REPRODUCIBILITY OF CRITICAL CHLORIDE THRESHOLD LEVELS FOR STAINLESS STEEL REINFORCEMENT

Similar documents
Threshold Chloride Concentration of Stainless Steels in Simulated Concrete Pore Solution

Corrosion Properties of Enhanced Duplex Steel UNS S32304

IN SITU MONITORING OF REINFORCEMENT CORROSION BY MEANS OF ELECTROCHEMICAL METHODS

Galvanic Corrosion Between AISI304 Stainless Steel and Carbon Steel in Chloride Contaminated Mortars

EFFECT OF CHLORIDE CONCENTRATION IN SOIL ON REINFORCEMENT CORROSION

Galvanic coupling between carbon steel and stainless steel reinforcements. Qian, S.; Qu, D.; Coates, G. NRCC-48162

MEASUREMENT OF STEEL CORROSION IN CONCRETE

RAPID MACROCELL TESTS OF LDX 2101 STAINLESS STEEL BARS

Potentiodynamic Scanning (PDS) of Stainless Steel Karen Louise de Sousa Pesse

MMFX 2 (ASTM A 1035, GRADE 100) STEEL REBAR CORROSION PERFORMANCE TESTING IN ACCORDANCE WITH AASHTO MP 18M/MP MMFX Technologies Corporation

Advanced numerical design for economical cathodic protection for concrete structures

RAPID MACROCELL TESTS OF ENDURAMET 33, ENDURAMET 316LN, and ENDURAMET 2205 STAINLESS STEEL BARS

Re-alkalisation technology applied to corrosion damaged concrete

RAPID MACROCELL TESTS OF ENDURAMET 2304 STAINLESS STEEL BARS

Corrosion Behavior of Carbon Steel in Concrete Material Composed of Tin Slag Waste in Aqueous Chloride Solution

STAINLESS STEEL SELECTION FOR FLUE GAS DESULFURIZATION EQUIPMENT

The stainless steel appeal as alternative material for reinforcing bars to be applied in long lasting construction

RAPID MACROCELL TESTS OF ASTM A775, A615, and A1035 REINFORCING BARS

RAPID MACROCELL TESTS OF 2205 AND XM-28 REINFORCING BARS

A SIMPLIFIED METHOD TO ESTIMATE CORROSION RATES A NEW APPROACH BASED ON INVESTIGATIONS OF MACROCELLS Method to estimate corrosion rates

MEASURING THE EFFECTIVENESS OF MIGRATING CORROSION INHIBITORS (MCIs) BY ELECTROCHEMICAL TECHNIQUES

Corrosion Rate Measurement on C-Steel

Electrochemical study on corrosion of different metals exposed to deaerated geothermal brines containing CO 2

SANDVIK SPRINGFLEX SF SPRING WIRE WIRE

Galvanic corrosion evaluation of 6061 aluminum coupled to CVD coated stainless steel Elizabeth Sikora and Barbara Shaw 6/9/2016

Improvement of corrosion resistance of HVOF thermal sprayed coatings by gas shroud

Threshold Chloride Levels for Localized Carbon Steel Corrosion in Simulated Concrete Pore Solutions Using Coupled Multielectrode Array Sensors

Probabilistic Estimation of Corrosion Propagation Period for Prestressed Concrete Structures Exposed to Chlorides

Pitting susceptibility of AISI 304 stainless steel after cold rolling Olandir V. Correa a, Mara C.L. de Oliveira b, Renato A. Antunes c.

