Behaviour of Crevice Corrosion in Iron

Embed Size (px)

Citation preview

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    1/16

    Behaviour of crevice corrosion in iron

    Mohammed Ismail Abdulsalam *

    Chemical and Materials Engineering Department, King Abdulaziz University,

    P.O. Box 80204, Jeddah 21589, Saudi ArabiaReceived 23 May 2003; accepted 17 August 2004

    Available online 27 October 2004

    Abstract

    Crevice corrosion was investigated in iron exposed to a strong-buffered acetate solution

    (0.5M CH3COOH + 0.5M NaC2H3O2), pH = 4.66. The current and the potential gradient

    within the crevice were measured at crevice depth (L) = 7.35, 8, 10, and 15mm, for a crevice

    that was positioned facing the electrolyte in a downward position. A remarkable shift inpotential (>1.2V) in the active direction was measured inside the crevice cavity, when the

    potential at the outer surface was held at 800mV(SCE). Experimentation showed that there

    is a critical depth value, above which little changes occur on the transition boundary between

    passive and active regions on the crevice wall,xpass, and below whichxpasslocation shifts shar-

    ply towards the crevice bottom. Steeping of the potential gradient occurred with time indicat-

    ing enhancement of crevice corrosion, which was seen by the gradual increase in the current.

    These findings were in close agreement with the IR voltage theory and related mathematical

    model predictions. Morphological examination showed an intergranular attack around the

    active/passive boundary (xpass) on the crevice wall.

    2004 Elsevier Ltd. All rights reserved.

    Keywords: IR voltage theory; Iron; Crevice corrosion

    0010-938X/$ - see front matter 2004 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.corsci.2004.08.001

    * Tel.: +966 5568 2242; fax: +966 2695 1754.

    E-mail address: [email protected](M.I. Abdulsalam).

    Corrosion Science 47 (2005) 13361351

    www.elsevier.com/locate/corsci

    mailto:[email protected]:[email protected]
  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    2/16

    1. Introduction

    Crevice corrosion is a dangerous form of localized corrosion, which occurs as a

    result of the occluded cell that forms under a crevice on the metal surface. Well-known examples include flanges, gaskets, disbonded linings/coatings, fasteners, lap

    joints, weld zones and surface deposits. Systems relying on passive surface films

    for corrosion resistance can be particularly vulnerable to this form of corrosion.

    In these systems, which display active/passive transition in a corrosive environment,

    crevice corrosion can occur in the absence of pH change or chloride ion build-up in-

    side crevices. Examples of these were reported in iron [1,2], and nickel[3,4]. In these

    cases Pickering and co-workers showed that crevice corrosion is caused by the IR

    voltage drop which placed the local electrode potential existing on the crevice wall

    in the active peak region of the polarization curve. In addition, IR voltage drop

    mechanism has been shown to operate with other metals including; steel[5], and alu-minium[6]. Another proposed theory to explain the onset of crevice corrosion ad-

    dresses the change in the chemical composition of the electrolyte and the

    formation of a critical crevice solution with concentrations of H+ and Cl that are

    large enough to breakdown the passive film [7].

    Separation between the anodic and the cathodic reactions is necessary for the

    occurrence of crevice corrosion by the IR drop mechanism[8]. This condition prevails

    naturally for an open circuit experiment when an oxidant is added to the bulk solution

    where the potential at the outer surface (Esurf) is suddenly shifted from its open circuit

    condition in the active region into the passive region. Additionally, due to the oc-

    cluded nature of the crevice geometry, the separation can still occur when the crevice

    solution becomes depleted of oxygen and other passivating oxidants originally present

    in the bulk solution. Alternatively, in laboratory controlled experiments this condition

    is achieved by a potentiostat. The potentiostatic control offers the advantage of a more

    quantitative analysis. Another practical significance of this experimental set up is in

    anodic protection industries. Under the same logic, it was reported that applied poten-

    tial is unable to protect the entire structure due to the local electrode potential deep

    inside the crevice shifting to the active peak of the polarization curve [8,9].

