Sunteți pe pagina 1din 7

Section IV: Testing for Corrosion Types

J. R. Scully, Editor

Uniform Corrosion
J a m e s A. Ellor I a n d J o h n R e p p 2

TESTING FOR UNIFORM corrosion is one of the most c o m m o n corrosion tests employed to determine material suitability in a wide variety of applications. The following is intended to be an introduction into the subject of testing for uniform corrosion, its applications, basic methodology, and limitations. The reader is referred to applicable standards fully detailing the procedures. ASTM G 15, Terminology Relating to Corrosion and Corrosion Testing, defines uniform corrosion as: "corrosion that proceeds at about the same rate over a metal surface." The standard does not define the term "about." This discussion defines uniform corrosion as equivalent material loss across the continuous surface of a material. This very rarely occurs in practice. The degree to which the loss is indeed identical across a surface, as opposed to localized at some material feature, is a direct indication of compliance with this definition of uniform corrosion. The analyses described are not applicable to test specimens that show significant localized c o r r o s i o n - - f o r example, if only 20 % of the surface area appears attacked, or if certain sections have areas of metal loss ten times that of other areas. Should the corrosion pattern show such characteristics of nonuniformity, the researchers will need to consider these factors in their final analysis. Often, this might be addressed by running and reporting on both unif or m and n o n u n i f o r m corrosion tests.

BACKGROUND AND THEORY


An appreciation of the basic principles guiding uniform corrosion aids in discussion of test methods. In general, corrosion is the net result of chemical or electrochemical reactions occurring at the surface of a metal. An example is red-rust occurring on iron exposed to the atmosphere. In most atmospheres, bright iron or steel surfaces are not favored by n a t u r e - - n a t u r e instead favors iron oxide or "rust." Most simply, uniform corrosion results when no single region or area of a metallic surface is favored over another. Over an exposure period, the active corrosion areas are evenly distributed over a surface. In localized corrosion, some feature of the material, a local crevice, defect, or surface metallurgy, acts to promote one particular site as the most active region of corrosion. 1P.E., Vice President, Corrpro, Ocean City, NJ 08226. 2Manager, Corrpro/Ocean City Research Corporation, Ocean City, NJ 08226. 205

Two aspects of corrosion should be introduced: one is the tendency for a corrosion reaction and the second is the rate at which the reaction will occur. Often, one of these principles is considered without regard for the other, leading to improper conclusions. In uniform corrosion on a metallic surface, the tendency for a corrosion reaction results from the magnitude of the electromotive force of the corrosion cells comprising the corrosion process []. In some materials, the corrosion cell may arise from the heterogeneous nature of the surface, the often discussed "local cell." Conceptually the local cell is similar to a battery, often used to discuss corrosion reactions, where oxidation occurs at an anode and reduction at a cathode. Yet unlike a battery, which is comprised of different materials immersed in a c o m m o n electrolyte, a local cell on an otherwise uniform piece of material can be related to, among other factors, surface impurities, alloy phases, or n o n u n i f o r m stress distributions. On a pure, homogeneous surface, corrosion may still occur. Practically, the corrosion rate of a pure material is often lower than a nonhomogeneous material. Corrosion is believed to occur through simultaneous oxidation and reduction reactions taking place at a c o m m o n point [2]. In either case, the tendency for the reaction can be determined by inspection of the thermodynamic equilibrium potentials of the anodic and cathodic reactions. Consider the corrosion of iron in an acidic environment. The anodic reaction is Fe = Fe 2 + 2e . The cathodic reaction is 2H + + 2e- + 1/202 = H20. The standard thermodynamic equilibrium potential at unity Fe +2 concentration for the anodic reaction is -0.44V versus the normal hydrogen electrode (NHE). The thermodynamic equilibrium potential for the cathodic reaction at unit activity for H is 1.229 V versus N H E [3]. The overall cell electromotive force (EMF) is 1.669 V. If V > 0, then the reaction has a tendency to proceed as written. In this case, iron would go into solution and oxygen would be reduced. Similar calculations can be made for different metals in solutions with varying cathodic reactions to determine the tendency for corrosion. For instance, iron would still corrode in an oxygen-free acid. Here 2H + 2e- --- H 2 at 0.0 V (NHE) and Fe = Fe 2 + 2e- at -0.44 V (NHE) results in a cell potential of 0.44 V. Fontana and Greene, among many others, provide a good description of this type of analysis [4]. While such calculations can be made to investigate the tendency for corrosion, they do not indicate the rate of corrosion. Greater cell EMF values do not necessarily indicate

