You are on page 1of 15

Corrosion Science 47 (2005) 30533067 www.elsevier.

com/locate/corsci

Tafels law in pitting corrosion and crevice corrosion susceptibility


J.R. Galvele
*
Instituto Sabato, Universidad Nacional de San Martn, Av. Gral Paz 1499, B1650KNA San Martn, Buenos Aires, Argentina Available online 29 September 2005

Abstract The critical potentials for pitting are reviewed, under the scope of their relation with the crevice corrosion mechanism. A short review of the correlation between pitting corrosion and crevice corrosion with the Tafels law is made. Some problems observed in the mechanistic description of crevice corrosion are pointed out. The use of Tafels law measurements in the understanding of the mechanism of pitting is also mentioned. 2005 Elsevier Ltd. All rights reserved.
Keywords: A. Aluminium, iron; stainless steel; B. Modelling studies

1. Introduction To commemorate the centennial of Julius Tafels publication of his famous Tafels Law, the present review recalls the particular cases where this law was applied to pitting corrosion and crevice corrosion studies. As for other aspects of crevice corrosion and pitting corrosion, the reader is referred to more specic reviews of the problem [13].

Tel.: +54 11 6772 7390; fax: +54 11 6772 7404. E-mail address: galvele@cnea.gov.ar

0010-938X/$ - see front matter 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.corsci.2005.05.043

3054

J.R. Galvele / Corrosion Science 47 (2005) 30533067

Crevice corrosion and pitting corrosion are problems of undeniable signicance in practice and they are well documented in the literature [46]. Crevice corrosion could also be a nightmare, in the laboratory, when pitting corrosion testing is attempted on metals or alloys susceptible to crevice corrosion. The interference of crevice corrosion in pitting corrosion measurements explains why complex electrochemical cells have been developed [7] in order to avoid this problem. For certain metal-environment systems, particularly in NaCl solutions and in sea water, a strong correlation between pitting corrosion and crevice corrosion mechanisms has been reported by Wilde and Williams [8]. These authors have shown this correlation between the critical potentials for pitting of various commercial alloys, measured in the laboratory in NaCl solutions, and the susceptibility to crevice corrosion, after long exposures in sea water. The present review shows that various aspects of pitting corrosion, as well as the results reported by Wilde and Williams can be explained by using the Tafels law.

2. Critical pitting potentials Pitting corrosion is characterized by three critical potentials [9,10], the most relevant of them being the pitting potential, Ep (Fig. 1). The pitting potential can be dened, in a potentiostatic polarization curve, as the potential below which the metal surface remains passive and above which pitting corrosion starts to grow on the metal surface. Several methods have been used to measure the pitting potential, such as potentiostatic methods, potentiokinetic methods, galvanostatic methods, etc. Abundant literature was published, particularly in the 80s [9,10] on pitting potential measurements for iron, stainless steel, nickel, manganese, zinc, cadmium, alumin1x10
-3

Current density (A/cm2)

1x10

-4

1x10

-5

1x10

-6

1x10

-7

1x10

-8

Ep
-0.7 -0.6 -0.5 -0.4 -0.3

-0.8

Potential (V/SHE)
Fig. 1. Typical anodic polarization curve showing the presence of a pitting potential, Ep. Curve based on measurements for aluminium in NaCl solutions [10].

J.R. Galvele / Corrosion Science 47 (2005) 30533067

3055

ium, NiSn alloy, zirconium, titanium, tantalum, etc. Most of the solutions used were halide solutions, but references could also be found for pitting potentials in sulphate solutions, nitrate solutions, perchlorate solutions, etc. Since the pitting potential values measured with potentiostatic or potentiokinetic techniques, on various metals, particularly on stainless steels in chloride solutions, showed wide scattering, in 1971 Pessall and Liu [11] introduced a surface scratching technique, which has been adopted by numerous authors. A comparison of the various pitting potential measurement techniques could be found, for example, for zinc [12], cadmium [13] and iron [14]. As implied by the surface scratching technique, the main characteristic of the pitting potential is that the exposure of bare metal to the aggressive solution, below the pitting potential, leads to repassivation, and that it is only above the pitting potential that pitting corrosion begins to grow. This fact was conrmed on aluminium samples strained in chloride solutions and in nitrate solutions at various potentials [15,16]. The second characteristic potential for pitting corrosion was described by Pourbaix [17]. Pourbaix found that pits that start to grow at the pitting potential, will keep growing even when the potential is lowered below the pitting potential. According to Pourbaix, the pits will stop growing only when the potential of the metal is lower than a certain potential called repassivation potential, or protection potential, Er (Fig. 2). The existence of a repassivation potential was reported by several authors in a wide variety of metal-environment systems [9,10]. It was found that the repassivation potential varied with the amount of pit propagation that had taken place [8,18]. The deeper the pit the lower the Er. The third characteristic potential for pitting corrosion, present only in certain metalenvironment systems, was described by Schwenk [19] and is usually called pitting inhibition potential, Ei (Fig. 3). Schwenk reported that, for 18/10 chromiumnickel
-2

