Kelly2017 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Localized Corrosion: Crevice Corrosion

RG Kelly, University of Virginia, Charlottesville, VA, United States


JS Lee, Naval Research Laboratory, Stennis Space Center, MS, United States
© 2018 Elsevier Inc. All rights reserved.

Introduction 291
Stages of Crevice Corrosion 292
Incubation 292
Stabilization/Propagation 293
Stifling 294
Mechanistic Descriptions to Rationalize Phenomenology/Stages 294
Critical Crevice Solution (CCS) 294
Critical Potential Drop (IR*) 294
Combinations of Critical Chemistry and Ohmic Drop Mechanisms 295
Critical Factors of Crevice Corrosion 296
Environment 296
Geometry 296
Techniques for Crevice Corrosion Evaluation and Materials Selection 297
Nonelectrochemical Methods 297
Electrochemical Methods at Open Circuit 297
Electrochemical Methods With Applied Signal 298
Mathematical Prediction of Crevice Corrosion 298
Summary 299
References 299

Introduction

Crevice corrosion is a specific type of localized corrosion that involves the creation of a physically occluded region at a particular site
on the metal surface by a crevice former (Fig. 1). The material creating the crevice can be metal (e.g., as part of a bolted joint) or
nonmetal (e.g., a polymeric washer). Underdeposit corrosion can be considered to be a type of crevice corrosion in which the crevice
former is typically an inert solid that has precipitated from solution or been deposited on the surface by settling from suspension. In
all cases of crevice corrosion, the occluded region develops a chemistry that is substantially more aggressive than the bulk solution
due to a combination of processes: (a) separation of dominant anodic and cathodic reaction sites, (b) hydrolysis of metal cations
created by metal dissolution within the crevice leading to a lowering of pH there, (c) large increases in crevice chloride concentration
due to the migration of chloride ions from the bulk as required order to maintain charge neutrality, and (d) limitations on mass
transport due to the physical occlusion. This combination of factors creates a local solution composition that supports high-rate
metal dissolution within the crevice. After a period of time, enough mass loss has occurred to cause failure (e.g., by creation of
a leak path at a flange). Crevice corrosion morphology is often characterized by regions of unattacked metal near the crevice mouth
followed by areas of severe penetration somewhat deeper into the crevice, with the deepest regions suffering less attack (Fig. 1). This
morphology is referred to as intermediate attack and its existence represents an important means to identify the contributions of
different crevice corrosion mechanisms.
Crevice corrosion shares many commonalities with pitting corrosion, another form of localized corrosion. In both types of
localized corrosion, an alteration of the local environment is a necessary condition for substantial damage to accumulate. In
some cases, crevice corrosion is thought to start via the formation of pits within the occluded site.1–3 The primary difference between
crevice corrosion and pitting is the requirement for crevice corrosion to occur inside a physically occluded region. Pitting can occur

Fig. 1 A schematic of a crevice formed by a former and substrate where L is the crevice length and G is the crevice gap. The region of greatest
substrate penetration (shaded area) is often at some distance intermediate of L with unattacked regions near the crevice mouth and deep with the
crevice. Adapted from Lee, J. S. Using Modeling and Microfabrication for Insights Into Mechanisms Controlling Location of Crevice Attack.
Department of Materials Science and Engineering, University of Virginia, Charlottesville, VA, 2011, pp. 234.

291
292 Localized Corrosion: Crevice Corrosion

on a boldly exposed surface, although there are some instances in which it can initiate at the interface (i.e., crevice) between an
inclusion and the matrix.4
Any passive alloy is fundamentally susceptible to crevice corrosion due to the loss of its protective oxide in the occluded region.
The environmental conditions required for crevice corrosion for alloys vary widely, but are in close correlation to the protectiveness
and robustness of their passive films. Crevice corrosion of stainless steel is probably the most widely studied alloy system43,63 due to
its wide industrial use in structures in which the mechanical designs create crevices, such as flanges on the ends of pipes. Other alloy
systems that commonly suffer from crevice corrosion are aluminum alloys,5–8 nickel alloys,9–13 and even titanium alloys14,15 under
sufficiently severe conditions.

Stages of Crevice Corrosion

The time course of crevice corrosion can be divided into three stages: (a) incubation, (b) propagation, and (c) stifling. Phenome-
nologically, during the incubation period no localized attack is observable macroscopically. Instead, during this time the chemical
conditions inside the occluded region are evolving as briefly described earlier. Once the critical conditions are achieved, propagation
occurs. Propagation typically involves moderate rates of attack ( 10 mA/cm216,17) with the area of attack expanding until a steady
state is achieved.18 Stifling of the attack, including complete repassivation, can occur at any time if the Critical Crevice Solution
(CCS) composition is not maintained within the crevice. This loss of the CCS can occur if, for example, the crevice mouth opens,
increasing the mass transport to the point that the dissolution rate cannot replace the CCS at a sufficient pace, or if the potential of
the sample falls below the repassivation potential.19

