Sie sind auf Seite 1von 56

http://www.nrc-cnrc.gc.

ca/irc

Laboratory Study of Corrosion Performance of Reinforced Steels for use in Concrete Structures

IRC-RR-284
Zhang, J.Y.; Qian, S.Y.; Baldock, B.

September 2009

The material in this document is covered by the provisions of the Copyright Act, by Canadian laws, policies, regulations and international agreements. Such provisions serve to identify the information source and, in specific instances, to prohibit reproduction of materials without written permission. For more information visit http://laws.justice.gc.ca/en/showtdm/cs/C-42 Les renseignements dans ce document sont protgs par la Loi sur le droit d'auteur, par les lois, les politiques et les rglements du Canada et des accords internationaux. Ces dispositions permettent d'identifier la source de l'information et, dans certains cas, d'interdire la copie de documents sans permission crite. Pour obtenir de plus amples renseignements : http://lois.justice.gc.ca/fr/showtdm/cs/C-42

LABORATORY STUDY OF CORROSION PERFORMANCE OF DIFFERENT REINFORCING STEELS FOR USE IN CONCRETE STRUCTURES

National Research Council of Canada Institute for Research in Construction

Research Report IRC-RR-284

Authors:

Jieying Zhang Research Officer Shiyuan Qian Research Officer Bruce Baldock Technical Officer

September 2009

Disclaimer "Her Majesty the Queen in right of Canada, as represented by the Minister of Industry ("Canada"), does not warrant or guarantee the completeness of the information and data in this report and does not assume any responsibility or liability with respect to any damage or loss arising from the use or interpretation of information and data. The information and data in this report is intended to convey experimental results and should be used only as a guide.

Copyright 2009, Her Majesty in right of Canada. All rights reserved. Reproduction by any means is not permitted without the written approval of the National Research Council of Canada.

ii

TABLE OF CONTENTS
ABSTRACT 1. 2. . ..1

Introduction ................................................................................................................. 2 Experimental Program ................................................................................................ 3 2.1 Testing in Concrete Prisms ....................................................................................... 4 2.1.1 Preparation of concrete prisms........................................................................... 4 2.1.2 Sampling of concrete prisms and steel bars ....................................................... 4 2.1.3 Measurement set-up ........................................................................................... 5 2.2 Testing in Electrochemical Cell ................................................................................ 6 2.2.1 Electrodes and solution ...................................................................................... 6 2.2.2 Electrochemical techniques ............................................................................... 7 2.2.3 Pitting resistance measurement .......................................................................... 7 2.3 Additional Testing of Material Properties ................................................................ 8

3.

Corrosion Performance of Sandblasted Steels in Concrete Prisms ............................ 9 3.1 Effect of Chloride Concentration .............................................................................. 9 3.1.1 Corrosion performance of carbon steel .............................................................. 9 3.1.2 Corrosion performance of Cr steel ................................................................... 12 3.1.3 Corrosion performance of stainless steels........................................................ 14 3.2 Effect of Steel Type ................................................................................................ 15

4.

Corrosion Performance in Electrochemical Cells ..................................................... 16 4.1 Testing in Saturated Ca(OH)2 solution of pH=12.6 ................................................ 16 4.2 Pitting Corrosion Resistance ................................................................................... 18

5.

Effect of Surface Condition ...................................................................................... 20 5.1 Literature Review on Effect of Mill Scales ............................................................ 20 5.2 Experimental Observations ..................................................................................... 21

6. 7.

Summary and Future Research Needs ...................................................................... 24 References ................................................................................................................. 26

iii

LIST OF TABLES
Table 1. Concrete mix batches and properties .................................................................. 28 Table 2. Sampling of concrete prisms for corrosion resistance measurement .................. 28 Table 3. Chemical compositions of steel .......................................................................... 29 Table 4. ASTM C876-91 criterion for corrosion probability of steel in concrete ............ 29 Table 5. Typical corrosion condition of steel in concrete ................................................. 29 Table 6. Calculated PREN and measured chloride concentration for pitting corrosion ... 30

iv

LIST OF FIGURES
Figure 1. Schematic configuration of concrete prism with embedded steel bars ............ 30 Figure 2. Temperature cycle controlled in the environmental chamber ........................... 30 Figure 3. SEM images of microstructures of different types of steels ............................. 31 Figure 4. Corrosion potential of carbon steel in concrete prism ....................................... 32 Figure 5. Corrosion rate of carbon steel in concrete prism ............................................... 32 Figure 6. Carbon steel in concrete at various levels of corrosion-induced damage ......... 33 Figure 7. Corrosion potential of Cr steel in concrete prism .............................................. 34 Figure 8. Corrosion rate of Cr steel in concrete prism ...................................................... 34 Figure 9. Performance of Cr steel in concrete prism with 1.5% of chlorides ................... 35 Figure 10. Photos of concrete prisms with Cr steel at 3.0% of chlorides ......................... 36 Figure 11. Corrosion potential of SS 2205 in concrete prism........................................... 37 Figure 12. Corrosion rate of SS 2205 in concrete prism................................................... 37 Figure 13. Corrosion potential of SS 316LN in concrete prism ....................................... 38 Figure 14. Corrosion rate of SS 316LN in concrete prism ............................................... 38 Figure 15. Corrosion performance of CS and Cr steel in chloride-free concrete ............. 39 Figure 16. Corrosion potential of steels in concrete prisms with 1.5% of chlorides ........ 39 Figure 17. Corrosion rate of steels in concrete prisms with 1.5 % of chlorides ............... 40 Figure 18. Corrosion potential of steels in concrete prisms with 3.0 % of chlorides ....... 40 Figure 19. Corrosion rate of steels in concrete prisms with 3.0 % of chlorides ............... 41 Figure 20. Corrosion potential of steels in concrete prisms with 4.5 % of chlorides ...... 42 Figure 21. Corrosion rate of steels in concrete prisms with 4.5% of chlorides ................ 42 Figure 22. Corrosion potential of steels in concrete prisms with 6.0 % of chlorides ....... 43 Figure 23. Corrosion rate of steels in concrete prisms with 6.0 % of chlorides .............. 43 Figure 24. Corrosion potential and rate of Cr steel in saturated Ca(OH)2 solution ........ 44 Figure 25. Corrosion potential and rate of carbon steel in saturated Ca(OH)2 solution. . 44 Figure 26. Corrosion rate of steel in saturated Ca(OH)2 solution .................................... 45 Figure 27. Potential dynamic scans of Cr steel in saturated Ca(OH)2 solution.. 45 Figure 28. Potential dynamic scans of carbon steel in saturated Ca(OH)2 solution......... 46 Figure 29. Potential dynamic scans of SS 2205 in saturated Ca(OH)2 solution. ............. 46 Figure 30. Corrosion performance of sandblasted and non-sandblasted carbon steel at 1.5% of chlorides ...................................................................................................... 47 Figure 31. Corrosion performance of sandblasted and non-sandblasted carbon steel at 3.0% of chlorides ...................................................................................................... 47 Figure 32. Corrosion performance of sandblasted and non-sandblasted Cr steel at 1.5% of chlorides .................................................................................................................... 48 Figure 33. Corrosion performance of sandblasted and non-sandblasted Cr steel at 3.0% of chlorides .................................................................................................................... 48

Figure 34. Corrosion performance of sandblasted and non-sandblasted Cr steel at 4.5% of chlorides .................................................................................................................... 49 Figure 35. Corrosion performance of sandblasted and non-sandblasted Cr steel at 6.0% of chlorides .................................................................................................................... 49

vi

ABSTRACT
This report presents the experimental program and the results of a two-year investigation of the corrosion performance of carbon steel (ASTM A 615), a low-carbon chromium steel (ASTM A 1035/A 1035M-07) produced by MMFX Technologies Corp. and referred to as Cr steel in this report, stainless steels 316LN, 304LN and 2205 for use in reinforced concrete structures, with the emphasis on the performance of Cr steel. The investigation was carried out in both concrete prisms and electrochemical cells and performed by different electrochemical techniques including half-cell potential, linear polarization resistance, potential dynamic, and AC impedance measurement. Concrete prisms containing five different concentrations of chloride ions were stored in an environmental chamber with 80% of relative humidity and temperature cycled between 25oC and 45oC to accelerate the corrosion process. In concrete prisms containing 1.5% of chlorides (by weight of cement), active corrosion started on sandblasted Cr steel after about ten months, while active corrosion started almost immediately on carbon steel. The corrosion rate of sandblasted Cr steel was found to be lower than that of carbon steel for the same amount of chlorides in concrete prisms, and did not increase significantly when the chloride concentration was increased from 3.0% to 6.0%. In the saturated calcium hydroxide solution with pH=12.6, the chloride threshold of Cr steel was about 5.0% (by weight) and that of carbon steel was about 1.0%. Pitting corrosion was observed to occur on carbon steel at 0.06% concentration of chlorides, and occurred at 0.18%0.3% of chlorides on Cr steel. It was found, however, that the corrosion resistance of non-sandblasted Cr steel (with mill scales, as-received) embedded in concrete prisms was lower than sandblasted Cr steel. The time to initiation of active corrosion at 1.5% of chlorides was five months shorter on non-sandblasted Cr steel compared to sandblasted Cr steel. The corrosion rate of non-sandblasted Cr steel was higher than that of sandblasted Cr steel, by about half an order of magnitude at 1.5% and 3.0% of chlorides and by about one order of magnitude at 4.5% and 6.0% of chlorides. It is important to investigate further the detrimental effect of mill scales on the corrosion performance of Cr steel in order to propose an appropriate protection method, and thus to utilize the full potential of Cr steel. It was found that the three types of stainless steel had the highest corrosion resistance regarding all the above mentioned performance criteria in both concrete prisms and electrochemical cells, followed by Cr steel, and then carbon steel.