1x10

Current density (A/cm2)

1x10

-3

1x10

-4

1x10

-5

1x10

-6

Er
0.2 0.4 0.6

Ep
0.8 1.0

0.0

Potential (V)
Fig. 2. Typical anodic polarization curve showing the presence of a pitting potential, Ep, and a repassivation potential, Er. Arbitrary values used, examples of curves could be found in Ref. [8].

3056

J.R. Galvele / Corrosion Science 47 (2005) 30533067


1x10
-2

Current density (A/cm2)

1x10

-3

1x10

-4

1x10

-5

Ep

Ei

1x10

-6

0.0

0.5

1.0

1.5

Potential (V/SHE)
Fig. 3. Typical anodic polarization curve showing the presence of a pitting potential, Ep, and an inhibition potential, Ei. Curve based on measurements for stainless steel in NaNO3 + NaCl solutions [20].

stainless steel in chloride solutions containing nitrates, above the pitting potential there was a pitting inhibition potential, above which the steel became passive. This observation was later conrmed by Leckie and Uhlig [20]. The existence of a pitting inhibition potential was described [9,10] in systems such as iron in sulphate solutions, iron in perchlorate solutions, iron and nickel in mixtures of chloride plus nitrate solutions, in bromide plus perchlorate solutions, etc. While both the pitting potential, Ep and the pitting repassivation potential, Er, have important practical applications, this does not seem to be the case with the pitting inhibition potential, Ei. Keitelman and Galvele [21] studied the anodic behaviour of high purity iron in a 0.5 M sodium sulphate solution. They measured the pitting potential of iron in pH 9.0 and pH 10.0 solutions. The pitting potential was found to be very close to the corrosion potential measured in a Na2SO4, pH 2.7, pit-like solution. On the other hand, the pitting inhibition potential measured in a 0.5 M Na2SO4 solution, pH 10.0, was very close to the passivation potential found for the metal in the pit-like solution. The practical use of the pitting inhibition potential was limited, because susceptibility to crevice corrosion was observed at potentials above the pitting inhibition potential. Keitelman and Galvele [21] concluded that the pitting inhibition potential could not be used to stop pitting corrosion, because in those areas where a high potential drop could be found, like in a crevice, the pitting inhibition potential was not reached and crevice corrosion developed. 3. Origin of the critical potentials for pitting corrosion As early as 1951, Edeleanu and Evans [22] showed that in the case of aluminium the pH in the solution inside the pits, was lower than the pH in the bulk solution.

J.R. Galvele / Corrosion Science 47 (2005) 30533067

3057

This observation was later conrmed by various authors, not only for aluminium but also for a wide variety of metals and alloys [8,2327]. It was on the basis of these observations that Van Muylder et al. [24] advanced the rst explanation for the pitting potential. Their work was centred on pitting corrosion of copper and they suggested that the pitting potential of copper, in neutral or alkaline solutions, was equal to the equilibrium potential of Cu/Cu++ in the acid solution inside the pit. They assumed that the same explanation applied to other metals. Van Muylders [24] model was able to explain the pitting of copper, but failed to explain the pitting potential of metals such as iron, aluminium, zinc, etc., or when reducible anions were present, such as it was the case with aluminium in nitrate solutions. Neither could this model explain the fact that buer solutions acted as pitting inhibitors, increasing the pitting potential of numerous metals in corrosive solutions. An alternative model [12,13,15,2835] was developed to explain the pitting potential of metals such as zinc, cadmium, aluminium or iron, and the pitting potentials found in the presence of reducible anions or pitting inhibitors. Contrary to the thermodynamic approach used by Van Muylder et al. [24], the authors of this alternative model assumed that the pitting potential was the resultant of the kinetic reactions inside the pit. The model was based on the assumption that the aggressive ions were anions of strong acids, the only requirement being that they would not produce insoluble products when in contact with the metal and that the metal, while corroding in aqueous solutions, would react with water, producing localized acidication. While the nal result was the same, perhaps from a mechanistic point of view it would have been more appropriate to emphasize that, instead of a localized acidication, a depletion of HO, ions was being produced. The localized acidication was the result of the two following reactions: rst, the dissolution of the metal at the bottom of the pit Me Mez ze followed by the hydrolysis equilibrium of the metal ions inside the pit: Mez zH2 O $ MeOHz zH 2 1