Incubation
The incubation period begins when both the occluded region and the boldly exposed area outside become wet and the potential is
high enough. Fontana and Greene20 described the processes occurring inside the occluded site in the situation in which the material
is spontaneously passive in the bulk solution of interest (i.e., it does not have an active–passive transition in that solution). One
example of such a system would be austenitic stainless steel in neutral NaCl solution. Under these conditions, there is initially
no environmental difference between the surfaces inside and outside the crevice. Both metal dissolution and oxygen reduction occur
at both locations. However, while oxygen can be replenished to the outer, boldly exposed surface through diffusion from the bulk
through a diffusion layer on the order of hundred microns, oxygen within the occluded area becomes depleted as the diffusion
distance is on the order of mm. This depletion represents the critical first step in the initiation of crevice corrosion: separation of
anode and cathode. Once the oxygen within the crevice is depleted, that region will become the primary anode for the sample,
with the primary cathode being the area outside the crevice. The (typically) much larger area outside the crevice then establishes
a highly unfavorable cathode:anode area ratio, and higher rate dissolution will occur in the crevice. This dissolution will lead to
an increase in the metal ion concentration within the crevice, and without the production of hydroxyl ions by oxygen or water
reduction locally, hydrolysis of these metal ions will lead to acidification. In addition, charge conservation will be maintained
by the migration of chloride (and other anions) into the crevice by the potential gradient. Thus a lower pH, higher chloride concen-
tration solution composition will develop within the crevice. These conditions will lead to higher dissolution rates, higher metal
chloride concentrations, and lower pH values in the crevice.
This cascade continues until at some point the CCS is reached. Reviews of the CCS for different alloy systems can be found
elsewhere,21–23 but all CCS share the commonality of being low pH (typically < 2), high chloride (typically > 6 M). The CCS specific
for a material is that combination of pH and metal chloride concentration that destabilizes the passive film and prevents its
reformation. The interfacial electrochemistry changes from spontaneously passive to active dissolution as shown in Fig. 2. When
coupled to the large, external cathode, the dissolution rate can now increase, leading to the stable propagation stage.

Fig. 2 Schematic polarization curves (A) before crevice initiation (spontaneously passive surface), and (B) after crevice activation. Note that if the
critical conditions are not maintained, the polarization behavior can revert from (B) to (A) via repassivation.
Localized Corrosion: Crevice Corrosion 293

Stabilization/Propagation
Stable propagation can be deemed to occur when either (a) the applied current abruptly rises (for systems under potentiostatic
control in a laboratory), or (b) the open circuit potential (OCP) dramatically drops (for systems under open circuit conditions
with a limited external cathode surface). These two cases are shown in Fig. 3. For potentiostatic experiments, the current typically
rises (albeit with some fluctuations) until it reaches a pseudo-steady state. For OCP systems with limited external cathodes, the
activation inside the crevice demands higher current from the external surface which it can only supply by polarization to more
negative potentials. Under OCP conditions, the large potential drop can be preceded by a series of smaller, transient potential drops.
After the large potential drop, there are typically very few fluctuations.
In both cases, the aggressive solution in parts of the crevice has exceeded the CCS composition, thus allowing higher rate
dissolution unimpeded by the passive film in those areas. A primary limitation on the dissolution rate is the ohmic drop between
the crevice mouth and high-rate dissolution region due to the current passage through a restricted geometry. Even seemingly small
currents (100 mA) can lead to substantial (> 0.1 V) ohmic drops due to the existence of crevice gaps on the order of microns or less.
During the propagation stage, the surface dissolves under activation control, leading to the formation of crystallographic facets
(Fig. 4). This morphology is indicative of dissolution occurring under activation control rather than under a salt film. Under
activation control, different crystallographic planes will dissolve at different rates, leading to the faceted appearance.

Fig. 3 (A) Apparent current versus time for crevice corrosion initiation on three different samples of 304 L in 17 mM NaCl with each point post-test
area corrected,16 (B) Open circuit potential versus time for a Type 410 stainless steel with a crevice former in 0.6 M NaCl.

Fig. 4 Crystallographic attack within a crevice of Ni200 with the mouth held at þ0.6 V(SCE) and exposed to 0.5 M H2SO4. Crevice gap ¼ 93 mm.24
294 Localized Corrosion: Crevice Corrosion

In some cases, widespread stable pitting occurs in the crevice initially.25 These pits then coalesce into an active region.1 At early
stages of crevice corrosion that initiates by this mechanism, the crystallographic facets may be much harder to see as they will be
within the pits themselves.

Stifling
Crevice corrosion is a metastable corrosion system. It relies on a delicate balance between high-rate dissolution inside the crevice
producing the aggressive solution fast enough to replace its loss by diffusion. In addition, the dissolution must occur sufficiently
deep in the crevice to avoid causing so much attack at the mouth that the diffusion restriction is lost. In addition, it requires a suffi-
cient cathodic current to be available either from the external surface or a potentiostat to both meet the dissolution current demand
while also maintaining the potential of the system above its repassivation potential. If any of these conditions are not met, the
process of stifling will begin.
Fundamentally, stifling occurs when the rate of metal chloride transport out of the crevice exceeds the rate at which it can be
produced by dissolution within the crevice. As the concentration of metal chloride decreases, the dissolving surface pH rises,
lowering the dissolution rate further. In the reverse of the initiation process, the cascade continues with lower dissolution rates
leading to lower metal chloride concentrations and higher local pH. If nothing intercedes, eventually the solution chemistry within
the crevice will be less severe than the CCS and the surface will repassivate, ending the crevice corrosion propagation. Causes for
stifling include the formation of corrosion products at the crevice mouth, the loss of adequate cathodic current due to competition
with other anodic sites, changes in reduction kinetics, or the loss of geometric mass transport restriction because of dissolution at the
crevice mouth. Stifling due to corrosion product formation occurs because the crevice mouth is the point at which the highly
concentrated metal chloride solution meets a near neutral or mildly alkaline pH solution with very low solubility for metal
ions. Large-scale precipitation of metal hydroxides can lead to sufficient large ohmic drop that the crevice potential falls below
the repassivation potential. If other anodic sites develop (e.g., pits on the surface, other crevice areas) that demand substantial
cathodic current, then it is possible for the cathodic area to distribute its current in such a way that the CCS cannot be maintained
for a given crevice. Such stifling does not happen under potentiostatic control because a properly operating potentiostat will supply
as much current as necessary to keep ALL anodic sites dissolving.
Reinitiation of crevice corrosion is certainly feasible, if the conditions develop. In some cases, repeated initiation/propagation/
stifling cycles occur, particularly for alloys that are under conditions that are borderline for supporting crevice corrosion. Such
instances manifest themselves with large, slow swings in potential for systems under OCP condition, and somewhat faster
oscillations of current for systems under potentiostatic control.

