than 10%. Hence, for this paper only critical section analysis (non-spatial analysis) is considered. Illustrative example For illustrative purposes, the corrosion model developed will be used to determine the effect that corrosion has on the RC beam shown in Figure 13. The beam has 30 3 60 cm2 dimensions, with 3D16 as tension reinforcement and D13 space at 150 mm as shear reinforcement. The design concrete strength f c is 20 MPa and 9 the yield strength of the reinforcement fy is 465 MPa. The beam was designed to carry flexural moment owing to service dead load of 7.43 t m and required flexural moment (Mu ) of 13.0 t m. The beam was also designed to carry shear force owing to service dead load of 30.0 t and required shear force (Vu ) of 42 t. Figure 14 shows that 25 years since corrosion initiation, there is a 90% probability that flexural moment of

Probabilistic analysis
Using probabilistic analysis, the statistical parameters of maximum pit-depths distribution are combined with statistical parameters of reinforced concrete beams (i.e. beam dimension, concrete strength, steel yield strength, cover thickness, workmanship quality, in situ strength factor, model error for flexure and shear and also corrosion rate) to determine the effect of corrosion on flexural and shear strength of reinforced concrete beam. Monte
Table 2. Statistical parameters used in this study
Parameters F cyl 9 fy kw ki H (beam depth) B (beam width) Cover thickness Model error (flexure) Model error (shear) icorr Pitting model parameter Mean f c + 7.5 MPa 9 465 MPa 0.53 (1.2 0.00816 3 kw f cyl ) 9 Hnom 3.2 mm Bnom + 2.5 mm Cbnom + 1.6 mm 1.01 1.15 5.0 A/cm2 o-exp 1.68

Probability density

COV s6 0.10 0.078 0.10 s 6.4 s 3.7 s 11.1 0.046 0.125 0.3 o-exp 2.99

Distribution Lognormal Beta Normal Normal Normal Normal Normal Normal Normal Normal Gumbel

Reference Attard and Stewart (1998) Mirza and MacGregor (1979a) Stewart (1995) Stewart (1995) Mirza and MacGregor (1979b) Mirza and MacGregor (1979b) Mirza and MacGregor (1979b) Ellingwood et al. (1980) Ellingwood et al. (1980) Stewart and Rosowsky (1998) Table 1

Note: kw degree of workmanship; ki in situ strength factor; s standard deviation

96

Magazine of Concrete Research, 2010, 62, No. 2

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Pitting corrosion model for reinforced concrete structures in a chloride environment


V service 12 D13-150 10 10 10 10 10 10 10
1

Vu icorr 50 A/cm2

Probability density

80 60 40 20

No corrosion
2

T T

25 years

600

50 years

3D16

00

100 20 30 40 Vn: t 50 60 70

Figure 15. The effect of corrosion on shear strength of a RC beam 30 3 60 cm2


35 300 10 10 10 10 10 10 10
1

icorr-exp
1

50 A/cm2

Figure 13. RC beam 30 3 60 cm2


Probability density
35 30 10 10 10 10 10 10 10
1 1 1 1

30 25 20 15 10 50 00

T T

50 years (pitting) 50 years (general)

M service T 25 years

Mu icorr 50 A/cm2

25 years (general) 25 years (pitting)

Probability density

25 20 15 10 50 00

T
1

50 years

No corrosion

100 00 50 100 Mn: t m 150 200

1 2

100 00 50 100 150 Mn: t m 200 250

Figure 16. Comparison between pitting corrosion model and general corrosion model.

Figure 14. The effect of corrosion on flexural strength of a RC beam 30 3 60 cm2

the beam is under the required moment. After 50 years the probability of exceeding the required moment increases to 99%. For the same period the probability of exceeding the service moment is close to 30%. For no corrosion, there is a 5% chance that the required moment (Mu ) will be exceeded. Figure 15 demonstrates the effect that corrosion has on shear strength of the RC beam shown in Figure 13. The figure shows that 25 years since corrosion initiation there is a 10% probability that shear strength of the beam is under the required shear strength. After 50 years the probability of exceeding the required shear strength increases to 65%. For service load condition, the probability that the shear strength of the beam will be exceeded is negligible. Figures 14 and 15 clearly indicate that corrosion has a significant effect on flexural and shear strength of a RC beam. Finally, the pitting corrosion model is compared with the general corrosion model to determine the effect of corrosion on flexural strength of the RC beam. The comparison between the two corrosion models is shown
Magazine of Concrete Research, 2010, 62, No. 2