Reactions (1) plus reaction (2) gave the rate of production of protons. Simultaneously, there was a series of reactions leading to the consumption of protons. Numerous authors reported the evolution of hydrogen from active pits [10]. This observation indicated that the reaction of hydrogen evolution was taking place 2H 2e H2 NO 10H 8e 3H2 O NH 3 4 3 4

If reducible anions were present, they could consume protons by reactions such as

Another factor that will limit the localized accumulation of protons is the presence of weak acid anions (L). The presence of soluble salts of weak acids will reduce the presence of protons inside the pit, by the following equilibrium: H L $ HL 5

3058

J.R. Galvele / Corrosion Science 47 (2005) 30533067

If the diusion of protons from the acidic solution inside the pit to the bulk solution is also taken into account, the conclusion is that, in order to keep a localized acidication, the pitting potential, Ep, must be equal to or higher than the corrosion potential of the metal in the pit-like solution, E . c This model of pitting allowed to explain numerous experimental observations related to pitting corrosion. In the particular case of pitting of aluminium it explained why the pitting potential of aluminium in nitrate solutions was so high (+1.70 V (SHE)) while in NaCl solutions it was only 0.53 V (SHE). Wexler and Galvele [15] reported the evolution of gaseous nitrogen from the pits of aluminium in nitrate solutions. This observation was conrmed by Bargeron and Benson [36]. The evolution of nitrogen indicated that the following reaction was taking place inside the pit: 2NO 12H 10e N2 6H2 O 3 6

Pitting of aluminium in nitrate solutions could only take place at potentials at which the nitrate ion was thermodynamically stable. According to Pourbaixs diagrams the thermodynamic stability of nitrate ions was reached at potentials above +1.2 V (SHE). The model of localized acidication for pitting received serious criticisms, the most important of them being that the changes in composition expected in small pits were insignicant and could not explain pitting initiation [29]. As a result of those criticisms, a series of publications was made studying the transport phenomenon inside a pit [28,29,33,34]. The most important conclusion of these studies was that there was a critical x i value (Figs. 4 and 5), above which important changes in the ion composition inside the pit took place. x being the depth of the pit, in cm and i the current density circulating in the pit, in A/cm2. For all the
100

80

Fe[II] (%)

60

Fe(OH)2 (aq) Fe + Fe(OH) Fe(OH)2 (s)


2+

40

20

0 -8 1x10

1x10

-7

1x10

-6

1x10

-5

1x10

-4

1x10

-3

x . i (A/cm)
Fig. 4. Distribution of iron corrosion species as a function of the product of the depth x and the current density i in a unidirectional pit, for iron in a pH 10 solution [33].

J.R. Galvele / Corrosion Science 47 (2005) 30533067


100

3059

80

Al[III] (%)

60

Al(OH)3 (s) Al ++ Al(OH) + Al(OH)2 Al(OH)3 (aq)


3+

40

20

0 -8 1x10

1x10

-7

1x10

-6

1x10

-5

1x10

-4

1x10

-3

x . i (A/cm)
Fig. 5. Distribution of aluminium corrosion species as a function of the product of the depth x and the current density i in a unidirectional pit, for aluminium in a pH 8 solution [33].