Mechanistic Descriptions to Rationalize Phenomenology/Stages

Phenomenological descriptions of crevice corrosion help rationalize qualitatively the experimental observations. Mathematical
models based on mechanistic descriptions offer the possibility of assistance with material selection, component design, mitigation
strategies, and service life prediction. The crevice corrosion models to date have been constructed assuming that one or more of the
processes described earlier dominates and/or that one stage of damage is most important. The earliest models focused on capturing
the time course of the CCS before initiation occurs. Later models focused on the stabilization of crevice corrosion by ohmic drop for
material systems that exhibit an active–passive transition in the crevice solution (IR*). There are also models which seek to combine
both aspects into a more unified computational framework.

Critical Crevice Solution (CCS)


The CCS mechanism of crevice corrosion was proposed by Fontana and Greene26 and expressed mathematically for stainless steel in
seawater by Oldfield and Sutton.27 The CCS model is primarily concerned with the manner in which the occluded geometry of
a crevice restricts the mass transport of species (diffusion and migration only) into and out of the occluded region resulting in a solu-
tion chemistry that is much different to the bulk solution as described earlier (Fig. 5). The model predicts the highest corrosion rates
within the crevice MUST occur at the deepest region (farthest from the bulk solution) of the crevice. The main drawback of the
model is that it ignores the potential dependence of corrosion rate when it assumes the passive current density is low enough
(mA/cm2) as to not result in an appreciable potential drop within the crevice. However, after crevice corrosion initiates, the current
flowing through the crevice increases by orders-of-magnitude invalidating this assumption. Therefore, the CCS model is not predic-
tive for crevice corrosion propagation rates. It also cannot predict the occurrence of intermediate attack, which is the most common
form of crevice corrosion damage distribution.

Critical Potential Drop (IR*)


The IR* model by Pickering28–31 focuses on how the restricted geometry causes ohmic potential drops in the solution within the
occluded site. At locations where the potential falls to a critical value defined as Epass (also referred to as Ecrit), the current density
increases substantially as the passive film is no longer stable and active dissolution commences (Fig. 6). It must be noted that the
Localized Corrosion: Crevice Corrosion 295

Fig. 5 The CCS model is primarily concerned with the manner in which the occluded geometry of a crevice restricts the mass transport of species
(diffusion and migration only) into and out of the occluded region resulting in a solution chemistry that is much different to the bulk solution. The
chemistry generally becomes more aggressive with time (t) as the dissolved metal ions hydrolyze, (A) decreasing the pH in the crevice, and
(B) electroneutrality considerations lead to the migration of chloride ions into the crevice. The net result is a low pH, high chloride concentration
solution. When the solution reaches critical values in terms of pH and chloride concentration, the stainless steel cannot maintain its passive film, and
onset of crevice corrosion begins. The model predicts the highest corrosion rates occur at the deepest region (farthest from the bulk solution) of the
crevice. Adapted from Lee, J. S. Using Modeling and Microfabrication for Insights Into Mechanisms Controlling Location of Crevice Attack.
Department of Materials Science and Engineering, University of Virginia, Charlottesville, VA, 2011, pp. 234.

Fig. 6 The IR* theoretical model by Pickering28–31 focuses on how the restricted geometry causes potential drops in the solution within the
occluded site. At locations where the potential falls to a critical value defined as Epass (also referred to as Ecrit), the resulting anodic current density
causes high levels of metal dissolution within the crevice. The model has two main drawbacks. The model can only be applied to metal/electrolyte
systems which exhibit active/passive behavior (i.e., Epass/Ecrit is present in the electrochemical boundary condition). Also, the model does not account
for chemical distribution changes within the crevice. As a result, the model predicts that spontaneously passive metal/electrolyte systems (e.g., 316
stainless steel in neutral chloride solutions) will remain passive. Adapted from Lee, J. S. Using Modeling and Microfabrication for Insights Into
Mechanisms Controlling Location of Crevice Attack. Department of Materials Science and Engineering, University of Virginia, Charlottesville, VA,
2011, pp. 234.

model can only be applied to metal/electrolyte systems which exhibit active/passive behavior in the solution within the crevice
(i.e., Epass/Ecrit is present in the electrochemical boundary condition). Also, the model does not account for chemical distribution
changes within the crevice. As a result, the model predicts that spontaneously passive metal/electrolyte systems (e.g., 316 stainless
steel in neutral chloride solutions) will remain passive. In addition, the incubation time for the onset of crevice corrosion cannot be
predicted because the potential distribution inside the crevice is instantaneously established when an electric field is applied. In
contrast, development of chemical species distributions occurs over time periods several orders-of-magnitude higher.