EFFECT OF SACRIFICIAL ZINC ANODES ON CURRENT DENSITY FOR EFFECTIVE CATHODIC PROTECTION OF RCC

EVALUATING MITIGATION METHODS FOR CHLORIDE INDUCED CORROSION OF REINFORCED CONCRETE STRUCTURES

CLEANING AND PREVENTION OF INORGANIC DEPOSITS IN PLATE HEAT EXCHANGERS USING PULSATING CURRENT

Insititue of Paper Science and Technology Central Files CORROSION CONTROL ANNUAL RESEARCH REVIEW HANDOUT BOOK

The Effect of Calcium Nitrite on the Electrochemical Characterization of 3003 Aluminum Alloys in Sea Water. H.Mogawer and R.Brown

Chemical characteristics The chemical composition is shown in Table 1. Outokumpu EN ASTM

Ulf NürnbergerN. Contribution of the group Metallic

A Study of Performance of Aluminium Anode by Addition of Zinc in Sea Water

STRESS CORROSION CRACKING OF STAINLESS STEELS IN HIGH PRESSURE ALKALINE ELECTROLYSERS

Factors Influencing Materials Selection in Condensing Economizers

Analysis of the Rapid Chloride Migration test

Corrosion Resistance of Alternative Reinforcing Bars: An Accelerated Test

Laboratory Experiments in Corrosion Engineering II

INITIAL EVALUATION OF CORROSION PERFORMANCE AND ALKALI REACTIVITY OF CERAMIC-COATED REINFORCING BARS

Erosion corrosion of mild steel in hot caustic. Part II: The evect of acid cleaning

ROLES OF PASSIVATION AND GALVANIC EFFECTS IN LOCALIZED CO 2 CORROSION OF MILD STEEL

Sensitization & Corrosion Behaviour of Austenitic Stainless Steel 304 & 316

A CORROSION MANAGEMENT AND APPLICATIONS ENGINEERING MAGAZINE FROM OUTOKUMPU 3-4/2013. The two phased optimization of duplex stainless steel

Application Note CORR-4

The Evaluation of Corrosion Behavior of AISI 347 Stainless Steel to ASTM A335 Low Alloy Steel Dissimilar Welds

SANDVIK 3RE60 TUBE AND PIPE, SEAMLESS

Corrosion Mechanisms, Monitoring, and Service Life Enhancement of Reinforced Concrete Structures

Corrosion and Passivation Behaviour of Three Stainless Steels in Differents Chloride Concentrations

IPC TECHNICAL PAPER SERIES NUMBER 74 LOCALIZED CORROSION OF STAINLESS STEELS IN PAPER MACHINE WHITE WATER DAVID F. BOWERS FEBRUARY, 1979

INVESTIGATION OF THE CORROSION PROPAGATION CHARACTERISTICS OF NEW METALLIC REINFORCING BARS

LASER SURFACE MELTING OF 17-4 PH PRECIPITATION-HARDENABLE STAINLESS STEEL Paper 1203

SCC Initiation for X80 Pipeline Steel Under AC Application in High ph Solution

13 Practical and Economic Aspects of Application of Austenitic Stainless Steel, AISI 316, as Reinforcement in Concrete

Steels suitable for galvanizing

CHLORIDE INDUCED WATER CONDUCTING POTENTIODYNAMIC INVESTIGATIONS

Cathodic protection in reinforced concrete structures affected by macrocell corrosion: a discussion about the significance of the protection criteria

Electrochemical behavior of stainless steel UR45N and carbon steel in solutions simulating the concrete interstitial electrolyte

DURABILITY of CONCRETE STRUCTURES

Testing of Randolph Street Bridge

1. CORROSION OF REINFORCEMENT

SIGNIFICANCE OF HYDROGEN EVOLUTION IN CATHODIC PROTECTION OF STEEL IN SEAWATER

Comparison of Inhibitors MCI. Carbonation- Induced Corrosion

Corrosion Monitoring of Steel Bars in Port Concrete Structures

Pitting Corrosion of Some Stainless Steel Alloys Preoxidized at Different Conditions

Corrosion Effects of Cr and Ni in Thermo-Mechanical Treated Steel Bar in Marine Environments

GERMANN INSTRUMENTS. GalvaPulse. for Corrosion Rate, Half-Cell Potentials and Electrical Resistance

Pitting corrosion of type 304 stainless steel and sulfate inhibition effect in chloride containing environments