The corrosion of steel in reinforced concrete (RC) structures built in aggressive environments is one of the main causes of deterioration that costs billions of dollars in maintenance and rehabilitation in North America. The ingress of chloride ions (chlorides) onto the surface of the reinforcement is the major cause of corrosion in RC structures. The corrosion resistance of steel is often characterized by how much chlorides it can resist before corrosion is initiated, which is referred to as the chloride threshold value of the steel (Tuutti 1982). After corrosion is initiated, the corrosion rate in the propagation stage determines how fast the expansive corrosion products accumulate, which in turn determines the progress of corrosion-induced damage such as cracking of the concrete cover (Andrade and Alonso 1996). Therefore, both the chloride threshold value in the initiation stage and corrosion rate in the propagation stage should be considered to be the two main indicators of the corrosion resistance of reinforcing steel that govern the service life of RC structures. The use of corrosion-resistant reinforcing steels is one of the approaches that have been adopted to improve the durability of concrete structures for more than three decades (Clemena and Virmani 2003). Stainless steel (SS) refers to iron based alloys with a minimum of 12% chromium (Cr) and 4.5-26% nickel (Ni) by weight (Jones 1992). There are various types of stainless steel with different microstructures determined by their chemical compositions and the manufacturing steps involved. Low carbon, chromium steel (ASTM A 1035/A 1035M-07) is a new type of steel that offers a higher corrosion resistance than carbon steel but at a lower cost than stainless steel, as it has a lower content of Cr than stainless steel. This steel is referred to as Cr steel in this report. The service life of concrete structures in corrosive environments is generally defined as the sum of a corrosion initiation stage and a corrosion propagation stage. The corrosion initiation stage corresponds to the transport process of chlorides and/or carbon dioxide onto a steel surface, while the steel remains passivated with a corrosion rate lower than a defined level, e.g. 0.1 A/cm2 (Gonzlez et al. 1980). The onset of corrosion of the reinforcing steel is assumed to start when the concentration of chlorides at the level of the reinforcement has reached a chloride threshold value. Once the corrosion of the reinforcing steel has started, the remaining service life of concrete structures depends largely on the corrosion rate, because it determines the rate at which the corrosion products are being formed, and subsequently, their volume and corresponding tensile stresses generated in the concrete cover and its failure.

The quantitative values of corrosion resistance of different types of steel however, are quite variable and the reported data in the literature are highly scattered. For instance, the chloride threshold of conventional carbon steel was reported with varying values that could differ by more than one order of magnitude (Glass and Bluenfeld 1997); the chloride threshold of SS 316 was reported to be 1.67 to 50 times higher than that of carbon steel (Bertonini et al. 1996; Trejo and Pillai 2004). The results of most studies so far on the corrosion performance of MMFX microcomposite reinforcing steel (identified as low carbon, chromium steel, ASTM A 1035/1035M-07) have been documented in a report by AMEC (2006). Darwin et al. (2002) tested bare and mortar wrapped MMFX microcomposite reinforcing steel by rapid macrocell test. It was found that the steel exhibits a macrocell corrosion rate between one-third and two-thirds that of conventional reinforcing steel. Scully and Hurley (2007) studied and concluded that MMFX microcomposite reinforcing steel after surface treatment by pickling has significantly higher chloride threshold than carbon steel, but MMFX microcomposite reinforcing steel as received without surface treatment was not better than carbon steel in their laboratory chloride threshold initiation tests. Hartt et al. (2009) found in their investigation that the chloride threshold for corrosion initiation of MMFX microcomposite reinforcing steel under rolled condition was about four times greater than that of carbon steel. Furthermore, there are few studies that have considered simultaneously all the following important criteria related to the corrosion performance of reinforcing steels in RC structures, namely: 1) chloride threshold in corrosion initiation stage; 2) corrosion rate in corrosion propagation stage; 3) performance in concrete and in simulated concrete pore solution; and 4) pitting corrosion resistance. The objective of this study is to investigate the corrosion performance of a low carbon, chromium steel (ASTM A 1035/A 1035M07) produced by MMFX Technologies Corp., which is referred to as Cr steel in concrete structures through laboratory testing and its comparison with those of conventional carbon steel and stainless steel 2205, 316LN and 304LN, in order to provide a comprehensive understanding of its expected corrosion performance in RC structures. This report presents experimental set-up details and results after the two-year investigation including: 1) the results of measurements of the corrosion potential and corrosion rate of both sandblasted and non-sandblasted steel bars in concrete prisms; and 2) chloride threshold and corrosion rates of the polished steel samples in the simulated concrete pore solution, as well as pitting corrosion resistances.

Two main types of testing were carried out in this study on the corrosion performance of Cr steel and four other types of steel: 1) testing in concrete prisms; and 2) testing in 3

electrochemical cells containing electrolytes simulating pore solutions of concrete. The testing in electrochemical cells enables fast comparative studies of the performance of different types of steel, and the results can provide a more fundamental understanding of the corrosion characteristics of steel, in addition to an assessment of chloride threshold values. On the other hand, the testing in concrete will provide information on the corrosion performance that is more similar to that in actual concrete structures. Each of the two types of testing has its own merit and their results complement each other (Page and Treadway 1982). 2.1 Testing in Concrete Prisms 2.1.1 Preparation of concrete prisms The concrete prisms had a size of 200 mm 150 mm 80 mm and two parallel No. 5 (ASTM A 615, diameter of 15.9 mm) steel bars were embedded in each prism, as illustrated in Figure 1. The steel bars included carbon steel (CS), Cr steel produced by MMFX Technologies Corp., stainless steel 316LN (SS 316LN) and stainless steel 2205 (SS 2205). The bars were prepared to be 230 mm in length; one end was drilled and an 18 gauge wire was attached to the bar using a crimp fit connection and a stainless steel 8-32 bolt. Both ends were coated with epoxy resin and covered with a shrinkable tubing, to prevent chipping of the epoxy. Such a preparation was necessary to eliminate the edge effect, and the total length of a bar exposed to concrete is 150 mm. Surface treatments included sandblasting of all types of steel bars. For Cr steel and CS, some additional steel bars were prepared under as-received condition or non-sandblasted (NS) for comparison purposes, which were identified as Cr steel-NS and CS-NS.

The concrete mix design was the same for all the prisms with CSA GU (previously CSA type 10) Portland cement and weight ratio of water: cement: sand: aggregate = 0.5:1:2:3 , but with varying concentrations of chloride ions, expressed in percentage by weight of cement. As seen in Table 1, ten batches of concrete mix were made with chloride concentrations varying from 0%, 1.5%, 3.0%, 4.5%, and 6.0%. The concrete mixtures were cast into plexiglass molds, and after 24 hours, the concrete prisms were de-molded and then were cured for 28 days in a concrete curing room with 95% 5% relative humidity (RH) and 22 oC 2 oC. The 28-day compressive strengths of the cylindrical specimens with chloride concentrations of 0% and 6.0% are also listed in Table 1.

2.1.2 Sampling of concrete prisms and steel bars The concrete prisms were designed for the measurement of the main corrosion performance indicators, including corrosion potential, and corrosion rate. For the measurements of corrosion potential and rate, 30 prisms were prepared and the variables 4

included: i) steel type, ii) chloride concentration, and iii) initial steel surface condition, as shown in Table 2. For each steel type and chloride concentration, two prisms containing four replicates of steel bars were prepared (except one prism containing only one bar). For the low chloride concentrations of 0% and 1.5%, no prisms were prepared for the stainless steels because of their relative high corrosion resistance; and likewise, for the high chloride concentrations of 4.5% and 6.0%, no prism was prepared for carbon steel because of its relative low corrosion resistance. All the steel bars in the prisms in Table 2 (section A) were first sandblasted and immersed in a saturated calcium hydroxide, Ca(OH)2, solution for 24 hours before they were cast into concrete. These two steps were done first to eliminate the effect of initial conditions of these bars obtained from different manufacturing sources, and secondly to ensure a relatively uniform initial passive condition before they were cast into concrete containing chlorides. The above surface treatment of the steel bars is necessary to compare the performance of different types of reinforcing steel; however, this procedure is not practical on the construction site of concrete structures. In order to address this difference, some additional prisms were prepared which contained Cr steel-NS bars under as-received condition without sandblasting. As shown in Table 2 (section B), two chloride concentrations, 4.5% and 6.0%, were selected for Cr steel-NS bars. More nonsandblasted Cr steel-NS bars were later added into the program after the preliminary experimental findings, as well as non-sandblasted CS-NS bars to allow for a more comprehensive comparison. Another set of 16 bars were prepared and cast into 8 concrete prisms, including four non-sandblasted Cr steel-NS bars and four non-sandblasted CS-NS bars at 1.5% and 3.0% of chlorides, respectively, as listed in Table 2 (section C).