Combinations of Critical Chemistry and Ohmic Drop Mechanisms


Both the CCS and IR* models attempt to predict the conditions under which active crevice corrosion occurs. The relative strengths
and weaknesses of each model have already been stated. In attempts to maximize the advantages while minimizing the disadvan-
tages of each model, efforts have been attempted to bring the CCS and IR* models together.12,32–36 The general approach taken for
296 Localized Corrosion: Crevice Corrosion

all models that combine CCS and IR* is to use each model’s strengths. At the onset of exposure to a spontaneously passive
environment, current out of the crevice is governed by the passive current density; therefore the ohmic potential drop is minimal
and restricted mass transport results in the development of the required aggressive chemistry inside the crevice according to the CCS
theory. As the crevice solution becomes more aggressive, the metal dissolution rate, and thereby the current flowing through the
crevice increases, resulting in a substantial potential drop. At this point, the IR* criteria becomes significant, and location and
magnitude of attack become potential dependent and are governed by the electrochemical boundary condition. As the crevice
solution changes, the representative boundary conditions are modified. Such an approach is applicable for alloys that form an
active–passive transition in the CCS, such as nickel-based alloys.1,9,12,33,37 For alloys that instead exhibit essentially activation
controlled kinetics such as austenitic stainless steels16,17 or aluminum alloys,38,39 ohmic drop serves to limit the damage rate. Under
these conditions, intermediate attack occurs due to the gradient in chemical composition. Near the mouth of the crevice, the CCS
cannot be maintained against diffusion, whereas at some point deeper into the crevice, the diffusion barrier is sufficient to create
and maintain the CCS. At this location, the activation controlled kinetics lead to a high rate of dissolution. Deeper yet into the
crevice, the increased ohmic drop acts to decrease the dissolution rate until the OCP of the system is reached and the cathodic
focusing effects described by Stewart32 become more important.
Metastable pitting has been proposed as a critical factor in the initiation2,3,40 and stabilization1 of crevice corrosion. Stockert and
Boehni2 suggested that metastable pitting is stabilized by the occluded geometry of a crevice providing the necessary chemistry for
passivity breakdown. Based upon experimental and mathematical examinations, Laycock and coworkers3,40 surmised that passive
dissolution alone would not result in a significant potential drop for IR* to be achieved; rather metastable pits were required for
crevice corrosion initiation. Similarly, Kehler and Scully1 proposed that metastable pit coalesce provided the necessary CCS for
depassivation within a crevice resulting in the necessary IR* for crevice corrosion stabilization.
As discussed earlier, combination of CCS and IR* mechanisms can accurately predict crevice corrosion in many alloy/electrolyte
systems. However, Hodges41 observed that for austenitic stainless steel in acidic chloride environments for crevice gaps below 3 mm,
the site of greatest attack down the crevice length was independent of gap size and in direct contrast to IR*. For the same system,
modeling results by Lee42 qualitatively predicted the location of attack and that neither IR* nor CCS theories models were suitable;
rather a chemistry-dependent potential-current behavior was the factor that controlled the spatial distribution of crevice corrosion
attack in this system.

Critical Factors of Crevice Corrosion


Environment
Environmental factors that affect or control crevice corrosion include temperature, fluid velocity, oxidizing power, chloride content,
seawater variations, sulfide, body fluids, cathodic protection, and galvanic coupling. In-depth treatment of each factor has previ-
ously been detailed by Sedriks.43 In general, increase in temperature, velocity, oxidizing power (e.g., chlorine concentration),
and chloride concentration increases an alloy’s susceptibility to crevice corrosion. Coupling to a less noble material offers sacrificial
cathodic protection while coupling to a more noble alloy increases the likelihood of crevice corrosion. It should be noted that while
chloride concentration is an important factor, the presence of oxyanions (e.g., sulfate) can counteract the aggressiveness of the
chloride ion to a passive layer. Therefore, the ratio of chloride to oxyanions is often cited as a controlling environmental factor.

Geometry
A scaling law describes the effect of crevice geometry on chemical, potential, and current distributions of an occluded region. For
corrosion conditions to remain constant as crevice geometry is changed, these distributions must remain constant on a normalized
length scale. A scaling law is a factor of two geometric measures of a crevice that must be maintained as the scale of a crevice is
altered. Most often, the geometric factors involve some distance into the crevice (x) and the crevice gap (G).
Xu and Pickering29 developed an analytical solution for the region of greatest attack as a function of x [xcrit] with the crevice
length (L) as one of its variables. Abdulsalam and Pickering44 simplified this model by checking that the aspect ratio of crevice
length over gap (L/G) was greater than the critical value (L/G)crit for stable corrosion to take place. The scaling law scaling law
xcrit/G was shown to be valid for G > 10 mm, but monotonicity was lost at smaller gaps.
DeJong45 and Lee et al.24 proved that the scaling law in the IR* controlled Ni200/H2SO4 system was x2crit/G (Fig. 7). These scaling
laws were confirmed by experiments utilizing rigorously defined crevice geometries at the scale of real crevices (< 10 mm). The
advantage of using micron-scale crevices is the resulting corrosion morphology more closely resembles that found in real crevices.41
The disadvantages involve the difficultly in fabrication of rigorously defined geometries at the micron-scale41 and in situ distribu-
tion measurements during active crevice corrosion.46
Numerous researchers have used scaling laws to increase the sizes of their crevices to facilitate in situ measurements.7,28,44,47–53 A
disadvantage of using scaled crevices is that the resultant change in geometry due to active dissolution has much less of an effect on
distributions inside the crevice than for micron-scale crevices. For example, assume two crevices having gaps of 1 mm and 100 mm
undergo active corrosion resulting in 1 mm penetrations in each crevice. The gap for the 1 mm crevice will have increased by 200%,
while the larger crevice will only result in a gap increase of 1%.
Localized Corrosion: Crevice Corrosion 297

Fig. 7 Comparison of experimental and computational data of xcrit for Ni200 in 0.5 M H2SO4 as a function of gap plotted on quadratic scale xcrit
and a linear scale for gap.24

Techniques for Crevice Corrosion Evaluation and Materials Selection

Methods for the determination of crevice corrosion resistance are often divided into three categories: (1) nonelectrochemical, (2)
electrochemical with at open circuit no applied signal, and (3) electrochemical with applied signal. Further details and guidance can
be found in a review by Sridhar et al.54 in addition to each individual standard.55–58