metals studied, the critical x i value was of the order or x i = 106 A/cm. Since it was known [29] that at pit initiation the current density could reach values of the order of 19 A/cm2, it was clear that a aw in the surface oxide lm could give enough length for the diusion path to reach the critical x i value. Since the model did not restrict the nature of x, Figs. 4 and 5 could be applied without distinction to pitting corrosion as well as to crevice corrosion. The initiation of pits in surface aws, as well as the relation between pitting and crevice corrosion was supported by Wood et al. [37]. These authors have shown that pitting, on both aluminium and stainless steels starts from aws in the oxide lm and propagates by the ion transport mechanism described above. The existence of aws, as precursors of pitting was also shown, by electrochemical methods, for zirconium in various environments [38,39]. Wood et al. [37] also studied the evolution of crevice corrosion of AISI type 304 austenitic stainless steel under a 13 mm diameter 0.3 mm glass microscope cover slide, at constant potential, in an air saturated 5% NaCl solution, pH 8.0. The development of crevice corrosion was followed with an appropriate optical system. They observed that, while at 600 mV (SCE) pits nucleated on all the metal surface, at 400 mV (SCE) pits nucleated only under the glass cover and spread laterally to take the shape of crevice corrosion. According to these authors, their study conrmed the close link between pitting and crevice corrosion in the case of stainless steels, and it strongly suggested that crevice corrosion is really no more than lateral pitting occurring within an occluded area. From their study it is concluded that the only function of IR drops was to favour the nucleation of pits near the border of the glass cover, and that they made it dicult to nucleate in the centre. Another conclusion from the transport process studies [28,29,33,34] was the nature of the pitting potential. It was concluded that the pitting potential had no

3060

J.R. Galvele / Corrosion Science 47 (2005) 30533067

relation with thermodynamic equilibrium, as suggested by Van Muylder et al. [24], but it was given by the following equation: Ep E g / c E c 7

Ep being the pitting potential, the corrosion potential in the pit-like solution, g the polarization needed to reach the critical x i value, and / the electrical potential induced by the migration of the aggressive anions to the bottom of the pit. The validity of Eq. (7) was tested in numerous publications [9,10,35] and it was conrmed also in an independent work by Newman et al. [40].

4. Critical x i values and Tafels law Studies as shown in Fig. 4 were made for Fe, Ni, Co, Zn and Cd. In all these cases, the metal dissolved from pits as a divalent cation. As shown in Fig. 4, for iron in a pH 10 solution, for low x i values the main, and practically the only corrosion species was insoluble Fe(OH)2,S. The proportion of soluble species is practically nil. Nevertheless, there is an x i value above which there is an abrupt change in the relation of the corroded species. At a value of x i close to 106 A/cm, the proportion of insoluble species, Fe(OH)2,S drops sharply, and it is replaced by soluble Fe2+. From Fig. 4 it is evident that this is a critical x i value above which the solid corrosion species are replaced by soluble corrosion species. This is a necessary, and frequently sucient, condition for pitting to start to grow. The low value for the critical x i indicates that pitting could start to grow in any small defect in the oxide lm. The critical x i values found for iron [33] were pH 10; pH 11; pH 12; x i 5:0 107 A=cm x i 5:0 106 A=cm x i 5:0 105 A=cm

The critical x i value increases one order of magnitude each time the pH increases one unit. If the x value, e.g. thickness of the oxide lm, is constant, the current density to start pitting will also increase one order of magnitude every time the pH increases one unit. If the current density, in the pit-like solution, follows a logarithmic relation to the potential, we can conclude that the relation between pH and pitting potential would be of the type: Ep A B pH 8

where A and B are constants. According to this result B should be close to the Tafel slope for iron in the pit-like solution. Such results were conrmed for iron in a NaCl solution [10,14], for cadmium in NaCl solution [41] and for cadmium in a Na2SO4 solution [41]. In the case of trivalent metals, like Al3+, the presence of a further step of hydrolysis slows down the eect of the external pH on the x i value. As shown in Fig. 5, the transition from passivity to pitting is not as sharp as for divalent metal ions,

J.R. Galvele / Corrosion Science 47 (2005) 30533067

3061

Fig. 4. Consequently, for comparison reasons, the critical x i value chosen was the point where the concentrations of solid Al(OH)3 and soluble Al3+ were equal, i.e. the intersection point of both concentration curves. The results found were [33] pH 5; pH 8; pH 10; x i 1:4 106 A=cm x i 2:0 106 A=cm x i 3:1 106 A=cm

