You are on page 1of 27

Journal of ASTM International, Vol. 5, No. 2 Paper ID JAI101241 Available online at www.astm.

org

Gerald S. Frankel1

Electrochemical Techniques in Corrosion: Status, Limitations, and Needs


ABSTRACT: The corrosion of metals occurs primarily by electrochemical processes involving metal oxidation and simultaneous reduction of some other species. The fundamental understanding of these processes has allowed the development of a number of electrochemical techniques for the study of the corrosion phenomena and assessment of the corrosion rate. In fact, electrochemical techniques are so ingrained in the eld that many practitioners think of corrosion rates rst in terms of current density rather than thickness or mass loss per unit time. Standard approaches for electrochemical corrosion rate determination are commonly used in the eld for on-line monitoring of systems and facilities. Electrochemistry also provides powerful tools for developing fundamental understanding of corrosion phenomena. However, there are some limitations to the abilities of current electrochemical techniques and some needs for the future. This paper describes the status of electrochemical techniques, their limitations, where nonelectrochemical methods are required, and future needs in the eld. KEYWORDS: electrochemical tests, corrosion

Introduction As a result of the development of the fundamental understanding of corrosion electrochemistry, fast and accurate potentiostats, and computer technology, a suite of electrochemical techniques exists for the study of corrosion. These techniques provide the technologist with the ability to monitor corrosion rates in service, giving early warning of conditions that could adversely affect performance and integrity. They also provide the experimentalist with the ability to determine corrosion rate with high sensitivity, assess rate controlling mechanisms, and in some cases make life predictions. Furthermore, variations in electrochemical techniques and innovative cell designs allow researchers to probe mechanisms and develop new and improved materials. A number of excellent reviews describe the existing electrochemical techniques in detail and provide instructions on their proper use 15 . It is not the purpose of this paper to present a comprehensive summary of electrochemical methods. Interested readers are referred to other works. Instead, the focus will be on the limitations of available techniques for assessing corrosion and the need for improved methods. A brief summary of the available test methods will be given, including different types of exposure and some experimental design considerations. A generic view of corrosion problems will then be presented, using a rubric that includes the exposure of uncoated or coated metals in bulk solutions or under atmospheric conditions. The status, limitations, and needs of the available methods will be discussed in the context of that matrix. The focus will be on electrochemical methods, but a variety of non-electrochemical methods and mixed methods will be addressed. Owing to the limitations of electrochemistry, it is critical to utilize a variety of approaches when studying complicated systems. A number of important corrosion phenomena will not be addressed in this review. Stress effects and environmentally assisted cracking as well as oxidation and high temperature corrosion will not be covered. However, the electrochemistry of stress corrosion cracking is critical and still not understood. Furthermore, some high temperature corrosion topics such as hot corrosion are of immense concern in power generation applications and are also not well understood.
Manuscript received May 17, 2007; accepted for publication January 17, 2008; published online February 2008. Presented at ASTM Symposium on Advances in Electrochemical Techniques for Corrosion Monitoring and Measurement on 2223 May 2007 in Norfolk, VA; S. Papavinasam, N. Berke, and S. Brossia, Guest Editors. 1 Fontana Corrosion Center, The Ohio State University, Columbus, OH 43210.
Copyright 2008 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.

2 JOURNAL OF ASTM INTERNATIONAL

Available Test Methods The applicability of various test methods depends on the exposure conditions. For the purposes of this review, two exposure conditions will be considered: immersion in solution and atmospheric exposure. Most full immersion environments are known in some detail a priori, such as seawater, process streams, or concrete pore solution. Atmospheric exposure conditions, in contrast, are much less well understood. The electrolyte responsible for atmospheric corrosion can be a thin layer in equilibrium with humid air and inuenced by atmospheric pollutants or surface contamination from sources such as road salt. In outdoor exposure, dew, and precipitation play a major role in determining the local environment. Time-of wetness, dened as the time with T 0 C and RH 80 %, has been shown to be a critical parameter for outdoor atmospheric corrosion 6 . However, wet layers can exist on contaminated surfaces at much lower humidities owing to deliquescence of the salts on the surface 7 . As a result of the complexity of the atmospheric corrosion environment, testing conditions are usually less connected to real exposures. The most common atmospheric exposure test is ASTM B117 Standard Practice for Operating Salt Spray Fog Testing Apparatus, 8 . However, the relevance of salt spray test performance to real environments, including marine environments, has been discussed widely. In fact, the Introduction section of the B117 standard states 8 , Prediction of performance in natural environments has seldom been correlated with salt spray results when used as stand alone data. Correlation and extrapolation of corrosion performance based on exposure to the test environment provided by this practice are not always predictable. Correlation and extrapolation should be considered only in cases where appropriate corroborating long-term atmospheric exposures have been conducted. Various cyclic tests have been suggested to be more representative 9 . Microelectronic devices are often tested under conditions of high temperature and humidity T/H with the added application of a bias 10 . The prediction of lifetime in service from lifetime in more aggressive test conditions requires knowledge of acceleration factors. The acceleration factors for microelectronic devices in T/H testing have been studied in some detail 10 , but acceleration factors for most corrosion tests relative to atmospheric exposures are not known, although various correlations have been made 9 . More comments on this situation will be provided below. Non-Electrochemical Measurements A number of non-electrochemical measurement techniques can be used to assess corrosion rate. Weight loss measurement, considered by some to be the gold standard of corrosion testing is certainly the easiest. However, there are important issues to consider even for weight loss measurements. First, since mass can be measured easily only to about 0.1 mg, the sensitivity of weight loss measurements is limited. Other issues include end-grain attack leading to different corrosion rates on different exposed faces, crevice corrosion associated with hanging or supporting the sample, and waterline attack if the sample extends beyond the surface. Finally, weight loss measurements are usually performed after long exposure times so they provide an average rate over time as well as over the exposed surface. It should be noted that the quartz crystal microbalance can provide rapid and very sensitive submonolayer measurements of weight loss for thin lm samples 1 . Corrosion rate also can be determined by dimensional changes of the exposed samples or analysis of the solution for dissolved species. Both of these approaches also have sensitivity limitations similar to weight loss. However, solution analysis can be particularly powerful for studying the corrosion of alloys because of the ability for chemical differentiation, which is only possible using electrochemistry by rotating ring-disk measurements. A technique that has had more application in corrosion rate monitoring than in corrosion science involves the change in electrical resistance ER of a probe sample. The reduction of the cross-sectional area of a probe by corrosion is accompanied by a proportionate increase in the electrical resistance, which can be tracked easily. A major advantage of the ER technique is its applicability to a wide range of corrosive conditions including environments having poor conductivity or non-continuous electrolytes such as vapors and gases. However, ER monitoring typically requires a relatively long exposure period for a detectible difference in probe resistance and electrically conductive deposits can affect the measurements. Thierry et al. compared the use of weight-loss coupons, ER, and polarization resistance measurements in cooling water applications 11 . They found that, with the increasing extent of localized corrosion, the ER measurements indicated corrosion rates in excess of those observed on the coupon. In addition, changes in

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 3

corrosivity of the water systems could be detected within a few hours by polarization resistance techniques, whereas the ER technique required a few days to measure the changes. Electrochemical Measurements As mentioned above, it is not within the scope of this paper to describe in detail the wide range of electrochemical methods available for studying corrosion. In this section, the primary electrochemical methods used for determining corrosion rate are addressed broadly. Potentiodynamic polarization PDP over a potential range about 200 250 mV from the open circuit potential OCP results in a polarization curve that can be analyzed for corrosion rate, provided that the rates of other anodic reactions such as those associated with redox reactions are small in comparison which is a requirement of all electrochemical assessments of corrosion rate . Typically presented in a semi-logarithmic plot, polarization curves provide corrosion rate by extrapolation of the linear cathodic and/or anodic regions to the corrosion potential or by tting to the following equation 1,5,12 : inet = icorr exp 2.3 E Ecorr ba exp 2.3 E Ecorr bc 1

where inet is the current measured as a function of applied potential E, Ecorr is the corrosion potential, icorr is the corrosion current density, and ba and bc are the anodic and cathodic Tafel slopes, respectively. The equation represents an idealized form of the electrochemical data for the case of a mixed electrode in which there is only one anodic and one cathodic reaction taking place on the corroding surface. Both reactions must be controlled by activation polarization and Ecorr must be far from both reversible potentials. Most commercial corrosion analysis software packages contain the capability to t data to this equation, but it has no agreed-upon name, nor does the technique of tting data to this equation. It has been proposed to call this equation the Wagner-Traud equation 13 . Wagner and Traud, in their 1938 paper published in English in 2006 14 , came close to deriving this equation, but did not take the nal steps to do so. They certainly at that time could not have envisioned the power of computers to utilize this equation for determination of corrosion rates. PDP over a wide range of potential generates more information about the system than just the corrosion rate. For instance, information can be obtained about the proximity of the OCP to regions of passivity or localized corrosion susceptibility. It is often possible to view more of the anodic polarization curve by de-aerating the solution, which can reduce the corrosion potential. PDP is a tool for laboratory investigations, not corrosion rate monitoring, as it involves perturbation of the potential relatively far from the steady-state corrosion potential. The corrosion rate also can be determined from the polarization resistance RP using the Stearn-Geary equation provided that the polarization resistance is similar to the charge transfer resistance and if the Tafel slopes are known 2,4 . The most common way to determine RP is by the linear polarization resistance LPR method, in which the potential is scanned about 5 10 mV relative to the corrosion potential. The slope dE/di , at the zero current potential is a measure of RP. An easier approach is a two-point measurement at potentials above and below the OCP. Or even simpler, a single measurement can be made at a potential either above or below the OCP and the slope dE/di can be determined using the net current, potential point of 0, OCP since the i-E curve must go through this point. These simplied analyses assume that the polarization response is perfectly linear, and error will result if there is any deviation from linearity. The LPR method has been put to considerable use in corrosion monitoring as it involves relatively little potential perturbation. However, accurate assessment of corrosion rate requires knowledge of the Tafel slopes, which must be determined separately or assumed. On the other hand, calculated corrosion rates are usually not wrong by more than a factor of 23 if the Tafel slopes are both assumed to be 100 mV/ dec. The electrochemical impedance spectroscopy EIS technique involves the application of a timevarying voltage and measurement of the current response. The ratio of two gives the frequency-dependent impedance. Several books and papers have been written on EIS and its application to corrosion 1521 ; details will not be given here. Sufce it to say that the impedance represents a fuller description of the transfer function for the response of a system to a perturbation, relative to DC methods. The low frequency limit of the impedance magnitude can be related to RP and thus the corrosion rate using the Stearn-Geary equation. Again, the Tafel slopes are required to do so. Constant phase elements CPEs are used widely

