Vous êtes sur la page 1sur 7

Types of Data

D a v i d C. S i l v e r m a n ~

~------

:=::== . . . . . .

::::::::::::::::::::::::::::::::::::::

CORROSION TESTS ARE performed to provide information on material degradation in specific environments, information that is not available from other sources. Corrosion testing can be divided into two broad categories, electrochemical and nonelectrochemical. Within these categories, test results are presented in a n u m b e r of ways, ranging from numerical output to qualitative examination of the test specimen. Both types of data are represented in each category. A very complete listing of data formats for collection and compilation for computerized databases is presented in ASTM G 107, Guide for Formats for Collection and Compilation of Corrosion Data for Metals for Computerized Database Input. The information in that standard includes a significant portion of the type of test data that might be recorded during corrosion testing. Since required data are test-dependent, not every test would include all of the data listed, but the standard does provide a reasonable, first pass checklist and should be used to ensure that needed information is not overlooked. The purpose of this chapter is to provide an overview of types of corrosion data for metals and alloys that might be obtained from or be relevant to different types of corrosion tests. The examples discussed are not meant to be all-inclusive. They are meant to provide a flavor of the types of data that might be recorded. The reader is strongly encouraged to refer to the references and cited standards for a more complete discussion of the types of data relevant to the test discussed.

ELECTROCHEMICAL TEST DATA


The m a i n variables that are measured in an electrochemical test are the voltage and the current. The goal is to translate this information into a corrosion rate or some other information that describes the corrosion process. ASTM G 3, Practice for Conventions Applicable to Electrochemical Measurements in Corrosion Testing, provides guidelines for conventions for reporting a myriad of electrochemical data for the more c o m m o n tests.

electrode measured at open circuit in an electrolyte. The voltage is measured relative to a reference electrode. For example, ASTM G 69, Practice for Measurement of Corrosion Potentials of A l u m i n u m Alloys, describes how to measure the corrosion potential of a l u m i n u m alloys, and ASTM G 82, Guide for Development and Use of a Galvanic Series for Predicting Galvanic Corrosion Performance, describes the development and use of a practical galvanic series for predicting relative corrosion performance. Very often, the attempt is made to relate the measured potential to potentials calculated from thermodynamic data, e.g., as presented in Pourbaix Diagrams [1], in order to understand better the corrosion mechanism. One reason this approach requires caution is that the measured corrosion potential is at best a steady state potential in which reactions are not at equilibr i u m and the potential on the Pourbaix diagram is a thermodynamic or equilibrium potential. Confusion sometimes exists with the sign convention when reporting voltage with respect to the reference electrode. The thermodynamic electrochemical potential of a reaction written as oxidizing (M-->Mn++ne-, where M is the metal and e is the electron) has the opposite sign as when the same reaction is written as reducing (Mn++ne-~M). The recommended approach is to use the Stockholm sign invariant convention. I n this convention, the positive direction (increasing potential) implies increasingly oxidizing conditions at the electrode surface (noble). The negative direction (decreasing potential) implies increasingly reducing conditions at the electrode surface (active). For example, the potential of the gold couple with its ion would be more positive than the couple of zinc with its ion. Care must be taken when recording the potential of a specimen in an electrolyte relative to a reference electrode so that the recorded potentials are consistent with this convention. One can always make sure that the connection is correct, i.e., the leads are hooked up properly, by checking against a known couple (see ASTM G 3).

Current or Current D e n s i t y Potential


One very common electrochemical measurement is the corrosion potential. This potential is the voltage of the corroding 1principal Consultant, Argentum Solutions, Inc., 14314 Strawbridge Ct., Chesterfield, MO 63017, e-marl: dcsflverman@argentums~176176 web site:www.argentumsolutions.com. Measurement of current or current density is the most c o m m o n output of electrochemical corrosion tests. This quantity is usually related either to the corrosion rate or to some features of the corrosion process, such as surface redox reactions that can change the corrosion characteristics. The relationship between the measured current and the corrosion rate or m e c h a n i s m depends to some degree on