While for divalent metal ions an increase of only one unit in the pH of the bulk solution produced an increase as big as one order of magnitude in the x i value; for aluminium an increase of 5 pH units produced just a twofold increase in the x i value. It can be concluded that the critical x i value, for aluminium, is practically independent of the external pH. This conclusion is in agreement with the observation made by Kaesche [42], who found that the pitting potential of aluminium was not aected by the solution pH. The critical x i value can also be changed by the presence of weak acid salts in the bulk solution. As shown by ion transport analysis [29] an increase of one order of magnitude in the concentration of the weak acid salt, produced also an increase of one order of magnitude in the x i value. If the current density, in the pit-like solution, follows a logarithmic relation to the potential, we can conclude that the relation between weak acid salt concentration, Cb, and pitting potential would be of the type Ep A B log C b 9 where A and B are constants. According to this result B should be close to the Tafel slope for the metal in the pit-like solution. Such results were reported for iron in borate containing NaCl solutions [14], and also for zinc in borate containing NaCl solutions and in borate containing NaClO4 solutions [43]. In the case of zinc it was found that when the borate concentration exceeded that of the aggressive salt, the values of the pitting potential exceeded those predicted by Eq. (9), but it was shown that Eq. (7) was still valid [34,35].

5. Flaws, passivity breakdown and metastable pitting If aws, as suggested by Wood [37], really exist, and if to start pitting a critical x i value has to be reached, it is reasonable to expect that, below the pitting potential, incipient processes of pit initiation should exist. Galvele et al. [38] studied the repassivation rate of various metals and alloys. The current decay, when measuring repassivation rates, could be described by a relation of the type i A tb where i is the current density, t is the time and A and b are constants. A value of b = 1 is typical of a high eld lm formation reaction, and it is usually related to the formation of a compact highly protective passive lm. Galvele et al. [38] studied the repassivation rate for zirconium in various halide solutions, as well as for brass in pH 12, 1 M NaNO2 solution. They observed that

3062

J.R. Galvele / Corrosion Science 47 (2005) 30533067

at low potentials the value of b was 1. Then there was a region of potentials where an initial value of b = 0.5, typical of a diusion controlled process, was followed by a repassivation rate of b = 1. The authors identied this potential region as the region where passivity breakdown took place. They also concluded that the repassivation rate technique was very convenient for detecting the passivity breakdown. No other references were found in the literature where this line of work was followed. Using dierent techniques, abundant work is found on the subject of formation of metastable pitting, which should be closely related to the phenomenon of passivity breakdown mentioned above. In the particular case of stainless steels, Pistorius and Burstein published an extensive work [4446]. Metastable pitting was also studied for other metals, and a full description of the subject could be found in the review published by Frankel [2]. An analysis of metastable pitting or of passivity breakdown would fall outside the scope of the present work.

6. Crevice corrosion The way crevice corrosion mechanisms are treated in the literature causes certain concern to the present author. To reduce the number of possibilities, the present consideration will be restricted to those systems where a passivating species is depleted from the crevice. After such a depletion, corrosion inside the crevice will start. But the way corrosion will proceed, for example if the potential is changed, will depend on the nature of the aggressive species present inside the crevice. For certain anions, like chloride ion, the corrosion of the metal generally increases monotonically with the potential. On the other hand, with anions like sulphate ion, and metals like iron, an active to passive transition is found when the potential is increased. Nevertheless, when considering the mechanisms involved, it seems to the present author that both the nature of the anions, and their role, are not suciently emphasized. The author has taken into account two recent papers. One by Wang and Newman [47], and the other by Al-Zahrani and Pickering [48]. The rst paper studies crevice corrosion of Type 316L stainless steel in alkaline solutions. The second studies crevice corrosion of spontaneously passive iron in 0.2 M Na2SO4 + 0.025 M K2CrO4 (pH 9) solution. In the rst system, when the passivating species, OH ions, are consumed, the metal is exposed to a chloride solution where pitting could occur. In the second one, when the chromate is reduced, iron is exposed to a sulphate solution which, as described above, shows a pitting potential and a pitting inhibition potential. The present author nds it dicult to accept that these dierences are irrelevant. The discussion of this problem, though signicant as it might be, falls outside the scope of the present work. The only reason to mention it here is to use it as an introduction to the next section. In the next section a classical work by Wilde and Williams [8] will be analysed from the point of view of Tafels law. In this case crevice corrosion in NaCl solutions is considered. No evaluation is made of the validity of Wilde and Williams technique as a general tool for detecting crevice corrosion susceptibility.