4 JOURNAL OF ASTM INTERNATIONAL

FIG. 1The simplied Randles circuit with CPE. in the analysis of EIS corrosion data. The extra tting parameter associated with the non-ideal capacitance of a CPE improves the t to the data. The CPE is a mathematical construct of convenience. However, it is not surprising that a physical structure such as an electrochemical interface does not behave exactly like a combination of standard circuit elements, and no rationale need be given for the use of CPEs. The simplied Randles circuit with a CPE shown in Fig. 1 is commonly used to represent many corroding interfaces. EIS is a particularly useful technique for low conductivity electrolytes as the ohmic resistance is determined explicitly. It also provides a good description of the response of paint-coated samples and is sensitive to early stages of coating failure. One main difculty with the technique is the proper selection of an equivalent circuit. An equivalent circuit should always be based on a physical model of the corroding system; addition of circuit elements simply to improve the t is unacceptable. However, a number of complex circuits could be rationalized as the detailed nature of the physical system often is not known. There is broad agreement about the two time constant model Fig. 2 used to represent a defective coating, which is one of the main applications of EIS. On the other hand, good coatings typically exhibit one time constant during the early stages of exposure, and it often is not clear exactly when use of the two-timeconstant defective coating model should be implemented when evaluating a time series of spectra. The electrochemical noise EN technique involves the measurement of electrochemical events i.e., current or potential transients or both simultaneously produced by the corrosion process. EN has been reviewed by several investigators 2129 . The most common approach is to measure current noise utilizing a zero resistance ammeter of two identical electrodes shorted together and the potential noise between the pair and a reference electrode RE or a third identical electrode. The ratio of the root mean squared deviation of the potential and current uctuations is one measure of the noise resistance. Alternatively, the data can be transformed into the frequency domain to generate a power density spectrum or evaluated using wavelet analysis. One problem with EN is the proper approach for accounting for the exposed area A of the sample. Cottis has described how current noise amplitude is proportional to A1/2, whereas potential noise amplitude is proportional to A1/2 26 . He recommends simply reporting the area and not normalizing the noise resistance by area. However, to determine a corrosion rate, area normalization is required, so the usefulness of EN for determination of absolute corrosion rate is questionable. On the other hand, EN is particularly appealing for in situ monitoring as no applied perturbation is required. Even though the absolute corrosion rate cannot be obtained, the EN character is quite different for passive conditions low noise , metastable pitting random events of short duration , and stable pitting individualized events of longer duration so it can be useful for assessing the onset of localized corrosion or stress corrosion cracking for a stressed sample. Changes in conditions can be detected, triggering closer inspection or sampling of passive probes immersed in the environment. A multi-pronged strategy is being used to monitor the corrosion of a steel tank holding liquid radioactive waste at the Hanford Site 30 . A probe made from a thick-walled berglass pipe inserted into the tank holds multiple samples for measurements of LPR, ER, and EN at different positions within the tank. LPR is performed using two nominally identical electrodes with no reference electrode. EN is performed using three nominally identical electrodes, one being a pseudo-reference electrode. Samples with different conguration are tested, including a bullet-shaped electrode, two concentric rings, and stressed and unstressed C-ring samples for sensing stress corrosion cracking SCC . The probe also contains reference

FIG. 2Nested two-time-constant model used to represent defective coatings.

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 5

electrodes designed to monitor the potential of the tank since all of the rate monitoring will be done on special samples and not the tank. A second passive probe has been installed with stressed and unstressed samples that will be removed periodically for inspection. This approach to monitoring takes advantage of each of the multiple techniques utilized. Electrochemical frequency modulation EFM is the last electrochemical method for assessing corrosion rate to be discussed here. EFM uses harmonic analysis whereby a corroding system is perturbed by a potential waveform consisting of the combination of two sine waves of different frequencies 31 . A nonlinear system will respond to this excitation with a current that has components at harmonic and intermodulation frequencies. In practice, the method is very similar to an EIS measurement. A reference electrode and a counter-electrode CE as well as a computer-controlled potentiostat are needed to apply the waveform and measure the response. As with EIS, the magnitude of the excitation potential signal is on the order of 10 20 mV. The harmonic and intermodulation frequencies of the two excitation frequencies should not overlap and the frequencies should be low enough to avoid capacitive inuences. Assuming that both the anodic and cathodic reactions obey Tafel kinetics and taking the rst few terms of a Taylor series expansion of the exponential form of the current response, it can be shown that the intensities of the current response at certain harmonic and intermodulation frequencies create a set of equations with the variables being the corrosion rate and the anodic and cathodic Tafel slopes. These equations can be solved to determine values of these parameters. EFM directly generates a value of the corrosion rate unlike the LPR and EIS methods, which require knowledge of the Tafel slopes for the determination of the corrosion rate from the polarization resistance through the Stern-Geary equation. Furthermore, idealized responses contain components at various frequencies having intensities of xed ratios. Therefore, by comparing the ratios from real data with theoretical values it is possible to determined so-called causality factors, which are measures of the quality of the analysis. Data analysis methods exist to handle the cases of a both anodic and cathodic reactions being under activation control exhibiting Tafel kinetics , b passive systems, and c cathodic reaction under diffusion control. However, the data analysis requires that one of these conditions be preselected. Ku and Mansfeld recently took a critical look at the EFM technique 32 . They evaluated several different electrochemical systems with EFM as well as EIS and PDP. EFM agreed well with the other techniques when the frequency range was entirely within the low frequency DC limit of the impedance, which is usually easy for systems with high corrosion rates. When the EFM frequency range was within the capacitive regions of impedance, the estimated corrosion rate was higher than that predicted by the other techniques. Ku and Mansfeld recommended using EFM with caution 32 , but the same can be said of all electrochemical techniques. An interesting development in electrochemical testing over the past decade has been the miniaturization of corrosion cells by the use of microcapillaries 3336 . In this approach, a capillary pulled to a diameter typically from 1 100 m is used as the electrochemical cell. Note that considerable skill is required to fabricate and handle capillaries of diameter less than about 15 m. The capillary is attached to a holder that ts into an optical microscope objective carousel. The microscope allows for precise positioning of the end of the microcapillary on a particular spot on the surface. A silicone coating on the end of the microcapillary prevents leakage and the reference and counter-electrodes are positioned in a connected reservoir. This approach has been useful for studying the effects of inclusions such as MnS in steel 36 and intermetallic particles in Al alloys 34 . Essentially all of the electrochemical techniques described above can be used with this cell, with the advantage that the microcell geometry provides high spatial resolution. Experiment Design Considerations Several factors must be taken into account when designing an electrochemical experiment for the study of corrosion rate or mechanism. Decisions need to be made regarding sample selection, surface preparation, masking to expose a certain area, specics of the experimental cell, selection of the appropriate test environment, and choice of the best technique. A detailed discussion of these factors can be found elsewhere 1 and will not be covered here. However, it is important to note that the design of the experiment and the preparation of samples often are the most critical parts of electrochemical experimentation.

6 JOURNAL OF ASTM INTERNATIONAL

FIG. 3Potentiodynamic polarization curve for Fe in 0.5 M H2SO4. Solid line measured curve; dashed line t to Eq 1. Problems to Study and Approaches A wide range of problems exist in the eld of corrosion. The application of protective coatings is one of the main approaches to corrosion protection, and the issues associated with coated samples are quite different than for uncoated electrodes. Furthermore, there are two broadly different types of exposure conditions: immersion in an electrolyte or solution and atmospheric exposure. Therefore, corrosion problems can be classied by a matrix considering samples that are coated or uncoated and circumstances of immersion in solution or exposure to atmospheric conditions. Presently, the four possible combinations of this matrix will be considered in turn. For each, the unlled needs for electrochemical testing will be discussed. All the experiments were performed at room temperature. Metal in Solution For an uncoated metal immersed in an electrolyte, all of the electrochemical techniques described above are usually applicable. However, the choice of technique depends on the form of corrosion exhibited: uniform active dissolution, passivity, or localized corrosion. Special studies on the effects of stress, inhibitors or microbes, for instance, might require yet other approaches. The conditions of active dissolution, passivity, and localized corrosion will be discussed in turn. Uniform Active DissolutionIt is instructive to make a comparison between many of the techniques described above for a model system representing uniform active dissolution, Fe in 0.5 M H2SO4. Two nominally identical 99.9 % Fe electrodes of area 0.63 cm2 were immersed in 0.5 M H2SO4 along with a Pt CE and SCE RE. The following series of experiments was performed: 1 LPR on each electrode, 2 EN using the two electrodes, 3 another set of LPR measurements, and then on one of the electrodes 4 EFM, 5 EIS, and nally PDP over a relatively wide potential range. PDP was performed last as it alters the potential furthest from OCP. For a system such as Fe in sulfuric acid, all of the electrochemical techniques work quite well. The polarization curve generated by PDP is shown in Fig. 3. The linear portion of the cathodic branch extends for more than a decade of current density allowing accurate manual determination of the Tafel slope: 99 mV/ dec. In contrast, the anodic portion of the curve is more complicated, even for this simple system, and a bend in the curve is observed about 75 mV above OCP. The mechanism of iron dissolution to form ferrous ions involves several steps 37 and can change with potential. Determination of the anodic Tafel slope near OCP is difcult owing to the bend in the curve and extrapolation of this region to OCP intersects at a different current density than the extrapolation of the cathodic portion. In such an example, the value determined by extrapolation of the cathodic region is probably more accurate, 204 A / cm2 in this case. The anodic Tafel slope is determined to be 53 mV/ dec by manual tting. An alternative and probably more accurate approach is to t the data to Eq 1, the equation representing ideal activation-controlled polarization 38 . The limits of the tting must be properly set, however, and the bend in the anodic curve requires an upper limit of the t to be quite close to the corrosion potential. The results of the t, shown as the dashed curve in Fig. 3, are anodic and cathodic Tafel slopes of 110 mV/ dec

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 7

FIG. 4Linear polarization experiment for Fe in 0.5 M H2SO4. Tangent is drawn at point where i

0.

and 103 mV/ dec, respectively, with a corrosion rate of 216 A / cm2. At high potentials, the current density becomes independent of potential owing to mass transport limitations and salt lm precipitation. The rst part of the bending of the curve, however, results from ohmic potential drop in solution between the reference and working electrodes. This can be minimized by current interrupt methods, but leads to instability at higher potentials during the onset of passivity. LPR results in a rather linear relationship near OCP Fig. 4 , and an RP of 107 or 67 cm2 can be determined easily. EIS is also quite easy with this system, because the low frequency limit is reached at a frequency of about 1 Hz Fig. 5 . Fitting the data to the simplied Randles circuit with a CPE shown in Fig. 1, an RP of 107 or 67 cm2 is found, which is identical to that determined by LPR. As mentioned above, the calculation of corrosion rate from this value using the Stern-Geary equation requires Tafel slopes, but the answer will not be much different assuming Tafel slopes of 100 mV/ dec. Using the Tafel slopes determined by tting the PDP to Eq 1, the corrosion rate from LPR and EIS data is determined to be 343 A / cm2. This value is relatively close to the rate determined from the PDP data, especially considering that most corrosion rate measurements are only reproducible to within about a factor of 2 at best. Noise analysis at a sampling rate of 20 Hz for 101.2 s a total of 2024 points results in the data shown in Fig. 6. After trend removal, the power spectral density of the current and potential traces were obtained and the noise resistance calculated from Rn = EPSD / IPSD 1/2 was found to be 172 . This resistance is similar to the polarization resistances determined by LPR and EIS, but as described above, area normalization and the determination of corrosion rate

FIG. 5EIS data for Fe in 0.5 M H2SO4. (a) Nyquist plot and (b) Bode plot. The lines are ts to the simplied Randles circuit shown in Fig. 1.