    UnderIR drop mechanism controlled crevice corrosion, metal dissolves inside the

    crevice and the anodic current flows through the crevice electrolyte to the outer sur-face where the oxidant is reduced. The resulting IR voltage translates into an elec-

    trode potential on the crevice wall, E(x), that shifts in the less noble direction with

    increasing distance,x, into the crevice[10,11]. Recently, this concept was formalized

    [12,13], the results being in accordance with an earlier model for cathodic polariza-

    tion of a crevice[14].Walton et al.[15]developed a reactive transport based theoret-

    ical model and showed a good prediction to the measured potential distribution for

    crevice corrosion systems operating by the IR drop mechanism. It follows that the

    corrosion rate on the wall of the crevice is strongly position dependent as a result

    of the steep potential gradient in the depth direction of the crevice [1618]. There-

    fore, it is important to study the potential distribution inside the crevice and its rela-tion to the polarization curve in order to obtain a better understanding of the

    mechanism by which crevice corrosion occurs.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1337

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    3/16

    Experimental studies on the IR drop mechanism of crevice corrosion showed a

    transition from passive to active dissolution on the crevice wall as results of the crev-

    ice corrosion process[14,10,12,1921]. This transition boundary appeared at a cer-

    tain distance into the crevice,xpass, which is located at the Epassvalue on the crevicewall. The appearance ofxpasson the crevice wall indicates that the IR voltage drop

    inside the crevice is enough to shift the potential at the bottom of the crevice in the

    active region of the polarization curve, thereby creating active crevice corrosion. In

    accordance with theIRvoltage theory,Epassis located in the active/passive transition

    region of the polarization curve. The transition boundary was seen to be a straight

    horizontal line when the resistance of the electrolyte inside the crevice is uniform

    throughout the crevice cavity. The location ofxpass on the crevice wall is predicted

    by the relation[1,22]:

    Du Ex0 Epass IRxpass 1

    where Du* is the critical potential drop, Ex=0 is the passive applied potential at

    the crevice mouth, I is the current flowing out of the crevice, R= q/A, q is the elec-

    trolyte resistivity, and A is the cross-sectional area of the electrolyte column in the

    crevice.

    Analysis of the data is straightforward when the polarization curve for the crevice

    solution does not change during the experiment. The latter can be approached by

    using relatively open crevices with the upside down orientation with the outer sur-

    face facing downward in the solution (Fig. 1). It was shown that this crevice set-up

    keeps the pH value inside the crevice nearly the same as for the bulk solution,

    whereas it increased by a factor of four for the right side up orientation for which

    convective mixing did not occur[3]. Hence, the upside down orientation helps hold

    the pH constant due to the convective mixing of the crevice solution with the bulk

    solution [3,4]. The more dense corrosion products can easily move downward out

    of the crevice cavity in the direction of gravity, effectively maintaining a dilute ion

    concentration and the bulk solution pH. A similar finding of effective mixing was re-

    ported in an artificial crack [23].

    Most available studies on the IRdrop mechanism of crevice-corrosion address the

    effects of the oxidation power, gap-opening dimension, electrolyte composition andtemperature, while very few discuss the effects of the crevice depth. This paper de-

    scribes the characteristics of crevice corrosion of iron in an acetate buffered solution

    (constant pH) at room temperature, and addresses the role of the crevice depth.

    In order to keep the composition of the electrolyte from changing an artificial crevice

    with an upside down orientation was used instead of the upside up orientation

    reported previously[2,5,18,24]. The experiments were performed at different crevice

    depths, using an electrochemical microprobe technique to measure the potential

    distribution inside the crevice. Commercially pure iron known as Carpenter Electric

    Iron which has low-carbon content was used. Electric Iron is known for having good

    direct current soft magnetic properties after heat treatment. It has been used inelectromechanical relays, solenoids, magnetic pole and other flux-carrying

    components.

    1338 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    4/16

    2. Experimental

    The material used in this work was Carpenter Electrical iron; a low carbon com-

    mercially pure iron of a composition (wt%): C:0.012, Mn:0.10, Si:0.11, P:0.006,

    S:0.009, Cr:0.14, Ni:0.04, Mo:0.02, Cu:0.03, V:0.07, Fe:bal. The heat treatment con-

    dition was annealing at 843 C for 1h in dry hydrogen and cooled at 65.5 C per

    hour. Rectangular specimens were cut to the size of 20 15 5mm. This size fits

    with the size of the groove made on the Teflon block part of the electrode. The exper-

    iments were performed in a strong-buffered acetate solution Fe/HAcNaAc (0.5M

    CH3COOH + 0.5M NaC2H3O2). It was prepared with reagent grade sodium acetate

    (NaC2H3O2 3H2O), acetic acid (CH3COOH) and double distilled water. The meas-

    ured pH was 4.66, while the conductivity was: j= 0.03Scm1, at 24C. The exper-

    iments were carried out at room temperature.