2.1.3 Measurement set-up After 28 days in the curing room, the concrete prisms were placed in an environmental chamber, in which the relative humidity (RH) was set at 80% and the temperature was cycled between 25oC and 45oC, as shown in Figure 2. It provided an environment for accelerated corrosion of the embedded steel bars. The corrosion measurements were scheduled every three weeks for the first 12 weeks, then every four weeks for the following 40 weeks, and every eight weeks for the remaining time. During each scheduled week of testing, the environmental chamber was set at 25C and 80% R.H. until all the measurements were completed. For each set of measurements, the prisms were removed from the chamber for less than four hours. The linear polarization resistance (LPR) measurements were carried out using a Solartron Electrochemical Interface. The measurements were initiated at the open circuit potential (corrosion potential), Ecorr, scanned to 15 mV below Ecorr and then to 15 mV above Ecorr, and finally terminated at 15 mV below Ecorr with a scanning rate of 0.01 mV/s. At the 5

same time, the polarization current density (i) was recorded. The polarization resistance RP was calculated as the slope of a potential-current density plot at the potential (E) near Ecorr, i.e.: dE . (1) RP di The corrosion rate was calculated from the measured Rp by the Stern-Geary relation (Jones 1992) as follows: B icorr , (2) Rp where B is a constant determined by the polarization properties of both cathodic and anodic Tafel slopes. In this report, a constant of B=26 mV was used for all the corrosion rate calculations for simplicity (Andrade and Gonzlez 1978). The AC impedance measurements were performed by using a Solartron Electrochemical Interface coupled with a Frequency Response Analyzer (FRA). The measurements were carried out at the open circuit potential with a 5 mV sinusoidal signal scanned over the range of frequencies from 10,000 Hz to 1 Hz for the concrete IR drop correction and to 0.01 Hz for the characterization of the concrete/steel interface, including concrete resistance, interfacial resistance and capacitance. For LPR and AC impedance measurements, a stainless steel plate served as a counter electrode to the steel bars, and a Cu/CuSO4 electrode (CSE) as a reference electrode. In the concrete prisms, all measured potentials obtained are presented against CSE.

2.2 Testing in Electrochemical Cell 2.2.1 Electrodes and solution The corrosion performances of Cr steel, carbon steel (CS) and three types of stainless steel, SS 2205, SS 304LN and SS 316LN, were studied in electrochemical cells (EC). The steels were machined into electrode samples that were 15 mm in length and 13 mm in diameter, from the same batch of steel bars used in the concrete prisms. The electrode samples were threaded so that another CS or SS rod could be screwed into them to make a good electrical connection. The rods were isolated from the solution by a glass tube. The electrode sample and the end of glass tube were then embedded in epoxy resin leaving each a steel surface with a fixed area of 133 mm2 exposed to the solution. The electrode samples were polished with #600 silicon carbide papers, degreased by acetone and washed by de-ionized water, and then immersed in the saturated Ca(OH)2 solution with a pH of 12.6 for two days for the formation of passive film before testing.

The electrochemical experiments were carried out in electrochemical cells with a saturated Ca(OH)2 solution (pH=12.6) to simulate the concrete pore solution. For preparing solutions, de-ionized water was used and high purity argon and oxygen were used in some experiments to purge or increase the oxygen concentration in the solution. 2.2.2 Electrochemical techniques The electrochemical experiments included the following techniques: linear polarization resistance, potential dynamic for pitting test, AC impedance and open-circuit potential measurements. All tests were conducted in three-compartment electrochemical cells, where i) the working electrode was the steel sample; ii) the counter electrode was made of platinum foil or mesh with a large surface area; and iii) the reference electrode was a saturated calomel electrode (SCE). For all the experiments carried out in the electrochemical cells, the measured potentials are reported relative to SCE (e.g. if Ecorr of a reinforcing steel is measured to be 300 mV vs. SCE, it would be 360 mV vs. CSE which is used in ASTM C876-91 for concrete structures.) A Luggin capillary was used to reduce the IR drop between the tip of the reference and surface of the working electrode. The measurements were carried out using a Solartron 1480 MultiStat or Solartron SI 1287 Electrochemical Interface, which was controlled by a PC computer using CorrWare software. Similar to the measurement in the concrete prisms, AC impedance spectroscopy was measured by a Solartron system, from which the solution resistance between tip of reference and working surface in electrochemical cell was determined and then used for the IR correction, which was for the calculation of a more precise Rp. The linear polarization resistance technique was used to determine the electrochemical polarization resistance (Rp) and the corrosion rate (Icorr) of the steel electrodes. The potential of the steel electrode was scanned at a slow rate of 0.01 mV/sec, 15 mV around open-circuit potential. The open circuit potential was recorded before each measurement. A sufficient time was left between the measurements to allow the steel sample to fully depolarize. 2.2.3 Pitting resistance measurement Pitting corrosion is a process involving the formation of fine-sized pits (e.g. < 0.1 mm) in a passive surface of a metal as a consequence of the aggressive action of a component, such as chloride ion in the solution (electrolyte). It is a form of highly localized corrosion, which involves loss of passivity around pitting holes. It usually occurs at high potential, where passivity is expected to dominate. The most commonly pitting agent in an aqueous electrolyte is chloride ion. Many metals undergo pitting corrosion in solutions with sufficient chloride concentration, including carbon steel and stainless steel.

The pitting corrosion performance was evaluated by potential dynamic scans. Each scan was started at open circuit potential, proceeded anodically (the potential was increased towards more positive direction) at a rate of 0.2 mV/second up to 770 mV vs. SCE, and then reversed its direction (the potential was scanned towards more negative direction) back to the open-circuit potential. During the potential scan, the corresponding current was monitored. If pitting occurred during scanning, the current would no longer follow the normal pattern but would increase more rapidly or not decrease even when the potential is decreasing to more negative values. The potential at which this abnormal pattern of current occurs is the pitting potential. It is known that the pitting potential decreases with increasing activity or concentration of chlorides (Szklarska-Smialowska 1986; Bertolini et al. 1996). The electrodes for pitting corrosion measurements were made from CS, Cr steel, SS 316LN, SS 304LN and SS 2205 the same way as those used in the electrochemical cell tests. They were first polished using #600 silicon-carbide papers and then immersed in saturated Ca(OH)2 solution of pH=12.6 for two days, before the open circuit potential and linear polarization resistance were measured in this solution with no chlorides. Afterwards, NaCl was added in small increments (the size of the increment was dependent on the type of steel) into the solution and dissolved by stirring the magnetic stirrer. An electrode was left in the solution at that chloride concentration for several hours to reach equilibrium and a stable open-circuit potential value, before LPR and potential dynamic scan were conducted. After one scan is completed, the chloride concentration was then increased, and the same measurement process was repeated until a pitting corrosion was observed, and this concentration will be the concentration for pitting corrosion for the specific steel. 2.3 Additional Testing of Material Properties The corrosion resistance of reinforcing steel depends on its chemical composition and microstructure. Information on the chemical composition and microstructures could be helpful to identify and characterize the steel bars under testing and provide additional explanation of the corrosion data in terms of material properties. A Hitachi S-4800 ColdField Scanning Electron Microscope (SEM) was used to acquire images of the microstructures of steel. The images were acquired at beam energy of 2 keV and a current of 10 A. An Energy Dispersive Spectrometer (EDS) provided elemental analysis of the image area. Table 3 lists the chemical compositions of the five types of steel. A small section of rebar was taken from each steel, and small disks of steel were cut in cross-section 3 mm thick using a low speed-wafering saw with a diamond blade. The disks were polished on a lapping wheel with the surfaces ground using silica carbide grinding compound from 80 to 600 grit, and then were polished with alumina polishing

compound from 15 m to 1 m using appropriate polishing cloths. The polished disks were stored in a closed container in a nitrogen environment to prevent oxidation. For the microstructure characterization, one disk for each type of steel was etched for 7.5 minutes, rinsed with isopropyl alcohol, and then etched for another 7.5 minutes. The etched disks were rinsed with isopropyl alcohol, dried and stored in a nitrogen environment. The images of microstructures of the four steels are shown in Figure 3. As expected, Figure 3 shows that the carbon steel has two phases of ferrite and pearlite, Cr steel has a martensitic structure, SS 316LN has an austenitic structure, and SS 2205 has two phases of austenite and ferrite.

The experimental results of the two-year investigation of the corrosion performance of sandblasted Cr steel, CS, SS 316LN and SS 2205 in concrete prisms, including the effect of chloride concentration and steel type, are presented in this section. 3.1 Effect of Chloride Concentration

3.1.1 Corrosion performance of carbon steel The corrosion performance of carbon steel (CS) in chloride-contaminated RC structures has been more extensively studied and is better known than other types of steel. Results of the corrosion performance of CS are first presented to serve as a reference for other types of steel. Figure 4 presents the measured corrosion potentials (often referred to as half-cell potential) of the CS bars, denoted by Ecorr, as a function of time in the concrete prisms containing three different chloride concentrations, namely: 0%, 1.5%, and 3.0%. Figure 5 presents their respective corrosion rates, denoted by Icorr, as a function of time. On these curves as well as on the other curves to be presented, each datum represents an average value measured each time on four replicates of steel bars, and an error bar of standard deviation is also presented for each datum point in order to show the variation among the replicates. Table 4 and Table 5 list the guidelines for interpreting the values of Ecorr and Icorr, respectively, in terms of the corresponding corrosion state. As shown in Figure 4 and Figure 5, the three different chloride concentrations of concrete yielded distinct values of Ecorr and Icorr measured on the embedded CS bars. The general trend is that as the chloride concentration increases, Ecorr becomes more negative and Icorr becomes higher, which means that the risk of corrosion increases. This trend is expected, and more specifically:

1. For the concrete containing no chlorides, the Ecorr were mostly between 300 mV and 200 mV for the first three months, and then remained above 200 mV till the end of the two-year monitoring program. The Icorr values showed that the CS bars remained in the passive state. 2. For chloride concentration of 1.5%, the Ecorr showed a continuous increase for the first five months, and then remained between 350 mV to 300 mV for the remaining period of the monitoring program. At the same time, the Icorr showed a continuous decrease from about 0.5 A/cm2 to 0.1 A/cm2, ranging from moderate to low corrosion rate in the first five months, only to increase again slightly to above 0.5 A/cm2 with time, and remained at a moderate corrosion rate throughout the remaining period. Figure 6(a) shows the photo of the prism taken at the age of 21 months. Only minor cracks can be observed on the surface of the prism. 3. For chloride concentration of 3.0%, the Ecorr were between 600 mV and 400 mV with time. The Icorr varied from 1 A/cm2 to 2 A/cm2 (high corrosion rate) over the first six-month period. The most notable characteristic is that Icorr started to increase after about six months, and reached a value close to 100 A/cm2 by the end of the first year. At that time, the concrete cover was found completely delaminated, as seen in the photo presented in Figure 6(b) and the monitoring was finished for these samples. The finding of time-increase of the corrosion rate of the CS bars in concrete prisms containing 3% of chlorides is interesting and could be important, considering that most of the service life prediction models take into account only an average value of corrosion rate in the propagation stage, which is also normally much lower than what was found (100 A/cm2). This finding calls for further study (e.g. a study under a non-accelerated corrosive environment) to confirm its possible generalization. Also, it is worthwhile to continue the monitoring after the end of the two-year program in order to collect more data, especially to see if there is significant corrosion acceleration in concrete containing 1.5% of chlorides, as well as in those concrete prisms containing 3.0% and more of chlorides for the Cr steel bars. This will help gain a better understanding of the corrosion performance throughout the whole propagation stage until the delamination of concrete. These results also showed that for each chloride concentration, both the half-cell potential (Ecorr) and LPR measurements (Icorr) showed the progress of corrosion with time on the CS bars. They provided complementary data in that the potential became more negative as the corrosion rate generally increased. The ASTM C876-91 guidelines group Ecorr into three levels of corrosion risk (Table 4) and the empirical estimation of corrosion rate by

10

Andrade and Alonso (1996) (Table 5) has been widely used to associate Icorr to the corrosion state. Corrosion states interpreted by Ecorr and Icorr under these guidelines, however, are not necessarily in agreement; for example, one corrosion state interpreted by Ecorr can mean several corrosion states interpreted by Icorr. The CS bars were in a passive state as measured by Icorr in the concrete prisms without chlorides, while they were categorized as in intermediate corrosion risk because the measured Ecorr were below 200 mV in the first three months and in the low corrosion risk afterwards. The CS bars in the concrete prisms containing 1.5% of chlorides had a low to moderate corrosion rate as measured by Icorr, while the Ecorr identified their corrosion states as high corrosion risk in the first three months and intermediate corrosion risk afterwards. It can be seen that when Ecorr measured by the half-cell potential technique falls between 350 mV to 200 mV, which corresponds to an intermediate corrosion risk, the actual corrosion state can be either passive, low corrosion rate, or moderate corrosion rate. This is because the corrosion potential is a thermodynamics parameter influenced by many factors, such as oxygen content, ion concentration and temperature. For example, one of the reasons for the increase of Ecorr in the concrete with no chlorides may be due to the continuing growth of the passive film, hydration and drying in the concrete at the concrete/steel interface. Therefore, LPR measurement should be considered for the determination of the actual corrosion state when the half-cell potentials fall into the intermediate corrosion risk. Furthermore, the ASTM C876-91 guidelines for interpreting the potential readings were developed for carbon reinforcing steel, and it may not be applicable for other types of steel. As a result, the following interpretation of the corrosion potentials for the Cr steel as well as stainless steels will not be limited to the ASTM C876 guidelines but will be considered in conjunction with the corrosion rate measurements for a more appropriate assessment of corrosion state. The symmetric error bars for each datum of Ecorr or Icorr represent the variations of the corrosion performance among the replicates of the steel bar. Note for Icorr, only one side of the error bars are presented; otherwise the log-scale would make the other side appear misleadingly unsymmetrical. For the same concrete material and chloride concentration, design, and environmental exposure, the variation must be caused by the heterogeneous nature of concrete materials and heterogeneous nature of steel at the microstructural level (as seen in Figure 3) on which chloride ions induce the corrosion activity. The variation as shown in Figure 4 and Figure 5 seemed to be random with time regardless of the various chloride concentrations, which may be due to the variation of the local change of environment with time at the steel bar/concrete interface.

11

3.1.2 Corrosion performance of Cr steel Figure 7 and Figure 8 show the corrosion potentials (Ecorr) and corrosion rates (Icorr), respectively, of the sandblasted Cr steel bars in the concrete prisms containing five chloride concentrations, namely, 0%, 1.5%, 3.0%, 4.5%, and 6.0%, as a function of time. The patterns of change in Ecorr or Icorr with chloride concentration are similar to those of the CS bars that Ecorr decreased while Icorr increased with increasing chloride concentration. More specifically, for the Cr steel bars in concrete prisms with no chlorides, the Ecorr became more positive with time and reached values above 100 mV in about one year. At the same time, the Icorr started from below 0.1 A/cm2, indicating that these bars were in the passive state, and then decreased further with time to below 0.01 A/cm2. On the other hand, for the Cr steel bars in concrete prisms containing high chloride concentrations of 3.0%, 4.5% and 6.0%, the average Icorr were initially between 0.1 and 1 A/cm2, indicating that these bars were corroding at a low or a moderate rate, and they showed slight increase with time but stayed between 0.1 and 1 A/cm2 over the two-year monitoring program. The Ecorr of the bars at these concentrations were between 400 mV and 550 mV over time. The performance of the Cr steel bars at 1.5% of chlorides is the most notable. Both its Ecorr and Icorr are plotted in Figure 9 for closer examination. The Icorr at the beginning of the monitoring were below 0.1 A/cm2 (passive state) and decreased further with time for the first five months, indicating that the passive film continued to build up, while the Ecorr values became more positive from about 350 mV to 250 mV. For comparison, the CS bars at same chloride concentration had active corrosion at a low corrosion rate. After about five months, however, the Icorr started to increase significantly and Ecorr showed a significant decrease until about the age of ten months, when the Icorr reached the value of 0.1 A/cm2, corresponding to active corrosion. At the same time, the Ecorr was about 400 mV. This was the transition period of the Cr steel bars from passive to active corrosion. After that, the Icorr fluctuated with time between 0.1 A/cm2 to 0.5 A/cm2, within a range of low corrosion rates. For comparison, the CS bars at this stage had a moderate corrosion rate for the same chloride concentration. As can be seen in Figure 9, the advance of active corrosion of the Cr steel bars was clearly captured by the measurements of both Ecorr and Icorr. The chloride concentration of 1.5% by weight of cement can therefore be defined as the chloride threshold for the Cr steel in this experiment. The chloride exposures of real concrete structures can be from light, moderate, heavy, to severe, depending on surface chloride concentration. For bridge decks in North America, heavy chloride exposure is estimated to have an average 12

surface chloride concentration of 5.3 kg/m3 (Weyers et al. 1993), which is equivalent to about 1.5% by weight of cement. It would take time for the chlorides on the surface to penetrate onto the embedded steel/concrete interface, depending on the design of concrete structures such as cover depth, concrete permeability, etc. Therefore, the chloride concentration of 1.5% by weight of cement mixed into concrete prisms used in this experiment was equivalent to a highly corrosive environment for the steel bars. The results obtained at 1.5% of chlorides also showed that Cr steel had better corrosion resistance than carbon steel in two aspects: 1) longer corrosion initiation stage before active corrosion; and 2) lower corrosion rate in the corrosion propagation stage. It should be noted that the period of ten months before active corrosion was observed in the environmental chamber, corresponding to an accelerated corrosion environment. The time to active corrosion should be much longer under field conditions, considering also the time it would take for chlorides to penetrate onto the embedded steel surface. The results of the comparison of the corrosion performance of the Cr steel and CS bars can be summarized as follows: 1. Carbon steel: as the chloride concentration increased from 1.5% to 3.0%, the Icorr of the CS bars increased significantly from moderate to high corrosion rate. Furthermore, the Icorr at 3.0% of chlorides had accelerated after about six months while the Icorr at 1.5% of chlorides had only slight increase over the two-year period. 2. Cr steel: the Icorr of the Cr steel bars, however, did not increase as significantly as those of the CS bars when the chloride concentration was increased from 3.0% up to 6.0%. The Icorr at these concentrations were all below 1 A/cm2, which is within a range of low to moderate corrosion rate. It seems that the corrosion rate of Cr steel bars is not sensitive to heavier chloride exposures, up to 6.0% of chlorides in this study over the two-year monitoring period. This means that the service life would be possibly extended over the propagation stage after the initiation of active corrosion, even for relatively heavy chloride exposures compared with the CS steel. Figure 10(a) shows a photo of a concrete prism containing the Cr steel bar at 3.0% of chlorides taken after 21 months. Only minor cracks can be observed in this concrete prism, which are similar to the cracks in the prism with CS bars at 1.5% of chlorides. It should be noted that the above results were obtained on the sandblasted Cr steel rebars. The corrosion performance of asreceived Cr steel rebars will be discussed in Section 5. These preliminary findings after the two-year monitoring period call for further studies: 1) to continue the monitoring of the corrosion performance after the two-year program, in