The Electrochemical Behavior of Carbon Steel Wires of Nepal in Different Environments

Electrochemical studies of Aluminium 6061 / Titanium Diboride Metal matrix Composites

Corrosion Mitigation Systems for Concrete Structures

Effect of Heat Treatment on the Corrosion Behaviour of GTM- SU-718 Superalloy in NaCl, HCl and H 2 SO 4 environments

AN OVERVIEW OF THE USE OF GALVANIZED REINFORCEMENT IN CONCRETE CONSTRUCTION

Crack Initiation and Crack Propagation of Pre-corroded Ni-16Cr Alloy in 4.5%NaCl Aqueous Solution

FINAL REPORT. Gerardo G. Clemeña, Ph.D. Research Scientist Virginia Transportation Research Council

CHAPTER 2 MATERIALS AND METHODS

Causes and Remediation of Corrosion Failure of Duplex Stainless Steel Equipment in a PVC Plant

Sandvik SAF 2205 (Billets)

SULPHIDE STRESS CORROSION CRACKING OF WELDED PLATES

Available online at ScienceDirect. Procedia Engineering 184 (2017 )

DRAFT EVALUATION OF CORROSION RESISTANCE OF MICROALLOYED REINFORCING STEEL

P-6. Optimization of Citric Acid Passivation for Corrosion Resistance of Stainless Steel 440C. Chi Tat Kwok 1,2 and Wen He Ding 1.

Durability of Marine Concrete with Mineral Admixture and Marine Aquatic Organism Layer

Tutorial Corrosion II. Electrochemical characterization with EC-Lab techniques

Cathodic Protection of Steel in Concrete Using Conductive Polymer Overlays

CHAPTER 7 STUDIES ON CORROSION RESISTANCE OF STAINLESS STEEL CLADDINGS

Effect of silica fume on the corrosion performance of reinforcements in concrete

Effects of Corrosion on Reinforced Concrete Beams with Silica Fume and Polypropylene Fibre

A REVIEW OF EXPERIENCES WITH AL-6XN AND ZERON 100 IN AIR POLLUTION CONTROL EQUIPMENT. *Devin M. Wachowiak and Jason D. Wilson

THE USE OF GALVANOSTATIC PULSE MEASUREMENTS TO DETERMINE CORROSION PARAMETERS Galvanostatic Pulse Measurement of Corrosion

DURABILITY AND SERVICE LIFE PREDICTION OF GFRP FOR CONCRETE REINFORCEMENT

THE ELECTROCHEMISTRY OF CORROSION Edited by Gareth Hinds from the original work of J G N Thomas

The influence of Mg 17 Al 12 phase volume fraction on the corrosion behaviour of AZ91 magnesium alloy. Andrzej Kiełbus* and Grzegorz Moskal

CORROSION MONITORING OF PORT STRUCTURES

C Si Mn Ni Cr Cu Mo N P S

SANDVIK SPRINGFLEX STRIP STEEL

ONLINE MONITORING OF CORROSION UNDER CATHODIC PROTECTION CONDITIONS UTILIZING COUPLED MULTIELECTRODE SENSORS

Transcription:

REPRODUCIBILITY OF CRITICAL CHLORIDE THRESHOLD LEVELS FOR STAINLESS STEEL REINFORCEMENT S. Randström, M. Almén, R. Pettersson Outokumpu Stainless AB Avesta Research Centre SE-774 22 Avesta Sweden sara.randstrom@outokumpu.com M. Adair Outokumpu Stainless Ltd. ASR Rod Mill Sheffield S9 3XG United Kingdom KEY-WORDS: Stainless steel reinforcement, Critical Chloride Threshold Level, Chlorides ABSTRACT Corrosion resistant reinforcement is crucial in order to reduce the repair cost of corroding concrete structures. In the hunt for lower cost and high performance reinforcement, a measure of the chloride tolerance is needed to select the proper material based on the surface chloride content, the concrete type and the surrounding temperature. Critical Chloride Threshold Level (CCTL) measurements have been performed over the past 40 years to investigate the chloride tolerance of different reinforcement materials, but a standardized method that gives a fast and reliable response still does not exist. In this article, a potentiostatic in-solution method, which has been used by other researchers, is used for determining the CCTL of different stainless steel grades. The reproducibility of the method and the influence of the condition of the stainless steel reinforcement on the CCTL are investigated. The results indicate that there is some scatter in the results also for stainless steels and it is better to define the CCTL as a probability distribution based on a number of replicates for each stainless steel grade instead of a discrete number. INTRODUCTION Stainless steel reinforcement is a maintenance-free solution for corroding concrete structures due to the superior corrosion resistance of stainless steels compared to carbon steels or other metallic reinforcing bars. The stainless steel reinforcement is used selectively, with only the outermost layers of the carbon steel reinforcement replaced with the more corrosion-resistant reinforcement. This makes the investment attractive, with only a minor increase in the initial cost, and a much lower maintenance cost during the predicted life-time. Determination of the corrosion resistance of steels in concrete is very different to corrosion testing for many other applications. The diffusion rate of chlorides is comparatively slow, the presence of the sponge-like concrete affects transport rates of both gases and liquids and the alkaline environment increases the chloride tolerance, so long as the ph does not drop due to carbonation. As a result, accelerated testing is needed, where some parameters are assumed to either a) not affect the corrosion resistance or b) affect the corrosion resistance to a controlled degree. Over the years, the corrosion resistance of the stainless steels rebars has been determined by many different methods. Mainly two different ways to determine the corrosion resistance have been used, castin testing and testing in the concrete pore-solution. An extended discussion of these methods can be found in an earlier paper (Randström and Adair 2009). Cast-in testing, which resembles the real application most, can be advantageous since the method enables that chlorides slowly diffuse into the concrete structure, as they do in a real structure. However, using this method, the testing times are often some years or even longer, and the exact chloride level at the position where the corrosion started can sometimes be difficult to determine. The onset of corrosion on stainless steels can also be difficult to

detect when the rebar is cast-in (Hurley and Scully, 2006). Frequently, the use of Linear Polarisation Resistance (LPR) is used to determine whether the stainless steel is active or still passive. This method, requires a significant part of the material to be active, and is therefore better to use on low-alloyed steel where chlorides cause activation of a larger surface than on stainless steel rebars, where chlorides cause localized corrosion. In addition, the corrosion potential is often measured to determine whether corrosion has occurred. The corrosion potential measurement is an easy, non-destructive method to detect possible corrosion in structures and bridges made of carbon steel. According to the standard ASTM C876,which is a standard for half-cell potential measurements of reinforcement in concrete, the probability that the carbon steel reinforcement is corroding if the potential is below -280 mv SCE (mv versus the standard calomel electrode) is over 90%. However, if the oxygen level is low, the above mentioned potential can be measured without any corrosion occurring. For laboratory-experiments, where shorter times are investigated, the oxygen level can be low due to completely water-filled pores, so it is strongly recommended that corrosion is confirmed by visual inspection and/or weight loss measurements. For stainless steels a recent report (Sederholm and Almqvist, 2008) showed that cast-in stainless steels can have potentials more negative than -280 mv SCE for several months without showing any signs of corrosion by visual inspection and weight loss measurements. For the purpose of determining chloride levels and for ranking corrosion resistance, a potentiostatic test in synthetic pore-solution can be more useful. In this case, a solution similar to that found in the pores of the concrete is used. The composition of such pore-solutions is mainly based on hydroxides of sodium, potassium and calcium (Bertolini et al., 2004). Although these tests do not take the solid concrete matrix into account, they can predict the corrosion resistance and the chloride tolerance in this alkaline environment in a fast and efficient way. A potentiostatic method has proved useful to determine the chloride tolerance since onset of the local corrosion attack is immediately detected by the increase in current (Bertolini et al., 1996), (Hurley and Scully, 2006). In this method, the rebar specimen is polarized to a fixed potential (often +200 mv SCE ) in a chloride-free synthetic pore-solution. The chlorides are thereafter added incrementally until corrosion occurs. A corrosion test using this method takes around two weeks to perform and different solutions can be used. EXPERIMENTS Specimens Stainless steel rebars with diameters ranging from 8-25 mm were used in four different stainless steel grades, with both austenitic and duplex microstructures. The nominal compositions of these grades are seen in Table 1 and the rebars were all commercially produced by Outokumpu. The rebars were cut in 10 cm lengths and a hole (4 mm) for electrical contact was drilled in one of the ends of the rebar. For investigation of the surface effect, some of the rebars were first sand-blasted with aluminium silicate. All rebars were thereafter ultrasonically cleaned in acetone for 20 min, and covered with Lacomit (Agar Scientific) in both ends.