in Figure 16. Figure 16 shows that the pitting corrosion model produces less reduction of flexural strength than the general corrosion model. For example after 50 years since corrosion initiation, the pitting corrosion model gives mean flexural strength of 8.5 t m, whereas the general corrosion model gives mean flexural strength of 6.0 t m. This result contradicts the result obtained by Stewart (2004), which indicates that pitting has a more detrimental effect on the flexural strength of the RC beam. However, the data used by Stewart (2004) are very limited in number and also the specimens used to obtain the pitting data are small in size. Hence, the results obtained from this study are more accurate as it derives from larger specimens and the number of pitting data is statistically acceptable (72 pit-depth data).

Sensitivity analysis
A sensitivity analysis was conducted to study the relative importance of the variability of each random variable on the calculated response (e.g. flexural and shear strength of reinforced concrete beam). Each parameter is first treated as deterministic variables (equal to its mean). The relative sensitivity of each parameters is 97

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Darmawan determined by considering the effect of (1 standard deviation) change from these mean values on the final results, while all other parameters are treated as random variables. The sensitivity analysis identified yield strength, pitting model parameters, model error for flexure, corrosion rate and cover thickness as the most important parameters influencing the flexural strength of a RC beam, see Table 3. For shear strength, the most important parameters are model error for shear, pitting model parameters, corrosion rate, cover thickness and yield strength. Following these results, it is expected that structural reliabilities are also most influenced by these important parameters. Hence, in order to obtain more accurate predictions of structural reliability, more effort should be directed at improving the accuracy of these parameters, including the pitting model statistical parameters. Note that for this study, the corrosion initiation stage is not considered. sion rate (icorr-exp ) used to develop the distribution of maximum pit-depths at time To-exp, see Figure 17. Therefore, to reach the same degree of corrosion (i.e. loss of steel) a real structure requires a longer time of exposure (To ). It should be noted that measured values of corrosion rates icorr are average values for the overall steel surface: they relate to general (uniform) corrosion. For reinforcing steel this relationship is 1 A/cm2 % 11.6 m/y. Hence, using the general corrosion case with a time-invariant corrosion rate (constant icorr with time) To can be expressed as  To  icorr-exp To exp icorr (6)

Conclusion
This paper has described the accelerated corrosion test designed to obtain a statistical parameter of pitdepths distribution. From the statistical analysis of pitting data it is concluded that the probability distribution of maximum pit-depths at the completion of an accelerated corrosion test of length of test To-exp for a given corrosion rate icorr-exp and length of wire Lo can be described using the Gumbel distribution (EV-type I). These statistical parameters of maximum pit-depths distribution are combined with statistical parameters of RC beams to determine the effect of corrosion on flexural and shear strength of RC beam. From the analysis it can be concluded that pitting corrosion has a less detrimental effect on flexural and shear strength of a RC beam than general corrosion.

where icorr is the measured corrosion rate in the actual structure. If general corrosion is considered, then the decrease of the bar diameter D, the loss of cross-sectional area A and the volume of corrosion product V (see Figure 18) for a given corrosion rate icorr (A/cm2 ) and period of exposure To (years) from corrosion initiation can be determined as (e.g. Thoft-Christensen and Hansen, 1993) D(To ) 0:0232icorr To
D

(7)

icorr-exp

Time

Appendix
In real structures subject to corrosion, the corrosion rate (icorr ) is much smaller than the accelerated corroTable 3. Variation of flexural and shear strength by varying each parameter by
Parameters Mn (flexural strength) 1.07% 9.31% 0.40% 0.57% 0.67% 0.08% 2.58% 4.65% 3.54% 5.04% Vn (shear strength) 0.59% 0.87% 0.22% 0.78% 0.27% 0.46% 1.85% 14.25% 2.13% 2.31%