8 JOURNAL OF ASTM INTERNATIONAL

FIG. 6EN data for Fe in 0.5 M H2SO4. (a) Original current and potential noise. (b) Current and potential noise detrended and ltered with 0.01 1 Hz bandpass lter. (c) Fast Fourier transform (FFT) of the current and potential noise. (d) Rn determined by the square root of the ratio of the potential and current FFTs. from Rn is problematic. The EFM technique also works quite well with this system, and it provides values of Tafel slopes. The EFM data for perturbation frequencies of 0.2 and 0.5 Hz are shown in Fig. 7. The raw EFM data in Fig. 7 a show the modulated current response from the two perturbation frequencies and the frequency spectrum in Fig. 7 b clearly displays the response at the harmonic and intermodulation frequencies. Analysis of the data results in a corrosion rate of 200 A / cm2 and anodic and cathodic Tafel slopes of 84 mV/ dec and 112 mV/ dec, respectively. The causality factors are 1.926 and 2.903, which are very close to the theoretical values of 2 and 3. The results of the various techniques are given in Table 1. It is clear that the available techniques provide a number of different approaches for determining corrosion

FIG. 7EFM data for Fe in 0.5 M H2SO4. (a) Raw data. (b) Frequency spectrum.

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 9


TABLE 1Analysis of electrochemical methods for determination of corrosion rate for Fe in 0.5 M H2SO4. Electrode areas 0.63 cm2. LPR using ba and bc from PDP EIS using ba and bc from PDP EFM active dissolution model 84 112

EN

ba, mV/ dec bc, mV/ dec R P, RP, cm2 icorr, A / cm2

PDP extrapolating cathodic region 53 99

PDP tting to Eq 1 110 103

107 67 343

107 67 343

172 200 204 216

rate, but the results for this system are all relatively close. It should be noted that the corrosion rates for 99.9 % Fe in 0.5 M H2SO4 determined by these electrochemical methods are about 10 times lower than the published rate for steel at room temperature in this concentration 5 wt% of sulfuric acid as determined by weight loss measurements, 1200 mpy or 2.5 mA/ cm2 39 . The corrosion rate of about 200 A / cm2 is representative of values measured using electrochemical techniques on Fe in this solution by more than 200 lab groups overseen by the author during lab exercises in university courses and short courses for professionals over the past ten years. Weight loss measurements performed over the period of days by the same groups typically result in somewhat higher values. These differences between electrochemical and weight loss measurements are interesting and deserve consideration. Electrochemical methods, when interpreted correctly, generate an accurate representation of the instantaneous corrosion rate for the conditions of the measurements. On the other hand, as mentioned above, weight loss measurements are the gold standard of corrosion rate measurements and generate an average corrosion rate across the sample surface for the measurement period. These methods might provide different results for various reasons. One difference could be from crystallographic texture. The electrochemical experiments were performed on the sides of a drawn wire or the rolling surface of a plate. During immersion experiments, end grain attack was observed to be faster than on these surfaces. The rate of attack at the water line has also been observed to be much higher than the rate well below the surface. This discussion highlights some of the issues associated with corrosion rate testing. Weight loss experiments typically do not differentiate the rates on the different sides of an exposed sample or at a waterline and electrochemical measurements might not be performed on all surfaces or consider water line attack. PassivityThe experimental techniques described above also can be applied to passive samples for determination of corrosion rate as will be shown in the following examples. For passive samples in an environment such as neutral aerated solutions, both the anodic and cathodic oxygen reduction reactions are easily polarizable and the open circuit potential can be quite variable. The same exercise described above for the Fe/ H2SO4 system involving the application of multiple techniques was applied to a passive system, AA2024-T3 in 1 M Na2SO4. The areas of the electrodes were 1.1 cm2. The polarization curve is given in Fig. 8. The anodic current only becomes independent of potential at several hundred mV noble to the OCP. At lower potentials the slope is high but not innite: about 257 mV/ dec, as determined by manual tting. This can be considered a sort of Tafel slope even though the electrode was not dissolving actively. The cathodic Tafel slope, probably associated with oxygen reduction with diffusion limitation inuences, is also high: 246 mV/ dec. Extrapolations of the two linear portions almost intersect at the corrosion potential, resulting in a corrosion rate of about 0.2 A / cm2. Fitting Eq 1 results in anodic and cathodic Tafel slopes of 152 mV/ dec and 190 mV/ dec, respectively, and a corrosion rate of 0.13 A / cm2. The linear polarization experiment resulted in a curve that is less linear than for the active dissolution case, as shown in Fig. 9, but still can be analyzed easily for determination of polarization resistance: 79 500 or 72 300 cm2. The impedance spectrum looks like a single time constant, but it trails to higher impedances at the lowest frequencies Fig. 10 . It can be tted with different equivalent circuits, and two have been utilized in this analysis. The rst uses separate charge transfer resistances in parallel for the anodic and cathodic reactions with a Warburg impedance in series with the cathodic resistance Fig. 11 21 . A parallel CPE is used to address the capacitance. The second circuit is the nested two-time-constant model that is commonly used for defective coatings Fig. 2 .

10 JOURNAL OF ASTM INTERNATIONAL

FIG. 8Potentiodynamic polarization curve for AA2024-T3 in 1 M Na2SO4. Solid line measured curve; dashed line t to Eq 1. Both t the data rather well, and the extracted polarization resistance is similar in the two cases: 1.71 105 and 2.10 105 or and 1.55 105 cm and 1.91 105 cm2, respectively. Using the Tafel slopes from tting the PDP data to Eq 1, the corrosion rate can be determined to be 0.42 A / cm2 and about 0.18 A / cm2 for the LPR and EIS data, respectively. The analysis of the EFM data depends on the type of corrosion assumed. For active, diffusion controlled, and passive dissolution, the corrosion rates from EFM are 0.60, 2.2, and 2.2 A / cm2, respectively. These values are overestimates of the actual rate. As described above, EFM will overestimate the corrosion rate if the excitation frequencies are in the capacitive region, as is the case for this example. The causality factors are 1.89 and 3.07, which are again close to the theoretical values of 2 and 3 31 . It is interesting that the active model gives the rate closest to that determined by the other methods. The EN data result in a value of RP that is considerably less than that determined by the other methods: 40 700 . The results of all the tests are shown in Table 2. There is much more distribution in the data than for the active dissolution case. The data from the PDP and EIS methods were close. It is not clear why the polarization resistance measured by LPR was less than half that measured by EIS. Furthermore, there is no agreement regarding the best equivalent circuit for a passive metal. Advanced experimental techniques have been used to study passivity by measuring the current transients associated with passive lm formation. For an oxide that can be removed by cathodic reduction, the experiment is a simple one. Following a cathodic reduction treatment, the potential is stepped into the passive region and the anodic current is monitored as a function of time. The current will decay approximately exponentially 4046 , so, to obtain data over a wide range of current and time, it is necessary to have a high resolution analog-to-digital converter or a series of data collection devices operating at

FIG. 9Linear polarization experiment for AA2024-T3 in 1 M Na2SO4. Tangent is drawn at point where i 0.

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 11

FIG. 10EIS data for AA2024-T3 in 1 M Na2SO4. (a) Nyquist plot and (b) Bode plot. The lines are ts to the circuit shown in Fig. 11. different current and time scales 4345 . In the latter case, it is necessary to piece together the different time segments, which is not easy. With a wide dynamic range of data, it is possible to plot log current density versus log time to determine the kinetics of the passivation process. If the sample area and peak current are large, a limiting value and subsequent small shoulder in the decay can be observed. This has been attributed to ohmic potential drop associated with the high initial current 47 . The ohmic potential drop decreases as the current decreases, so the potential at the electrode surface is not constant with time in that situation. Therefore, it is desirable to minimize ohmic potential drop by using a conductive electrolyte and an electrode conguration that minimizes the area of the exposed metal. Most alloys of technical interest cannot be reduced cathodically, but repassivation can be studied by exposing an area of unpassivated metal. Many different techniques have been used to depassivate or bare small areas of metal for repassivation studies, including scratching, breaking, guillotining, impinging particles, and incident laser irradiation. Burstein and co-workers have developed and extensively used the scratching and guillotining approaches 4345,47 . Scratching is easily accomplished using a hardness indentation stylus and a rotating disk electrode. The stylus is pulsed against the surface to cut a groove into

FIG. 11Equivalent circuit with separate charge transfer resistances in parallel for the anodic and cathodic reactions, a Warburg impedance in series with the anodic resistance, and a parallel CPE.
TABLE 2Analysis of electrochemical methods for determination of corrosion rate for AA2024-T3 in 1 M Na2SO4. Electrode areas 1.1 cm2. LPR using ba and bc from PDP EIS using ba and bc from PDP

EN

ba, mV/ dec bc, mV/ dec R P, RP, cm2 icorr, A / cm2

EFM passive model Innity 312

PDP extrapolation 257 246

PDP tting to Eq 1 152 190

79 500 72 300 0.42

190 000 173 000 0.18

40 700
2.2 0.2 0.13

12 JOURNAL OF ASTM INTERNATIONAL

FIG. 12Schematic of (a) sample cross section and (b) breaking electrode cell [49]. The arrow represents the position where the sample is struck to initiate fracture. the sample. A simpler approach is to scratch the surface manually with a sharp scribe. In this case, the scratch can be more reproducible if the sample has a short dimension, such as is achieved by mounting a thin foil on edge into an epoxy mount. The manual scratching across the thin sample can then be accomplished relatively quickly and reproducibly 48 . To study the earliest stages of repassivation where the current densities are the highest, it is necessary to create the fresh area as quickly as possible and to minimize the area of the fresh metal. A very small area of fresh metal can be created extremely quickly on the order of s by the thin lm breaking experiment 49 . In this approach, a thin lm deposited onto a brittle substrate such as glass or Si is suspended into the solution Fig. 12 . Breaking of the thin lm electrode results in the creation of a fresh metal area of size equal to the cross section of the thin lm. Current densities on the order of 1000 A / cm2 were measured using this technique on Al thin lms 49 . Localized CorrosionElectrochemical testing can be used to assess the susceptibility to localized corrosion by determination of critical potentials, i.e., the breakdown, repassivation, and corrosion potentials, as well as the critical pitting or crevice temperature 1,50 . The critical potentials are typically measured by cyclic potentiodynamic polarization as described elsewhere 1,50 . The relative values of the critical potentials can be used as an assessment of the susceptibility to localized corrosion 50 . This assessment, however, is complicated by the several factors, including the broad dispersion of values measured in many systems. An approach for determining a more accurate value of repassivation potential is the scratching technique described above 51 . By purposely removing the passive lm, the initiation stage of pitting is eliminated and the repassivation potential can be measured directly. A sample can be scratched during potentiostatic polarization at various potentials or periodically during potentiodynamic polarization. At potentials below the repassivation potential, the current will quickly decay to the passive current level following scratching. When the sample is scratched at a potential above the repassivation potential, the current will increase continuously. An electrochemical equivalent to the mechanical scratching experiment involves stepping the potential for a short period to a high potential E1 to initiate localized attack, and then stepping down to a low potential E2 52 . If the localized attack repassivates and the current decays