Table 1 Nominal composition of the stainless steel grades used Steel grade EN ASTM Microstructure Cr Ni Mo C 304L 1.4307 304L Austenitic 18.1 8.1-0.02 316L 1.4436 316L Austenitic 16.9 10.7 2.6 0.02 LDX 2101 1.4162 S32101 Duplex 21.5 1.5 0.3 0.03 2304 1.4362 S32304 Duplex 23 4.8 0.3 0.02 Method The prepared rebars were hung in a QuickFit beaker and a ring counter-electrode of platinum was placed around it. A Luggin capillary was placed close to the hanging rebar and connected to a standard calomel electrode (SCE). All potentials will hereafter be referred to this reference electrode. 1 L of pore solution, with a composition of 2.61 g/l NaOH, 9.04 g/l KOH and 0.17 g/l Ca(OH) 2 (hereafter referred to as OPC solution) was poured into the beaker. The rebar was adjusted so that the water line was at the Lacomit-covered area. Air, scrubbed through a wash bottle containing 3g/L Ca(OH) 2 in water was bubbled through the solution to promote stirring and saturate the electrolyte with oxygen. The setup is seen in Figure 1. Counter Electrode Air bubbling Luggin cappillary Rebar Specimen Figure 1Schematic view of the cell setup The open circuit potential (OCP) was measured for 1.5 h before the experiment started (OCP part), and the potential was thereafter increased from 25 mv below the open circuit potential to the final potential,+200 mv SCE with a scan rate of 10 mv min -1 (Linear Polarisation part). The potential was thereafter held at this potential throughout the experiment (Chronoamperometric part). Chlorides were added after at least 10 h had passed at this potential. This potential has been selected by other research groups and is believed to give conservative results. It corresponds to an overpotential of more than

400 mv, since open circuit potentials in this alkaline environment normally are around -200 mv SCE (Hurley and Scully, 2006). Chloride additions were made by replacing a certain amount of the synthetic pore solution with a solution of the same composition but with a sodium chloride concentration of 5 M. These additions were made every 12 th hour. The volume replaced was selected so that the chloride concentration in the electrolyte was increased by 0.1 M every chloride addition. The CCTL was defined as the chloride content at which the current density exceeded 10 µa cm -2 for 5 h. When CCTL was reached, a sample of the electrolyte was transferred to an air-tight bottle to measure ph and chlorides. ph was measured after the experiment with a ph meter (Jenway 3345 Ion meter equipped with a Schott ph electrode - Blue line) or a combined ph /chloride measurement unit (Mettler Toledo SevenMulti S80 equipped with an InLab Routine Pro ph-electrode and a DX235-Cl/InLab Reference electrode for chloride measurements). RESULTS AND DISCUSSION An example of the OCP and chronoamperometric part of the electrochemical experiment is seen in Figure 2. The increase in current density in this case started within 2 hours after the last addition of chlorides, resulting in a chloride concentration of 6.4 wt% which was determined as the CCTL. The very long incubation time of the corrosion supports the theory that polarization curves may be a too fast method to determine CCTL, a discussion earlier brought up by (Hurley and Scully, 2006). In their study, the CCTL determined by a potentiodynamic method was higher compared to when a potentiostatic method with incremental chloride additions was used. It is therefore important to note what kind of method that has been used to determine the CCTL when data is compared. Figure 2 Left: Open circuit potential (corrosion potential) of a rebar specimen in this experiment. Right: Example of curve during the chronoamperometric part of the experiment, from the first chloride addition. The graph illustrates a 316 rebar specimen. The time which chloride additions were made are marked with red diamond markers. In the test, the corrosion at the stainless steel rebar typically starts as a spot where corrosion products are deposited, as seen in Figure 3. The corrosion propagates and a larger area is covered with corrosion products, as seen in the right part of Figure 3. The propagation is sometimes very slow, and the corrosion at the spot can stop so the current decreases to a low value. Therefore, the CCTL was defined as the chloride concentration where propagating corrosion occurred.