206 C O R R O S I O N TESTS AND STANDARDS MANUAL


higher corrosion rates. The corrosion rate is the parameter of concern in most practical problems. The corrosion reaction itself is complex, consisting of n u m e r o u s steps. The overall reaction rate, and thus, the corrosion rate, can be no faster than the slowest of these steps (the rate determining step). Each portion of the reaction sequence imparts a resistance to the corrosion reaction, i.e., limits the corrosion current resulting from the thermodynamically favored anodic and cathodic reactions. The key rate-controlling steps are commonly referred to as concentration control and activation control [4]. Concentration control implies that the rate determining step is the rate at which reactants or products move to or from the corroding surface. Concentration control is also referred to as mass transfer or diffusion control. Common illustrations of concentration control include the ability of a reactant, e.g., dissolved oxygen, to diffuse to a steel surface in an aqueous electrolyte or the ability of a reaction product, e.g., hydroxide, to diffuse from a steel surface. High supply rates of reactants favor a higher corrosion rate; low diffusion rates of products from a surface hinder corrosion rate. Activation control refers to the rate of oxidation and reduction that takes place at a metallic surface. Activation control is also referred to as charge transfer control. For zinc in an acid environment, one reaction product is typically hydrogen gas. Given that the solution is sufficiently acidic, there is a large abundance of hydrogen ions in the solution; thus, mass transfer of hydrogen ions to the zinc surface is not usually the rate-controlling step. The ratecontrolling step can be the transfer of electrons from the zinc to the hydrogen ions to form adsorbed atomic hydrogen on the zinc surface. Charge transfer rates can vary as material properties in different chemical environments (e.g., catalytic properties). Other factors may influence uniform corrosion rates, such as the development of oxide layers on a metallic surface. In reality, such layers limit the corrosion rate by affecting diffusion and imparting concentration control. operational parameters for testing convenience provides for erroneous results. Any testing method should attempt to determine which of these parameters is key and to simulate that parameter as closely as possible during the test. It should be noted that there is often a desire to accelerate corrosion tests by increasing the corrosive stress factors, e.g., temperature or chemical concentrations. Before accepting the results from such tests, it is wise to confirm their validity, especially in light of the importance discussed above of matching the key operational conditions. In accelerated tests, it is suggested that one confirm (1) the acceleration factor and (2) that novel or u n n a t u r a l corrosion processes do not result. R u n n i n g a series of tests where the stress factors are incrementally increased and the resultant corrosion rate determined can assess acceleration factors. Severe discontinuities in this relationship or the formation of novel corrosion products or corrosion morphologies can be indications that novel or u n n a t u r a l corrosion processes are occurring. In such cases, the acceleration factors should be decreased.

M A S S LOSS T E S T S
The simplest uniform corrosion tests are those based on coupon immersion and mass loss tests. The tests provide data on the uniform metallic corrosion rate most often expressed in English units as mils per year (mpy) or in metric units as millimeters per year (mm/y). While there are various standards for such tests, the predominant methods are those prescribed by ASTM G 1, "Practice for Preparing, Cleaning, and Evaluating Corrosion Test Specimens." This standard contains a comprehensive reference list of other ASTM standards that also apply, depending on the type of tests of interest to the investigator. These mass loss tests consist of preparing metallic coupons, cleaning them before testing, weighing them before exposure, exposing them to the corrosive media, post-test removal of visible corrosion products, and reweighing. The corrosion rate is calculated from the mass loss measured, converted to a volume of metal loss by the material density, and finally to a corrosion rate by dividing this volume by the material surface area and the test time. The reader is directed to ASTM G 1 for the procedure details. The following comments apply to these standard procedures and should be considered. The test coupon size selected should be large enough to minimize the ratio of edges to general surface area. On some sensitive materials, corrosion often preferentially takes place at an edge, thereby limiting the measurement of "uniform" corrosion. There are no advantages to using small coupon sizes other than the size of the test equipment. Specimens should be as large as practically possible considering the size of the test chamber, material cost, and capacity of an analytical balance for coupon mass measurements. This author has used samples as small as 2.5-cm by 5-cm by 0.3-cm, and as large as 15-cm by 10-cm by 0.6-cm. The volume of the test media should be sufficient to ensure that the corrosion products formed or dissolved within the media do not significantly affect its properties. One may make an estimate of the required volume by estimating the