13

order to further confirm this key finding of the performance of Cr steel bars though extending the propagation stage, which would have potential beneficial impact on the service life of concrete structures; 2) service life modeling in order to assess quantitatively its beneficial impact by taking into account the lower corrosion rate in the propagation stage, in addition to the extended initiation stage due to higher chloride threshold. 3.1.3 Corrosion performance of stainless steels Figure 11 and Figure 12 present the corrosion potentials (Ecorr) and corrosion rates (Icorr) respectively, of SS 2205 bars in the concrete prisms containing four different chloride concentrations, namely, 1.5%, 3.0%, 4.5%, and 6.0% as a function of time. All the Ecorr values, over time, remained in a consistently narrow range of about 100 mV and 150 mV after the first six months, except for two measurements at 1.5% of chlorides which had greater fluctuations. All the Icorr values were also in the narrow range below A/cm2 (passive state) at the beginning and fluctuated over time with overall values below A/cm2. At two measurements, the standard deviations of Ecorr of the four replicates at 1.5% of chlorides are noticeably high; e.g., a high deviation of 100 mV of one measurement is followed by a negligible deviation of less than 10 mV. This might be due to temporary dryness of the samples when the measurement was conducted, because the measured Icorr at the same time were also found to have relatively higher standard deviations. Even so, in these cases the Icorr of each of the four replicates was still lower than 0.1 A/cm2 and stayed passive, and therefore the corrosion performance of SS 2205 bars can be considered consistent and not sensitive to chloride concentration up to 6.0% with no active corrosion initiated over the two-year monitoring program. It means that its chloride threshold value in concrete was higher than 6.0% by weight of cement in this experiment. Figure 13 and Figure 14 present the Ecorr and Icorr, respectively, for SS 316LN bars in the concrete prisms containing three different concentrations of chloride ions, namely, 3.0%, 4.5%, and 6.0% as a function of time. The average Ecorr increased from about 280 mV to above 150 mV over the first seven months and remained around this value thereafter. Also, the average Ecorr for these different chloride concentrations remained within a narrow range of less than 50 mV except in the last few measurements. The average Icorr remained in a narrow range within a half order of magnitude, from about 0.02 A/cm2 (passive) at the beginning and then fluctuated over time around 0.01 A/cm2 (passive), except a couple of measurements for the bars at 6.0% of chlorides which showed relatively greater deviation. This shows that all the SS 316LN bars performed consistently and were not sensitive so far to chloride concentrations up to 6.0% with no initiation of active corrosion.

14

It was noted in the first six months that the Ecorr values of SS 316LN were more negative than those of SS 2205, but their Icorr were lower. With time, the Ecorr of SS 316LN slightly increased to about the same values as those of SS 2205, but their average Icorr remained slightly lower than those of SS 2205. These observations demonstrate again that the comparison of half-cell potentials should only be conducted among steel bars of the same type.

3.2 Effect of Steel Type In order to compare the corrosion performance of Cr steel with those of CS and SS, the half-cell potentials, Ecorr, and corrosion rates, Icorr, of these steel bars as presented above, are re-grouped so that each figure shows the different types of steel at each chloride concentration. Figure 15 shows the corrosion potentials (left Y-axis) and corrosion rates (right Y-axis) of the CS and Cr steel bars in the concrete containing no chloride ions. The Ecorr values of all the CS and Cr steel bars had similar values and changed with time in a similar manner. The Icorr of both types of steel also had similar values. The Cr steel bars had slightly lower corrosion rates than the CS bars after the first three months, which may be due to high content of chromium (Cr) element in this steel. It should be noted that the standard deviations of both steels with no exposure to chlorides were low throughout the two-year monitoring period, and this will be also used as a reference for other concentrations. For concrete containing chlorides, the local chloride concentrations would not be uniform along the embedded steel bar and the distribution pattern along the bar would be different among the bar replicates. Hence, the measured corrosion performance might be associated with higher deviations than that of the bars with no chlorides in concrete. Figure 16 and Figure 17 present the Ecorr and Icorr, respectively as a function of time, of the CS, Cr steel, and SS 2205 bars in the concrete prisms containing 1.5% of chlorides. Unlike the case of concrete with no chlorides, the Ecorr values covered three distinct ranges for the three types of steel, respectively, with overall values becoming more positive in the ascending order of CS, Cr steel, and SS 2205 within the first six months. The Icorr values, however, were in two distinct ranges: the corrosion rates of Cr steel and SS 2205 being similar and CS having a higher Icorr by more than one order of magnitude. After six months, the pattern changed, mainly because a significant change occurred on the Cr steel bars at this chloride concentration, as discussed in the previous section. The Ecorr of the Cr steel bars decreased rapidly with time, crossed over the Ecorr of the CS bars, and became even more negative. The Icorr of the Cr steel bars increased rapidly, but did not cross over and remained lower than the Icorr of the CS bars. This shows again that

15

although the half-cell potentials were more negative for Cr steel in this case, it didnt necessarily mean that it had higher corrosion rates. Figure 18 and Figure 19 present Ecorr and Icorr, respectively, as a function of time, for CS, Cr steel, SS 2205 and SS 316LN bars in the concrete prisms containing 3.0% of chlorides. Two distinct groups of Ecorr can be easily observed. The stainless steel bars had more positive Ecorr values over time as expected. On the other hand, both the CS and Cr steel bars have shown much more negative Ecorr values (lower than 400 mV), and the CS bars had more negative Ecorr by about 50 mV to 100 mV than those of the Cr steel bars. It is also observed that the standard deviations of Ecorr of the Cr steel and CS bars in active corrosion are much higher than those of SS bars, and also higher than those in the concrete prisms with no chlorides. The Icorr were different by several orders of magnitude for the different steels. Compared with the SS bars in the passive state, the Cr steel bars were at a low corrosion rate over the time, and the CS bars were at high corrosion rate with a narrow range of 1 A/cm2 to 2 A/cm2 until about six months of age when a significant increase of corrosion rate occurred, as previously discussed. Figure 20 and Figure 21 present the Ecorr and Icorr, respectively, of the Cr steel, SS 2205, and SS 316LN bars in the concrete prisms containing 4.5% of chlorides as a function of time. Again, both the SS 2205 and SS 316LN had Ecorr values almost identical after about seven months. The Ecorr of the Cr steel bars were around 500 mV over time. Both types of SS bars showed much more positive Ecorr over time than those of the Cr steel bars. Both types of SS bars were in the passive state over time, while the Cr steel bars were at a low to moderate corrosion rate over time. Similar behavior in each type of the tested steels were observed in concrete containing 6.0% of chlorides, as shown in Figure 22 and Figure 23.

4.1 Testing in Saturated Ca(OH)2 Solution of pH=12.6 The corrosion of Cr steel and four other types of steel, CS, SS 316LN, SS 304LN and SS 2205, were tested in electrochemical cells containing saturated Ca(OH)2 solution. Sodium chloride (NaCl) was added step by step at certain intervals, while the monitoring of corrosion potential, Ecorr, was continuous and the corrosion rate, Icorr, was measured several times for each chloride concentration. As a result, the change in the corrosion state, especially from a passive state to an active state, can be captured, and the chloride concentration at which it occurs will be identified as the chloride threshold.

16

Figure 24 shows corrosion potential (Ecorr) and corrosion rate (Icorr) recorded on an Cr steel electrode with chloride concentration (top X-axis), as well as with time (bottom Xaxis). The Ecorr was initially at around 300 mV vs. SCE (equivalent to 360 mV vs. CSE as used in ASTM C876-91), when the experiment started in the saturated Ca(OH)2 solution with no chlorides. As Ecorr increased within a short period of time, Icorr decreased sharply from about 0.03 A/cm2 (passive state) to 0.007 A/cm2, indicating a continued formation and growth of passive film on the surface. When chlorides were introduced into the solution, Icorr started to fluctuate slightly, clearly showing the interference of chloride ions with the passivated surface, but the general trend of Icorr was still decreasing even after three consecutive additions of chlorides. This means that the Cr steel remained passivated even when the chloride concentration reached 4.5%. Shortly after the chloride concentration reached 4.9%, Ecorr dropped suddenly by a significant amount of 170 mV, down to about 420 mV (equivalent to 480 mV vs. CSE) and Icorr had a significant increase by almost two orders of magnitude from 0.005 A/cm2 (passive state) to 0.25 A/cm2 (active corrosion). The dramatic changes in both Ecorr and Icorr had clearly captured the moment when the Cr steel surface was depassivated and the active corrosion was initiated. For this Cr steel electrode, corrosion was initiated at a chloride concentration of 4.9%, which was identified as the chloride threshold for this sample in this test. After the active corrosion occurred, the monitoring of Ecorr and Icorr continued in order to assess the further development of corrosion in the propagation stage, as the chloride concentrations continued to increase. As shown in Figure 24, Icorr continued to increase with chloride concentration, reaching 5 A/cm2 at 6.4% of chlorides, while the Ecorr reached 520 mV vs. SCE (equivalent to 580 mV vs. CSE). Figure 25 shows the progress of Ecorr and Icorr recorded on a CS electrode with chloride concentration and time. Similar to the Cr steel when there were no chlorides, Ecorr increased while Icorr decreased with time, indicating the continued build-up of the passive film in an environment of pH=12.6. When the chloride concentration reached 0.61%, the Ecorr had a relative large drop (from about 220 mV to 350 mV) and then increased again, and at the same time, the Icorr showed a relative large increase from 0.01 A/cm2 to 0.1 A/cm2 (both were passive though) before it started to decrease again. Both curves clearly showed the interference of chlorides with the passive film and then the recovery of the passive film at this small concentration of chlorides. When the chloride concentration was increased to 1.2%, Ecorr dropped significantly towards 500 mV, and at the same time, Icorr had a significant increase by two orders of magnitude from 0.035 A/cm2 (passive state) to 3.35 A/cm2 (active state). The

17

continuous monitoring of the CS steel also clearly captured the deterioration of the passive film and initiation of active corrosion at the chloride concentration of 1.2%, which is the chloride threshold of this CS electrode. The continuous monitoring of the corrosion potential and corrosion rate for both the Cr steel and CS electrode (Figure 24 and Figure 25) clearly showed how the passive film was developed and deteriorated with increasing chloride concentrations, as well as how the active corrosion was initiated. The chloride thresholds of all the other electrode replicates were obtained in the same way, and the chloride threshold of one type of steel will be reported as the average of the replicates. Figure 26 summarizes the corrosion performance with increasing chloride concentrations in terms of the corrosion rate for CS, Cr steel, SS 2205, SS 304LN and 316LN. The average chloride threshold of carbon steel is about 1.0%, while the average chloride threshold of Cr steel is around 5.0%. The stainless steels exhibited the highest chloride thresholds; their corrosion rates remained below 0.01 A/cm2 (10 times below the low corrosion rate of 0.1 A/cm2) even when the chloride concentration was increased to 21%, indicating no active corrosion was initiated on these stainless steels. Note the chloride threshold is really a function of many factors besides the chloride concentration, and one of them is temperature as shown in the previous section. The reported chloride threshold values were based on the testing in the electrochemical cells at room temperature.