Figure 3 The onset of corrosion attack on a stainless steel rebar when using the potentiostatic method. The right image is taken approximately 30 minutes after the left image. Note that a high potential was applied to the specimen, forcing the corrosion to occur rapidly The CCTL values determined by this potentiostatic method for the four different grades are seen in Figure 4. The figure shows test results from commercially produced rebars, and the corrosion resistance shown is therefore representative for stainless steel rebars that are delivered to the customer. Figure 4 CCTL values of the different stainless steel grades The results indicate that there is no significant difference in corrosion resistance between 316L, LDX 2101 and 2304, while 304L has a slightly lower CCTL. The results correspond fairly well with those from other researchers who have used a similar potentiostatic method, as can be seen in Table 2. The largest deviation compared to other authors is seen for LDX 2101, which clearly has a corrosion resistance comparable to 316L in this study, but was reported to have a corrosion resistance lower than 316L in the study by (Hurley and Scully, 2006) and (Bertolini et al., 2009). In the case of (Hurley and Scully, 2006), the LDX 2101 it was stated that the rebar was laboratory pickled, which could be the reason for the difference in corrosion resistance.

Table 2 CCTL values in wt% chlorides for the investigated stainless steel rebar grades. Comparisons with other groups using a similar method have also been added, with additional data of low-chromium alloys not included in this study. Numbers in brackets represent the numbers of rebar tested. This investigation Bertolini et al. 1996 Bertolini et al. 2009 Hurley and Scully 2006 Stainless steel Commercially Not mentioned Commercially Commercially product produced and [No. of specimens produced and produced rebar, pickled rebar tested not stated] pickled rebar laboratory pickled (Outokumpu) ph (after test) 12.8-13.3 12.6 12.6 12.6 Carbon steel <0.4 [3] 0.1-0.6-0.04-0.05 [3] Fe-9%Cr - - - 0.4-0.6 [3] 410-2 - - 304L 2.8-10.6 [7] 5.0 6.5-10 (304LN) [2] 316L 5.7-6.8 [6] 5.5 >10 (316LN) [2] 2.8-11.29 (316LN) [9] LDX 2101 -type 5.7-10.6 [9] - 3.5-6 [2] 1.1-1.4 [3] 2304 4.6-10.6 [5] 10 7.5-8 [2] - The pickling time has an effect on the corrosion resistance in potentiostatic tests. The chloride tolerance of three different surface types for LDX 2101, seen in Figure 5 was investigated and the result can be seen in Figure 6. The over-pickled surface, shown in the left image in Figure 5, had many micro-crevices in the outermost layers of the stainless steel rebar and clearly has a lower corrosion resistance compared to normally pickled LDX 2101. These types of micro-crevices are typical when stainless steels are overpickled particularly if they contain precipitates which increase the pickling rate at the grain boundaries. Grain boundary precipitates are introduced by faulty heat-treatments or when carbon contents are too high. This could explain the low corrosion resistance of the 2101-type stainless rebar tested by (Bertolini et al. 2009), since the carbon content was 0.049, well above the normal carbon content for LDX 2101. It can therefore be stated that the pickling procedure and melt composition is critical in order to obtain satisfactory performance of the material. It is moreover important to consider the surface condition when rebars in different grades are compared, and it is vital for a producer to deliver material with good surfaces with optimal corrosion resistance. A sand-blasted surface, included in Figure 5 and Figure 6, was found to have the same CCTL as a normally pickled surface. If such an operation is used, it should be verified that the surface after treatment is reasonably smooth without embedded particles. As with all sand-blasting of stainless steels, it is also critical to ensure that the blasting sand has not been contaminated by previous use on carbon steel. Such contaminated sand can very quickly lead to rust speckling on the stainless steel because of embedded carbon steel particles.