UNIFORM CORROSION TESTS


The uniform corrosion test methodologies outlined below are intended to provide the user with an overview to the process. They are centered on tests in aqueous applications, but can also be adopted to monitor uniform corrosion concerns in an atmospheric exposure. Uniform corrosion tests are only useful if the test results demonstrate that the corrosion was indeed uniform. The tests described below discuss the general procedures and areas of concern. The test results will be valid to the extent that the test simulates the potential material in-service application. Factors that affect the in-service corrosion rate by influencing the controlling thermodynamic and/or kinetic parameters that are not accounted for by the experimental design limit the usefulness of the test. Typical operational parameters of concern include, but are not limited to, exact material composition and forming techniques, service temperature, electrolyte chemistry, electrolyte velocity, time-ofwetness (wet/dry cycles), exposure duration, and mechanical forces (e.g., abrasion or stress). Often, neglect of the key

CHAPTER 1 7 - - U N I F O R M C O R R O S I O N T E S T I N G
total volume of material possibly dissolved into the media based on an approximate corrosion rate and the size of the test coupon. ASTM G 31, "Practice for Laboratory Immersion Corrosion Testing of Metals" also addresses this same issue. The exposure test(s) should be designed such that the expected mass loss of coupons is within detection limits of the measurement instrumentation. Variations to coupon dimensions and/or test duration can accomplish this. Coupons should always be large enough so that the mass loss expected over the test duration is resolvable above m i n i m u m , but does not exceed m a x i m u m detection limits. With low corrosion rates, larger coupons are necessary. Alternatively, the duration of exposure may also be changed to obtain detectable mass loss. However, this method is typically not favored, since significantly increased test durations may result. Coupons must be identifiable before and after testing. Most simply, the coupons can be identified by a mark made directly onto the sample. However, with some notchsensitive materials such as stainless steels that may have a tendency to favor localized corrosion over uniform corrosion, such a mark can modify the sample corrosion behavior. For the most sensitive materials, even a tag attached to the sample can affect the corrosion. In these cases, the materials should be identified by location within a test environment. Of course, the larger the test sample, the less of an influence some small mark will have on the overall results. The material surface should be representative of the finish expected in-service. If the exact service finish is not known, the test surface finish should be noted. For stainless steel or a l u m i n u m alloys, where corrosion p h e n o m e n a can attack specific elements of the metallic structure preferentially, the coupon forming process should also be noted (e.g., annealed, heat-treated, cold-rolled, cast, etc.). ASTM G 1 provides specific details on the accuracy of the coupon weighing procedures before and after testing. Data to the fifth significant digit are suggested. While this is good practice, the practitioner should also consider the expected corrosion rate. Often in situations where the corrosion rate is expected to be quite high, data are not required to the fifth significant digit. This can be a significant time-saver if m a n y coupons are to be tested and must be weighed on an analytical balance. The precision of electronic digital scales often makes this an unnecessary consideration. Removal of visible corrosion products, after testing, is accomplished by cleaning of the coupon surface. Cleaning is accomplished by chemical stripping, mechanical removal (i.e., abrasive blasting or brushing), and electrolytic methods to remove corrosion products. This author's experience suggests that the best methods are abrasive blasting and chemical stripping. These techniques are more likely to remove all of the visible corrosion products without harming the metallic surface. With respect to the chemical procedures described by ASTM G 1, the immersion time suggested for coupons in the alternative chemical baths should be used as a guideline only. There are m a n y times when tenacious corrosion products can take significantly longer to remove than that expressed by the standard. All corrosion products must be removed without resulting in additional corrosion of the metallic substrate for an accurate mass loss measurement.