J.R. Galvele / Corrosion Science 47 (2005) 30533067

3063

7. Correlation between repassivation potential and crevice corrosion The existence of a critical x i parameter, described in Section 4, has another important implication. It explains the existence of a pitting repassivation potential, Er, lower than the pitting potential, Ep. As mentioned above, Er is lower the deeper the pits are allowed to grow, and this was easily explained by the use of the critical x i value [29]. Once the critical x i value is reached, at a certain potential, the pit will start to grow. If, after allowing the growth of the pit, the potential is reduced, the current density, i, will decrease, and pitting will stop once the critical x i value is reached again. Nevertheless, the nal x value will be larger than the initial one, and consequently the current density value required to stop pitting will be lower. The current density is directly related to the electrode potential, and the dierence between the i value reached at Ep and the new i value required at Er will depend on the E = f(i) relation, which in many cases is the Tafels Law. Wilde and Williams compared the value of the dierence Ep Er with the susceptibility to crevice corrosion for a series of alloys. Since the value of Er was function of the size of the pit, Wilde and Williams [8] used a standardised method to measure the Ep Er dierence. They worked in a 3.5% NaCl solution, and determined the value of Ep with a potential scanning rate of 0.600 V/h. The value of Er was determined after reversing the sweep at 0.2 A/cm2. Crevice corrosion was measured as weight loss on samples exposed for 4.25 years in sea water. Fig. 6 schematically shows the results found by these authors. Wilde and Williams found that the larger the Ep Er dierence the more susceptible the alloy was to crevice corrosion.

300

250

410

200

Ep - Er (mV)

430 150 304 100 316 50 Hastelloy C 0 0 5 10 15 20 446

USS 100

25

30

Weight loss (mg/cm2)


Fig. 6. Comparison between the dierence Ep Er, measured potentiokinetically in a NaCl solution, and crevice corrosion measured after long exposures in sea water. Schematic representation based on Wilde and Williams [8].

3064
1x10

J.R. Galvele / Corrosion Science 47 (2005) 30533067


1

(a)

Current density (A/cm2)

1x10

(b)

1x10

-1

aprox. 0.050 V 1x10


-2

1x10

-3

aprox. 0.800 V

1x10

-4

0.0

0.5

1.0

1.5

2.0

Potential (V)
Fig. 7. Susceptibility to crevice corrosion in NaCl solutions based on the Tafel slope, inside the pit, for iron (a) [14], and the pseudo-Tafel slope for stainless steel (b) [49]. The potential scale is arbitrary, and was chosen for illustrative purposes. In this example the following assumptions are made: critical value x i = 106 A/cm, pit size x = 106 cm, crevice size x = 104 cm. In case (a) the dierence in potential required to start pitting corrosion from that to start crevice corrosion will be of only 50 mV. In case (b) crevice corrosion will start at a potential 800 mV lower than that required to start pitting.

At this point it is important to notice that the technique developed by Wilde and Williams contains an implicit assumption. It assumes that the depth of the pits is the same in all the samples under comparison, and that this condition is achieved when the current density reaches 0.2 A/cm2. Nevertheless, this condition will be true only when the number of pits in all the samples tested is similar. The reason for Wilde and Williams correlation could be found more clearly in Fig. 7. We will assume that the current density, when pitting starts, is of the order of 1 A/cm2. We will also assume that curve (a) shows the Tafel curve for iron in a pit-like solution in NaCl, and curve (b) is a pseudo-Tafel curve for stainless steel; this last value being based on [49]. If the critical x i value is of the order of x i = 106 A/cm, pitting could start on any aw of 106 cm. If the pit is allowed to grow up to 104 cm, to reach again the critical x i value the current density will have to drop to 102 A/cm2. In the case of curve (a) the dierence Ep Er will be of only 50 mV. On the other hand, the dierence Ep Er, for the curve (b) will be of 800 mV. If, instead of a pit of 104 cm, we have a crevice of the same size, in case (a) the dierence between pitting and crevice corrosion will be of only 50 mV, a value usually obscured by the corrosion potential oscillations in practical cases, and it could be concluded that the metal (a) is not susceptible to crevice corrosion. On the other hand, for the metal (b), crevice corrosion will start at a potential 800 mV lower than that required for pitting, and this metal will be more susceptible to crevice corrosion than to pitting.