59
Copyright* 2005 by ASTM International www.astm.org

60

C O R R O S I O N T E S T S AND S T A N D A R D S MANUAL
Related Information from Electrochemical Measurements, provides the methodology and equations. The method is summarized as follows. The slope actually contains contributions from the polarization resistance (shown) and the uncompensated resistance. The latter term contains several contributions, two of which are the electrical resistance in the solution between the reference electrode sensing point and the working electrode and the electrical resistance of the leads and measuring circuit. This u n c o m p e n s a t e d solution resistance m u s t be subtracted from the resistance calculated from the slope to obtain the actual polarization resistance. The corrosion current is estimated from this corrected value. The procedure is to divide the Stern Geary constant by the polarization resistance. ASTM G 102 and I ~ Refs 2 and 3 provide details on how the Stern Geary constant might be estimated. If the current m e a s u r e m e n t is in current, e.g., amp, and not current per unit area of working electrode, e.g., amp/cm 2, the corrosion current density is estimated from the corrosion current by dividing it by the exposed specimen surface area. The specimen area must be measured. Finally, the corrosion rate as a mass loss is obtained from the corrosion current density. The corrosion current density is divided by the density of the alloy and is multiplied by the alloy equivalent weight and a constant to obtain the corrosion rate in the proper units. The alloy equivalent weight is the reciprocal of the sum taken over all of the elements in the alloy of the valence of the element times its mass fraction divided by its atomic weight. The reader is referred to ASTM G 102 for the details and justification of using the alloy equivalent weight. Thus, more information than that in Fig. 1 is needed for the corrosion rate to be estimated from polarization resistance measurements. A second type of presentation of current is as the plot of voltage versus logarithm of the current density (current divided by the surface area). ASTM G 5, Standard Reference Test Method for Making Potentiostatic and Potentiodynamic Anodic Polarization Measurements, and ASTM G 61, Test Method for Conducting Cyclic Potenfiodynamic Polarization
i.2

~
o
u.i r

Stope- P~

II

(-) Cathods162 enl~F CuJ~'

AnodLc Currenl:

0enss

FIG. lmHypothetical polarization resistance or linear polarization plot (from ASTM G 3).

the technique used to obtain the current. The type of data obtained and the format by which the information is displayed depend on the test. Some tests display voltage versus current, others display voltage versus the logarithm of the current, others display information about the impedance (voltage divided by the current), and others examine fluctuations in the current a r o u n d some average value. These different types of current data have different uses and m a y require additional types of data or analysis programs to translate the data into usable information. Polarization resistance or linear polarization measurements are used to determine the polarization resistance, a quantity related to the corrosion rate. Current is measured as a function of vokage in the vicinity of the corrosion potential. The two values are plotted as voltage versus current. The polarization resistance is the slope of the curve at the corrosion potential. The importance of this quantity lies in the fact that it is (usually) inversely proportional to the corrosion current, the proportionality constant being related to the Tafel slopes. ASTM G 59, Practice for Conducting Potentiodynamic Polarization Resistance Measurements, describes a m e t h o d by which these measurements are made a n d the polarization resistance is calculated. Figure 1 shows a hypothetical plot as a function of current. A very extensive discussion of the historical development, theory and assumptions, and pitfalls provides an excellent background to ASTM G 59 [2]. Additional experimental artifacts have also been discussed [3]. Data in addition to the slope as estimated from a curve, such as in Fig. 1, are required when calculating the corrosion rate from polarization resistance measurements. ASTM G 102, Practice for Calculation of Corrosion Rates and

I.O

At.I.OYC-2/1; UNSNaOZ?S .i

LQ

o
0.2

-o.s

TYPE STAINLESS 304 STEEL


UNS $30400

|o 0

io I

io,2

io ]

CNKT U R NO(NSITY ( ] , / ~ t / c m 2 )

FIG. 2mRepresentative cyclic potentiodynamic polarization curves (from ASTM G 61).