207

Furthermore, the standard does not describe a correction procedure for mass loss of the base metal. The standard suggests that the mass loss be plotted as a function of the n u m b e r of cleaning cycles. The breakpoint in this curve is regarded as the corrosion mass loss; further mass losses are suggested to be the mass loss of base material. The standard urges that a cleaning procedure be selected such that the slope of this latter curve is "zero," i.e., the cleaning procedure is having no effect on the base metal. This is not always possible. It is suggested that the cleaning cycle time be standardized and the n u m b e r of cleaning cycles recorded. After cleaning, a clean coupon of the same base material should be subject to the same cleaning procedure and cleaning cycle. The mass loss of this material should be recorded as a function of cleaning cycles. For each cleaning cycle of the corroded specimens, an adjustment should be made for the mass loss of the base metal as a result of the cleaning procedure. Some may argue that the base material mass loss per cleaning cycle is not the same for an already clean sample and a sample covered with corrosion products. This is a valid point; yet, if the base material is significantly affected by the cleaning solution, a correction should be attempted and the procedure described is appropriate. If the base material is not significantly affected, application of this correction factor will have no effect on the test results. Abrasive grit blast methods for removal of corrosion products have increased in popularity in recent years. Some test specifications specifically list this as the preferred cleaning method [5]. Fine abrasive grit (the author has used 120 grit glass bead) is used in an air-powered blast cabinet. The surface of the coupon is blasted until removal of all visible corrosion products is observed. Mass loss and corrosion rates are then determined by standard methods. The use of abrasive grit blasting has several advantages over chemical stripping. This cleaning method eliminates the use and disposal of hazardous chemicals during the cleaning processes. Eliminating hazardous chemicals also reduces the need for personal protective equipment (PPE) and reduces potential worker health and safety issues. This method also reduces the a m o u n t of time necessary to remove all corrosion products both during the cleaning and disposal operations. The author has used this method to clean carbon steel coupons following exposure tests. Similar results were obtained from samples using this and the chemical stripping cleaning methods. Use of this method on uncorroded test samples resulted in negligible mass loss of sound base metal. Before using this method on other materials, similar comparisons between cleaning methods and removal of base metal are suggested. The duration of the test can have a significant effect on the test data. Most materials will corrode most rapidly in the early stages of exposure to an environment before oxide films are developed that may inhibit the corrosion rate by limiting the diffusion rate of reacting species to and from the metallic surface. Thus, tests of 24-h duration extrapolated to 30-day or 1-year behavior may be extremely conservative. Where time allows, multiple test durations may be evaluated to determine the actual effect of exposure time on uniform corrosion.