    The crevice experimental set-up and procedure were similar as described previ-

    ously [4,5,20]. The exposed Fe crevice wall (depth: L= 1cm; width: w= 0.5cm)

    and outer (0.5cm 2cm) surface were polished to 0.05lm A12O3. In the electrode

    preparation process, all other surfaces in contact with the Teflon mount and edges

    of the specimen were sealed with a resin to prevent crevice corrosion between these

    materials and the specimen. Plexiglas formed the other walls of the crevice which hada gap (opening) dimension,a = 0.3mm. A schematic sketch of the electrode is shown

    inFig. 1. The electrical connection was made with an insulated copper wire soldered

    a

    w

    Crevice

    wall Iron

    L

    x

    Crevice

    Cavity

    Teflon

    Plexiglas

    Luggin

    microprobe

    SS Screw

    Crevice mouth

    L

    Outer surface

    w

    w

    Crevice

    wall Iron

    L

    x

    Teflon

    Fig. 1. Schematic diagram of the electrode assembly used in crevice-corrosion experiments.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1339

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    5/16

    to the backside of the sample. The sample was positioned so that the crevice

    mouth faced downward in the electrolyte (upside down orientation). This orienta-

    tion allows corrosion products to come out of the crevice by the gravity effect

    [3,4]. The use of a strong buffered solution also helped to minimize electrolytecompositional changes inside the crevice, such as acidification, due to the accumu-

    lation of corrosion products. In the previous work on crevice corrosion for Ni in

    1 N H2SO4 it was shown that gravity was sufficient to remove corrosion products

    out of the crevice [4]. In that very low pH system the chemical changes if they

    occurred, would have been deacidification rather than acidification, and this is

    a further step to retard acidification [3]. In the present system, which is suscepti-

    ble to acidification, a buffered solution will provide the assurance against

    acidification.

    In the experimental setup the solution level inside the crevice was maintained be-

    low the top of the Teflon block but above the Fe sample. Experiments were con-ducted by keeping the outer surface (bottom area of the iron specimen shown in

    Fig. 1) potential, Esurf, at 800mV(SCE) in the passive region of the polarization

    curve for this system. This was done by using a potentiostat. The potential distribu-

    tionE(x) on the crevice wall was measured using a fine glass microprobe of 0.03mm

    outside diameter and a saturated calomel electrode (SCE) via Luggin probe. The

    potential measurements were performed by using a voltmeter that had high input

    impedance in order to avoid creating potential drop effect in the very thin micro-

    probe. Thexpasslocation andEpassvalue were measured in situ using the microprobe

    connected to a three-directional micromanipulator with the aid of a macro lens

    viewer.

    The morphological characteristics of the crevice wall for some selected speci-

    mens were subsequently examined under both an optical microscopy system and

    a scanning electron microscopy (SEM). This was done after the experiment,

    where images of the characteristic features of the corroded-surface profile were

    recorded.

    The anodic polarization behaviour of the flat iron specimen without a crevice

    was investigated in a 0.5M CH3COOH + 0.5M NaC2H3O2 buffer solution. Prior

    to the polarization the specimen was mounted in epoxy resin, metallographically

    polished to mirror-like surface and soldered to a copper wire for the electrical con-nection. Slow potentiodynamic DC polarization scans were run by using a potenti-

    ostat and a three electrode flat specimen cell (Model K0235, EG&G, Princeton).

    The polished rectangular specimen surface (2 cm2) exposed to the solution was

    positioned side ways in the bulk solution. Two graphite electrodes were used as

    counter electrodes. An electrolyte volume of 1000ml was used, and a saturated cal-

    omel electrode in conjunction with a Luggin probe was used as a reference

    electrode.

    For the purpose of investigating the effect of the scan direction on the polarization

    behaviour two scans of opposite direction were conducted in a deaerated solution at

    a scan rate of 0.15mV/s. One, was done in the passive-to-active direction, from 1.2 to0.7V(SCE). The other scan was done on another freshly prepared specimen in thereverse active-to-passive direction.

    1340 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    6/16

    3. Results and discussion

    3.1. Anodic polarization behaviour in a buffered acetate solution

    Fig. 2, shows the anodic polarization curves for iron in a deaerated buffered ace-

    tate solution (Fe/NaAcHAc system) for two scans of opposite direction. Both

    polarization curves exhibit the active/passive transition with a large active peak.

    The open circuit potential (Eoc) was630mV(SCE) for both scans. The Epassvalue:165mV(SCE), was measured inside the crevice by placing the tip of the microprobeat the visible boundary between the passive and active regions on the crevice wall.