4.2 Pitting Corrosion Resistance A simple, theoretical way of rating the pitting corrosion resistance of different types of stainless steel is to compare their Pitting Resistance Equivalent Numbers, referred to as PREN, which is calculated based on their chemical compositions (SzklarskaSmialowska 1986). The following typical linear formula is used to calculate the PREN: PREN = Cr + m Mo + n N (3)

Where m and n are the factors for molybdenum and nitrogen. The most commonly used version of the formula is: PREN = Cr+ 3.3 Mo + 16 N (4)

Table 6 shows a range of calculated PREN values for the three types of stainless steel used in this study. As can be seen, SS 2205 has the highest PREN, therefore is expected

18

to have the highest pitting corrosion resistance. Since Cr steel is not a stainless steel, a PREN calculated from its composition does not necessary reflect its pitting resistance; however, it was still calculated to be 8.6 to 10.6, just to have a rough estimate, but not to be used to compare with PREN of stainless steel. The actual pitting corrosion resistance was tested for these steels, including carbon steel. Figure 27 shows the pitting scans of one Cr steel electrode at different chloride concentrations. The scan was initiated at open-circuit potential (around 0.125 V vs. SCE) and scanned anodically towards 0.77 V vs. SCE. The direction of the scan was then reversed after reaching a potential of 0.77 V, and continued to scan back (lower) to the open-circuit potential. The current increased gradually until around 0.6 V and then increased at a relatively faster rate. In general, pitting corrosion in chloride containing solutions is characterized by a minimum potential, called the pitting potential (the turning point), above which the pitting is observed to occur and below which the electrode is in essence passive. The pitting potential often decreases with chloride concentration or activity. Similar scan patterns were recorded in the solutions containing 0, 0.061% and 0.12% of chlorides, respectively. However a different scan curve started to appear in the solution containing 0.18% of chlorides in which the current had a significant increase from the normal scan curve and remained at a high value even after the scan direction was reversed, indicating that pitting corrosion was initiated at this chloride concentration. Figure 28 and Figure 29 show similar potential dynamic scans performed on the CS, and SS 2205 electrodes, respectively. The overall shapes of the curves without pitting corrosion are somewhat different as a result of different anodic processes on different steel surfaces. However, the basic patterns are similar when pitting corrosion occurs: e.g. the current increased significantly and remained at this large value for most of the scan period. Table 6 also summarizes the chloride concentrations for pitting corrosion to initiate in saturated Ca(OH)2 solutions. The pitting corrosion was observed at the lowest chloride concentration of 0.06% on carbon steel. It was initiated on Cr steel at chloride concentrations in the range of 0.180.3%. On SS 304LN and SS 316LN, the pitting corrosion was initiated at the chloride concentration in the range of 3.0-7.3% and 3.69.7%, respectively. Regarding SS 2205, as the chloride concentration was increased to 21%, no pitting corrosion was observed. Obviously, SS 2205 had the highest pitting corrosion resistance measured in this test and this conforms to what was predicted by its PREN as shown in Table 6. It is noted that the pitting corrosion occurred in a relatively wide range of chloride concentrations for the SS 304LN and SS 316LN. However, it occurred in a relatively

19

narrow range for CS, Cr steel and SS 2205 steels. Pitting corrosion usually initiates at a very small point on the steel surface, where an alloy is likely depleted locally or some harmful elements are enriched. It then propagates quickly while most of the surface remains passive. For example, the presence of significant quantities of sulphur as an impurity in ferrous and nickel-based alloys is generally degrading for passivity. Sulphur becomes incorporated into the oxide film or accumulated at the metal/oxide interface; the element retards formation of the oxide film and catalyses dissolution of iron and nickel, leading to the pitting corrosion that appears in a wide range of chloride concentrations.

5.1 Literature Review on Effect of Mill Scales It is important to compare the corrosion performance of different types of steel by using polished samples; however, in real construction most of un-coated steel rebars are milled scaled on surface. Mill scales are natural iron oxides (FeO, Fe2O3 and Fe3O4) that are formed on the steel surface during the heat treatment of the fabrication processing or exposure to air. The removal of mill scales is not required for carbon steel rebars while it is recommended for most stainless steel products by ASTM A 380. Mechanical or chemical methods are used for surface treatment, generally known as polishing, to remove the mill scales. Sandblasting and polishing by sand papers are commonly used mechanical methods, and pickling is a chemical treatment for which ASTM A 380 recommends the use of an aqueous solution of nitric and hydrofluoric acid at 60 C. The effect of the removal of mill scales on the corrosion performance of carbon steel however, is found highly controversial in literature. Mohammed and Hamada (2006) reported that the chloride threshold value of polished carbon steel is two time of that of mill-scaled (as-received condition) carbon steel. Pillai and Trejo (2005) found on the other hand that the polishing of carbon steel (ASTM A 615) reduced its threshold value by about 40% by using an accelerated chloride penetration test. Mahllati and Saremi (2006) observed that the presence of mill scales will reduce the protective effect of the passive film, which implies that the chloride threshold value will also be lower for steel with mill scales. Once the active corrosion occurs, however, they found that the effect of mill scales on the corrosion rate is negligible because the main factor for corrosion rate is chloride ions. The effect of the surface treatment of stainless steel rebars on their corrosion performance in concrete is also highly controversial. Scully and Hurley (2007) reported that in the simulated pore solutions, the presence of mill scales will reduce the chloride threshold values of stainless steel 316 LN, 316L clad and 2102 LDX to that of the carbon steel, while Pillai and Trejo (2005) found that the polishing of stainless steel 304 increased its

20

threshold value by 24% but polishing 316 LN actually lowered its chloride threshold value by about 36%. On the other hand, Hansson and Mammolii (2008) found a marginally beneficial effect from pickling stainless steel 316LN and 2205 by tests in concrete and synthetic pore solutions. The controversy comes most from the different polishing methods and different corrosion performance evaluation methods and criteria. Few mechanistic studies and explanation were found in the literature. Mohammed and Hamada (2006) found that the interface between concrete and steel was less porous for the polished rebars, which might provide additional corrosion protection from the concrete side for the polished rebars due to the reduced electrical resistance. Mahllati and Saremi (2006) concluded that the passive film has less electric resistance on the steel surface with mill scales. For the relatively new steel type of the low carbon chromium steel (ASTM A 1035/A 1035M-07, referred to as Cr steel), there are even less studies on the effect of surface treatment. Pillai and Trejo (2005) found that the chloride threshold value of polished microcomposite steel, whose chemical compositions were almost the same as the Cr steel, was about 25% higher than that of microcomposite steel with mill scales. Scully and Hurley (2007) reported that in the simulated pore solutions, the presence of mill scales will reduce the chloride threshold values of MMFX microcomposite reinforcing steel (ASTM A 1035/A1035M-07) from a Cl/OH ratio of 4.9 to that of carbon steel (in a range of 0.25 to 0.34).

5.2 Experimental Observations To compare Cr steel with different types of steel, it is necessary to do sandblasting of all the steel bars as they came from different sources with different unknown surface conditions. The objective of using the Cr steel-NS bars was originally to assess if different surface conditions would make a difference at the two high chloride concentrations. It was found that the corrosion potentials, Ecorr, of the non-sandblasted Cr steel bars (Cr steel-NS) were similar to those of the sandblasted Cr steel bars, however the Cr steel-NS bars had significantly higher corrosion rates, Icorr, at 4.5% and 6.0% of chlorides for the first six months of monitoring. After this preliminary finding, it was decided to add more non-sandblasted Cr steel-NS bars into the program, as well as non-sandblasted CS-NS bars for the sake of comparison. Another 16 bars were prepared and cast into concrete prisms, four replicates of nonsandblasted Cr steel-NS bars and non-sandblasted CS-NS bars at 1.5% and 3.0% of chlorides, respectively, as listed in Table 2. By the end of the two-year investigation program, these samples reached their one-year age.