Figure 5 Surface of over-pickled (left), normally pickled (middle) and sand-blasted rebar (right). Figure 6 Effect of surface modification (over-pickling and sand-blasting) of LDX 2101 The scatter seen in the results in Figure 4 as well as in the study of (Hurley and Scully, 2006), brings up the discussion whether it is correct to define the CCTL value as a discrete number or if it is better to define it as a probability that the stainless steel rebar corrodes. If it is assumed that the data values are normally distributed, a probability plot, can be made as shown in Figure 7. This plot emphasizes a number of the points made qualitatively earlier in the paper. Firstly it is seen that at the 90% probability level, indicated with a grey circle, there is good correspondence between the CCTL levels for 316L, LDX 2101 and 2304, while that for 304L is significantly lower. Secondly the slope of the curves, which is related to the standard deviation in the measurements, is similar for 304L, LDX 2101 and 2304, but somewhat lower for 316L. This type of plot is very common for e.g. life-time predictions and is strongly recommended as an assessment tool for CCTL data because of the scatter seen in the measurements.

Figure 7 Probability graph assuming that the CCTL results are normally distributed A chloride threshold based on the concept of a statistical distribution has earlier been introduced for carbon steels cast in concrete (Glass 1995), (Browne 1982), and mentioned in Concrete Society s Technical Report 61 (Bamforth, 2004) suggesting that corrosion of carbon steel should be considered as an increased probability for corrosion rather than a discrete value. The probability classification proposed by (Browne 1982) is shown in Table 3. Table 3 Probability classification of carbon steel rebar seen in Concrete Society Technical Report 61 (Bamforth 2004), proposed by (Browne 1982) Chloride (% weight of cement) Risk of corrosion <0.4 Negligible 0.4-1.0 Possible 1.0-2.0 Probable >2.0 Certain The threshold limit for carbon steel differs between regions. In the USA, the value of 0.2% chlorides per mass of cement is often regarded as the limit, while 0.4% is often mentioned as the limit in Europe. This seems to correspond to a probability that the reinforcement is not corroding of 90% and 80%, respectively (Bamforth, 2004). Using these probability percentage for the stainless steel reinforcement investigated in this study, especially for LDX 2101, that has been tested nine times; the difference in CCTL using these to criteria would be around 1wt%. Although the difference is not very large, it is recommended that a 90% probability criterion is used, since this is a more conservative result, and since it underlines the need for a large number of tests to determine eventual differences between stainless steel grades.