208

CORROSION TESTS AND STANDARDS MANUAL are usually developed from theoretical behavior of metals in an environment and are checked through actual mass-loss data for materials in the test environment. However, when evaluating a material in a new environment or novel operating condition, the assumption that similar conversion-"constants" apply can be misleading and result in erroneous conclusions. In such cases, a limited n u m b e r of mass loss tests (or other physical tests) should be r u n in parallel to the electrochemical tests to confirm the validity of the conversion"constants." Often, if the test is only intended to develop the relative order of magnitude of corrosion, as opposed to an exacting value, this is not necessary and historical "constants" may be used. With the advent of advanced electronics and computerization, electrochemical techniques have evolved rapidly. The most common technologies today are the polarization resistance technique, electrochemical impedance, and Tafel extrapolation. Regardless of the technique used, each relies on the same basic principles: in each test, a metallic coupon in an electrolyte is subject to an electrical perturbation. This perturbation is the application of a current from an external source (power supply). This current stimulates the surface corrosion reactions. The voltage (potential) response of the coupon is measured and correlated with the current applied--a galvanodynamic test. Conversely, the coupon potential is controlled and correlated with the requisite c u r r e n t - - a potentiodynamic test. In either case, the resultant current is representative of the rate determining mass transfer or charge transfer rate. This may be related to the corrosion rate. Standard test procedures are defined within ASTM standards: ASTM G 59, "Practice for Conducting Potentiodynamic Polarization Resistance Measurements"; G 5, "Standard Reference Test Method for Making Potentiostatic and Potentiodynamic Anodic Polarization Measurements"; G 106, "Practice for Verification of Algorithm and Equipment for Electrochemical Impedance Measurements"; and G 102, "Practice for Calculation of Corrosion Rates and Related Information from Electrochemical Measurements." Each of these methods describes a standard procedure or practice for the test method. A complete discussion of the technologies is beyond the scope of the current text. For the current text, the focus is on the application of the most simple and most widely used of these techniques, the polarization resistance measurement, ASTM G 59. The parameters discussed are, however, applicable concerns for all electrochemical tests. The ASTM G 59 standard experiment described determines the uniform corrosion rate of 430 stainless steel in 1.0 N H2SO 4 solution, purged with oxygen-free hydrogen. The procedure described is adequate for learning the basic polarization resistance techniques, equipment calibration, and operator training. However, as acknowledged by the standard, the techniques are not applicable to all material and environment combinations. The reader should consider the following modifications to the standard procedure when dealing with materials other than the calibration standards. Consider a metal in a flowing natural water. The standard procedure suggests establishing a coupon in a standard "polarization cell." This exposure is valid to the extent that the in-service application of the material is similar to a

One last factor that should be considered in the design of mass loss tests is metallic plated specimens, e.g., galvanized steel. It is often difficult to distinguish between the corrosion of the zinc-galvanizing and the base steel and difficult to clean the zinc corrosion products without damaging the steel. For this example, it is suggested that the test design include both bare steel and solid zinc coupons as additional materials. Coupons of the base metal and plating materials should be evaluated for other samples, e.g., a l u m i n u m and nickel for nickel-plated a l u m i n u m samples. The corrosion rates of the homogeneous materials can aid in distinguishing the corrosion of the base metal and plating materials. After completion of the test, the corrosion rate may be calculated from the mass loss as follows: Corrosion Rate = (K W)/( A x T x D) where K = constant, T = time of exposure, h, A = area, cm a, W = mass loss, g, and D = density, g/cm 3. K = 3.45 106 for a corrosion rate in mpy. Other constants include K = 8.76 104 for a corrosion rate in mm/y. Refer to ASTM G 1 for a complete listing of constants. Common material densities include about 7.9 g/cm 3 for most steels and m a n y stainless steels, 8.9 g/cm 3 for copper alloys, and 2.7 g/cm 3 for aluminum. Refer to ASTM G 1 for a detailed listing of material densities. It should be realized that the utility of the tests is limited if the corrosion is not uniform. If there is extensive localized corrosion, the mass loss data will greatly underestimate the maximum corrosion rate, as it will distribute the corrosion over the entire surface of the coupon. Where localized corrosion may be a concern, the investigator may use a micrometer or other type of sensitive instrument to measure the coupon thickness before and after test. Multiple measurements should be made to ensure that the coupon surface is well-characterized. The thickness loss in mils divided by the test duration in years should be compared to the corrosion rates calculated by the above formula. The degree to which the two results agree indicates the degree to which the coupon did indeed undergo uniform corrosion.

ELECTROCHEMICAL TESTS
Electrochemical tests are often preferred to mass loss studies as a method of measuring uniform corrosion. Such tests are preferred because in theory they: (1) provide a realtime measurement of the metallic corrosion rate, (2) can provide time-corrosion rate data on a single coupon, and (3) are rapid to perform. Disadvantages of the electrochemical tests include the requirements for comparatively expensive equipment (versus mass loss tests) and higher levels of technical expertise for data analysis. Furthermore, the data reduction requires the use of conversion-"constants," factors applied to the results of the electrical measurements to convert the data to a corrosion rate. These "constants"