Nonelectrochemical Methods
ASTM G4855 standard details the procedures in testing crevice (and pitting) corrosion resistance of stainless alloys. Crevice corrosion
is evaluated using Test Methods B, D, and F where artificial crevices are formed on alloys specimens of interest with nonmetal crevice
formers. Method B uses 6 wt.% ferric chloride as the oxidizing and corrosive electrolyte to produce localized corrosion under the
crevice formers. Temperature is maintained at 22 or 50 C with an exposure time of 72 h. Crevice corrosion resistance is evaluated by
mass-loss, maximum penetration depth, and attack area. Methods D (nickel-based, high chromium alloys) and F (stainless alloys)
use acidified ferric chloride solution to determine the critical crevice temperature. Multiple crevice assemblies are used as formers for
increased statistical analysis (Fig. 8). ASTM G7857 is similar to G4855 with the exception that seawater is used as the chloride-
containing electrolyte.

Electrochemical Methods at Open Circuit


Electrochemical techniques used to study crevice corrosion include those where no external signal is applied (open circuit) and ones
that involve perturbing the system with an applied signal. The corrosion potential (Ecorr) of a crevice exposed to an electrolyte is
measured against a reference electrode at open circuit. This technique is the most basic electrochemical measurement. However,
its advantage over applied signal techniques is that it mimics what happens in real systems where crevice corrosion can initiate,
propagate, and repassivate in one exposure. Sudden decrease in Ecorr often indicates active crevice corrosion, but knowledge of
the metal/electrolyte system is required for interpretation. However, no information regarding reaction rate can be determined using
open-circuit measurements since current does not flow between the crevice and the reference electrode. Open circuit measurement
can be combined with ASTM G48 Standard55 to provide transient Ecorr measurements during evaluation of crevice corrosion.
Current flow during crevice corrosion at open circuit can be measured using a remote crevice assembly59 and following galvanic
corrosion testing guidance of ASTM G71.58 Anode (creviced) and cathode (noncreviced) are connected through a zero-resistance
ammeter and the current (and potential if using a reference electrode) are monitored. Care must be taken in the geometric design
of the remote crevice assembly including the anode:cathode ratio. Anodic and cathodic reactions often cannot be assumed to occur
exclusively on the respective samples, that is, anodic reactions occur on noncrevice samples and/or cathodic reactions take place on
boldy exposed surfaces of the creviced sample or within the crevice itself. For these reasons, remote crevice assemblies often under
estimate metal loss during open-circuit testing.
298 Localized Corrosion: Crevice Corrosion

Fig. 8 Multiple crevice assembly with 316 L substrate.55 Adapted from Lee, J. S. Using Modeling and Microfabrication for Insights Into
Mechanisms Controlling Location of Crevice Attack. Department of Materials Science and Engineering, University of Virginia, Charlottesville, VA,
2011, pp. 234.

Electrochemical Methods With Applied Signal


Perturbation of the system by application of an external signal in the form of a voltage or a current (against a counter electrode)
provides information both in terms of system potential but also rate of reaction. When the applied signal is a constant voltage,
the technique is referred to as potentiostatic. This method is widely used in the study of crevice corrosion. Sedriks43 provided
numerous experimental examples using the potentiostatic method in his discussion of crevice corrosion in stainless steels. Holding
the voltage of a system potentiostatically provides the current as a function of exposure time. The advantage of the potentiostatic
method is that it is easy to set-up and interpretation of the resulting current can provide information on crevice initiation and
propagation, where increase in current often signals the onset of active crevice corrosion. The disadvantage of this method is
that it is capable of providing cathodic currents (against a counter electrode) much greater than a local cathode could supply.
Therefore, rates of reaction are often over estimated in comparison to open-circuit exposures limited by the anodic or cathodic
reactions. Cyclic potentiodynamic polarizations are also used to evaluate critical potentials of crevice corrosion as described by
ASTM G61.56
Application of a constant current to the system is referred to as the galvanostatic method. This method is used less frequently in
the study of crevice corrosion but is employed by some researchers.60 The advantage of this technique is that it more closely
resembles open-circuit exposures in that the cathodic current is limited by the cathode size. The disadvantage is constant current
is required providing marginal information about reaction rates.

Mathematical Prediction of Crevice Corrosion


Computational and mathematical models have been employed in attempts to identify the mechanisms for crevice corrosion with
varying degrees of success. By nature, computational models are more complex and require numerical solutions specified by
a convergence tolerance. Mathematical models are much simpler constructs because they can be solved analytically with exact
solutions. Assumptions in modeling crevice corrosion include: dilute solution theory, idealized geometry, simplified chemical
reaction treatment, simplified electrochemical boundary conditions, and steady-state results. Due to these assumptions, the current
understanding of crevice corrosion is limited and often does not accurately predict crevice attack observed for in-service alloys such
as stainless steels.
Each model has an underlying mechanism which governs the structure of the model, whether it be CCS, IR*, modified versions
of these, or a combination of the two. Oldfield and Sutton27 based their mathematical model on CCS theory for study of initiation
times of crevice corrosion in stainless steels. The model was later used by Lee and Kain61 to predict the geometric parameters
required for crevice corrosion initiation in a variety of stainless steels (Fig. 9). Galvele62 presented a diffusion-based mathematical
model for pitting of divalent metals. The model predicted a criterion for stable pitting corrosion based on the product of current
density and pit depth (x  i). Xu and Pickering29 developed a computational model for the study of active/passive systems based
Localized Corrosion: Crevice Corrosion 299

Fig. 9 Effect of crevice gap and depth on the initiation of crevice corrosion. The areas below the line for each alloy indicate gap/depth combinations
where initiation is predicted to occur.61

on IR* theory. Gartland34 deployed a computational model based on IR* theory which accounts for changes in crevice solution
chemistry over time (CCS theory). The model was used to predict crevice corrosion initiation and propagation of stainless steels
in chloride environments. Stewart32 developed a computational model that combined CCS and IR* theories to examine the effects
of cathodic reactions inside crevices. Using this model, Lee et al.24 validated a governing quadratic scaling law for crevice corrosion
of nickel in sulfuric acid, a system which was constructed to ensure that it followed the ohmic drop model. The aforementioned
models all exhibit predictive capabilities.