    This value is within the passive-to-active transition region of the bulk solution polar-

    ization curve (Fig. 2). The current at 165mV(SCE) inFig. 2is approximately onetenth of the peak current and noticeably larger than the passive current consistent

    with easy visible observation of this boundary on the crevice wall. This agreementof the measured Epass and Epass of the bulk solution polarization curve (in Fig. 2)

    indicates that the composition of the crevice solution had not deviated significantly

    from that of the bulk solution. The measured value for Epass inside the crevice is in

    agreement to others reported in the literature for this system [2].

    In the passive region, the current was lower for a scan in the passive to active

    direction (reverse scan); while the two scans almost coincide in the active region.

    A similar finding was reported for duplex stainless steel in aerated acidic/chloride

    solution [21]. However, this is a different finding than was seen in nickel/1N

    H2SO

    4 system, where a difference of 92mV in the value of E

    pass was observed be-

    tween the two scans (at a given current density) [4]. One possible interpretation

    0.0001

    0.001

    0.01

    0.1

    1

    10

    100

    -1000 -500 0 500 1000 1500

    Potential, E, (mV,SCE)

    Currentdensity

    ,i,(mAcm-2)

    Reverse Scan

    Forward Scan

    Epass=-165mV(SCE

    )

    E*=-460mV(SCE)

    ipeak(-283 mV, 18.43 mA cm-2

    )

    Fig. 2. Potentiodynamic polarization curves for iron in deaerated buffered acetate solution (0.5 M

    CH3COOH + 0.5M NaC2H3O2) for scans in the active-to-passive and passive-to-active direction. The scan

    rate was 0.15mV/s. Area: 1.89cm2.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1341

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    7/16

    for this is that the characteristic of the oxide film forming on the surface of the iron

    in NaAcHAc solution is different from the one forming on nickel in 1N H2SO4. In

    nickel a relatively stable appearing (dense gold-coloured) film was observed, while

    the film on the iron was weak and easily dissolved from the surface, leaving a roughsurface. In the works reported by Abdulsalam and Pickering on crevice corrosion in

    nickel, it was shown that Epassdepends on the direction of the scan[4,20].The value

    forEpassobtained through the reverse scan was found to be more in agreement with

    the experimentally measured value inside the crevice cavity[4,10]. However, for the

    crevice corrosion system in this paper there is little effect of the scan direction on the

    value ofEpass.

    3.2. Crevice corrosion characteristics

    With the selected crevice parameters and conditions, crevice corrosion com-menced at the start of the experiment and continued throughout the test period.

    The potential at the crevice bottom (EL) jumped from approximately

    622mV(SCE), a value around Eoc, to more positive potentials as the crevice depthdecreased. The change in surface appearance started to occur within a few seconds

    after the potentiostat was turned on, mA currents were measured immediately. This

    indicates that the conditionL > Lc, whereLcis the critical crevice depth was met.Lcis the smallest crevice depth that results in crevice corrosion for the test conditions,

    and is defined as the distance x at which the activepassive transition coincides with

    the crevice depth,L, for the given metal/electrolyte system[12,25]. Also, since a large

    active peak exists on the anodic polarization curve, no induction period was neces-

    sary and therefore crevice corrosion started immediately[1,17,26]. This implies that

    theIR voltage drop was enough at the start of the experiment for the applied poten-

    tial at the outer surface to shift part of the crevice wall atLinto the active peak of the

    bulk solution polarization curve.

    At the onset of crevice corrosion, the upper mirror-like part of the crevice wall

    lost its lustre and became attacked.Fig. 3shows photographs of the on-going exper-

    iment of crevice corrosion of iron in the buffered acetate solution (L= 7.35, 8 and

    15mm), taken at dissimilar times throughout the experiment. The glass microprobe

    appears inside the cavity. The morphology of the surface under the action of crevicecorrosion as viewed through the clear Plexiglas can be divided into three main re-

    gions. The first (lowest) region is the passive region (un-attacked) that extends from

    the crevice mouth to thexpassboundary. In this passive region, a small band of light

    etching was observed at x 6 xpass that formed within the initial seconds of crevice

    corrosion. The second region starts at x> xpass is the severely attacked region that

    extends to xlim. The location of xlim cannot be determined visually, since it is not

    associated with an obvious change in the surface morphology, but it can be esti-

    mated from the potential profile, E(x), as the distance at which Ebecomes independ-

    ent of x, (Fig. 4). The third region is the etched region extending from xlim to the

    crevice bottom (x= L), over which the potential is nearly constant. Similar crevicewall morphology to these three regions was reported for crevice corrosion in nickel

    [4,20], and duplex stainless steel [21].