CHAPTER 2 - - T Y P E S OF DATA
Measurements for Localized Corrosion Susceptibility of Iron-, Nickel-, or Cobalt-based Alloys, show examples of these types of plots. They are generated by ramping the voltage at a preset rate in the anodic direction with respect to the corrosion potential. Sometimes, the voltage ramp is then reversed so that the potential is driven at a slow rate back to the corrosion potential. This procedure of using a forward and reverse ramp of voltage is called cyclic potentiodynamic polarization. Figure 2 shows an example of a cyclic potentiodynamic polarization scan. One of the purposes of performing this type of experiment and presenting the data in this m a n n e r is to estimate the susceptibility of an alloy to undergo localized corrosion in the form of crevice corrosion and pitting. Certain relationships among "characteristic" voltages, e.g., the difference between the "repassivation potential" and the corrosion potential and the difference between the "pitting potential" and the corrosion potential, as well as other features related to the current, can be used to make a j u d g m e n t about corrosion [4]. Both quantitative and qualitative information constitute the data that are used to interpret the polarization scan. The electrochemical reactivation technique provides a different way of presenting and using current (see ASTM G 108, Test Method for Electrochemical Reactivation (EPR) Test Method for Detecting Sensitization of AISI Type 304 and 304L Stainless Steels). This test, like the one above, involves a voltage ramp, but in this case, the voltage ramp is from the passive region of the alloy to the active region. The required test result is the integration of the current or the total charge passed. This charge is related to the degree of sensitization of the alloy. In this case, the charge is divided by both the surface area and often the average grain size so it, too, must be measured. Coupling of dissimilar metals can create a current caused by the potential difference between the two alloys. Such galvanic currents are measured by a zero resistance ammeter placed in an external circuit that connects the two alloys immersed in an electrolyte (see ASTM G 71, Guide for Conducting and Evaluating Galvanic Corrosion Tests in Electrolytes). Sometimes, the current when divided by the surface area can be used to calculate the corrosion rate from Faraday's law. Often when using this type of test the coupled specimens have different surface areas so the surface area of each specimen is a type of data. If corrosion is localized on the alloy surfaces, however, conversion of current to a massloss based corrosion rate may not be meaningful.

61

divided by the surface area, and the impedance has the units of ohm-cm 2. Details are provided in ASTM G 106, Standard Practice for Verification of Algorithm and Equipm e n t for Electrochemical Impedance Measurements. The impedance is usually plotted in one of two ways, as the inverse of the imaginary component versus the real component, as in Fig. 3a, or as a pair of plots in which the magnitude of the impedance and the inverse of the phase angle are plotted as function of the logarithm of the frequency, as in Figs. 3b and 3c (see ASTM G 106 and G 107). A plethora of literature exists that discusses how to relate the structure of the impedance spectrum to the corrosion mechanism and corrosion rate. In the limit of zero frequency, the impedance becomes equal to the polarization resistance discussed above. Situations arise in which this value is not inversely related to the corrosion rate [5]. The reader is referred to two recent symposia on electrochemical impedance spectroscopy that provide excellent "snapshots" of the state of the art in determining corrosion rates and mechanisms from the impedance spectra and provide information on the additional types of data needed to make the connection [6,7]. In addition, Ref 3 provides an overview of practical applications of this technology, as well as those mentioned previously [3].

Electrochemical Noise
The technique that has been labeled as electrochemical noise is the m e a s u r e m e n t and analysis of fluctuations in potential or current that arise from uncontrolled variations in a corrosion process. This technique has received much attention both for laboratory studies and on-line monitoring for localized corrosion (pitting, crevice corrosion, and stress corrosion cracking), coating degradation, and general corrosion [8]. Though no ASTM standard yet exists with respect to measurement procedure or analysis algorithm, a brief discussion of types of data associated with this technique is warranted because of the attention that this technique has received. The recorded data themselves are current and voltage recorded at constant time intervals. The method used to analyze the data points is as important as the data points themselves. The methods used for expressing the data fall into two categories, time domain techniques and frequency domain techniques. The two methods are related because frequency and time are the reciprocals of each other. The analysis technique influences the data requirements. Reference 9 provides a brief overview of the various mathematical methods and a multitude of additional references. Specialized transforms (Fourier) can be used to transfer information between the two domains. Time domain measures include the normal statistical measures such as mean, variance, third moment, skewness, fourth moment, kurtosis, standard deviation, coefficient of variance, and root m e a n square as well as an additional parameter, the ratio of the standard deviation to the root m e a n square value of the current (when measuring current noise) used in place of the coefficient of variance because the m e a n could be zero. An additional time d o m a i n measure that can describe the degree of randomness is the autocorrelation function of the voltage or current signal. The m a i n frequency domain