Summary

Crevice corrosion is distinguished by the requirement that the attack occur on a surface that is physically occluded from the bulk
solution. It occurs on materials that passivate in the solution of interest. The physical occlusion leads to an altered set of conditions
inside that crevice with a more aggressive solution forming due to the hydrolysis of dissolved cations, which leads to a marked
change in the electrochemical kinetics and local potential. The galvanic couple so formed, between the still passive surface outside
the crevice and the activated surface within the crevice, has a large cathode-to-anode ratio which leads to high-rate attack in the
crevice which leads the characteristic damage morphology. Several mechanisms have been suggested to explain the phenome-
nology, and in most cases, it is likely a combination of both critical chemistry and ohmic drop that control the distribution of
damage inside the crevice. A brief review of the range of testing methods is presented as is the state of the art in mathematical
modeling.

See also: Corrosion in Electronics; Corrosion in Pressurized Water; Corrosion of Steel in Concrete: New Challenges.

References

1. Kehler, B. A.; Scully, J. R. Role of Metastable Pitting in Crevices on Crevice Corrosion Stabilization in Alloys 625 and 22. Corrosion 2005, 61, 665–684.
2. Stockert, L.; Boehni, H. Susceptibility to Crevice Corrosion and Metastable Pitting of Stainless Steels. Mater. Sci. Forum 1989, 44–45, 313–327.
3. Laycock, N. J.; Stewart, J.; Newman, R. C. The Initiation of Crevice Corrosion in Stainless Steels. Corros. Sci. 1997, 39, 1791–1809.
4. Stewart, J.; Williams, D. E. The Initiation of Pitting Corrosion on Austenitic Stainless-Steel - on the Role and Importance of Sulfide Inclusions. Corros. Sci. 1992, 33, 457–474.
300 Localized Corrosion: Crevice Corrosion