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 13

at E2, then the cycle is repeated, but the value of E2 is increased during each step until repassivation at E2 no longer occurs. The highest value of E2 at which the repassivation occurs is then the repassivation potential. This value represents repassivation of relatively small pits or crevices. Localized corrosion of Al alloys is strongly related to microstructural features on the scale of m to tens of m. The inuence of local heterogeneities cannot be studied by most scratching methods because there is no control over the placement of the scratch relative to the microstructure. However, microstructures are visible in atomic force microscopes AFMs , particularly by using the scanning Kelvin probe force microscope SKPFM technique 53,54 and pressure applied by an AFM tip on the surface of a sample during in situ contact mode rastering is a form of scratching on the micro scale. An AFM tip can be placed directly on a feature of interest as determined by the topography or by ex situ SKPFM and the use of ducial marks. Furthermore, the force applied by a tip can be controlled exactly. In situ rastering of an AFM tip on a surface at high forces has been termed AFM scratching 5557 . AFM scratching involves in situ contact mode rastering of an area at high force followed by reducing the force and increasing the scanned area size to assess the effect of the scratching. AFM scratching of pure Al and AA2024-T3 in chloride solution with and without the addition of chromate or vanadate provided understanding on the critical concentrations of inhibitor needed for repassivation of the scratched surface under various conditions 56,57 . A more complicated technique, called the Tsujikawa Hisamatsu electrochemical THE method, has been described for the determination of a conservative repassivation potential associated with the repassivation of well-developed crevice corrosion 58,59 . Interest in the THE method has been spurred recently because of efforts to predict the behavior of very corrosion resistant alloys for very long times in nuclear waste storage applications 60 . This method has several steps. A potentiodynamic polarization scan to a xed current is used to initiate attack. This is followed by a galvanostatic hold at that current to grow the attack, and then stepwise decreasing of the potential under potentiostatic control to determine the repassivation potential, which is dened as the potential at which the current does not increase during the potential hold period. Typically, this technique is performed using a sample to which a multiple crevice assembly 61 has been attached because the repassivation potential for a deeply grown crevice should be the lowest repassivation potential, and thus the most conservative from a design perspective. In other words, if the corrosion potential were to remain below the repassivation potential determined by the THE technique on a creviced sample, then one could be relatively condent that localized corrosion would not occur during exposure conditions. Several issues exist with the THE technique. First, the repassivation potential determined by this method might be overly conservative for many applications 59 . Furthermore, the repassivation potential could depend on the hold current density and the potential step period. Finally, the determination of the repassivation potential by the criterion described above is not always clear. The current has been observed to stay constant for several potential steps and then increase again at a lower potential. The study of metastable pitting has provided unique opportunities for understanding pitting corrosion since metastable pits involve initiation, growth, and repassivation in short, discrete, and plentiful events 62 . It is easier to address the stochastics of pitting with the numerous metastable pits generated in a few experiments than by performing many experiments that generate stable pits. The signal associated with each metastable pitting event can be analyzed to determine pit current density, which allows determination of the electrochemical kinetics of dissolution in pits and assessment of the rate-controlling steps. Observation of metastable pitting current transients requires low background current because the current associated with metastable pits is small. The smaller the background signal, the smaller the events that can be resolved. It is critical to avoid artifacts associated with crevice corrosion, which will swamp the metastable pitting signal. To get a low background current, it is necessary to use electrodes with small surface areas. Data collection must be at a high enough rate to distinguish the individual events. Electrochemical monitoring of corrosion is useful for sensing the onset of localized corrosion through a decrease in RP or a change in the EN response. However, for localized corrosion the unknown active area makes it impossible to determine the current density from RP even using assumed Tafel slopes. Another problem is that a considerable amount of hydrogen evolution can occur within pits, especially for Al and Al alloys. It has been found that about 15 % of the total anodic dissolution current is consumed locally by hydrogen evolution in Al pits and crevices 63,64 . This means that the current measured at an applied potential is only a fraction of the true anodic dissolution current. Finally, the determination of pit growth

14 JOURNAL OF ASTM INTERNATIONAL

FIG. 13Schematic of the lower part of the cell for foil penetration experiments [71]. kinetics at open circuit is a challenge, because by denition, no net current is passed at open circuit. In many monitoring applications, knowledge of localized corrosion initiation is sufcient and the exact rate of penetration or growth is not required. However, in fundamental studies of localized corrosion susceptibility or mechanism, or for lifetime prediction of components, the rates of localized corrosion growth and the potential dependence are of interest. A variety of experimental approaches address the problems listed above in different ways. In general, they involve either the formation of a single localized corrosion site, or non-electrochemical means for determination of the growth kinetics. One of the reasons that metastable pits are so interesting is that metastable pit current transients represent individual pitting events. Other techniques for forming a single localized corrosion site include the exposure of a small area, laser irradiation of a small spot, implantation of an activating species at a small spot, or the use of articial or single pit electrodes. Other than the articial pit electrode technique, in which the whole exposed area is active, the other techniques listed still require an assumption regarding the geometry of the active pit surface to determine the pit area and thus a pit current density. Articial pit electrodes are formed by imbedding a wire in an insulator such as epoxy 6567 . The local environment within an articial pit electrode crevice should be identical to that formed in real pits. The whole exposed area is active so the measured current can be converted easily to pit current density. Furthermore, articial pit electrodes have an ideal one-dimensional geometry allowing for easy modeling of transport. Articial pit electrodes have been used extensively to study Fe and stainless steel behavior. Polarization curves for articial pit electrodes of FeCrNi alloys with and without 2.7 % Mo in a 1 M Cl bulk solution showed that anodic polarization curve in the pit environment for the Mo-containing alloy was shifted to higher potentials 68,69 . The change in the corrosion potential was approximately equal to the change in pitting potential measured on a standard electrode by potentiodynamic polarization 68,69 . This indicates that the effect of Mo on the pitting potential can be explained solely by its inuence on the pit dissolution kinetics. Non-electrochemical techniques are useful for the study of localized corrosion growth because they eliminate several problems: the need to determine the current from a single site, assumptions regarding active surface area, and complications associated with hydrogen evolution within the pits or intergranular regions 63,64 . The foil penetration method is a non-electrochemical method that measures the time for localized corrosion to perforate foils of varying thickness as a means to determine the growth rate of the fastest growing localized corrosion site 7072 . The penetration time is determined by exposing one side of a foil sample to an aqueous environment Fig. 13 . A piece of lter paper and then a Cu foil are pressed against the back of the sample. Penetration of localized corrosion on the other side is sensed by a decrease in resistance between the sample and Cu foil resulting from wetting of the lter paper by the localized corrosion environment. This technique has been applied successfully to the study of pitting, crevice corrosion, and intergranular corrosion in stainless steel, Al and Al alloys 7072 . It even can assess localized corrosion kinetics under open circuit conditions 71 . Another non-electrochemical approach to pitting involves the study of 2-D pits in thin lm samples 64,7275 . Pits in thin metallic lms with thickness on the order of 10 1000 nm rapidly penetrate the metal, reach the inert substrate, and proceed to grow outward in a 2-D fashion with perpendicular sidewalls. The measurement of pit wall velocity from the analysis of magnied images of the growing 2-D pits Fig. 14 provides a simple and direct means for determination of pit current density via Faradays law, with no need for assumptions. Since the pit depth is limited by the metal lm thickness, pits in thin lms grow at steady state with no increase in ohmic path or diffusion length with time as is the case for pits in bulk samples. As a result, 2-D pits in thin lms exhibit a pit current density that is constant with time at a given applied potential. This steady-state aspect of pitting in thin lm samples allows unambiguous

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 15

FIG. 14Schematic drawings of the experimental setup and of the nature of 2-D pit growth in thin lms [75]. determination of the current-density/potential relationship. The typical form for the polarization curve of a thin lm pit includes a region at low potentials where the current density is almost linearly dependent on applied potential followed by a region at high potential where the pit current density reaches a limiting value, and pitting is mass-transport limited. Another consequence of the constant pit depth is the nding that the repassivation potential and the lowest pit current density at which a pit can grow are extremely reproducible. It is therefore possible with pits in thin lms to assess accurately the critical conditions for pit stability. One nal advantage of working with thin lms produced by physical vapor deposition is that the full range of compositional variety can be achieved with a single phase structure typically nanocrystalline or amorphous because of the non-equilibrium nature of the as-deposited microstructure. The series of standard electrochemical methods described above was performed on a system undergoing pitting corrosion, AA2024-T3 with electrode area of 1.1 cm2 in 1 M NaCl. The PDP curve is given in Fig. 15. Since the sample was pitting at open circuit, the effective anodic Tafel slope was very low and the current rapidly increased at potentials above the OCP. In standard electrochemical approaches the active pit area is not known. Therefore, the reported current density is a nominal current density, normalized to the whole exposed area, and not representative of the actual rate of dissolution in the pits. By tting to Eq 1, the apparent anodic and cathodic Tafel slopes were found to be 17 mV/ dec and 390 mV/ dec, respectively, and a nominal corrosion rate of 1.67 A / cm2 was determined. Corrosion was under cathodic control and the corrosion potential was pinned at a value close to the pitting potential. Figure 16 shows the linear polarization results. The large difference in anodic and cathodic Tafel slopes resulted in a very

FIG. 15Potentiodynamic polarization curve for AA2024-T3 in 1 M NaCl. Solid line measured curve; dashed line t to Eq 1.

16 JOURNAL OF ASTM INTERNATIONAL

FIG. 16Linear polarization experiment for AA2024-T3 in 1 M NaCl. Tangent is drawn at point where i 0. non-linear curve that is not amenable for easy determination of polarization resistance. An estimate results in Rp = 812 or 738 cm2. The EIS plots shown in Fig. 17 are also somewhat difcult to analyze. Owing to the non-steady-state nature of an electrode undergoing pitting, the stability requirement for the determination of a valid impedance is not fullled. The impedance magnitude decreases at frequencies lower than about 0.1 Hz and the phase angle becomes positive so that the data in the Nyquist plot lie in the rst quadrant. This is likely the result of the increasing pit area and corrosion current with time during the EIS measurement. A number of different equivalent circuits can be used to t these data. A specic model for tting EIS spectra for a sample undergoing localized corrosion was proposed by Shih and Mansfeld 76 . In that model circuit components representing the pitted area are in parallel to components representing the passive area. That model, however, will not result in decreasing impedance at low frequencies. The entire EIS spectrum in Fig. 17 can be tted well to a two-time-constant nested model with CPEs if a negative exponent is allowed for the inner CPE. Given the non-steady-state nature of the data, it is probably more reasonable to ignore the data below 0.1 Hz. The remaining data can be tted to a nested two-time-constant model with positive CPE exponent. The best t is for an inner CPE exponent of 1, which does not make sense, so the exponent was set to 1. The values of the pore and double layer resistances are 3610 and 2010 , respectively, for a total low frequency impedance of 5620 or 5110 cm2. Using the apparent Tafel slopes from the PDP curves, nominal corrosion rates of 1.15 A / cm2 and 7.9 A / cm2 for the EIS and LPR data, respectively, can be determined. Despite the difculties in the analyses, these values are quite close to the result of Tafel extrapolation. The Rp from

FIG. 17EIS data for AA2024-T3 in 1 M NaCl. (a) Nyquist plot and (b) Bode plot. The lines are ts to the circuit shown in Fig. 11.