    1342 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    8/16

    Epass=165mV(SCE) was measured with the microprobe tip at xpass and wasfound to be a constant value independent of time or L. This finding indicates that

    the convective mixing of the bulk and crevice solution is maintaining the initial (bulksolution) anodic polarization curve, and that the electrolyte composition inside the

    crevice is not changing [3,4,20]. This is in agreement with the literature where it

    Fig. 3. In situ photographs of the corroding iron crevice wall (upside down orientation) in buffered acetate

    solution (0.5M CH3COOH + 0.5M NaC2H3O2), showing the location ofxpassand the distinctive regions

    that appeared during crevice corrosion. (a) 40min, (b) 15min, and (c) 7min. Magnification 5.5.

    -600

    -400

    -200

    0

    200

    400

    600

    800

    -1 0 1 2 3 4 5 6 7 8 9 10

    Distance into crevice,x, (mm)

    Potential,E,

    (mV,S

    CE)

    Epass= -165 mV(SCE)

    Eapplied= 800 mV (SCE)

    L = 8.0 mm

    3.5 hr

    15 min

    12 hr

    Direction of motion

    ofxpasswith time xpass

    xpass

    xpass

    Fig. 4. Variation of the potential distributionE(x) with time inside the crevice for iron in buffered acetate

    solution (0.5M CH3COOH + 0.5M NaC2H3O2), whenL= 8.0mm.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1343

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    9/16

    was shown thatEpasschanges by approximately 150mV for the Ni/1N H2SO4, as the

    bulk solution pH changes from 0.3 to 2 when the bulk solution was saturated in Ni2+

    ion[3].

    In agreement with the observation reported in earlier works, xpass moved withtime towards the crevice mouth [4,10,21]. Fig. 4 shows this behaviour through the

    E(x) distribution measured using the glass microprobe at different times for

    L= 8mm. The location ofxpassis indicated by the arrow atEpass, and listed inTable

    1. These values are in close agreement with the measured ones physically at the xpassboundary. The advancement ofxpasstowards the crevice mouth with time is consist-

    ent with an increasing current. The measured current increased from the onset of

    applying Esurf from 1.26mA initially to 2.23mA at t= 12h. Similar behaviour for

    other systems was reported in the literature [24]. It is worth mentioning here that

    while the active region on the crevice wall is located in the region between Epassand E* the most contributing part to the current is located only at a distance thatis slightly greater than xpass. At this location, the peak current density will exist in

    agreement with the shape of the active peak of the polarization curve (Fig. 2). This

    is in agreement with the potential gradient showing a steep gradient in the region

    where the peak current is expected to exist. With time, the gradient at this region be-

    comes steeper (Fig. 4), thereby, the current also increases with time. This was dis-

    cussed in more detail and verified both experimentally and theoretically in earlier

    works [20]. This change on E(x) with time was less pronounced when L= 15mm.

    In addition, with time Elim became more negative (Fig. 4).

    It is interesting to note that while in this work for the crevice with L = 10mm the

    current was measured as 1.2mA after 40min, in other work this corresponds to

    1.34mA for a crevice in the right-side up orientation [2]. This is expected since in

    the right side up orientation, acidification is possibly occurring which increases the

    size of the active peak in the polarization curve, thereby, increasing the crevice cor-

    rosion current[4,24]. In addition ELwas about 70mV more negative with the right-

    side up orientation. This is explained by the increase of the electrolyte resistance that

    is allowed with this crevice orientation.

    The effect ofL on the initialxpassvalue (attP 0), measured directly by physically

    placing the tip of the microprobe in situ at the xpassboundary seen through the Plexi-

    glas with the help of a macro lens viewer, is shown inTable 2,for L = 7.35, 8, 10 and15mm, being 4.5, 3.3, 3.5 and 3.7mm, respectively. These measurements are in close

    agreements with the values obtained using the initially measured E(x) distributions

    Table 1

    Time dependency ofxpass and Elim inside the crevice for iron in the buffered solution, L= 8.0mm

    Time, h xpassa (0.1), mm xpass

    b (0.1), mm Elima (1), mV(SCE)

    0.25 3.3 3.4 4713.5 2.6 2.7 505

    12.0 2.4 2.4 514a Measured inside the crevice by using microprobe.b Estimated from the potential profile at Epass (Fig. 4).

    1344 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    10/16

    shown in Fig. 5, at the intersection of E(x) curve and Epass line. When LP 8 mm

    (more specifically in the range from L= 8 to 15mm) the location ofxpass at tP 0was further into the crevice as the crevice depth, L, was increased. However, the

    amount of change ofxpasswas very moderate (0.4mm for DL= 7mm), in accordance

    with the moderately decreasing initial current, I, with L, Table 2. These results are

    consistent with the predication of the IR voltage drop theory, shown in Fig. 6 by

    the good agreement between the experimental value for xpass and the calculated

    one according to Eq. (1).