Compared to cast-in rebars, the environment (ph, temperature, conductivity, chloride and oxygen levels) is more strictly controlled in the pore-solution test. The scatter of the results therefore indicates that it is related to the variability in corrosion initiation and propagation processes. Such variability could explain the scatter of data both seen in cast-in experiments (Alonso et al, 2006) and other pore-solution tests (Hurley and Scully, 2006) when a significant number of specimen have been investigated. The potentiostatic method used in this article differs from the real situation in the concrete structure regarding transport rates of oxygen and chlorides but nevertheless gives an idea what to expect in terms of corrosion resistance. The applied potential of +200 mv SCE in the potentiostatic tests corresponds to a large overpotential of around 400 mv, which is believed to give conservative results. On the other hand, the solid concrete matrix in cast-in tests creates crevices, which could decrease the corrosion resistance. Recently, (Bertolini et al., 2009) suggested that potentiostatic tests were misleading since they lacked correlation to cast-in tests. However, this conclusion was based on two measurements in pore-solution and corrosion potentials plus LPR measurements of sometimes only single specimens in concrete blocks with different chloride levels. Of course, it is difficult to predict what will happen in a real application, and there will likely be some discrepancies between pore-solution tests, cast-in tests and the real application, as there is for all comparisons between lab-scale testing and the industrial application. Nevertheless, for the understanding of the chloride tolerance of stainless steel in this application it is very important to start with a few controlled parameters and thereafter move on by adding more unknown parameters, with control of the scatter and limitations of the method used at each step. CONCLUSIONS In this study, stainless steel rebars in four different grades have been tested. At least five different specimens were tested for each grade, to allow assessment of the scatter of the method. The results indicate that 316L, LDX 2101, and 2304 have similar corrosion resistance while 304L has slightly lower corrosion resistance. The results are generally in agreement with what other researchers have found except for the fact that this study shows that LDX 2101 has similar corrosion resistance to 316L. It has also been demonstrated that over-pickling of the LDX 2101 material decreases the corrosion resistance, which could explain the lower corrosion resistance reported for this grade. The scatter of the results is quite large even though the parameters are carefully controlled in this method. There is therefore a risk that making only one or two measurements can give completely misleading results. Even when the surface is normalized by sand-blasting, the scatter seems to remain. Therefore, it is much more relevant to describe the chloride tolerance of stainless steel rebars as a probability distribution rather than as a discrete, critical value. This is in line with the approach used for carbon steel that has been cast in concrete. It would appear that the value of 0.2% chloride by mass of cement used for carbon steel in the USA would correspond to 90% probability and the European figure of 0.4% chlorides correspond to 80% probability. The analysis of this research would indicate that a value of 90% probability should be used for stainless steel rebar. REFERENCES Alonso C., Castellote M., Andrade C. (2002) Chloride Threshold dependence of pitting potential of reinforcements, Electrochim. Acta, 47, 3469-3481 Bamforth P.B. (2004) Enhancing reinforced concrete durability, Concrete Society Technical Report 61, The Concrete Society, Surrey

Bertolini L, Bolzoni F, Pastore T, Pedeferri P. (1996) Behaviour of stainless steel in simulated concrete, Br. Corr. J. 31, 218-222. Bertolini, L., Elsener, B., Pedeferri, P., Polder, R. (2004) Corrosion of steel in concrete Wiley- VCH, Weinheim Bertolini L., Gastaldi M. Corrosion resistance of austenitic and low-nickel duplex stainless steel bars, Eurocorr 2009, 6-9 September 2009, Nice Browne R.D. (1982) Design prediction of the life of reinforced concrete in marine and other chloride environments, Durability of Building Materials, Vol. 3, Elsevier Scientific, Amsterdam Glass GK (1995) The determination of chloride binding relationships, International RILEM Workshop on Chloride penetration into concrete St-Remy-Les-Chevreuse, October 15-18 1995 Eds. Nilsson L.O. and Olivier, J.P. Hurley MF, Scully JR (2006) Threshold Chloride Concentrations of Selected Corrosion-Resistant Rebar Materials Compared to Carbon Steel, Corr. Sci. 62, 892-904. Randström S., Adair M. (2009) Critical Chloride threshold levels for stainless steel reinforcement in pore solutions, Paper no. 046, Corrosion and Prevention 2009, Coffs Harbour, 15-19 November 2009, Australia Sederholm, B., Almqvist, J. (2008) Stainless steel in concrete Galvanic effects on carbon steel, Report no. KIMAB 2008-133, Swerea KIMAB, Stockholm