Impedance
Measurement of impedance of a corroding electrode has become important in corrosion prediction for such diverse applications as coatings and corrosion rate estimation in low conductivity media. As most commonly practiced, an electrode is subjected to a small amplitude (e.g., 5-10 mV) sinusoidal variation in the voltage of varying frequency, usually about the corrosion potential. The current is measured. The applied voltage is divided by this measured current. Since both the voltage and current have a sinusoidal component with respect to time and are usually out-ofphase, the division results in the impedance, which itself has real and imaginary contributions. Often, the current is

62

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


EQUlVALgNTOROJIT
100.0 -

~o
C O

J~

~.0-

50.0. &

A A

4=
&

_=

& A

==

_E
!

25.O-

&

==
A

0.0
0.0

|
,,~.0

='.0

=o'.o

7s'.o ~o;.o

Real Resistance (ohm)

measure is the power spectral density of the corrosion potential noise or the current noise. The power spectrum is the Fourier Transform of the autocorrelation function and is the distribution of power of the signal in the frequency domain. The power spectral density is usually estimated by the Fast Fourier Transform algorithm or the m a x i m u m entropy algorithm. Another transform that is presently being investigated but has not found widespread use is the wavelet transform [IO]. This transform has been proposed to overcome some of the shortcomings of the more traditional Fourier transform by using a family of basis functions that are localized in both the time and frequency domains. Debate remains as to which measures provide the most information, so all are found in the literature. The types of data m e n t i o n e d in the previous paragraph pertain to the results from analyzing fluctuations in either the voltage or current signal individually. One additional analysis is to divide the standard deviation of the voltage signal by the standard deviation of the current signal to obtain the noise resistance [11]. This parameter can be equivalent to the polarization resistance discussed i n the previous section.

I
:IE Q

,~ 10=".
& &

NONELECTROCHEMICAL
/,

TEST DATA

==
A&&AID.&&&&A&&AA &

"10 o Q.

10'

_E
lO0

1r

,I

'"i)

'l'J,

t ll,,~'ll

....

....

lo"

Fr~uenc~ (h=)

Though a significant fraction of metallic corrosion involves electrochemical processes, much of corrosion testing involves techniques that do not have as the m a i n goal the m e a s u r e m e n t of current or voltage. These techniques are used to examine a n u m b e r of different forms of corrosion. Sometimes the data that result from the tests are quantitative. Sometimes the data are qualitative. Pictures of either the actual corroded sample or magnified portions of the corroded sample are often very useful and must be considered as types of data.

Corrosion Rate from Mass Loss


60-

==
A
A

c-

& A

&

&

OrO"20" et !
A A * ^A~ A~I~II' A

A
A A
& & &

A AI)'II'AI~'A~

o
o:'
r

i. , i i H . I

, i, ~ , , ~

.....

lo'

..... ;7

.....

le

Frequency (hz)

FIG. 3--Plot of electrochemical impedance spectrum in: (a) Nyquist format, inverse of the imaginary impedance versus real impedance; (b) Bode format, logarithm of the impedance magnitude versus logarithm of the frequency; (c) Bode format, inverse of the phase angle versus logarithm of the frequency (from ASTM G 106).

The most c o m m o n method for estimating the corrosion rate is from the mass loss of a metal specimen of known dimensions immersed in a fluid for a k n o w n amount of time. The weight of the specimen is obtained before and after the exposure. The corrosion rate is obtained by dividing by the exposed area, the time, and the density. The procedure is outlined in ASTM G 31, Practice for Laboratory Immersion Corrosion Testing of Metals. An example of a typical vessel for conducting such a test is shown in that standard. Different units exist for expressing the corrosion rate (mils/year for penetration rate or mg/dm2/day for mass loss rate). Though a n u m b e r of assumptions are made about the test and specimen when making this estimate, reasonably reliable values can be obtained when corrosion is not localized but extends reasonably uniformly across the specimen. Guidelines have been presented as to the error associated with measurements of weight, time, and surface area when these are the only sources of error [12]. In the case of rectangular panels or coupons, the area may be calculated from the exposed flat sides plus the four edges. Many investigators have chosen not to include the edge area, though its inclusion is fairly easily done. In the

C H A P T E R 2 - - T Y P E S O F DATA
case of a wire specimen, the exposed area decreases as corrosion proceeds. In order to take this effect into account, a somewhat different approach must be used to calculate the mass loss per unit area.