CHAPTER 1 7 - - U N I F O R M C O R R O S I O N T E S T I N G
polarization cell. If the electrolyte is expected to flow past the metallic surface, this will influence the local mass transfer rate, and data from a polarization cell will be of little utility if either the anodic or cathodic reaction is affected by mass transport. Agitation in the polarization cell by a stirring mechanism is perhaps better than testing u n d e r completely quiescent conditions, yet is no guarantee that the data will be representative of the actual field conditions. Experience and correlation of the test data with field results provide the best confirmation of the test setup. The test setup concerns are not limited to electrolyte flow but include electrolyte chemistry, temperature, material preparation, etc. As with mass loss tests, the closer one comes to duplicating the in-service conditions, the more confidence can be placed in the resultant data. For a more complete discussion of uniform corrosion testing u n d e r mass transport controlled conditions, the reader is referred to the chapter on electrochemical testing in this manual. The standard procedure suggests that sample potential be monitored through 55 m i n of exposure before initiating the scan. This is not always an adequate time for coupon stabilization. The coupon potential versus time should be monitored for a sufficient time (even greater than 24 h if necessary) to ensure stability before beginning the scan. Stability is a relative term. However, consider that the testing will polarize (perturb) the material only a few millivolts (10-20) around its corrosion potential in a period of a few minutes. This suggests that stability is a variation of - 1 - 2 mV (less than 10 %) per the anticipated polarization resistance test period. While some may argue that the requirement for a long stabilization period eliminates the ability to get rapid data early in the exposure, it should be recalled that the procedure relies on measuring the coupon response from an external perturbation. If the natural perturbations are of the same magnitude as those imposed on the sample (external), the test data are difficult to interpret. In some cases, if the natural perturbation can be documented to be a consistent factor, this input can be subtracted from the effects of the external perturbation, and polarization resistance tests can be r u n before complete stabilization. Such data interpretation is beyond the current discussion. The standard procedure suggests that the polarization scan be initiated by polarizing the material 30 mV active. The scan is then r u n through the corrosion potential (previously measured stabilized potential) to 30 mV noble to this value. This range may be appropriate for the standard test, but can be significantly too m u c h for m a n y materials in a natural environment. In many natural environments, the actual external perturbation should be limited to -+10 mV about the corrosion potential. It is also interesting to note that the standard procedure suggests that the scan be initiated by shifting the coupon potential 30 mV active and then scanning in a noble direction at 0.6 V/h. The standard prescribes the scan rate from the cathodic to the anodic side of the corrosion potential but not a time for polarizing the material 30 mV active. This appears contradictory. Scan rate is very important in these experiments. If the sample is "immediately" driven to a potential 30 mV active with a potentiostatic input, the sample surface can be strongly influenced. In m a n y "real-life" experiments the procedures described are not adequate. It is

209

suggested that the scan be started from the corrosion potential at a standard scan rate in the active direction until the sample is polarized 10 mV active (versus 30 mV). At this point, the scan should be reversed in polarity, and the material polarized until it is 10 mV noble to the initial corrosion potential. At all times the scan rate is kept the same. The polarization resistance value is obtained as described by ASTM G 59. With respect to the actual value of scan rate, the ASTM standard suggests a value of 0.6 V/h. Again, this is adequate for the standard test b u t may not be for a wider variety of materials. The primary objective of the scan rate is to p e r t u r b the sample yet m i n i m i z e any "capacitive" effects at each point along the scan. If a p o t e n t i o d y n a m i c scan is r u n too fast, the a m o u n t of c u r r e n t required to "hold" the indicated potential is too high. Thus, the calculated polarization resistance value is too low. One simple way of determining an appropriate scan rate is to use a scan rate slow enough so that if the scan is stopped (and the applied potential or current is maintained constant), the current does not decrease with time u n d e r a potential controlled scan, nor does the potential continue to shift u n d e r a current controlled scan. A little trial and error will determine the optimal scan rate. Often times it can be much faster than 0.6 V/h, especially in well-aerated, flowing solutions. Under quiescent conditions with oxidized materials, the necessary scan rate is m u c h slower than 0.6 V/h. A detailed discussion on effects of scan rate, ohmic voltage error, choice of over-potential, and other factors appears in the chapter on electrochemical test methods in this manual. Under any conditions, the choices of scan rate and magnitude of the potential scans about the corrosion potential dictate the individual test duration. With multiple samples and frequent testing, a significant a m o u n t of labor and time can be consumed in the testing. The test designer must balance the alternatives of the test matrix with the available testing time and budget. Following acquisition of the polarization resistance value, a corrosion rate can be calculated by ASTM G 102. Briefly, Corrosion rate = (K x i~or~XEW)/(D) where Corrosion rate = mm/y, icon. m corrosion current density, pA/cm 2, K = 3.27 x 10 -3, m m g/~A cm yr, EW= equivalent weight = atomic weight of the primary alloy element/number of electrons involved in the oxidation process, D = material density, g/cm 3, ico~ = B/Rp where B = Stern-Geary constant, mV, and Rp = polarization resistance, g2-cm2. The derivation of the value for the term B is a central focus of ASTM G 102 and need not be repeated within the current text. As an approximate value for quick calculation only, B often is about 20 mV. E W values include 9 for alum i n u m alloys, 50 for copper alloys, and 27 for carbon steel.