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 17


TABLE 3Analysis of electrochemical methods for determination of corrosion rate for AA2024-T3 in 1 M NaCl. Electrode areas 1.1 cm2. LPR using ba and bc from PDP EIS using ba and bc from PDP EFM active dissolution model 116 170

EN

ba, mV/ dec bc, mV/ dec R P, RP, cm2 icorr, A / cm2

PDP tting to Eq 1 17 390

812 738 7.9

5620 5110 1.15

7723 6.7 1.67

LPR again seems to be low resulting in a high corrosion rate. However, the t to the LPR data was very problematic. The EN data were analyzed with the same approach described above, resulting in a value of Rn = 7723 , which is reasonable. EFM seems to fail in this system as the causality factors are 3.20 and 2.65, which are much different than the theoretical values of 2 and 3, respectively. Nonetheless, the value of corrosion rate determined from the EFM analysis is 6.7 A / cm2, which is still close to the values determined from the other methods. The results of all of the analyses are summarized in Table 3. Unmet Needs for Electrochemical Testing of Samples Immersed in SolutionThe analysis given above shows that all of the electrochemical techniques work quite well to determine corrosion rate for the case of uniform active dissolution. The results are close, even if assumed values of the Tafel slopes are used. Many of the methods can be used for corrosion rate monitoring as only small or no perturbation from the steady-state open circuit potential is required. The available techniques seem to meet the needs. Some might express concern about the reproducibility of electrochemical measurements, which as indicated is typically within a factor of about 23. The measurements themselves are extremely accurate. The variability is a result of real variation in sample-to-sample reactivity in many environments. For passive metals, the available electrochemical techniques also are suitable for most needs. The plans to bury high level radioactive waste in Yucca Mountain have brought to corrosion science a new problem that still might require new methods. There is a need to predict passive lm stability on metal canisters for tens of thousands of years. Even ignoring the possibility of localized corrosion, the growth and stability of passive lms must be understood. The canisters will experience one long thermal transient heat up and cool down in an environment consisting of rock dust and humid air or dripping dilute groundwater. The thickness, composition, and stability of the passive lm must be known and understood. Electrochemical approaches to this problem are being taken, but will not be reviewed here. However, this problem is extremely challenging. Prediction and assessment of the localized corrosion phenomena present the greatest challenges to electrochemistry and corrosion science. It is easy to determine the critical potentials, i.e., the corrosion, breakdown, and repassivation potentials, for a given metal in a given environment. Comparison of the corrosion potential and a prediction of what the corrosion potential will be in the future with the breakdown and repassivation potentials can be used to make an assessment of the likelihood for localized corrosion to occur. In particular, if the corrosion potential will remain below a conservative estimate of the repassivation potential for deep pits or crevices, one can be condent that localized corrosion will not be an issue. However, if this is not the case, there is no agreed-upon approach for prediction of component lifetime. Corrosion scientists are often asked to quantify the benet of a reduction in environmental severity e.g., a decrease in chloride concentration or a change in alloy. The eld lacks the level of understanding available, for instance, in the eld of fracture mechanics where laws of similitude allow eld predictions from lab experiments. In some cases, such as in microelectronics, the mere initiation of a pit could represent a failure, whereas in other situations failure would require complete penetration of a component such as a storage vessel. The standard electrochemical approaches do not determine the localized current density in pits or crevices. Pit growth kinetics can be assessed using methods described above, but they are time consuming and labor intensive. Improved methods are needed to measure rates of localized corrosion and to connect those measurements to eld exposures. Another critical step that is not well understood is the transition of pits and crevices to cracks, which tend to propagate at much faster rates than pits or crevices.

18 JOURNAL OF ASTM INTERNATIONAL

FIG. 18Schematic drawing of droplet cell. Metal in Atmosphere As described above, atmospheric conditions could be adsorbed water layers in equilibrium with humidity and pollutants, or could involve precipitation, including acid rain. Under most atmospheric exposure conditions, however, the layer of electrolyte on the surface is extremely thin. Electrochemical measurements clarifying the kinetics and mechanisms of atmospheric corrosion are complicated by problems associated with immersing a reference electrode RE and counter-electrode CE into the thin layer. A rather simple way to mimic atmospheric corrosion is by the droplet cell, which has small dimensions and allows the study of resistive electrolytes such as pure water 77 . A circular area of a at sample is exposed through a hole in a piece of protective tape Fig. 18 . A circular piece of lter paper of the same size is placed securely into the exposed hole and a small typically 10 20 l droplet of water is placed on the lter paper using a calibrated pipetter. This wet lter paper acts as the electrolyte. A piece of woven Pt mesh is placed on top of the wet lter paper, and a reference electrode is held against the back of the Pt counter-electrode. This simulates atmospheric corrosion, in which a thin water layer forms on the surface. As in atmospheric corrosion, soluble species on the sample surface and pollutant gases in the air are dissolved into the water droplet, which provides some conductivity. Mansfeld and Kenkel described the use of atmospheric corrosion rate monitors, composed of interdigitated ngers of different metals, that track the galvanic current owing through the cross section of stacks of alternating dissimilar metal plates 78 . For monitors exposed to the weather, the maximum current was found to ow at the point just before complete drying owing to the increase in oxygen diffusion rate with decreasing solution layer thickness. Mansfeld used similar devices composed of a single material to make measurements of polarization resistance 79 . The measurements were twoelectrode measurements with no reference electrode. Lyon and co-workers measured electrochemical kinetics in layers as thin as 50 m using an essentially bulk electrochemistry approach involving a capillary connection to a remote RE and CE 8084 . The problem of introducing an RE into the thin water layer responsible for atmospheric corrosion was solved by Stratmann and Streckel, who demonstrated that a Kelvin probe KP vibrating above a sample provides a measure of the corrosion potential 8587 . In brief, the changing capacitance between the vibrating probe and the sample generates a current if the potential of the probe and sample are different. The local potential is determined by adding a backing potential to the sample that either nulls out the signal or modies the signal in a way that can be analyzed. Stratmann and Streckel made corrosion rate measurements by monitoring the rate of decrease in the oxygen partial pressure 86 . They also used a modication of the KP apparatus to control the potential and measure the current to a coplanar counter electrode, thereby achieving the ability to measure polarization curves 87 . This arrangement might be called a Kelvin probe potentiostat KPP . They used the KPP to measure cathodic polarization curves on Pt under thin layers of 1 M Na2SO4. The limiting current density for oxygen reduction was found to vary inversely with the thickness of the electrolyte layer for layers with thickness between about 10 m and

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 19

100 m. A new approach for making KPP measurements has been described recently 88 . In this approach, the distance between the tip and sample is controlled, allowing simultaneous tracking of the sample topography or measurement of the thickness of an electrolyte layer as it dries. Electrochemical measurements under thin electrolyte layers were also reported by Nishikata, Tsuru, and co-workers 8993 . They rst performed EIS measurements on samples composed of two identical electrodes of steel, stainless steel, or Cu embedded in epoxy under thin electrolyte layers during cyclic wet-dry conditions 90,91 . They subsequently developed a three-electrode approach using a coplanar chloridized Ag RE embedded in epoxy between the Fe working electrode and the CE 89 . The RE allowed for the measurement of cathodic polarization curves, and they found the same dependence of limiting current density on layer thickness as did Stratmann et al. Nishikata, Tsuru, and co-workers also studied pitting corrosion of stainless steel under thin electrolyte layers 9294 . The corrosion rate associated with pitting was assessed by EIS during cyclic wet-dry cycling on different stainless steels 92 . Corrosion potential measurements were made on a second embedded working electrode using yet another sample design in which an agar-lled hole in the epoxy allowed ionic connection to a remote standard RE 93 . A decrease in the corrosion potential by hundreds of mV coincided with a sharp decrease in polarization resistance and the onset of pitting corrosion as a critical chloride concentration was reached during drying. The coplanar Ag/ AgCl RE and agar-lled hole used in the work of Tsuru et al. suffer from problems associated with changing Cl concentration and non-uniform current distribution. Exfoliation corrosion EFC is a form of localized corrosion that occurs in wrought Al alloys with elongated grain structures during exposure to aggressive atmospheres. The susceptibility to EFC typically is assessed by non-electrochemical techniques. ASTM Standard G34, Standard Test Method for Exfoliation Corrosion Susceptibility in 2xxx and 7xxx Series Aluminum Alloys, also known as the EXCO test, involves immersion in an oxidizing acidic chloride solution and comparison of the resulting surface to standard photographs 95,96 . Such test conditions are very aggressive and different than typical exposure conditions. Other test methods involve intermittent spraying of acidied salt solutions onto the surface of the specimens 97100 . The behavior of Al alloys in these accelerated environments has been correlated to long-term exposure in less aggressive natural environments 101103 . As a result, these tests are useful for assessing susceptibility to EFC attack. However, they do not provide quantitative measurements of susceptibility or growth kinetics, which are required for predictive modeling of corrosion development. Liddiard and co-workers have used a deection technique to quantify exfoliation extent and determine EFC kinetics 104 . In this technique, the effective remaining load-bearing section of a sample exhibiting EFC is determined from its compliance under four-point bending. The rate of EFC can be assessed from periodic measurements. The deection technique is valid only when the thinning of the specimen during corrosion is uniform. A new technique, exfoliation of slices in humidity ESH , was developed recently for the determination of EFC susceptibility and quantication of EFC kinetics 105,106 . In this technique, slices of plate are pretreated by potentiostatic polarization in chloride solution to develop localized corrosion sites. Subsequent exposure to high humidity after pretreatment of properly oriented and unconstrained samples results in the development of EFC at the edges of the slices. With time the EFC grows inwards and the EFC kinetics are determined by measuring the width of the central unattacked region of the samples. ESH test results have been shown to be representative of different EFC behavior of plates during outdoor exposure 106 . Unmet Needs for Electrochemical Testing of Samples in AtmosphereAt present there is no good electrochemical technique for electrochemical testing of samples exposed in atmosphere. Electrochemical methods for samples in corrosive atmospheres are difcult due to the thin layer electrolyte. A counterelectrode can be coplanar, but this can result in non-uniform current distributions. Placement of the reference electrode is particularly problematic, and the use of the Kelvin probe and coplanar reference electrodes such as Ag/ AgCl were discussed above. Standardized procedures for electrochemical testing of samples in atmosphere should be pursued. Furthermore, it is becoming evident that complex atmospheric chemistry can inuence the local environment for a sample exposed outdoors in ways that are not simulated in lab exposure tests. The interaction of sunlight of wavelength 243 nm and ozone results in the formation of uncharged OH radicals that can interact with chloride in seawater aerosols above or near the ocean to create ultimately Cl2 107 . The presence of a small amount of aggressive radicals or chlorine gas could have a large inuence on atmospheric corrosion. The understanding of the atmospheric chemistry is still evolving and the inu-