63

Flat panel: A - 2lw + 2lt +2wt


where l = length, w = width, t = thickness, A = exposed specimen area. (1)

Wire specimen radius loss:

~r = radius reduction from corrosion loss, d i = initial diameter before exposure, mi = initial mass, m f = final mass after exposure, Am = mass loss. The specimen must be cleaned after exposure in order to obtain mass loss results. In m a n y cases weighing the panel before cleaning is desirable so that an estimate of the corrosion product thickness or mass can be obtained. In some cases, removal of the corrosion products should be done nondestructively from at least a small area on some of the panels so that chemical analyses of the corrosion products may be carried out. Such removal may be done by scraping the surface or by using replicating tape. Replicating tape is not very effective in removing rust from steel specimens. ASTM G 1, Standard Practice for Preparing, Cleaning, and Evaluating Corrosion Test Specimens, provides additional information. The laboratory immersion test in which a specimen is exposed for a period of time and then removed for examination assumes that the corrosion rate remains constant with time. This assumption is not completely valid, even if the e n v i r o n m e n t remains static over the exposure time. The corrosion rate might decrease from a high rate during which passivation occurs or inhibitors adsorb, or the corrosion rate might increase as different kinetic processes establish themselves. When the a m o u n t of change is great, one might use a modified immersion test in which identical specimens are immersed for differing time periods in the same e n v i r o n m e n t [13]. The mass loss, exposure time associated with that mass loss, and area are used to calculate the corrosion rate for each interval. By comparing the rates during the different time periods, one can judge if the corrosion rate is constant, increasing, or decreasing. Again, the types of data needed are time, mass, and dimensions, but the exposure intervals are shorter than the total time of the test.

Localized Corrosion
Required data can sometimes be a mixture of qualitative observations and quantitative measurements that are combined to provide an overall picture of corrosion. As mentioned

above, certain features of the cyclic potentiodynamic polarization scan are used to estimate somewhat qualitatively the risk of localized corrosion in the form of pitting and crevice corrosion. Likewise, the results from coupon immersion tests used to estimate the risk of crevice corrosion and pitting also contain both types of information. I n the case of crevice corrosion, a c o m m o n procedure is to immerse a coupon u p o n which is m o u n t e d an artificial crevice former. ASTM G 78, Guide for Crevice Corrosion Testing of Iron-Base and Nickel-Base Stainless Alloys in Seawater and Other Chloride-Containing Aqueous Environments, describes a typical procedure. Typical observations include a recording of the general appearance of the specimen, its mass change, the m a x i m u m depth of corrosion in each crevice, the n u m b e r of crevices that had attack under them, and the severity of that attack. A picture sometimes becomes part of the data. This information might be used to rank alloy performance in the environment either qualitatively or even semiquantitatively. Similar types of qualitative and quantitative data are also combined when examining pitting. When a specimen is exposed to an environment, the depths of pits are often measured. However, the relationship between such depth and a pitting growth rate is not straightforward because pits do not necessarily grow at a continuous rate. Sometimes a sample is sectioned near the pit to examine its structure. ASTM G 46, Guide for Examination and Evaluation of Pitting Corrosion, provides a standard rating chart for pits in which the density of pits on a surface is rated on a scale of 1-5 (1 being the fewest pits, shallowest attack), as well as diagrams of different types of pit shapes. Comparing actual appearance to that shown in the standard can provide important information about the resistance of the alloy, how it is being attacked, and possible ways to overcome the attack. The appearance of the specimen is a type of data. ASTM G 46 discusses a n u m b e r of detection and analysis techniques that could be used for analyzing pitting and the types of data that each technique can supply. In addition, the standard provides a methodology by which a pitting probability might be estimated from an immersion test. Pitting corrosion is a p h e n o m e n o n that may or may not proceed at a constant rate. Pitting can initiate, propagate, stop, reinitiate, propagate, stop, etc., many times before the entire specimen is penetrated completely. The mere presence of pitting does not necessarily indicate failure. Only when the deepest pit has penetrated and a leak has occurred has the alloy actually failed. The time for such penetration to occur may be impossible to estimate. I n the laboratory, test time and specimen surface area can be far different from the expected field exposure time and surface area of the equipment being used. One method that has been proposed to bridge the large differences in space and time is using the statistical theory of extreme values [14,15]. The test and actual environments are assumed to be the same. Pit depths are measured on the laboratory specimen. The values are fit to a probability distribution function. If the fit is good, then by introducing the concept of return period, which is the ratio of the surface area of the object to the surface area of the specimen, the m a x i m u m pit depth can, in principle, be estimated for the object. The technique still requires more validation to determine the extent of usefulness.