    On the other hand, when L< 8mm (7.35mm), xpassshifted in the opposite direc-

    tion towards the crevice bottom for decreasing L, as shown in Fig. 5andTable 2.

    The amount of change in xpass here is much more pronounced (1.2mm for

    DL= 0.65mm). Reported data for Lcis included inFig. 6and it is shown to follow

    Table 2

    Depth dependency of the initially measured (within first 15min): xpass, Elim, and I for a crevice in iron

    exposed to acetate buffered solution

    Depth (L), mm xpassa

    (0.1), mm Elima

    (1), mV(SCE) I, mAL< 8

    7.35 4.5 418 1.4L> 8

    8.0 3.3 471 1.2610.0 3.5 528 1.215.0 3.7 600 1.05

    a Measured inside the crevice by using microprobe.

    -650

    -450

    -250

    -50

    150

    350

    550

    750

    950

    -1 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

    Distance into crevice,x, (mm)

    Potential,E,

    (m

    V,S

    CE)

    Epass

    xpass @ L = 10 mm

    Eapplied= 800 mV (SCE)

    L = 15.0 mm

    L = 7.35 mm

    L = 8.0 mm

    E@ ipeak, (Fig. 2)

    E*, (Fig. 2)

    EL @ L = 10 mm

    Fig. 5. Variation of the initial (within 15min) measured potential distributions with depth inside the

    crevice for iron in buffered acetate solution (0.5M CH3COOH + 0.5M NaC2H3O2).

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1345

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    11/16

    the same trend[12]. In this range it is clear that xpass deviates from the linear rela-

    tionship predicated by Eq.(1)as shown inFig. 6. This scenario observed experimen-

    tally is in agreement with theoretical analysis of the mechanism for a similarlocalized corrosion problem, formulated by using the electric field within the film

    on the surface for differentL values[25]. The model computation results showed that

    for a given crevice system there is a certain depth value, above which xpassvaries lit-

    tle, and below whichxpassincreases substantially toLc.Fig. 6is the experimental plot

    that agrees with the model result in reference [25].

    The initial measured current (within the first 15min) at different L are listed in

    Table 2. Little changes occurred on the current until L was decreased beyond

    8mm. These current values can be shown to be consistent with the initial potential

    distributions shown in Fig. 5, where the amount of the current is proportional to

    the range of the peak current region of the polarization curve that operates in the

    severely attacked region of the crevice wall. Therefore, the spread of this region

    can possibly favour the likelihood of the increase of the crevice-wall area that is

    exposed to potentials of the ipeak region, thereby, increasing the total dissolution

    (crevice corrosion) current, I. Similar current behaviour was reported for the effect

    ofa on the measured current in nickel/sulphuric acid crevice system [20].

    The potential behaviour atx= L,EL, is also shown inFig. 5, and listed inTable 2.

    The largerL is the more active EL, which is in agreement with theoretical model pre-

    dictions[25].Similar findings were also reported for the change ofEL with the ap-

    plied potential at the surface [4], and the crevice opening dimension [20]. Anotherobservation shown inFig. 5is that for the potential profiles for L= 8 and 15 there

    is a clear existence ofxlim, but at L= 7.35 this is not evident. In principle, the most

    0

    2

    4

    6

    8

    10

    12

    2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

    Crevice depth, L , (mm)

    xpass,

    (mm)

    Experimental data

    Calculated using Eq. (1)

    Lc= 6.7 (reference 12)

    Eapplied= 800 mV (SCE)

    No crevice corrosion

    detected

    Fig. 6. The effect of the crevice depth (L) on the location of the transition boundary (xpass). Shown also are

    the expectedxpass values according to theoretical prediction based on IR voltage theory.

    1346 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    12/16

    negative potential that can be measured at EL is the open circuit potential (Eoc)

    which is630mV(SCE) for the current system[1,4]. However it is usually observedthat a Elimis more positive than Eoc[3,10,20]. It is interesting to note that based on

    the current findings (Fig. 5), althoughElimappears to be constant with distance, it isactually still decreasing with increasing crevice depth, L. This is shown in Fig. 5

    where at L= 7.35mm, EL= Elim=418mV(SCE) which gradually decreased to600mV(SCE) when L increased to 15mm. Thus Elim is expected to approach Eocfor increasing L.