J.R. Galvele / Corrosion Science 47 (2005) 30533067

3065

8. Recent Tafels law measurements The present review described a limited number of cases where Tafels law was applied to localized corrosion, most of them published some time ago. The pit propagation process being electrochemical in nature, one would expect to nd more frequently relations of the type: i / exp E (Tafels law) between the current density and the potential. But, as pointed out by Frankel [2], the nonsteady state nature of the deepening of pits complicates the clear identication of the iE relationship. One remarkable exception is the work by Laycock and Newman [50]. The authors used articial pit electrodes prepared with AISI 302 and 316 stainless steel wires. The wires used were of 10, 50 and 500 lm diameters, and the measurements were carried out in 1, 0.3, 0.1 and 0.01 M NaCl solutions. Laycock and Newman measured the Tafel slopes inside the pits, and were able to identify two regimes of pit growth. At low potentials they found a mixed activation/ ohmic control, and at higher potentials they observed that the growth was under diffusion control with a metal salt lm present on the electrode surface. They were also able to explain, based on their measurements, the variation of the pitting potential with chloride concentration. As for the benecial eect of molybdenum alloying on the pitting resistance of stainless steel, they concluded that it was completely accounted by its eect on the anodic kinetics within the pit. Besides the results just described, Laycock and Newmans paper [50] contains a very thorough review on pitting research techniques and pitting mechanism theories.

9. Conclusions As shown in the present review, there are cases, both for pitting corrosion and for crevice corrosion where the use of Tafels Law allows us to explain: (a) The eect of pH on the pitting potential of divalent metals. (b) The eect of weak acid salts on the pitting potential of metals, when the concentration of the weak acid is low, and the aggressive solution acts as a supporting electrolyte. (c) The susceptibility to crevice corrosion of metals and alloys in NaCl solutions and in sea water, under conditions where those metals show a pitting potential. (d) The salt lm formation, and the electrochemical kinetics involved in pit propagation. (e) The mechanism by which molybdenum improves the pitting corrosion resistance of stainless steels. Acknowledgement The nancial support of the FONCYT, Secretara de Ciencia y Tecnologa, from Argentina, is acknowledged.