64

CORROSION

TESTS AND STANDARDS

MANUAL

If valid, it could provide a quantitative methodology for relating m e a s u r e d pit depths to equipment life. In terms of required data, one needs to know the surface area of both the laboratory coupon and the object being examined by the coupon and the depths of the pits on the coupon. Types of data recorded for detection of intergranular corrosion are both qualitative and quantitative. Qualitative data are visual. They are obtained by examining under magnification a m o u n t e d a n d polished specimen that is etched appropriately. The etched structure is usually c o m p a r e d to photographs of specimens suffering similar types of corrosion (see ASTM A 262, Practices for Detecting Susceptibility to Intergranular Attack in Austenitic Stainless Steels) to determine if intergranular corrosion is observed on the specim e n u n d e r examination [16]. The observed surface structure becomes the data. Characteristic information that might be recorded are evidence of grains missing from the structure, pits in the grain ends, and ditches between the grains. Quantitative d a t a relating to i n t e r g r a n u l a r corrosion are actually relative c o r r o s i o n rates between different materials or the same m a t e r i a l t r e a t e d differently. F o r example, the alloy u n d e r e x a m i n a t i o n is exposed to a strongly acidic solution (see ASTM G 28, Test Methods of Detecting Susceptibility to I n t e r g r a n u l a r Attack in Wrought, NickelRich, C h r o m i u m Bearing Alloys) for a specified length of time [16]. The c o r r o s i o n rate is calculated from the mass loss of the specimen, its area, and the time of exposure (see ASTM G 31). This rate is c o m p a r e d to that o b t a i n e d for a p r o p e r l y a n n e a l e d material. Large differences between the two c o r r o s i o n rates can m e a n the presence of i n t e r g r a n u l a r corrosion.

Velocity Sensitive Corrosion (Single Phase Flow)


Corrosion can be sensitive to fluid motion, and often, fluid motion is included when testing for corrosion susceptibility. W h e n fluid motion becomes part of the experimental protocol, measurements not only have to be m a d e under dynamic conditions, but the data recorded m u s t characterize the fluid m o t i o n in addition to corrosion rate or other indications of corrosion. This brief discussion is provided only to point out the type of information that must be obtained and recorded to integrate fluid motion into the estimation of corrosion. Though ASTM standards for testing for velocity sensitive corrosion do not yet exist, a recent NACE International Report reviews several methods for making these measurements [19]. While the techniques described all have different geometries, certain types of data are c o m m o n to all. Merely recording fluid velocity or agitation rate is not sufficient to characterize the relationship between fluid motion and corrosion. Fluid motion is often quantified in terms of the dimensionless group the Reynolds number (Re). Quantities such as fluid shear stress at the wall and the friction factor used to characterize the relationships between fluid flow field and the geometry are functions of this dimensionless group. If mass transfer controls or influences corrosion, the Schmidt number (Sc) and Sherwood number (Sh) are included in the description. These dimensionless groups are:
Re - (velocity)(characteristic_ length) (kinematic _ viscosity)

(3)