5. Hebert, K.; Alkire, R. C. Dissolved Metal Species Mechanism for Initiation of Crevice Corrosion of Aluminum I. Experimental Investigations in Chloride Solutions. J. Electrochem.
Soc. 1983, 130, 1007–1014.
6. Siitari, D. W.; Alkire, R. C. Initiation of Crevice Corrosion I. Experimental Investigations on Aluminum and Iron. J. Electrochem. Soc. 1982, 129, 481–487.
7. Alavi, A.; Cottis, R. A. The Determination of pH, Potential and Chloride Concentration in Corroding Crevices on 304 Stainless Steel and 7475 Aluminum Alloy. Corros. Sci.
1987, 27, 443–451.
8. Ramgopal, T.; Frankel, G. S. Role of Alloying Additions on the Dissolution Kinetics of Aluminum Binary Alloys Using Artificial Crevice Electrodes. Corrosion 2001, 57,
702–711.
9. Lillard, R. S.; Jurinski, M. P.; Scully, J. R. Crevice Corrosion of Alloy 625 in Chlorinated ASTM Artificial Ocean Water. Corrosion 1994, 50, 251–265.
10. Kehler, B. A.; Ilevbare, G. O.; Scully, J. R. Comparison of the Crevice Corrosion Resistance of Alloys 625 and 22 in Concentrated Chloride Solution from 60 To 95 C,
CORROSION/2000, NACE International: Houston, TX, 2000, 182/181–182/115.
11. Martin, F. J.; Natishan, P. M.; Lucas, K. E.; Hogan, E. A.; Groulleau, A. M.; Thomas, E. D. Crevice Corrosion of Alloy 625 in Natural Seawater. Corrosion 2003, 59, 498–504.
12. Shaw, B. A.; Moran, P. J.; Gartland, P. O. The Role of Ohmic Potential Drop in the Initiation of Crevice Corrosion on Alloy 625 in Seawater. Corros. Sci. 1990, 32, 707–719.
13. Klein, P. A.; Ferrara, R. J.; Kain, R. M. Crevice Corrosion of Nickel-Chromium-Molybdenum Alloys in Natural and Chlorinated Seawater, CORROSION/89, NACE International:
New Orleans, LA, 1989.
14. Yao, L. A.; Gan, F. X.; Zhao, Y. X.; Yao, C. L.; Bear, J. L. Microelectrode Monitoring the Crevice Corrosion of Titanium. Corrosion 1991, 47, 420–423.
15. Shoesmith, D. W.; Noël, J. J. 3.10 - Corrosion of Titanium and its Alloys A2 - Cottis, Bob. In Shreir’s Corrosion; Graham, M., Lindsay, R., Lyon, S., Richardson, T.,
Scantlebury, D., Stott, H., Eds.; Elsevier: Oxford, 2010; pp 2042–2052.
16. Brossia, C. S.; Kelly, R. G. Influence of Alloy Sulfur Content and Bulk Electrolyte Composition on Crevice Corrosion Initiation of Austenitic Stainless Steel. Corrosion 1998,
54 (2), 145–154.
17. Brossia, C. S. Effects of MnS Inclusions on Crevice Corrosion of Stainless Steels in Neutral Chloride Solutions. In Critical Factors in Localized Corrosion III: A Symposium in
Honor of the 70th Birthday of Jerome Kruger; Kelly, R. G., Frankel, G. S., Natishan, P. M., Newman, R. C., Eds.; The Electrochemical Society, Inc: Pennington, NJ, 1998; pp
326–338.
18. Shinohara, T.; Fujimoto, S.; Laycock, N. J.; Msallem, A.; Ezuber, H.; Newman, R. C. Numerical and Experimental Simulation of Iron Dissolution in a Crevice With a Very Dilute
Bulk Solution. J. Electrochem. Soc. 1997, 144, 3791–3796.
19. Sridhar, N.; Cragnolino, G. A. Applicability of Repassivation Potential for Long-Term Prediction of Localized Corrosion of Alloy 825 and Type 316 L Stainless Steel. Corrosion
1993, 49, 885–894.
20. Fontana, M. G.; Greene, N. D. Corrosion Engineering, 2nd ed.; McGraw-Hill Book Company: New York, 1978.
21. Turnbull, A. The Solution Composition and Electrode Potential in Pits, Crevices and Cracks. Corros. Sci. 1983, 23, 833–870.
22. Turnbull, A. Chemistry Within Localized Corrosion Cavities. In Advances in Localized Corrosion: Proceedings of the Second International Conference on Localized Corrosion;
Isaacs, H. S., Ed.; NACE International: Orlando, FL, 1987; pp 359–373.
23. Oldfield, J. W.; Kain, R. M. Prediction of Crevice Corrosion Resistance of Stainless Steels in Aqueous EnvironmentsdA Corrosion Engineering Guide. In Corros. Control
Low-Cost Reliab. Proc.dInt. Corros. Congr. 12th, 3B; 1993; pp 1876–1888.
24. Lee, J. S.; Reed, M. L.; Kelly, R. G. Combining Rigorously Controlled Crevice Geometry and Computational Modeling for Study of Crevice Corrosion Scaling Factors.
J. Electrochem. Soc. 2004, 151, B423–B433.
25. Stockert, L.; Boehni, H. Metastable Pitting Processes and Crevice Corrosion on Stainless Steels. In Advances in Localized Corrosion; Isaacs, H. S., Bertocci, U., Kruger, J.,
Smialowska, S., Eds.; NACE International: Houston, TX, 1990; pp 467–473.
26. Fontana, M. G. G. Corrosion Engineering, 3rd ed.; McGraw-Hill Book Company: New York, 1986.
27. Oldfield, J. W.; Sutton, W. H. Crevice Corrosion of Stainless Steels. I. A Mathematical Model. Br. Corros. J. 1978, 13, 13–22.
28. Pickering, H. W. The Significance of the Local Electrode Potential Within Pits, Crevices and Cracks. Corros. Sci. 1989, 29, 325–341.
29. Xu, Y.; Pickering, H. W. The Initial Potential and Current Distributions of the Crevice Corrosion Process. J. Electrochem. Soc. 1993, 140, 658–668.
30. Wang, M.; Pickering, H. W.; Xu, Y. Potential Distribution, Shape Evolution, and Modeling of pit Growth for Ni in Sulfuric Acid. J. Electrochem. Soc. 1995, 142, 2986–2995.
31. Pickering, H. W. Important Early Developments and Current Understanding of the IR Mechanism of Localized Corrosion. J. Electrochem. Soc. 2003, 150, K1–K13.
32. Stewart, K. C. Intermediate Attack in Crevice Corrosion by Cathodic Focusing, Department of Materials Science and Engineering, University of Virginia: Charlottesville, VA,
1999, 544.
33. Lillard, R. S.; Scully, J. R. Modeling of the Factors Contributing to the Initiation and Propagation of the Crevice Corrosion of Alloy 625. J. Electrochem. Soc. 1994, 141,
3006–3015.
34. Gartland, P. O. A Mathematical Model of Crevice Corrosion for Fe-Ni-Cr-Mo Alloys in Chloride Solutions, SINTEF: Trondheim, 1988.
35. Gartland, P. O. Modeling Crevice Corrosion of Fe-Cr-Ni-Mo Alloys in Chloride Solutions. In Proceedings of the 12th International Corrosion Congress - Corros. Control
Low-Cost Reliab., NACE International, Houston, TX; 1993; pp 1901–1914.
36. Gartland, P. O. Modeling of Crevice Processes. In Proceedings of the CORROSION/96 Research Topical Symposia, NACE International, Denver, CO; 1996; pp 311–339.
37. Kehler, B. A.; Ilevbare, G. O.; Scully, J. R. Crevice Corrosion Stabilization and Repassivation Behavior of Alloy 625 and Alloy 22. Corrosion 2001, 57, 1042–1065.
38. Lim, M. L. C.; Kelly, R. G.; Scully, J. R. Overview of Intergranular Corrosion Mechanisms, Phenomenological Observations, and Modeling of AA5083. Corrosion 2016, 72,
198–220.
39. Lim, M. L. C.; Scully, J. R.; Kelly, R. G. Intergranular Corrosion Penetration in an Al-Mg Alloy as a Function of Electrochemical and Metallurgical Conditions. Corrosion 2013,
69, 35–47.
40. White, S. P.; Weir, G. J.; Laycock, N. J. Calculating Chemical Concentrations During the Initiation of Crevice Corrosion. Corros. Sci. 2000, 42, 605–629.
41. Hodges, A. J. The effect of crevice geometry on crevice corrosion stability of 316 stainless steel, Department of Materials Science and Engineering, University of Virginia:
Charlottesville, VA, 2007, 138.
42. Lee, J. S. Using Modeling and Microfabrication for Insights Into Mechanisms Controlling Location of Crevice Attack, Department of Materials Science and Engineering,
University of Virginia: Charlottesville, VA, 2011, 234.
43. Sedriks, A. J. Corrosion of Stainless Steels, 2nd ed.; John Wiley & Sons, Inc: New York, NY, 1996.
44. Abdulsalam, M. I.; Pickering, H. W. The Effect of Crevice-Opening Dimension on the Stability of Crevice Corrosion for Nickel in Sulfuric Acid. J. Electrochem. Soc. 1998, 145,
2276–2284.
45. DeJong, L. A.; Kelly, R. G. The Demonstration of the Microfabrication of Rigorously Defined Crevices for the Investigation of Crevice Corrosion Scaling Laws. In Critical Factors
in Localized Corrosion III: A Symposium in Honor of the 70th Birthday of Jerome Kruger; Kelly, R. G., Frankel, G. S., Natishan, P. M., Newman, R. C., Eds.; The
Electrochemical Society, Inc: Pennington, NJ, 1998; pp 678–688.
46. Wang, X.; Kelly, R. G.; Lee, J. S.; Reed, M. L. Microfabrication of Crevice Corrosion Samples. Mater. Res. Soc. Symp. Proc. 2001, 657, EE5 31/31–EE35 31/36.
47. Bocher, F.; Presuel-Moreno, F.; Scully, J. R. Investigation of Crevice Corrosion of AISI 316 Stainless Steel Compared to Ni-Cr-Mo Alloys Using Coupled Multielectrode Arrays.
J. Electrochem. Soc. 2008, 155, C256–C268.
48. Kehler, B. A.; Scully, J. R. Predicting the Effect of Applied Potential on Crack tip Hydrogen Concentration in low-Alloy Martensitic Steels. Corrosion 2008, 64, 465–477.
(Reprinted from Proceedings of the CORROSION/2007 research topical symposium “Advances in Environmentally Assisted Cracking", 2007.
49. Cho, K.; Abdulsalam, M. I.; Pickering, H. W. The Effect of Electrolyte Properties on the Mechanism of Crevice Corrosion in Pure Iron. J. Electrochem. Soc. 1998, 145,
1862–1869.
Localized Corrosion: Crevice Corrosion 301