20 JOURNAL OF ASTM INTERNATIONAL

ence of the environment on corrosion and the electrochemistry of this phenomenon are totally unknown. A major need in the eld of atmospheric corrosion is a fundamental understanding of the mechanisms and acceleration factors associated with the critical parameters in lab exposures. Accelerated tests are used widely in the electronics industry, and the acceleration factors associated with temperature, humidity, and bias the typical stresses in the electronics industry are relatively well understood 10 . Parts or simulated parts are exposed to aggressive conditions and some property is measured or mean time to failure MTTF is determined. Electronic components are amenable to the measurement of failure time. The goal is to develop an acceleration factor AF for a given test, where AF= MTTF eld /MTTF test . Expressions exist for the AF for each of the stresses. The effects of the stresses are typically considered to be independent, though that is not always the case. Regardless, the total AF is usually a product of the individual AFs. This approach works relatively well for electronic components allowing extrapolation from the aggressive accelerated test environment to service environments, which are typically much more benign. Such understanding does not exist for atmospheric corrosion in general. For instance, there is no understanding that allows the behavior in ASTM B117 exposure to be used for prediction in some other environment. This understanding is complicated by factors such as the atmospheric chemistry described in the last paragraph. Nonetheless, it would be very valuable to have a scientic basis to predict corrosion rates and lifetimes for any environment from the results of an accelerated exposure test. Coated Metal in Solution Most of the techniques described above are not suitable for the study of metals covered with a protective organic coating. The primary electrochemical technique for this application is EIS. As mentioned above, EIS is well suited for painted samples as it is sensitive to the early stages of coating degradation, well before defects are visible by eye. However, it is a technique that is prone to artifact and misinterpretation, especially by ill-trained or inexperienced practitioners. So, for electrochemical evaluation of coated metals in solution, the major need is better training in the proper use and interpretation of EIS. However, even experienced practitioners must make critical decisions in the analysis of EIS data regarding the choice of equivalent circuit for tting of the data. It is possible for some data sets to develop multiple equivalent circuits that are reasonable and physically justiable. Sometimes it is not clear when a second time constant must be incorporated in a set of data taken, for instance, as a function of exposure time. Potentiostatic pulse testing PPT is a related technique that has been reported to be useful for monitoring the earliest stages of degradation of paint coatings 108,109 . A square pulse of 0.1 2.0 V is applied to a coated sample in a two-electrode cell with the reference electrode also acting as the counter-electrode. The pulse can be in the form of a square wave so that the frequency of the square wave and the sampling frequency will determine the limits of the frequency response. As in EIS, large amplitude signals can be used for systems with very protective coatings. The current response is tted to an equivalent circuit, as in EIS. Values of charge transfer Rct and pore or defect resistances Rd on the order of 1011 cm2 and double layer capacitance Cdl on the order of 1011 F / cm2 have been successfully determined with this technique 108 . This range of impedance is well beyond the typical range of standard EIS equipment. However, it is not possible to get all of the circuit components simultaneously by PPT when the time constants are very different 109 . On the other hand, the unavailable components e.g., the solution resistance Rs and coating capacitance Cc are typically not critical for the determination of the overall corrosion resistance or the early stages of coating degradation. PPT can be sensitive and cost efcient if a picoammeter is used instead of a potentiostat, and PPT data can be very useful for fast prediction of coating performance at the early stage of the coating failure. However EIS provides a fuller description of the coated electrode response. The Kelvin probe was described above as a tool that could be used for measuring or controlling potentials for samples under a thin layer of electrolyte. By using a small probe needle and appropriate scanning motors, it is possible to map potentials on a surface. The scanning Kelvin probe SKP is a useful tool for studying coated samples 110 . The Kelvin Probe measurement cannot be performed in full immersion, so SKP is not suitable for studying immersed samples. However, using a reservoir of solution adjacent to a coating delamination, it is possible to measure quantitatively the delamination rate of organic coatings at defects 110 . In this way, the behavior under immersion conditions can be simulated. SKP

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 21

potential proles have quantied delamination of organic coatings on galvanized steel and have led to the development of galvanizing methods with much lower delamination rates and thus improved corrosion protection 111 . One critical aspect of coatings, particularly for Al alloys, is the ability to heal defects spontaneously. This self-healing phenomenon, also called active corrosion inhibition, can be achieved by both pigments in organic paints as well as conversion coatings. The articial scratch cell is an approach that allows for clear assessment of various aspects of active corrosion inhibition 112 . An untreated surface representing a scratch is placed in close proximity to a treated or coated surface, separated by a thin layer of electrolyte, typically an NaCl solution. The two surfaces can be shorted electrically or not, and the thin electrolyte layer can be exposed to air at its edge or not, by using a partial shim or complete o-ring. In this fashion, the untreated surface is exposed to electrolyte that has been altered by exposure to the treated surface. By insertion of reference and counter-electrodes, it is possible to monitor the changes in corrosion resistance of both surfaces in situ. The initially coated sample might degrade with time, and the initially uncoated sample might improve with time as it interacts with the electrolyte. After the experiment the electrolyte can be analyzed to determine what has been released by the coated sample, and the surfaces of the electrodes can be analyzed. Surface analysis of the initially uncoated sample provides information regarding the transport of inhibiting species from the coated sample through the thin electrolyte layer. The articial scratch cell has been used to study chromate conversion coatings and Ce-containing coatings 112,113 . Unmet Needs for Electrochemical Testing of Coated Samples Immersed in SolutionThe EIS method is very well suited for the evaluation of coated samples immersed in solution. Sensitive, affordable, and user-friendly EIS systems have made the technique available for use by technologists. However, it is not as widely used for testing of coated samples as it could or should be, due largely to a resistance to change by practitioners in industry. It is likely that EIS will gain acceptance with time as more people use it and feel comfortable with it. As described above, the nested two-time-constant equivalent circuit is well suited for painted samples. An issue, however, is when to apply the two-time-constant model to analyze data for a sample that slowly degrades with time. Coated Metal in Atmosphere There are many applications in which painted samples are exposed to the atmosphere, including automotive, aerospace, and construction usages. This important corrosion exposure condition is the one with the fewest options for electrochemical measurements, or actually for any kind of meaningful measurements. The difculty is that this condition combines the complications of a coated sample with the limitations of atmospheric exposure. Another coating class that is of interest for corrosion protection is inorganic conversion coatings, which are often used in combination with organic overcoats but can also be used as free-standing protection without a topcoat. Furthermore, many coated systems have multiple layers, for instance, a sacricial protection layer covered by a conversion coating and then multiple organic layers. These layers can contain species that interact with the water permeating the coating and alter the local environment. So in the absence of a through-coating defect, the environment at the metal surface where corrosion occurs is different than the outer environment, either test or real life, that is measured or controlled. As described above, a common approach for examining the corrosion protection provided by a coating is exposure in an aggressive atmospheric environment such as the salt fog generated in a chamber as described by ASTM B117 8 . Samples are typically inspected after a period of exposure. The extent of corrosion is assessed by a metric such as the number or area of defects per unit area observable by eye, or the extent of delamination at a scribe. A pass/fail ranking is assigned based on some criterion. Salt fog testing has been criticized as not being representative of or correlating with many real world environments 98 and as quoted above, the B117 standard itself provides several strong statements about the appropriate limits for usage of test results. Electrochemical testing can be combined with salt fog testing to get a more quantitative measurement. In this approach, samples are exposed to the salt spray environment and periodically removed for testing by EIS. The EIS testing is done by immersion in solution using standard approaches and the sample is then returned to the salt spray environment. Such an approach leads to concerns regarding whether the periodic removal and immersion inuences the sample 93 .

22 JOURNAL OF ASTM INTERNATIONAL

EIS has also been investigated as a faster and possibly more discriminating alternative to salt spray testing. A matrix of inorganic conversion coatings on Al alloys was tested by salt spray exposure and EIS during immersion in 0.5 M NaCl 114 . The resistance to salt spray was assessed by the number of pits with an arbitrary value of ve pits on the 194 cm2 sample considered a failure. EIS was performed on a parallel set of samples and the coating resistance Rc was used as a gure of merit to assess coating performance. Both tests effectively assessed the extent of corrosion protection; the probability of achieving passing a 168-h salt spray exposure increased as Rc increased. For each alloy, threshold Rc values were proposed for which a given coating can be expected to attain a passing result in a 168-h salt spray test. Unmet Needs for Electrochemical Testing of Coated Samples Exposed to AtmosphereThis category has the most needs. All of the needs described above for metals exposed to atmosphere also are relevant here. However, the techniques developed for studying the behavior of metals under thin lm electrolytes are not applicable for coated samples. There is a severe lack of scientic underpinning to the prediction of lifetime for coated samples in atmospheric exposure conditions, which results largely from a lack of useful tools. Summary Electrochemistry techniques are invaluable for studies, monitoring, and prediction of metallic corrosion. They are available at affordable prices and with user-friendly software that makes their utilization straightforward. This review has considered the specic cases of uncoated and coated metal exposed to bulk electrolyte environments or atmospheric corrosion conditions. This simple matrix provides a useful framework for considering the status, limitations, and needs of electrochemical techniques in corrosion. The greatest limitations, and needs are in the areas of localized corrosion and atmospheric corrosion. In localized corrosion there is a need for methods to determine localized, not nominal, corrosion rates and for laws of similitude that will enable the prediction of component lifetime in the eld from short-term laboratory experiments. In atmospheric corrosion there is a need for sensitive, simple, and standardized approaches for determination of corrosion rates, both for uncoated and coated metal samples. Acknowledgments The author gratefully acknowledges Y. Zhai and S. Taira for performing the experiments reported herein and is indebted to R. Buchheit for his input and comments. References 1 Frankel, G. S. and Rohwerder, M., Experimental Techniques for Corrosion, in Corrosion and Oxide Films, Encyclopedia of Electrochemistry, Vol. 4, M. Stratmann and G. S. Frankel, Eds., Wiley-VCH, Weinheim, Germany, 2003. Mansfeld, F., The Polarization Resistance Technique for Measuring Corrosion Currents, in Advances in Corrosion Science and Technology, M. G. Fontana and R. W. Staehle, Eds., Plenum, New York, 1976. Scully, J. R., Electrochemical Methods for Laboratory Corrosion Testing, in Corrosion Testing and Evaluation: Silver Anniversary Volume, R. Baboian and S. W. Dean, Eds., ASTM International, West Conshohocken, PA, 1990. Scully, J. R., Polarization Resistance Method for Determination of Instantaneous Corrosion Rates, Corrosion (Houston), Vol. 56, 2000, pp. 199218. Kelly, R. G., Scully, J. R., Shoesmith, D. W., and Buchheit, R. G., Electrochemical Techniques in Corrosion Science and Engineering, Marcel Dekker, New York, 2003. Knotkova-Cermakova, D. and Barton, K., Corrosion Aggressivity of Atmospheres Derivation and Classication , in Atmospheric Corrosion of Metals, ASTM STP 767, S. W. Dean Jr. and E. C. Rhea, Eds., ASTM International, West Conshohocken, PA, 1982. Leygraf, C. and Graedel, T. E., Atmospheric Corrosion, Wiley-Interscience, New York, 2000. ASTM, Standard B117-97, Standard Practice for Operating Salt Spray Fog Testing Apparatus,