Stress Corrosion Cracking


The most c o m m o n form of evaluation is time to failure for stress corrosion specimens. Failure usually means observation of the first crack. In relatively brittle materials, small cracks propagate rapidly causing the material to break into pieces. For specimens with constant loads, the rate of crack propagation increases as cracking proceeds, and failure is usually manifested as breaking of the specimen. Most specimens, however, are stressed by means of a constant strain. In such cases, the rate of propagation decreases after the initial crack is formed and, consequently, the best measure of time to failure is the time to observation of the first crack. A n u m b e r of tests are available that are used to estimate the alloy most resistant to stress corrosion cracking in an environment or to estimate a relative ranking of resistance among alloys. These tests employ a n u m b e r of different types of specimens including U-bends, C-bends, precracked (Fracture Mechanics) specimens, tuning fork specimens, tension specimens, etc. The type of data generated depends on the specimen and the test. The reader is encouraged to consult two reviews [17,18] and the considerable n u m b e r of citations to which the articles refer. These references discuss the type of data obtainable in stress corrosion cracking tests in considerable detail, as well as the behavior of a n u m b e r of alloys. One of the reviews also contains references to a n u m b e r of the ASTM standards that discuss test methods for stress corrosion cracking [17].

Sc - (kinematic_viscosity) ( diffusivity ) S h - (mass_transfer_coefficient)(characteristic_length) ( diffusivity )

(4)

(5)

The dimensionless groups can usually be related by the following equation:


S h = a Re b Sc c

(6)

where a, b, and c are constants that d e p e n d on the geometry of the test apparatus. Equation 6 is usually assumed to model the relationship among these dimensionless groups and the variables a, b, and c are determined by curve-fitting to actual data for the geometry in question. Finally, when the corrosion rate is u n d e r complete mass transfer control, the corrosion rate can be derived from an equation of the form: Corrosion Rate = (Mass Transfer Coefficient) x (Concentration Gradient)

(7)

In theory, one could estimate the mass transfer controlled corrosion rate from knowledge of velocity, the characteristic dimension, and certain physical properties. In practice, such a simplified approach would not be prudent though

C H A P T E R 2 - - T Y P E S OF DATA differences between measured and calculated values might provide insight into the corrosion mechanism. The background and methodology for using these parameters in real situations to relate the laboratory geometry to the field and estimate corrosion rates are presented in detail elsewhere [20,21]. The equations point to the information deficiency if only velocity is recorded. Such physical properties as the diffusivity, viscosity (kinematic viscosity = absolute viscosity/density), and density need to be included in the recorded data. The characteristic length is a parameter that scales the geometry to the flow field and is used as a scaling factor for distance from the wall. Examples are the diameter of a rotating cylinder or the diameter of a pipe. The concentration gradient between the corroding surface and the bulk fluid can often be approximated by the bulk concentration (e.g., oxygen diffusion controlling) or the saturation concentration (e.g., steel corroding in concentrated sulfuric acid). Either should be measurable. The reader should consult the references for more details.

65

O T H E R DATA
Though this brief discussion has focused on the various types of corrosion data that might be recorded during certain corrosion tests or upon the analysis of the data, there also are noncorrosion data that should be recorded. Temperature, alloy type, heat number, surface condition, and e n v i r o n m e n t chemistry are a few examples. Many more exist and may be required to complete the analysis. ASTM G 107, and especially its Table 1, contains a reasonably comprehensive list of the type of data that might be recorded during a corrosion test. ASTM G 161, Standard Guide for Corrosion Related Failure and Analysis, and especially its Appendix Xl, list the type of qualitative and quantitative data that should be considered when analyzing corrosionrelated failures. Though not all of the data types listed are relevant to every test, the contents serve as a useful checklist. Experimenters are encouraged to consult these listings prior to setting up an experiment to help ensure that key data and observations are not omitted from the measurements or records during the testing. An important concept to keep in m i n d is that recording too m u c h information is always better than not recording enough information.