    Recently, Vankeerberghen et al. [12] developed a mathematical model for the

    potential drop into the crevice based on a Poisson-type field problem with non-linear

    boundary conditions that was described as a one-dimensional finite difference frame-

    work. The Poisson-type, second order differential equation used in the analysis is

    given by

    d2Ux

    dx2

    P

    r Six 2

    wherer is the conductivity of the solution, Sis the cross-sectional area,Pis the elec-

    trochemically active part of the perimeter, and U(x) is the potential in solution. The

    application of this model to a crevice corrosion system similar to the one used in this

    work showed for L= 10mm, xpass= 3.2mm. This value for xpass is in reasonable

    agreement with the experimentally determined one here, Table 2.

    3.3. Morphology of xpass boundary on the crevice wall

    After and during the experiment, examination of the crevice wall revealed more

    penetration and attack in the region just further than xpassinside the crevice. Similar

    observation was reported on the crevice wall for other crevice corrosion systems

    operating by the IR mechanism [1,2,4,10,12,2022,24]. Changes in the appearance

    of the crevice wall right at xpass were further examined under conventional optical

    microscopy and were photographed. Fig. 7 shows a photograph at 100 of the

    morphology of the xpass boundary after the experiment, which lasted 40min. The

    general feature is that an intergranular attack becomes more severe when movingin a direction away from the crevice mouth. The intergranular attack is usually

    caused by a composition difference at the grain boundary and prior heat treatment

    condition. Thus, for another iron or heat treatment no intergranular attack may be

    seen.

    Fig. 8shows scanning electron microscope (SEM) micrographs of the main fea-

    tures observed on the crevice-corrosion wall that included; xpass, severely attacked,

    and the etched (bottom) regions. At xpass, the attack is of intergranular nature; sim-

    ilar to that shown inFig. 7but at higher magnification with some scattered micropits

    are observed on the attacked grains. In this region and at xpassboundary the meas-

    ured electrochemical potential is162mV(SCE). From the polarization curve showninFig. 2, this potential is in the transition (active/passive region), in agreements with

    the physical observations in this region. Conventionally this area is associated with

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1347

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    13/16

    stress corrosion cracking, which generally coincides with intergranular attack at

    grain boundaries. Therefore, the IR voltage drop can bring the crevice wall in a

    potential region where the intergranular attack associated with stress corrosion

    cracking can take place. In the severely attacked region Fig. 8(b), it appears as a

    dense mountainous region with threadlike appearance. This changes to a porous

    etched surface showing a few scattered simple crystallographic pits in the etched re-

    gion located at the bottom of the crevice. Little penetration occurred in the etched

    region, Fig. 8(c), in agreement with the potential prevailing in this region being

    around Elim where the current is expected to be much lower than in the active se-

    verely attacked region.

    4. Conclusions

    Crevice corrosion occurred immediately without an induction time for a crevice iniron immersed in a strong-buffered acetate solution, under a condition where the

    crevice corrosion products were allowed to leave the crevice cavity by choosing a

    crevice in the upside down orientation. However, among other reported factors,

    the crevice geometry is important in determining the onset of crevice corrosion,

    where for the current crevice corrosion system, it is expected that no crevice cor-

    rosion will occur when L< 6.7mm.

    For a given crevice system there is a critical depth value, above whichxpassvaries

    little, and below which xpass increases substantially towards the crevice bottom.

    This continues until the depth L< Lc, where crevice corrosion will not occur

    immediately. Then, even if an induction period was allowed, crevice corrosion willbe in question for the current crevice geometry that does not allow corrosion

    products to accumulate inside the crevice.

    Fig. 7. Post-experiment greyscale photograph of the crevice wall atxpassboundary showing the corrosion

    attack for a crevice in iron exposed to acetate buffered solution. Esurf= 800mV(SCE), L= 0.85 cm and

    t= 40min. Magnification: 100.

    1348 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    14/16

    The active/passive boundary,xpass, on the crevice wall moved towards the crevice

    mouth (x= 0) for increasing time and increasing current flowing out of the crev-

    ice, in accordance with IR voltage theory and in agreement with other reportedfindings in the literature. The potential at xpasswas found to be constant irrespec-

    tive of time or crevice depth.

    Fig. 8. Post-experiment SEM micrographs of different regions on the crevice wall, inside a crevice in iron

    exposed to acetate buffered solution. Micrographs were taken after a crevice corrosion experiment with

    parameters: Esurf= 800mV(SCE), L= 0.8cm andt = 12h.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1349

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    15/16

    Potential distribution measurements inside the crevice, using a fine glass micro-

    probe, showed a large potential drop inside the cavity. A large potential drop

    of about 1.4V was measured at L= 15mm.