To-exp

To

Figure 17. Loss of diameter (D) against. time

D(To)/2

f cyl 9 fy kw ki H (beam depth) B (beam width) Cover thickness Model error icorr Pitting model parameter

D(To)/2 Do

Figure 18. General corrosion model Magazine of Concrete Research, 2010, 62, No. 2

98

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Pitting corrosion model for reinforced concrete structures in a chloride environment i h 2 Do Do 0:0232icorr To 2 4 i h 2 Do Do 0:0232icorr To 2 L V (To ) 4 A(To ) (8) (9) f a T , icorr , L f a (To-exp , icorr-exp , L). The distribution of maximum pitdepths at time T is therefore
0:54
: 0a54 a : e 0 54 ee

(18)

where Do is the initial diameter (mm) and L is the length of the steel bar (mm). Similarly, the decrease of the bar diameter D, the loss of cross-sectional area A and the volume of corrosion product V for period of exposure T can be expressed as D(T ) 0:0232icorr T (10) h i 2 D Do 0:0232icorr T 2 A(T) (11) 4 o i h 2 Do Do 0:0232icorr T 2 L V (T) (12) 4 The volume of corrosion products V should be equal for both general and pitting corrosion for the same corrosion rate and period of exposure. Then the existing pits at time To will change in volume by the ratio , hence Volume of pit at T 3 Volume of pit at To (13) where is the ratio of volume of corrosion products and is derived as V (T ) V T o h i (=4) Do 2 Do 0:0232icorr T 2 L h i (=4) Do 2 Do 0:0232icorr To 2 L T Do 0:0116icorr T T o Do 0:0116icorr To

where the parameters To-exp, icorr-exp and Lo , are taken from the accelerated corrosion tests. Note that Equation 18 is now independent of wire diameter. It should be noted that for T , To, the same assumptions as used in Figure 4 still apply (e.g. the shape or geometry of pit does not change with time), see Figure 19. However, for T , To, , 1.0. Many reliability analyses have assumed that the corrosion rate is constant with time (e.g Stewart and Rosowsky, 1998). However, it is expected that the formation of rust products on the steel surface will reduce the diffusion of the iron ions away from the steel surface. Also, the ratio between the anode and cathode area reduces with time. This suggests that the corrosion rate will reduce with time; rapidly during the first few years after initiation but then more slowly as it approaches a nearly uniform level (e.g. Vu and Stewart, 2000). Hence, corrosion rate can be expressed as a time-dependent variable as icorr (T ) icorr (1) 3 kT T > 1 year (19)

(14)

Since during the corrosion process the pit is assumed to change in depth only (i.e. length of the pit is held constant) and the geometry of the pit cross-section is unchanged with time, then Apit (T ) 3 Apit (To ) (15)

where icorr (1) is the corrosion rate at T 1 year, and k and are empirical factors. Using field data of RC slabs exposed to the environment reported by Liu and Weyers (1996), Vu and Stewart (2000) proposed that k 0.85 and 0.29. If k 1 and 0 then icorr (T) becomes constant with time (i.e. time-invariant corrosion rate). Accordingly for a timevariant corrosion rate, Equation 7 can be modified as " # D(T ) 0:0232
1 T

icorr (1)dT

icorr (T )dT
1

(20)

The cross-sectional area of a pit Apit can be calculated using a formula given by Val and Melchers (1997). However, for a wire, the relationship between pit depth (a) and cross-sectional area of pit (Apit ) can be expressed in simplified form as a k p Apit 0 54
:

Solving the integration in Equation 20 and substituting Equation 19 gives & ' k 1 T 1 D(T ) 0:0232 3 icorr (1) 1 1 (21) Following this, the expression for in Equation 14 is now given by
Length of pit a(T) a(To)

(16)

where kp is a factor equal to 1.22 and 1.35 for 5.03 mm and 4.3 mm diameter wires, respectively. It follows that a T 0 54 a To
:

(17)

Accordingly, the distribution of maximum pit-depth at time T for prestressing wire with length L and corrosion rate icorr is the distribution f a (To , icorr , L) trans: formed by the factor 0 54 . Note that f a (To , icorr , L)
Magazine of Concrete Research, 2010, 62, No. 2

Figure 19. Pit-growth model for T , To

99

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Darmawan ' 2 # k 1 T Do 0:0232 3 icorr (1) 1 1 1 # "  & '2 k 1 2 T 1 Do Do 0:0232 3 icorr (1) 1 1 o D2 o  & and To is given by 2 6 4 0 13 "

(22)

B 1 @ To exp ln 1

1 icorr-ex p 3 To exp C7 k 1 icorr (1) A5 k 3 icorr (1) (23)

The distribution of maximum pit-depth at time T for steel wires with length L and corrosion rate icorr (1) is then given by Equation 18 where is given by Equation 22.