REFERENCES
[1] Pourbaix, M., Atlas of Electrochemical Equilibria in Aqueous Solutions, National Association of Corrosion Engineers, Houston, TX, 1974. [2] Mansfeld, F., "The Polarization Resistance Technique for Measuring Corrosion Currents," in Advances in Corrosion

Science and Engineering, M. G. Fontana and R. W. Staehle, Eds., Vol. 6, Plenum Press, NY, 1976, p. 163. [3] Silverman, D. C., "Practical Corrosion Prediction Using Electrochemical Techniques," Chap. 68 in Uhlig's Corrosion Handbook, 2nd ed., R. W. Revie, Ed., John Wiley and Sons, 2000, p. 1179. [4] Rosen, E. M. and Silverman, D. C., "Corrosion Prediction from Polarization Scans Using an Artificial Neural Network Integrated with an Expert System," Corrosion, Vol. 48, No. 9, 1992, p. 734. [5] Gabrielli, C., Keddam, M., and Takenouti, H., "The Use of AC Impedance Techniques in the Study of Corrosion and Passivity," in Treatise in Materials Science Technology, J. C. Scully, Ed., Vol. 23, Academic Press, San Diego, CA, 1993, p. 395. [6] Electrochemical Impedance: Analysis and Interpretation, ASTM STP 1188, J. R. Scully, D. C. Silverman, and M. W. Kendig, Eds., ASTM International, West Conshohocken, PA, 1993. [7] Proceedings of the Second International Conference on Electrochemical Impedance Spectroscopy, Santa Barbara, CA, 1992, D. D. Macdonald, Ed., in EIectrochimica Acta, Vol. 38, 1993. [8] Eden, D. A., "Electrochemical Noise," Chap. 69 in Uhlig's Corrosion Handbook, 2nd ed., R. W. Revie, Ed., John Wiley and Sons, 2000, p. 1227. [9] Eden, D. A., "Electrochemical Noise - The First Two Octaves," Paper 386, presented at CORROSION~98, March, 1998, reprint available from NACE International, Conferences Division, P. O. Box 218340, Houston, TX 77218-8340. [10] Motard, R. L., Dai, X. D., Joseph, B., Silverman, D. C., "Improved Discrimination of Electrochemical Noise Signals Using Wavelet Analysis," Corrosion, Vol. 57, No. 5, 2001, p. 394. [11] Chen, J. F. and Bogaerts, W. F., "The Physical Meaning of Noise Resistance," Corrosion Science, Vol. 37, No. 11, 1995, p. 1839. [12] Freeman, R. A. and Silverman, D. C., "Error Propagation in Coupon Immersion Test," Corrosion, Vol. 48, No. 6, 1992, p. 463. [13] Fontana, M., Corrosion Engineering, McGraw-Hill Book Company, New York, 1986, p. 164. [14] Shibata, T., "Corrosion Probability and Statistical Evaluation of Corrosion," Chap. 22 in Uhlig's Corrosion Handbook, 2nd ed., R. W. Revie, Ed., John Wiley and Sons, New York, 2000, p. 367. [15] Ault, J. P. and Gehring, G. A., "Statistical Analysis of Pitting Corrosion in Condensor Tubes," in Corrosion Testing in Natural Waters, ASTM STP 1300, Vol. 2, ASTM International, West Conshohocken, PA, 1996. [16] "Failure Analysis and Prevention," Metals Handbook, Vol. 11, 9th ed., ASM International, Metals Park, OH, 1986. [17] Phull, B., "Evaluating Stress-Corrosion Cracking," Corrosion: Fundamentals, Testing, and Protection, ASM Handbook, Vol. 13A, Materials Park, OH, 2003, p. 575. [18] Corrosion--Metal~Environmental Reactions, L. L. Shreir, R. A. Jarman, and G. T. Burstein, Eds., Vol. 1, Chap. 8, ButterworthHeinemann, Oxford, 1998. [19] "State-of-the-Art Report on Controlled-Flow Laboratory Corrosion Tests," NACE Publication 5A195, NACE International, Houston, TX, 1995. [20] Silverman, D. C., "Rotating Cylinder Electrode-Geometry Relationships for Prediction of Velocity-Sensitive Corrosion," Corrosion, Vol. 44, 1988, p. 42. [21] Efird, K. D., "Flow-Induced Corrosion," Chap. 14 in Uhlig's Corrosion Handbook, 2nd ed., R. W. Revie, Ed., John Wiley and Sons, 2000, p. 233.

Vous aimerez peut-être aussi