3066

J.R. Galvele / Corrosion Science 47 (2005) 30533067

References
[1] [2] [3] [4] [5] [6] [7] [8] [9] Z. Szklasrska-Smialowska, Pitting and Crevice Corrosion, NACE International, TX, 2005. G.S. Frankel, J. Electrochem. Soc. 145 (6) (1998) 21862198. S.M. Sharland, Corros. Sci. 27 (3) (1987) 289323. L.L. Shreir, Localised corrosion, in: L.L. Shreir, R.A. Jarman, G.T. Burstein (Eds.), Corrosion, third ed., Butterworth-Heinemann Ltd., Oxford, 1995, pp. 1:1641:169. R.M. Kain, Crevice corrosion, in: J.R. Davis (Ed.), ASM Handbook, Corrosion, vol. 13, ASM International, Materials Park, OH, 1996, pp. 108113. D.A. Jones, Principles and Prevention of Corrosion, second ed., Prentice-Hall, NJ, 1996, pp. 199234. R. Ovarfort, Corros. Sci. 28 (2) (1988) 135137. B.E. Wilde, E. Williams, Electrochim. Acta 16 (11) (1971) 19711985. J.R. Galvele, Present state of understanding of the breakdown of passivity and repassivation, in: R.P. Frankenthal, J. Kruger (Eds.), Passivity of Metals, The Electrochemical Society, Princeton, NJ, 1978, pp. 285327. J.R. Galvele, Pitting corrosion, in: J.C. Scully (Ed.), Treatise on Materials Science and Technology, Corrosion: Aqueous Processes and Passive Films, vol. 23, Academic Press, London, 1983, pp. 157. N. Pessall, C. Liu, Electrochim. Acta 16 (11) (1971) 19872003. M.G. Alvarez, J.R. Galvele, Corrosion 32 (7) (1976) 285294. M.G. Alvarez, J.R. Galvele, Corrosion 127 (6) (1980) 12351241. M.G. Alvarez, J.R. Galvele, Corros. Sci. 24 (1) (1984) 2748. S.B. de Wexler, J.R. Galvele, J. Electrochem. Soc. 121 (10) (1974) 12711276. I.A. Maier, J.R. Galvele, J. Electrochem. Soc. 125 (10) (1978) 15941598. M. Pourbaix, Corrosion 26 (10) (1970) 431438. T. Suzuki, Y. Kitamura, Corrosion 28 (1) (1972) 16. W. Schwenk, Corrosion 20 (4) (1964) 129t137t. H.P. Leckie, H.H. Uhlig, J. Electrochem. Soc. 113 (12) (1966) 12621267. A.D. Keitelman, J.R. Galvele, Corros. Sci. 22 (8) (1982) 739751. C. Edeleanu, U.R. Evans, Trans. Faraday Soc. 47 (1951) 11211135. T. Hagyard, J.R. Santhiapillai, J. Appl. Chem. 9 (1959) 323330. J. Van Muylder, M. Pourbaix, P. Van Laer, Rapport Technique No. 127, CEBELCOR, Brussels, 1965. J.A. Smith, M.H. Peterson, B.F. Brown, Corrosion 26 (1970) 539542. R. Piccinini, M. Marek, A.J.E. Pourbaix, R.F. Hochman, in: R.W. Staehle, B.F. Brown, J. Kruger, A. Agrawal (Eds.), Localized Corrosion, NACE, Houston, TX, 1971, pp. 179183. T. Suzuki, M. Yamabe, Y. Kitamura, Corrosion 29 (1973) 822. J.R. Galvele, in: R.W. Staehle, H. Okada (Eds.), Passivity and its Breakdown on Iron and Iron Base Alloys, NACE, Houston, TX, 1976, pp. 118120. J.R. Galvele, J. Electrochem. Soc. 123 (4) (1976) 464474. C.J. Semino, J.R. Galvele, Corros. Sci. 16 (1976) 297306. I.L. Muller, J.R. Galvele, Corros. Sci. 17 (1977) 179193. I.L. Muller, J.R. Galvele, Corros. Sci. 17 (1977) 9951007. J.R. Galvele, Corros. Sci. 21 (8) (1981) 551579. S.M. Gravano, J.R. Galvele, Corros. Sci. 24 (6) (1984) 517534. A.D. Keitelman, S.M. Gravano, J.R. Galvele, Corros. Sci. 24 (6) (1984) 535545. C.B. Bargeron, R.C. Benson, J. Electrochem. Soc. 127 (11) (1980) 25282530. G.C. Wood, J.A. Richardson, M.F. Abd Rabbo, L.B. Mapa, W.H. Sutton, The role of aws in breakdown of passivity of aluminum and crevice corrosion of stainless steel, in: R.P. Frankenthal, J. Kruger (Eds.), Passivity of Metals, The Electrochemical Society, Princeton, NJ, 1978, pp. 973988. J.R. Galvele, R.M. Torresi, R.M. Carranza, Corros. Sci. 31 (1990) 563571. J.R. Galvele, Analysis of the critical factors in pit initiation and propagation tech, in: G.S. Frankel, R.C. Newman (Eds.), Proceedings of the Symposium on Critical Factors in Localized Corrosion, The Electrochemical Society Inc., Pennington, NJ, 1992, pp. 94108.

[10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37]

[38] [39]

J.R. Galvele / Corrosion Science 47 (2005) 30533067 [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] R.C. Newman, M.A.A. Ajawi, H. Ezuber, S. Turgoose, Corros. Sci. 28 (5) (1988) 471477. M.G. Alvarez, J.R. Galvele, J. Electrochem. Soc. 127 (6) (1980) 12351241. H. Kaesche, Z. Physik. Chem. NF 34 (1962) 87108. J. Augustynski, F. Dalard, J.C. Sohm, Corros. Sci. 12 (1972) 713. P.C. Pistorius, G.T. Burstein, Philos. Trans. R. Soc. Lond. A 341 (1992) 531559. P.C. Pistorius, G.T. Burstein, Corros. Sci. 36 (6) (1994) 525538. G.T. Burstein, P.C. Pistorius, Corrosion 51 (5) (1995) 380385. S. Wang, R.C. Newman, Corrosion 60 (5) (2004) 448454. A.M. Al-Zahrani, H.W. Pickering, Electrochim. Acta 50 (1617) (2005) 34203435. J.R. Galvele, J.B. Lumsden, R.W. Staehle, J. Electrochem. Soc. 125 (8) (1978) 12041208. N.J. Laycock, R.C. Newman, Corros. Sci. 39 (1011) (1997) 17711790.

3067

You might also like