References
Alonso C, Andrade C, Rodriguez J and Diez JM (1998) Factors controlling cracking of concrete affected by reinforcement corrosion. Materials and Structures 31(211): 435441. American Society for Testing and Materials (ASTM) (2007) ASTM C150-07. Standard Specification for Portland Cement. ASTM, West Conshohocken, PA. American Society for Testing and Materials (ASTM) (1999) ASTM G1-03. Standard Practice for Preparing, Cleaning, and Evaluating Corrosion Test Specimens. ASTM, West Conshohocken, PA, USA. Andrade C, Alonso C and Molina FJ (1993) Cover cracking as a function of rebar corrosion. Part 1: experimental test. Material and Structures 26(162): 453464. Attard MM and Stewart MGA (1998) Two parameter stress block for model for high strength concrete. ACI Structural Journal, ACI 95(3): 305317. Aziz PM (1956) Application of the statistical theory of extreme values to the analysis of maximum pit depth data for aluminium. Corrosion-NACE 12(10): 3546. Bamforth PB (1999) The derivation of input data for modelling chloride ingress from eight-year UK coastal exposure trials. Magazine of Concrete Research 51(2): 8996. Bazant ZP (1979) Physical model for steel corrosion in sea structures-theory. Journal of the Structural Division, ASCE ST6: 1137 1154. Bentz BE, Thomas MDA and Hooton RD (1999) An overview and sensitivity study of a multimechanistic chloride transport model. Cement and Concrete Research 29: 827837. Bond AP (1972) ASTM STP 516: Pitting Corrosion: A Review of Recent Advances in Testing Methods and Interpretation, Localized Corrosion: Cause of Metal Failure. ASTM, Philadelphia, PA, pp. 250261. Cady PD and Weyers RE (1983) Chloride penetration and deterioration of concrete bridge decks. Cement, Concrete and Aggregates 18: 4759. Darmawan MS and Stewart MG (2006) Effect of spatially variable pitting corrosion on structural reliability of prestressed concrete bridge girders. Australian Journal of Structural Engineering 6(2): 147158. Darmawan MS and Stewart MG (2007) Spatial time-dependent reliability analysis of corroding pretensioned prestressed concrete bridge girders. Structural Safety 29(1): 1631. Eldredge GG (1957) Analysis of pitting by extreme value statistics and its application to oil well tubing calliper surveys. CorrosionNACE. 123: 6776. Ellingwood B, Galambos TV Macgregor JG and Cornell CA (1980) ,