    The morphology ofxpassboundary showed an intergranular attack that gets moresevere when moving away from the crevice mouth. However, this is not conclusive

    for the current system as the intergranular attack is usually caused by a compo-

    sition difference at the grain boundary and prior heat treatment condition.

    Acknowledgments

    Acknowledgment is made to the Institute of Research & Consultation of King

    Abdulaziz University and to Saudi Basic Industries Corp. (SABIC) for their supportof this research. Professor Howard W. Pickering, The Pennsylvania State University,

    provided helpful comments.

    References

    [1] Howard W. Pickering, Corros. Sci. 29 (1989) 325;

    A. Valdes, H.W. Pickering, in Advances in Localized Corrosion, NACE-9, in: H. Isaacs, U. Bertacci,

    J. Kruger, Z. Smialowska (Eds.), Houston, TX, 1990, pp. 393401.

    [2] K. Cho, M.I. Abdulsalam, H.W. Pickering, J. Electrochem. Soc. 145 (1998) 1862.[3] M. Wang, H.W. Pickering, Y. Xu, J. Electrochem. Soc. 142 (1995) 2986.

    [4] M.I. Abdulsalam, H.W. Pickering, Corros. Sci. 41 (1999) 351.

    [5] E.A. Nystrom, J.B. Lee, A.A. Sagues, H.W. Pickering, J. Electrochem. Soc. 141 (1994) 358.

    [6] A.M. Abdullah, B.A. Shaw, H.W. Pickering, Crevice corrosion of aluminum 6XXX alloy, in: B.A.

    Shaw, R.G. Buchheit, J.P. Moran (Eds.), Proc. of Corrosion and Corrosion Prevention of Low

    Density Metals and Alloys, PV 2000-23, (ECS Phoenix Meeting), The Electrochemical Society,

    Penington, NJ, 2001, pp. 401410.

    [7] J.W. Oldfield, W.H. Sutton, Br. Corros. J. 13 (1978) 13.

    [8] H.W. Pickering, J. Electrochem. Soc. 150 (2003) K1.

    [9] W.D. France, N.D. Greene, Corrosion (Houston) 24 (1968) 247.

    [10] M.I. Abdulsalam, Corrosion 58 (2002) 364.

    [11] M.Z. Yang, M. Wilmott, J.L. Luo, Thin Solid Films 326 (1998) 180.[12] M. Vankeerberghen, M.I. Abdulsalam, H.W. Pickering, J. Deconinck, J. Electrochem. Soc. 150 (9)

    (2003) B445.

    [13] Y. Xu, H.W. Pickering, J. Electrochem. Soc. 140 (1993) 658.

    [14] B.G. Ateya, H.W. Pickering, J. Electrochem. Soc. 122 (1975) 1018.

    [15] J.C. Walton, G. Cragnolino, S.K. Kalandros, Corrosion Sci. 38 (1996) 1.

    [16] M.K. Sawford, B.G. Ateya, A.M. Abdullah, H.W. Pickering, J. Electrochem. Soc. 149 (2002) B198.

    [17] B.A. Shaw, P.J. Moran, P.O. Gartland, Corros. Sci. 32 (1991) 707.

    [18] A. Valdes, H.W. Pickering, in Advances in Localized Corrosion, NACE-9, in: H. Isaacs, U. Bertacci,

    J. Kruger, Z. Smialowska (Eds.), Houston, TX, 1990, pp. 393401.

    [19] H.W. Pickering, Corrosion 42 (1986) 125.

    [20] M.I. Abdulsalam, H.W. Pickering, J. Electrochem. Soc. 145 (1998) 2276.

    [21] J. Al-Khamis, H.W. Pickering, J. Electrochem. Soc. 148 (2001) B314.

    [22] B. DeForce, H. Pickering, J. Met., JOM 47 (1995) 22.

    [23] P.D. Peterson, A.P. Reynolds, M.A. Sutton, Corrosion 57 (2001) 693.

    1350 M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351

  • 8/12/2019 Behaviour of Crevice Corrosion in Iron

    16/16

    [24] K. Cho, H.W. Pickering, J. Electrochem. Soc. 138 (1991) L56.

    [25] Y. Xu, M. Wang, H.W. Pickering, J. Electrochem. Soc. 140 (1993) 3448.

    [26] R.S. Lillard, M.P. Jurinski, J.R. Scully, Corrosion 50 (1994) 251;

    R.S. Lillard, J.R. Scully, J. Electrochem. Soc. 141 (1994) 3006.

    M.I. Abdulsalam / Corrosion Science 47 (2005) 13361351 1351