21

Figure 30 and Figure 31 show the Ecorr (left Y-axis) and Icorr (right Y-axis) of both nonsandblasted and sandblasted CS bars for the chloride concentrations of 1.5% and 3.0%, respectively. It is observed that: 1) for 1.5% of chlorides, the Icorr of the non-sandblasted CS bars were higher than those of sandblasted and increased faster with time, while the average Ecorr were only marginally more negative; and 2) for 3.0% of chlorides, oppositely, the Icorr of the non-sandblasted CS bars were lower most of time than those of sandblasted bars, with the corresponding Ecorr being more positive. The results may be due to surface variation of nonsandblasted CS bars. Figure 32 to Figure 35 show Ecorr (left Y-axis) and Icorr (right Y-axis), as a function of time, for both the non-sandblasted and sandblasted Cr steel bars at chloride concentrations of 1.5%, 3.0%, 4.5%, and 6.0%. The following observations are made: 1. The initiation of active corrosion at 1.5% of chlorides occurred five month earlier on Cr steel-NS bars than on the sandblasted Cr steel bars. 2. The Icorr of the non-sandblasted Cr steel bars were higher than those of sandblasted Cr steel bars. The difference became larger with increasing chloride concentration, which was about half an order of magnitude at 1.5% and 3.0% of chlorides and was about one order of magnitude at 4.5% and 6.0% of chlorides. Figure 10(b) shows a photo of non-sandblasted Cr steel bars in concrete prisms containing 3.0% of chlorides taken at 14 months of age. It shows much more severe cracks than those of sandblasted Cr steel in concrete prisms containing 3.0% of chlorides as shown in Figure 10(a). 3. The Ecorr of the non-sandblasted Cr steel at 1.5% of chlorides were much more positive than those of sandblasted by about 100 mV to 200 mV. The Ecorr of the non-sandblasted Cr steel bars at 3.0% of chlorides were very similar to those of sandblasted over time. The Ecorr at 4.5% of chlorides were similar within the first six months, and thereafter the Ecorr of the non-sandblasted were more positive by about 100 mV. The Ecorr were similar within the first six month for 6.0% over the two-year monitoring period. These observations of Ecorr seemed not to follow any trend associated with chloride concentrations. The difference in performance measured by corrosion rate had an increasing trend with chloride concentrations, while the difference measured by the corrosion potential appeared not to have such a trend. These conclusions on the performance of non-

22

sandblasted Cr steel were based on the as-received surface condition of these tested bars. Note in field conditions as-received surface conditions of steel might be more complicated, possibly because of different storage conditions and handling before shipping, etc. A more systematic study is needed to study the corrosion performance of non-sandblasted Cr steel in terms of as-received surface conditions as used in field conditions.

23

The following provides a summary of the results obtained from the measurements in both electrochemical cells and concrete prisms, followed by discussions on future research. In the simulated concrete pore solution (pH=12.6), it was found that the average chloride threshold of carbon steel was about 1.0% by weight; the average chloride threshold of Cr steel was about 5.0%. The stainless steels exhibited the highest chloride threshold; no active corrosion was initiated even when the chloride concentration reached 21%. Pitting corrosion was observed to initiate on Cr steel at a range of chloride concentrations from 0.18% to 0.3%. Pitting corrosion was initiated on carbon steel, at a much lower chloride concentration around 0.06%. On stainless steels 304LN and 316LN, pitting corrosion was initiated at chloride concentrations in the range of 3.07.3% and 3.6-9.7%, respectively. For stainless steel 2205, no pitting corrosion was observed even when the chloride concentration reached to 21%. The sequence of pitting resistance of the tested steels conformed to what was predicted by their calculated PREN. The following observations were from the tests on sandblasted steel bars embedded in concrete prisms. o In chloride-free concrete, both Cr steel and carbon steel were in the passive state and no active corrosion was initiated over time. They demonstrated similar corrosion potential values. Their corrosion rates (corrosion current density) also had similar values, with Cr steel having a slightly lower rate than carbon steel. o In concrete containing 1.5% of chlorides, Cr steel remained in passive state before undergoing active corrosion at ten months of age. The progress of corrosion with time was clearly captured by both the potential and corrosion rate measurements. o As the chloride concentration increased from 1.5% to 3.0%, Cr steel showed lower corrosion rate than carbon steel. The corrosion rate of carbon steel increased significantly from a moderate to high corrosion rate. Furthermore, its corrosion rate increased after about six months, when the concrete cover was found completely delaminated. On the other hand, the corrosion rate of Cr steel remained low throughout the two years of monitoring. An interesting finding was that the corrosion rate of Cr steel did not increase significantly as the chloride concentration was increased from 3.0% to 6.0%. The corrosion rate was within a range of low to moderate at these concentrations. It appears that the corrosion rate of Cr steel is not sensitive to severe chloride exposures, up to 6.0% of chlorides during the two-year monitoring of this study.

24

Stainless steel 316LN and SS 2205 had shown the best corrosion performance in this study. Regardless of the chloride concentration up to 6.0% in concrete prisms, they remained in the passive state and their corrosion rates were consistently low in a narrow range of within half an order of magnitude around A/cm2. The corrosion rate of as-received Cr steel (non-sandblasted) was found to be much higher than that of sandblasted Cr steel, and the difference became greater with increasing chloride concentrations. The difference was about half an order of magnitude at 1.5% and 3.0% of chlorides, and increased to about one order of magnitude at 4.5% and 6.0% of chlorides. The initiation of active corrosion at 1.5% of chlorides concentration occurred five months earlier on non-sandblasted Cr steel than on sandblasted Cr steel. The performance of sandblasted Cr steel bars reflected its basic material properties, while the performance of non-sandblasted Cr steel was based on the as-received surface condition of these bars in this study only. Considering that the surface condition of steel rebars in field conditions might be more complicated, possibly because of different storage conditions and handling, etc., a more systematic study is needed to study the corrosion performance of non-sandblasted Cr steel, because it is the actual condition of steel that is used in the construction of concrete structures. The experimental findings provide valuable information on the corrosion performance of Cr steel in terms of chloride threshold value, corrosion rate in the propagation stage, and pitting corrosion resistance. Future research is needed on service life modeling and lifecycle cost analysis in order to assess the potential beneficial impact of using Cr steel as a more cost-effective material when compared to carbon steel and stainless steel.

25

1. AMEC Earth & Environmental (2006), Comparative performance of MMFX microcomposite reinforcing steel and other type of steel with respect to corrosion resistance and service life prediction in reinforced concrete structures, AMEC file VA06451. 2. Andrade C and Alonso C (1996). Durability design based on models for corrosion rates. The Modelling of Microstructure and Its Potential for Studying Transport Properties and Durability. Jennings H, Kropp J, and Scrivener K (eds.), Kluwer Academic Publishers, The Netherlands, pp.473-492. 3. Andrade C and Gonzlez JA (1978). Quantitative measurement of corrosion rate of reinforcing steels embedded in concrete by polarization resistance measurements. Werkstoffe und Korrosion. 29(8):515-519. 4. ASTM C 876-91 (1991). Standard test method for half-cell potentials of uncoated reinforcing steel in concrete. Annual Book of ASTM Standards. Philadelphia, USA: ASTM. 5. ASTM A 380 (2006), Standard Practice for Cleaning, Descaling, and Passivation of Stainless Steel Parts, Equipment, and Systems. Annual Book of ASTM Standards. Philadelphia, USA: ASTM. 6. ASTM A 1035/A 1035M (2007). Standard Specification for Deformed and Plain, Low-carbon, Chromium, Steel Bars for Concrete Reinforcement. Annual Book of ASTM Standards. Philadelphia, USA: ASTM. 7. ASTM A 615/A 615M (2009). Standard Specification for Deformed and Plain Carbon-Steel Bars for Concrete Reinforcement. Annual Book of ASTM Standards. Philadelphia, USA: ASTM. 8. Bertonini L, Bolzoni T, Pastore P, and Pedeferri P (1996). Behavior of stainless steel in simulated concrete pore solution. British Corrosion Journal. 31:218-226. 9. Clemea GG and Virmani YP (2003). Comparison of the corrosion resistance of selected metallic reinforcing bars for extension of service life of future concrete bridges in outdoor concrete blocks. Paper No. 03298, CORROSION/03, San Diego. 10. Darwin D, Browning J, Nguyen TV and Locke C Jr (2002). Mechanical and Corrosion Properties of a High-Strength, High Chromium Reinforcing Steel for Concrete. Report No. SD 2001-05-F, South Dakota Department of Transportation, USA. 11. EG & G Princeton Applied Research Application Note 140 Linear Polarization and Note 148 Tafel Plot. 12. Glass GK and Buenfeld NR (1997). Chloride threshold levels for corrosion induced deterioration of steel in concrete, in Chloride Penetration into Concrete, Nilsson LO and Ollivier JP (eds.). RILEM Publications. pp. 429-440. 13. Gonzlez JA, Algaba S and Andrade C (1980). Corrosion of reinforcing bars in carbonated concrete. British Corrosion Journal. 15 (3):135139.

26

14. Hansson CM and Mammoliti L (2008). Is it necessary to remove the mill scale from stainless steel reinforcing bars? Part 1: Tests in synthetic pore solution. Report HIIFP-015, Ministry of Transport Ontario,Canada. 15. Hansson CM and Mammoliti L (2008). Is it necessary to remove the mill scale from stainless steel reinforcing bars? Part 2: Tests in concrete. Report HIIFP-016, Ministry of Transport Ontario, Canada. 16. Jones DA (1992). Principles and Prevention of Corrosion, Second Edition, Prentice Hall. 17. Li LF, Caenen P, Daerden M, Vaes D, Meers G, Dhondt C, and Celis JP (2004). Mechanism of single and multiple step pickling of 304 stainless steel in acid electrolyte. Corrosion Science. 47:1307-1324. 18. Mahallati E and Saremi M. (2006), An assessment of the mill scale effects on the eletrochemcial characteristics of steelbars in concrete under DC-Polarization. Cement and Concrete Research. 36: 1324-1329. 19. Mohammed TU and Hamada H. (2006) Corrosion of steel bars in concrete with various steel surface conditions. ACI Materials Journal. 103(4): 233-242. 20. Scully J and Hurley M (2007). Investigation of the Corrosion Propagation Characteristics of New Metallic Reinforcing Bars, Report No. VTRC 07-CR9, Virginia Transportation Research Council, USA. 21. Page CL and Treadway KWJ (1982). Aspects of electrochemistry of steel in concrete. Nature. 297:109-114. 22. Pillai RG and Trejo D (2005). Surface Condition Effects on Critical Chloride Threshold of Steel Reinforcement. ACI Materials Journal. 102(2):103-109. 23. Szklarska-Smialowska Z (1986). Pitting Corrosion of Metals, NACE, Houston, Texas. 24. Trejo D and Pillai RG (2004). Accelerated Chloride Threshold TestingPart II: Corrosion-Resistant Reinforcement. ACI Materials Journal. 101(1): 57-64. 25. Tuutti K (1982). Corrosion of Steel in Concrete. Swedish Cement and Concrete Research Institute, Stockholm, Sweden. 26. Weyers RE, Prowell BD and Sprinkel MM (1993). Concrete bridge protection, repair, and rehabilitation relative to reinforcement corrosion: a methods application manual, SHRP-S-360.