Development of a probability based load criterion for American National Standard A58. National Bureau of Standards Special Publication 577. US Government Printing Office, Washington DC. Finley HF and Toncre AC (1964) Extreme value statistical analysis in correlation of first leak on submerged pipeline. Materials Protection, 3(9): 2934. Gonzales JA, Andrade C, Alonso C and Feliu S (1995) Comparison of rates of general corrosion and maximum pitting penetration on concrete embedded steel reinforcement. Cement and Concrete Research 25(2): 257264. Hausmann DA (1967) Steel corrosion in concrete. Material Protection 6(11): 1923. Hawn DE (1997) Extreme value prediction of maximum pits on pipelines. Materials Performance, 2932. Hong K and Hooton RD (1999) Effects of cyclic chloride exposure on penetration of concrete cover. Cement and Concrete Research 29(9): 13791386. Li CQ (2002) Initiation of chloride-induced reinforcement corrosion in concrete structural members-prediction. ACI Structural Journal 99(2): 133141. Liam KC, Roy SK and Northwood DO (1992) Chloride ingress measurement and corrosion potential mapping study of a 24-year-old reinforced concrete jetty structures in a tropical marine environment. Magazine of Concrete Research 44(160): 205215. Liu T and Weyers RE (1996) Time to cracking for chloride-induced corrosion in reinforced concrete. Corrosion of Reinforcement in Concrete Construction (Page CL, Bamforth PB and Figgs JW (eds)). The Royal Society of Chemistry, London, pp. 88104. Mehta PK (1991) Concrete in the Marine Environment. Elsevier Applied Science, Barking. Mirza SA and MacGregor JG (1979a) Variability of mechanical properties of reinforcing bars. Journal of the Structural Division. ASCE 105(ST5): 921937. Mirza SA and MacGregor JG (1979b) Variations in dimensions of reinforced concrete members. Journal of the Structural Division, ASCE 105(ST4): 751766. Nagesh M and Bhattacharjee B (1998) Modelling of chloride diffusion in concrete and determination of diffusion coefficients. ACI Materials Journal 95(2): 113120. Nakamura S and Kanako T (1975) Studies of initiation of corrosion in concrete containing salt. Corrosion 13(1): 22. Papadakis VG, Roumeliotis AP, Fardis MN and Vagenas CG (1996) Mathematical modelling of chloride effect on concrete durability and protection measures. Concrete Repair, Rehabilitation and Protection (Dhir RK and Jones MR (eds)). E & FN Spon, London, pp. 165174. Robert MB, Atkins C, Hogg V and Middleton C (2000) A proposed empirical corrosion model for reinforced concrete. Proceedings of the Institution of Structural Engineers, Structures and Buildings 140(1): 111. Sheikh AK, Boah JK and Hanse DA (1990) Statistical modelling of pitting corrosion and pipeline reliability. Corrosion-NACE. 46(3) 190197. Standards Association of Australia (SAA) (1997) AS 3972. Portland and Blended Cements. Standards Association of Australia. Homebush, New South Wales, Australia. Stewart MG and Rosowsky DV (1998) Structural safety and serviceability of concrete bridges subject to corrosion. Journal of Infrastructure System, ASCE 4(4): 146155. Stewart MG (2004) Effect of spatial variability of pitting corrosion and its influence on structural fragility and reliability of RC beams in flexure. Structural Safety 26(4): 453470.

100

Magazine of Concrete Research, 2010, 62, No. 2

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Pitting corrosion model for reinforced concrete structures in a chloride environment


Stewart MG (1995) Workmanship and its influence on probabilistic models of concrete compressive strength. ACI Materials Journal 92(4): 361372. Thoft-Christensen P and Hansen HI (1993) Optimal strategy for maintenance of concrete bridges using expert system. Proceedings. ICOSSAR 93, Rotterdam, 939946. Torres-Acosta AA and Martinez-Madrid M (2003) Residual life of corroding RC structures. Journal of Materials in Civil Engineering, ASCE 15(4): 344353. Tuutti K (1982) Corrosion of Steel in Concrete. Swedish Cement and Concrete Research Institute, Stockholm. Vajo JJ, Wei R, Phelps AC, Reiner L, Herrera GA, Cervantes O, Gidanian D, Bavarian B and Kappes CM (2003) Application of extreme value analysis to crevice corrosion. Corrosion Science 45(3): 497509. Val DV and Melchers RE (1997) Reliability of deteriorating reinforced concrete slab bridges. Journal of Structural Engineering 123(12): 16381644. Vu KAT (2003) Corrosion-induced Cracking and Spatial Time-Dependent Reliability Analysis of RC Structures. PhD thesis, The University of Newcastle, Newcastle, Australia. Vu KAT and Stewart MG (2000) Structural reliability of concrete bridges including improved chloride-induced corrosion models. Structural Safety 22(4): 313333. Vu KAT, Stewart MG and Mullard JA (2005) Corrosion induced cracking: experimental data and predictive models. ACI Structural Journal 102(5): 719734.

Discussion contributions on this paper should reach the editor by 1 August 2010

Magazine of Concrete Research, 2010, 62, No. 2

101

Delivered by ICEVirtualLibrary.com to: IP: 134.148.29.34 Sat, 01 May 2010 02:52:02

Potrebbero piacerti anche