4 5 6

7 8

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 23

9 10

11 12 13

14

15 16 17 18 19 20 21 22

23 24 25

26 27 28 29 30 31

32 33

Annual Book of ASTM Standards, Vol. 3.02, ASTM International, West Conshohocken, PA, 2000. Cyclic Cabinet Corrosion Testing, G. S. Haynes, Ed., ASTM STP 1238, ASTM International, West Conshohocken, PA, 1995. Frankel, G. S. and Braithwaite, J. W., Corrosion in Microelectronic and Magnetic Data-Storage Devices, in Corrosion Mechanisms in Theory and Practice, 2nd Ed., P. Marcus, Ed., Marcel Dekker, New York, 2002. Thierry, D., Thoren, A., and Leygraf, C., Corrosion Monitoring Techniques Applied to Cooling Water and District Heating Systems, Paper 463, in Corrosion/87, NACE, Houston, TX, 1987. Jones, D. A., Principles and Prevention of Corrosion, 2nd Ed., Simon and Schuster, Upper Saddle River, NJ, 1996. Frankel, G. S., Wagner-Traud to Stern-Geary; Development of Corrosion Kinetics, in Corrosion Retrospective, G. S. Frankel, H. S. Isaacs, J. R. Scully, and J. D. Sinclair, Eds., The Electrochemical Society, New York, 2002. Wagner, C. and Traud, W., On the Interpretation of Corrosion Processes Through the Superposition of Electrochemical Partial Processes and on the Potential of Mixed Electrodes, Corrosion, Vol. 62, 2006, pp. 844855. Electrochemical Techniques, R. Baboian, Ed., NACE, Houston, 1986. Corrosion Tests and Standards, R. Baboian, Ed., ASTM International, West Conshohocken, PA, 1995. Kendig, M. and Mansfeld, F., Corrosion Rates from Impedance Measurements: An Improved Approach for Rapid Automatic Analysis, Corrosion (Houston), Vol. 39, 1983, pp. 466467. Lorenz, W. J. and Mansfeld, F., Determination of Corrosion Rates by Electrochemical DC and AC Methods, Corros. Sci., Vol. 21, 1981, pp. 647672. Impedance Spectroscopy, J. R. Macdonald, Ed., John Wiley and Sons, New York, 1987. Tait, W. S., An Introduction to Electrochemical Testing for Practicing Engineers and Scientists, PairODocs Publications, Racine, WI, 1994. Cottis, R. A., Turgoose, S., and Newman, R. C., Corrosion Testing Made Easy: Impedance and Noise Analysis, B. C. Syrett, Ed., NACE, Houston, TX, 1999. Bertocci, U., Frydman, J., Gabrielli, C., Huet, F., and Keddam, M., Analysis of Electrochemical Noise by Power Spectral Density Applied to Corrosion Studies. Maximum Entropy Method or Fast Fourier Transform?, J. Electrochem. Soc., Vol. 145, 1998, pp. 27802786. Bertocci, U., Gabrielli, C., Huet, F., and Keddam, M., Noise Resistance Applied to Corrosion Measurements. I. Theoretical Analysis, J. Electrochem. Soc., Vol. 144, 1997, pp. 3137. Bertocci, U., Gabrielli, C., Huet, F., Keddam, M., and Rousseau, P., Noise Resistance Applied to Corrosion Measurements. II. Experimental Tests, J. Electrochem. Soc., Vol. 177, 1997, pp. 3743. Bertocci, U. and Huet, F., Noise Resistance Applied to Corrosion Measurements. III. Inuence of the Instrumental Noise on the Measurements, J. Electrochem. Soc., Vol. 144, 1997, pp. 2786 2793. Cottis, R. A., Interpretation of Electrochemical Noise Data, Corrosion (Houston), Vol. 57, 2001, pp. 265285. Huet, F., Listening to Corrosion, Interface (USA) 10, 2001, pp. 4043. Mansfeld, F. and Xiao, H., Electrochemical Noise Analysis of Iron Exposed to NaCl Solutions of Different Corrosivity, J. Electrochem. Soc., Vol. 140, 1993, pp. 22052209. Cottis, R. A., Sources of Electrochemical Noise in Corroding Systems, Russ. J. Electrochem., Vol. 42, 2006, pp. 557566. Hagensen, A. R. and Edgemon, G. L., A Multifunction Corrosion Monitoring System for Nuclear Waste Storage, Paper 07364, in Corrosion/2007, NACE, Houston, TX, 2007. Bosch, R. W., Hubrecht, J., Bogaerts, W. F., and Syrett, B. C., Electrochemical Frequency Modulation: A New Electrochemical Technique for Online Corrosion Monitoring, Corrosion (Houston), Vol. 57, 2001, pp. 6070. Kus, E. and Mansfeld, F., An Evaluation of the Electrochemical Frequency Modulation EFM Technique, Corros. Sci., Vol. 48, 2006, pp. 965979. Bohni, H., Suter, T., and Schreyer, A., Micro- and Nanotechniques to Study Localized Corrosion, Electrochim. Acta, Vol. 40, 1995, pp. 13611368.

24 JOURNAL OF ASTM INTERNATIONAL

34 35 36 37 38 39 40 41 42 43 44 45 46

47 48

49

50 51 52

53 54 55 56 57

58 59

Suter, T. and Alkire, R. C., Microelectrochemical Studies of Pit Initiation at Single Inclusions in Al 2024-T3, J. Electrochem. Soc., Vol. 148, 2001, pp. B36B42. Suter, T. and Bohni, H., Microelectrodes for Corrosion Studies in Microsystems, Electrochim. Acta, Vol. 47, 2001, pp. 191199. Suter, T., Peter, T., and Bohni, H., Microelectrochemical Investigations of MnS Inclusions, Mater. Sci. Forum, 1995, pp. 2540. Landolt, D., Corrosion and Surface Chemistry of Metals, EPFL Press, Lausanne, 2007. Kaesche, H., Metallic Corrosion, NACE, Houston TX, 1985. Fontana, M. G., Corrosion: A Compilation, The Press of Hollenback, Columbus, OH, 1957. Kruger, J. and Calvert, J. P., Ellipsometric-Potentiostatic Studies of Iron Passivity, J. Electrochem. Soc., Vol. 114, 1967, pp. 4349. MacDougall, B. and Cohen, M., Anodic Oxide Films on Nickel in Acid Solutions, J. Electrochem. Soc., Vol. 123, 1976, pp. 191197. Sato, N. and Cohen, M., The Kinetics of Anodic Oxidation of Iron in Neutral Solution, J. Electrochem. Soc., Vol. 111, 1964, pp. 512519. Burstein, G. T. and Marshall, P. I., Growth of Passivating Films on Scratched 304L Stainless Steel in Alkaline Solution, Corros. Sci., Vol. 23, 1983, pp. 125137. Burstein, G. T. and Newman, R. C., Anodic Behaviour of Scratched Silver Electrodes in Alkaline Solution, Electrochim. Acta, Vol. 25, 1980, pp. 10091013. Burstein, G. T. and Davies, D. H., Analysis of Anodic Films Formed on Cobalt in Bicarbonate and Borate Electrolytes, Corros. Sci., Vol. 20, 1980, pp. 989995. Bardwell, J. A., MacDougall, B., and Graham, M. J., Use of 18O/SIMS and Electrochemical Techniques to Study the Reduction and Breakdown of Passive Oxide Films on Iron, J. Electrochem. Soc., Vol. 135, 1988, pp. 413418. Burstein, G. T. and Davenport, A. J., The Current-Time Relationship During Anodic Oxide Film Growth Under High Electric Field, J. Electrochem. Soc., Vol. 136, 1989, pp. 936941. Kelly, R. G. and Newman, R. C., Conrmation of the Applicability of Scratched Electrode Techniques for the Determination of Bare Surface Current Densitites, J. Electrochem. Soc., Vol. 137, 1990, pp. 357358. Frankel, G. S., Jahnes, C. V., Brusic, V., and Davenport, A. J., Repassivation Transients Measured with the Breaking-Electrode Technique on Aluminum Thin-Film Samples, J. Electrochem. Soc., Vol. 142, 1995, pp. 22902295. Frankel, G. S., Pitting Corrosion of Metals: A Review of the Critical Factors, J. Electrochem. Soc., Vol. 145, 1998, pp. 21862197. Pessall, N. and Liu, C., Determination of Critical Pitting Potentials of Stainless Steel in Aquious Chloride Environments, Electrochim. Acta, Vol. 16, 1971, pp. 19872003. ASTM Standard F746-87, Standard Test Method for Pitting or Crevice Corrosion of Metallic Surgical Implant Materials, Annual Book of ASTM Standards, Vol. 13.01, ASTM International, West Conshohocken, PA, 2000. Schmutz, P. and Frankel, G. S., Characterization of AA2024-T3 by Scanning Kelvin Probe Force Microscopy, J. Electrochem. Soc., Vol. 145, 1998, pp. 22852294. Schmutz, P. and Frankel, G. S., Corrosion Study of AA2024-T3 by Scanning Kelvin Probe Force Microscopy and In Situ AFM Scratching, J. Electrochem. Soc., Vol. 145, 1998, pp. 22982306. Frankel, G. S. and Leblanc, P., Studies of Corrosion Using Scanning Kelvin Probe Force Microscopy and AFM Scratching, Corros. Sci. Technol., Vol. 31, 2002, pp. 419425. Iannuzzi, M. and Frankel, G. S., Inhibition of AA2024-T3 Corrosion by Vanadates: An AFM Scratching Investigation, Corrosion (Houston), Vol. 63, 2007, pp. 672688. Schmutz, P. and Frankel, G. S., Inuence of Dichromate Ions on Corrosion of Pure Aluminum and AA2024-T3 in NaCl Solution Studied by AFM Scratching, J. Electrochem. Soc., Vol. 146, 1999, pp. 44614472. Akashi, M., Nakayama, G., and Fukuda, T., in Corrosion/98, NACE, Houston, TX, 1998, Paper 98158. Shibata, T., Development in the Concept of Repassivation Potential as a Measure of Crevice Corrosion Susceptibility, in A Compilation of Special Topic Reports, F. M. G. Wong and J. H.