27

Figures and Tables

Table 1. Concrete mix batches and properties Batch # Cl-%(wt of 28 day cement) strength (MPa) 1 15.0 30.0 60.0 90.0 0 0 47.9 2 7.5 15.0 30.0 45.0 0.74 3.0 3 7.5 15.0 30.0 45.0 1.11 4.5 4 15.0 30.0 60.0 90.0 2.97 6.0 41.4 6 7.5 15.0 30.0 45.0 0.37 1.5 7 5.0 10.0 20.0 30.0 0.49 3.0 8 5.0 10.0 20.0 30.0 0.74 4.5 9* 15.0 30.0 60.0 90.0 0.74 1.5 10* 15.0 30.0 60.0 90.0 1.48 3.0 * Additional concrete prisms were added about one year after the others Water Cement Sand Aggregate (kg) (kg) (kg) (kg) NaCl (kg)

Table 2. Sampling of concrete prisms for corrosion resistance measurement Number of concrete prisms A: Sandblasted steel Chloride concentration Rebar Type 0.0% 1.5% 3.0% 4.5% 6.0% CS 2 2 2 Cr steel 2 2 2 2 2 SS 2205 2 2 2 2 SS 316 LN 2 2 2 B: As-received condition (non-sandblasted steel) Cr steel-NS 2 2 C: More as-received condition * Cr steel-NS 2 2 CS-NS 2 2 * Additional concrete prisms were added about one year after the other bars

28

Table 3. Chemical compositions of steel Element Fe Cr Ni Mo N C Mn Cu Si P S V B Other CS 97.8 Weight percentage (%) SS 2205 SS 316LN Cr steel 68.8 66.6 89.5 21.43 17.91 9.34 4.69 10.27 0.1 2.61 2.13 0.02 0.1 0.12 0.02 0.015 0.08 1.77 1.55 0.67 0.19 0.37 0.15 0.36 0.85 0.12 0.029 0.025 0.0008 0.001 0.003 0.0008 0.025 0.003 0.0024 0.11 0.015 SS 304LN* 17-19.5 8-11.5 0.12-0.22 <0.03 <2.00 <1.00 <0.045 <0.03

0.1 0.02

0.36 0.029 0.001 0.003 0.016

* As supplied by manufacture.

Table 4. ASTM C876-91 criterion for corrosion probability of steel in concrete Measured potential Ecorr (mV vs. CSE) Corrosion probability Low, 10% risk of corrosion Ecorr > 200 Uncertain * 350 < Ecorr < 200 High, 90% risk of corrosion Ecorr < 350 * Intermediate corrosion risk or increasing probability of corrosion

Table 5. Typical corrosion condition of steel in concrete Corrosion Rate High Current density value, Icorr ( A/cm2) 1.0 < Icorr

Moderate 0.5 < Icorr < 1.0 Low 0.1 < Icorr < 0.5 Passive Icorr < 0.1 * Adapted from Andrade and Alonso 1996.

29

Table 6. Calculated PREN and measured chloride concentration for pitting corrosion Type of steel Cr Mo N Calculated Tested Pitting PREN corrosion Cl (wt %) Carbon steel ------<0.06 Cr steel 8.0-11.0 -0.04 8.6-10.6 0.18-0.3 SS 304LN 17.0-19.5 NS 0.12-0.22 18.9-23.0 3.0-7.3 SS 316LN 16.5-18.5 2.0-2.5 0.12-0.22 25.0-30.3 3.6-9.7 SS 2205 21.0-23.0 2.5-3.5 0.10-0.22 30.8-38.1 21

Figure 1. Schematic configuration of concrete prism with embedded steel bars

Figure 2. Temperature cycle controlled in environmental chamber

30

(a) Carbon Steel

(b) Cr steel

(a) SS 316 LN

(b) SS 2205

Figure 3. SEM images of microstructures of different types of steel

31

Figure 4. Corrosion potential of carbon steel in concrete prism

Figure 5. Corrosion rate of carbon steel in concrete prism

32

(a)

Carbon steel with 1.5% of chlorides at 21 months of age

(b)

Carbon steel with 3.0% of chlorides at 12 months of age

Figure 6. Carbon steel in concrete at various levels of corrosion-induced damage

33

Figure 7. Corrosion potential of Cr steel in concrete prism

Figure 8. Corrosion rate of Cr steel in concrete prism

34

Figure 9. Performance of Cr steel in concrete prism with 1.5% of chlorides

35

(a)

Sandblasted Cr steel at 21 months of age

(b)

Non-sandblasted Cr steel at 14 months of age

Figure 10. Photos of concrete prisms with Cr steel at 3.0% of chlorides

36

Figure 11. Corrosion potential of SS 2205 in concrete prism

Figure 12. Corrosion rate of SS 2205 in concrete prism

37

Figure 13. Corrosion potential of SS 316LN in concrete prism

Figure 14. Corrosion rate of SS 316LN in concrete prism

38

Figure 15. Corrosion performance of CS and Cr steel in chloride-free concrete

Figure 16. Corrosion potential of steels in concrete prisms with 1.5% of chlorides

39

Figure 17. Corrosion rate of steels in concrete prisms with 1.5 % of chlorides

Figure 18. Corrosion potential of steels in concrete prisms with 3.0 % of chlorides

40

Figure 19. Corrosion rate of steels in concrete prisms with 3.0 % of chlorides

41

Figure 20. Corrosion potential of steels in concrete prisms with 4.5 % of chlorides

Figure 21. Corrosion rate of steels in concrete prisms with 4.5% of chlorides

42

Figure 22. Corrosion potential of steels in concrete prisms with 6.0 % of chlorides

Figure 23. Corrosion rate of steels in concrete prisms with 6.0 % of chlorides

43

Chloride Content
0% 3.6% 4.2% 4.5% 4.9% 5.2% 5.5% 5.8% 6.1% 6.4%

10 1 0.1 0.01 0.001 0.0001 0.00001 600

Corrosion Potential (mV vs SCE)

-200 -300 -400 -500 -600 0 100 200 300 400 Time (hour) 500

Corrosion Potential Corrosion Rate

Figure 24. Corrosion potential and rate of Cr steel in saturated Ca(OH)2 solution
Chloride Content
0
0% 0.61% 1.2% 1.8% 2.4% 3.0%

10 1 0.1

Corrosion Potential (mV vs SCE)

-200 -300 -400 -500 -600 0 100 200 300 Time (hour) 400

Corrosion Potential Corrosion Rate

0.01 0.001 0.0001

0.00001 500

Figure 25. Corrosion potential and rate of carbon steel in saturated Ca(OH)2 solution.

44

2 Corrosion Rate (A/cm )

-100

2 Corrosion Rate (A/cm )

-100

Figure 26. Corrosion rate of steel in saturated Ca(OH)2 solution


1.E-02 1.E-03
Cl 0% 0.06% 0.12% 0.18%

Current density (A/cm2)

1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 -0.2

0.2

0.4

0.6

0.8

Potential (V vs SCE)

Figure 27. Potential dynamic scans of Cr steel in saturated Ca(OH)2 solution.

45

1.E-01 1.E-02 1.E-03


2 Current density (A/cm )

Cl 0% Cl 0.06%

1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 -0.4

-0.2

0.2

0.4

0.6

0.8

Potential (V vs SCE)

Figure 28. Potential dynamic scans of carbon steel in saturated Ca(OH)2 solution.

1.0E-01 1.0E-02
Cl 12.1% Cl 12.7% Cl 13.3% Cl 15.2% Cl 17.0% Cl 18.8% Cl 21.2%

Current density (A/cm2)

1.0E-03 1.0E-04 1.0E-05 1.0E-06 1.0E-07 1.0E-08 1.0E-09 1.0E-10 -0.2

0.2 0.4 Potential (V vs SCE)

0.6

0.8

Figure 29. Potential dynamic scans of SS 2205 in saturated Ca(OH)2 solution.

46

Figure 30. Corrosion performance of sandblasted and non-sandblasted carbon steel at 1.5% of chlorides

Figure 31. Corrosion performance of sandblasted and non-sandblasted carbon steel at 3.0% of chlorides

47

Figure 32. Corrosion performance of sandblasted and non-sandblasted Cr steel at 1.5% of chlorides

Figure 33. Corrosion performance of sandblasted and non-sandblasted Cr steel at 3.0% of chlorides

48

Figure 34. Corrosion performance of sandblasted and non-sandblasted Cr steel at 4.5% of chlorides

Figure 35. Corrosion performance of sandblasted and non-sandblasted Cr steel at 6.0% of chlorides

49

Das könnte Ihnen auch gefallen