50. Abdulsalam, M. I.; Pickering, H. W. Effect of the Applied Potential on the Potential and Current Distributions Within Crevices in Pure Nickel. Corros. Sci. 1999, 41, 351–372.
51. Vankeerberghen, M.; Abdulsalam, M. I.; Pickering, H. W.; Deconinck, J. Determining the Critical Crevice Depth for Iron in a Sodium Acetate-Acetic Acid Buffer Solution.
J. Electrochem. Soc. 2003, 150, B445–B450.
52. Wolfe, R. C.; Weil, K. G.; Shaw, B. A.; Pickering, H. W. Measurement of pH Gradients in the Crevice Corrosion of Iron Using a Palladium Hydride Microelectrode.
J. Electrochem. Soc. 2005, 152, B82–B88.
53. Wolfe, R. C.; Pickering, H. W.; Shaw, B. A. Microprobe Study of pH During the Induction Period Preceding Crevice Corrosion. J. Electrochem. Soc. 2006, 153, B25–B32.
54. Sridhar, N.; Dunn, D. S.; Brossia, C. S.; Cragnolino, G. A.; Kearns, J. R. Crevice Corrosion. In Corrosion Tests and Standards: Application and Interpretation; Baboian, R., Ed.;
ASTM International: West Conshohocken, PA, 2005; pp 221–232.)
55. ASTM. Standard Test Methods for Pitting and Crevice Corrosion Resistance of Stainless Steels and Related Alloys by use of Ferric Chloride Solution, ASTM Standard
G48-03, ASTM International: West Conshohocken, PA, 2003, 185–195.
56. ASTM. Standard Test Methods for Conducting Cyclic Potentiodynamic Polarization Measurements for Localized Corrosion Susceptibility of Iron-, Nickel, or Cobalt-Based
Alloys, ASTM Standard G61-86(2003), ASTM International: West Conshohocken, PA, 2003, 237–241.
57. ASTM. Standard Guide for Crevice Corrosion Testing of Iron-Base and Nickel-Base Stainless Alloys in Seawater and Other Chloride-Containing Aqueous Environments,
ASTM Standard G78-01(2007), ASTM International: West Conshohocken, PA, 2003, 329–335.
58. ASTM. Standard Guide for Conducting and Evaluating Galvanic Corrosion Tests in Electrolytes, ASTM Standard G71-81(2003), ASTM International: West Conshohocken, PA,
2003, 269–273.
59. Lee, T. S. A Method for Quantifying the Initiation and Propagation Stages of Crevice Corrosion. In Electrochemical Corrosion Testing ASTM STP 727; Mansfeld, F.,
Bertocci, U., Eds.; ASTM International: West Conshohocken, PA, 1981; pp 43–68.
60. Jakupi, P.; Zagidulin, D.; Noel, J. J.; Shoesmith, D. W. Crevice Corrosion of Ni-Cr-Mo Alloys. In Critical Factors in Localized Corrosion 5: A Symposium in Honor of Hugh
Isaacs; Missert, N., Virtanen, S., Davenport, A. J., Ryan Mary, P., Eds.; The Electrochemical Society, Inc.: Pennington, NJ, 2007; pp 259–271.
61. Lee, T. S.; Kain, R. M. Factors Influencing the Crevice Corrosion Behavior of Stainless Steels in Seawater, CORROSION/83, NACE International: Anaheim, CA, 1983.
62. Galvele, J. R. Transport Processes and the Mechanism of Pitting Metals. J. Electrochem. Soc. 1976, 123, 464–474.
63. Szklarska-Smialowska, Z. Pitting and Crevice Corrosion, NACE International: Houston, TX, 2005.

You might also like