210 C O R R O S I O N TESTS AND S T A N D A R D S MANUAL


Refer to ASTM G 102 for more precise values and determination methods for multi-element alloys. Yet these data can be used to estimate the corrosion rate predicted by the Rp value obtained. It is also noted that Scully provides an excellent s u m m a r y of electrochemical tests in Section HI: Type of Tests in this text. environments. The primary advantage of the polarization resistance probes is that they can provide a rapid indication of the corrosion rate on a real-time basis. As with any polarization resistance technique, the procedure assumes that the predominant corrosion mode is uniform attack and not pitting. Polarization resistance-type probes are not suited for monitoring where pitting corrosion is of concern. On both types of probes, one additional concern has to be the presentation of the probes to the environment being monitored. In typical installations, the probes are used as through wall fittings in process piping. The flow cbal-ac:eristics imparted on the probes is not the same as that ob,crved on the pipe walls, and therefore, the corrosion rate~ nay not be similar. This is less of a problem in process vessels, yet is always a concern. ASTM G 96 describes additional precautions in the use of the probes.

OTHER TECHNIQUES
There are m a n y other techniques for measuring corrosion. Some of these may rely on chemical solution analysis or physical measurement of metal loss. One widely used additional technique is provided by ASTM G 96, Practice for On-Line Monitoring of Corrosion in Plant Equipment (Electrical and Electrochemical Methods). This guide covers two basic procedures. In one, a sample of the material, usually in the form of a continuous wire, is immersed into the envir o n m e n t of interest (liquid, air, solid, or multiphase). The electrical resistance through the wire is determined. As corrosion consumes the cross section of the wire, the resistance increases proportionately. The second procedure uses the polarization resistance technique described in ASTM G 59 to determine the corrosion rate in the environment. Both of these techniques have their utility and drawbacks. The resistance probe is limited because it is difficult to determine if the electrical resistance increases being monitored are the result of localized or uniform corrosion without visual inspection of the probe. However, the resistance probe is of great use in that it may be used in multiphase environments, whereas the polarization resistance probe is mostly limited to use in conductive aqueous

REFERENCES
[i] Uhlig, H. H. and Revie, R. W., "Corrosion Tendency and Electrode Potentials," Corrosion and Corrosion Control, An Introduction to Corrosion Science and Engineering, 3rd ed., John Wiley and Sons, New York, 1985. [2] Bockris, J. O'M. and Reddy, A. K. N., Modem Electrochemistry, Plenum Press, New York. [3] Handbook of Chemistry and Physics, 68th ed., R. C. Weast, Ed., CRC Press, Boca Raton, FL, 1987. [4] Fontana and Greene, Corrosion Engineering, McGraw-Hill Book Company, New York, 1987. [5] General Motors Corporation, "GM 9540P, Accelerated Corrosion Testing," September 1997.

S-ar putea să vă placă și