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 25

60

61

62 63 64 65 66

67 68

69 70 71 72 73 74 75 76 77

78 79 80 81 82

Payer, Eds., Dept. of Energy, Washington, DC, 2002, http://www.ocrwm.doe.gov/documents/ peer_2/index.pdf. Evans, K. J., Yilmaz, A., Day, S. D.l., Wong, L. L., Estill, J. C., and Rebak, R. B., Using Electrochemical Methods to Determine Alloy 22s Crevice Corrosion Repassivation Potential, JOM, Vol. 57, 2005, pp. 5661. ASTM Standard G79-01, Standard Guide for Crevice Corrosion Testing of Iron-Base and NickelBase Stainless Alloys in Seawater and Other Chloride-Containing Aqueous Environments, Annual Book of ASTM Standards, Vol. 3.02, ASTM International, West Conshohocken, PA, 2004. Frankel, G. S., Stockert, L., Hunkeler, F., and Boehni, H., Metastable Pitting of Stainless Steel, Corrosion (Houston), Vol. 43, 1987, pp. 429436. Akiyama, E. and Frankel, G. S., The Inuence of Dichromate Ions on Al Dissolution Kinetics in Articial Crevice Electrode Cells, J. Electrochem. Soc., Vol. 146, 1999, pp. 40954100. Frankel, G. S., The Growth of 2D Pits in Thin Film Aluminum, Corros. Sci., Vol. 30, 1990, pp. 12031218. Tester, J. W. and Isaacs, H. S., Diffusional Effects in Simulated Localized Corrosion, J. Electrochem. Soc., Vol. 122, 1975, pp. 14381445. Gaudet, G. T., Mo, W. T., Hatton, T. A., Tester, J. W., Tilly, J., Isaacs, H. S., and Newman, R. C., Mass Transfer and Electrochemical Kinetic Interactions in Localized Pitting Corrosion, AIChE J., Vol. 32, 1986, pp. 949958. Isaacs, H. S. and Newman, R. C., Dissolution Kinetics During Localized Corrosion, in Corrosion and Corrosion Protection, R. P. Frankenthal and F. Mansfeld, Eds., ECS, Pennington, NJ, 1981. Newman, R. C., The Dissolution and Passivation Kinetics of Stainless Alloys Containing MolybdenumI. Coulometric Studies of Fe-Cr and Fe-Cr-Mo Alloys, Corros. Sci., Vol. 25, 1985, pp. 331339. Newman, R. C., The Dissolution and Passivation Kinetics of Stainless Alloys Containing MolybdenumII. Dissolution Kinetics in Articial Pits, Corros. Sci., Vol. 25, 1985, pp. 341350. Zhang, W. and Frankel, G. S., Anisotropy of Localized Corrosion in AA2024-T3, Electrochem. Solid-State Lett., Vol. 3, 2000, pp. 268270. Sehgal, A., Frankel, G. S., Zoofan, B., and Rokhlin, S., Pit Growth Study in Al Alloys by the Foil Penetration Technique, J. Electrochem. Soc., Vol. 147, 2000, pp. 140148. Hunkeler, F. and Bohni, H., Determination of Pit Growth Rates on Aluminum Using a Metal Foil Technique, Corros. Sci., Vol. 37, 1981, pp. 645650. Frankel, G. S., Scully, J. R., and Jahnes, C. V., Repassivation of Pits in Aluminum Thin Films, J. Electrochem. Soc., Vol. 143, 1996, pp. 18341840. Frankel, G. S., Dukovic, J. O., Rush, B. M., Brusic, V., and Jahnes, C. V., Pit Growth in NiFe Thin Films, J. Electrochem. Soc., Vol. 139, 1992, pp. 21962201. Sehgal, A., Lu, D., and Frankel, G. S., Pitting in Aluminum Thin Films: Supersaturation and Effects of Dichromate Ions, J. Electrochem. Soc., Vol. 145, 1998, pp. 28342840. Shih, H. and Mansfeld, F., A Fitting Procedure for Impedance Sectra Obtained for Cases of Localized Corrosion, Corrosion (Houston), Vol. 45, 1989, pp. 610614. Brusic, V., Russak, M., Schad, R., Frankel, G., Selius, A., and DiMilia, D., Corrosion of Thin Film Magnetic Disk: Galvanic Effects of the Carbon Overcoat, J. Electrochem. Soc., Vol. 136, 1989, pp. 4246. Mansfeld, F. and Kenkel, J. V., Electrochemical Monitoring of Atmospheric Corrosion Phenomena, Corros. Sci., Vol. 16, 1976, pp. 111122. Mansfeld, F., Monitoring of Atmospheric Corrosion Phenomena with Electrochemical Sensors, J. Electrochem. Soc., Vol. 135, 1988, pp. 13541358. Cox, A. and Lyon, S. B., Electrochemical Study of the Atmospheric Corrosion of Mild Steel-I. Experimental Method, Corros. Sci., Vol. 36, 1994, pp. 11671176. Cox, A. and Lyon, S. B., Electrochemical Study of the Atmospheric Corrosion of Mild Steel-III. The Effect of Sulphur Dioxide, Corros. Sci., Vol. 36, 1994, pp. 11931199. Cox, A. and Lyon, S. B., Electrochemical Study of the Atmospheric Corrosion of Iron-II. Cathodic and Anodic Processes on Uncorroded and Pre-Corroded, Corros. Sci., Vol. 36, 1994, pp. 1177 1192.

26 JOURNAL OF ASTM INTERNATIONAL

Zhang, S. H. and Lyon, S. B., Anodic Processes on Iron Covered by Thin, Dilute Electrolyte Layers I Anodic Polarization, Corros. Sci., Vol. 36, 1994, pp. 12891307. 84 Zhang, S. H. and Lyon, S. B., Anodic Processes on Iron Covered by Thin, Dilute Electrolyte Layers II a.c. Impedance Measurements, Corros. Sci., Vol. 36, 1994, pp. 13091321. 85 Stratmann, M. and Streckel, H., On the Atmospheric Corrosion of Metals, Which are Covered with Thin Electrolyte Layers Part 1: Verication of the Experimental Technique, Corros. Sci., Vol. 30, 1990, pp. 681696. 86 Stratmann, M. and Streckel, H., On the Atmospheric Corrosion of Metals, Which are Covered with Thin Electrolyte Layers Part 2: Experimental Results, Corros. Sci., Vol. 30, 1990, pp. 697714. 87 Stratmann, M., Streckel, H., Kim, K. T., and Crockett, S., On the Atmospheric Corrosion of Metals, Which are Covered with Thin Electrolyte Layers Part 3: The Measurement of Polarization Curves on Metal Surfaces Which are Covered by Thin Electrolyte Layers, Corros. Sci., Vol. 90, 1990, pp. 715734. 88 Frankel, G. S., Stratmann, M., Rohwerder, M., Michalik, A., Maier, B., Dora, J., and Wicinski, M., Potential Control Under Thin Aqueous Layers Using a Kelvin Probe, Corros. Sci., Vol. 49, 2007, pp. 20212036. 89 Nishikata, A., Ishihara, Y., Hayashi, Y., and Tsuru, T., Inuence of Electrolyte Layer Thickness and pH on the Initial Stage of the Atmosphere Corrosion of Iron, J. Electrochem. Soc., Vol. 144, 1997, pp. 12441252. 90 Nishikata, A., Ishihara, Y., and Tsuru, T., An Application of Electrochemical Impedance Spectroscopy to Atmospheric Corrosion Study, Corros. Sci., Vol. 37, 1995, pp. 897911. 91 Nishikata, A., Yamashita, Y., Katayama, H., Tsuru, T., Usami, A., Tanabe, K., and Mabuchi, H., An Electrochemical Impedance Study on Atmospheric Corrosion of Steels in a Cyclic Wet-Dry Condition, Corros. Sci., Vol. 37, 1995, pp. 20592069. 92 Vera Cruz, R. P., Nishikata, A., and Tsuru, T., AC Impedance Monitoring of Pitting Corrosion of Stainless Steel Under a Wet-Dry Cyclic Condition in Chloride-Containing Environment, Corros. Sci., Vol. 38, 1996, pp. 13971406. 93 Vera Cruz, R. P., Nishikata, A., and Tsuru, T., Pitting Corrosion Mechanism of Stainless Steels Under Wet-Dry Exposure in Chloride-Containing Environments, Corros. Sci., Vol. 40, 1998, pp. 125139. 94 Tsutsumi, Y., Nishikata, A., and Tsuru, T., Anodic Behavior of Iron in Acid Solutions, J. Electrochem. Soc., Vol. 152, 2005, pp. B358B363. 95 ASTM, Standard G34-01, Standard Test Method for Exfoliation Corrosion Susceptibility in 2xxx and 7xxx Series Aluminum Alloys EXCO Test , Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA, 2004. 96 Lee, S. and Lifka, B. W., Modication of the EXCO Test Method for Exfoliation Corrosion Susceptibility in 7XXX, 2XXX and Aluminum-Lithium Alloys, in New Methods for Corrosion Testing of Aluminum Alloys, ASTM STP 1134, V. S. Agarwala and G. M. Ugiansky, Eds., ASTM International, West Conshohocken, PA, 1992. 97 Evans, D. G. and Jeffrey, P. W., Exfoliation Corrosion of AlZnMg Alloys, U.R. Evans Conference on Localized Corrosion, NACE-3, Houston, TX, 1974. 98 Haynes, G. S. and Baboian, R., Modied Salt Spray Fog Testing, Laboratory Corrosion Tests and Standards, in ASTM Standard G85-85 A2, STP 866, ASTM International, West Conshohocken, PA, 1985. 99 Lifka, B. W. and Sprowls, D. O., An Improved Exfoliation Test for Aluminum Alloys, Corrosion (Houston), Vol. 22, 1966, pp. 715. 100 Sprowls, D. O., Walsh, J. D., and Shumaker, M. B., Simplied Exfoliation Testing of Aluminum Alloys, in Localized Corrosion-Cause of Metal Failure, ASTM STP 516, ASTM International, West Conshohocken, PA, 1972. 101 Braun, R., Exfoliation Corrosion Testing of Aluminum Alloys, Br. Corros. J., London, Vol. 30, 1995, pp. 203208. 102 Lifka, B. W. and Sprowls, D. O., Relationship of Accelerated Test Methods for Exfoliation Resistance in 7xxx Series Aluminum Alloys with Exposure to a Seacoast Atmosphere, in ASTM Special Technical Publication 558, ASTM International, West Conshohocken, PA, 1973.

83

FRANKEL ON ELECTROCHEMICAL TECHNIQUES IN CORROSION 27

103 Sprowls, D. O., Summerson, T. J., and Loftin, F. E., Exfoliation Corrosion Testing of 7075 and 7178 Aluminum Alloys-Interim Report on Atmospheric Exposure Tests, ASTM International, West Conshohocken, PA, 1973. 104 Liddiard, E. A. G., Whittaker, J. A., and Farmery, H. K., The Exfoliation Corrosion of Aluminum Alloys, J. Inst. Met., Vol. 89, 1960-1961, pp. 377384. 105 Zhao, X. and Frankel, G. S., Effects of RH, Temper and Stress on Exfoliation Corrosion Kinetics of AA7178, Corrosion (Houston), Vol. 62, 2006, pp. 256266. 106 Zhao, X. and Frankel, G. S., Quantitative Study of Exfoliation Corrosion: Exfoliation of Slices in Humidity Technique, Corros. Sci., Vol. 49, 2007, pp. 920938. 107 Knipping, E. M., Lakin, M. J., Foster, K. L., Jungwirth, P., Tobias, D. J., Gerber, R. B., Dabdub, D., and Finlayson-Pitts, B. J., Experiments and Simulations of Ion-Enhanced Interfacial Chemistry on Aqueous NaCl Aerosols, Science, Vol. 288, 2000, pp. 301306. 108 Pistorius, P. C., in 14th International Corrosion Council, published on CD-ROM Capetown, S.A., 1999. 109 Kang, J. and Frankel, G. S., Potentiostatic Pulse Testing for Assessment of Early Coating Failure, Z. Phys. Chem., Vol. 219, 2005, pp. 15191538. 110 Rohwerder, M., Leblanc, P., Frankel, G. S., and Stratmann, M., Application of Scanning Kelvin Probe in Corrosion Science, in Analytical Methods in Corrosion Science and Engineering, P. Marcus and F. Mansfeld, Eds., CRC Press, Boca Raton, FL, 2006. 111 Stratmann, M., 2005 W. R. Whitney Award Lecture: Corrosion Stability of Polymer-Coated MetalsNew Concepts Based on Fundamental Understanding, Corrosion (Houston), Vol. 661, 2005, pp. 11151126. 112 Zhao, J., Frankel, G. S., and McCreery, R. L., Corrosion Protection of Untreated AA2024-T3 in Chloride Solution by Chromate Conversion Coating Monitored with Raman Spectroscopy, J. Electrochem. Soc., Vol. 145, 1998, pp. 22582264. 113 Buchheit, R. G., Mamidipally, S. B., Schmutz, P., and Guan, H., Active Corrosion Protection in Ce-Modied Hydrotalcite Conversion Coatings, Corrosion (Houston), Vol. 58, 2002, pp. 314. 114 Buchheit, R. G., Cunningham, M., Jensen, H., Kendig, M. W., and Martinez, M. A., A Correlation Between Salt Spray and Electrochemical Impedance Spectroscopy Test Results for ConversionCoated Aluminum Alloys, Corrosion (Houston), Vol. 51, 1998, pp. 6172.

You might also like