Serega Lv 04

Embed Size (px)

Citation preview

  • 7/31/2019 Serega Lv 04

    1/22

    Effect of the strain rate on stress corrosioncrack velocities in face-centred cubic alloys.

    A mechanistic interpretation

    S.A. Serebrinsky, J.R. Galvele*

    Comisioon Nacional de Energa Atoomica, Departamento Materiales, Instituto de Tecnologia, Avda.

    Libertador 8250, 1429 Buenos Aires, Argentina

    Received 28 August 2002; accepted 2 July 2003

    Abstract

    Constant extension rate tests on smooth samples, with strain rate (SR) values from 106 s1

    up to 20 s

    1

    , were used to study stress corrosion cracking (SCC) systems in face-centred cubicalloys. It was found that by increasing the SR a monotonic increase of the log CPR (crack

    propagation rate) takes place. It was also observed that the slope a in log CPR vs. log SR plots

    had different values for different SCC morphologies. Intergranular SCC is more steeply ac-

    celerated by SR, aIG 0:50.7, than transgranular SCC, aTG 0:20.3. The differences foundbetween intergranular SCC and transgranular SCC were analysed under the light of the

    available SCC mechanisms.

    2003 Elsevier Ltd. All rights reserved.

    Keywords: A. Stainless steel; Silver; Brass; C. Stress corrosion; Effects of strain

    1. Introduction

    Numerous variables have a strong effect on the stress corrosion cracking (SCC)

    process in metals and alloys, the strain rate (SR) being one of them [1,2]. In a recent

    publication Serebrinsky et al. [3] studied the effect of SR on the crack propagation

    rate (CPR) for a variety of alloyenvironment systems. In their study the authors

    covered a wide range of SR values, going from 106 s1 up to 20 s1. They found

    that, while an increase in the SR produced an increase in the CPR, the effect was

    *Corresponding author. Tel.: +54-11-6772-7390; fax: +54-11-6772-7404.

    E-mail address: [email protected] (J.R. Galvele).

    0010-938X/$ - see front matter 2003 Elsevier Ltd. All rights reserved.

    doi:10.1016/S0010-938X(03)00172-0

    www.elsevier.com/locate/corsci

    Corrosion Science 46 (2004) 591612

    http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/
  • 7/31/2019 Serega Lv 04

    2/22

    more significant for intergranular SCC than for transgranular SCC. The results

    showed that the slope a in log CPR vs. log SR plots was aIG 0:50.7 for inter-granular cracking, while it only reached values of aTG 0:20.3 for transgranular

    cracking. This difference in behaviour between transgranular cracking and inter-granular cracking, was confirmed by Alvarez et al. [4]. Alvarez et al. compared single

    crystals of Ag10Au alloy with polycrystalline samples of the same alloy. These

    authors strained the samples in both a 1 M HClO4 solution and a 1 M KCl solution,

    and found slopes similar to those reported by Serebrinsky et al. [3].

    In the present work a systematic study was made of all the cases studied by Ser-

    ebrinsky et al. [3] in their preliminary work. Only homogeneous face-centred cubic

    (fcc) alloys were considered in this study, assuming that similar plastic behaviour

    would be found when straining the different fcc alloys. The previously reported

    slopes for both intergranular and transgranular SCC were confirmed. An analysis

    was made of the experimental results, under the scope of the available SCC mech-

    anisms [5,6], and it was concluded that the surface mobility SCC mechanism ac-

    counted for the experimental observations.

    2. Experimental method

    2.1. Materials

    Only homogeneous alloys, with an fcc structure, were used in the present work.The samples were 0.8 mm diameter wires. All the samples were degreased in acetone

    and subjected to a thermal treatment, as described below. Before each test, the

    samples were again degreased with acetone. The alloys used were identified by their

    commercial name, except for the silver alloys, which were identified by their nominal

    atomic % composition. The analytical composition of the alloys used, in wt.%, were:

    type AISI 304 stainless steel (Cr 16.9, Ni 8.0, Si 0.7, Mo 0.4, C 0.08, S 0.03, Fe

    balance), Ag15Pd (Pd 15.5, Pb < 0.01, Cd < 0.01, Cu 0.030.1, Al 0.003, Si 0.001, Fe

    0.01, Mg 0.01, Pt < 0.005, Au < 0.005, Rh < 0.15, Ir < 0.015, Ag balance), and yellow

    brass (Al 0.002, Si < 0.02, Fe 0.05, Ti 0.02, Mg 0.0002, Mn 0.0005, Sn 0.02, Pb 0.02,

    Cr 0.0050.02, Ni 0.050.2, Zn 34.8 0.1, Cu balance). The composition of the alloysAgxAu, where x 2, 5 and 15 at.% is given in Table 1. In the present work thesealloys are referred to as AISI 304, AgPd, AgAu and brass, respectively.

    Table 1

    Chemical composition of the AgAu alloys used

    Alloy

    (at.%)

    Element (wt.%)

    Au Al Cu Pt Si Fe Mg Ag

    Ag2Au 3.99

  • 7/31/2019 Serega Lv 04

    3/22

    All the thermal treatments were carried out in argon atmospheres. The AISI 304

    was annealed for 30 min at 1100 C, and water quenched. The AgPd and AgAu

    alloys were annealed at 800 C for 1 h, and air-cooled. Brass was annealed for 24 h at454 C and water-quenched. No traces of b phase were detectable in the annealed

    brass samples.

    The mechanical properties of all the alloys were measured at strain rates ranging

    from 105 s1 up to 1 s1. The properties measured were: the conventional flow stress

    for 0.2% plastic strain (r0:2), the ultimate tensile strength (UTS) and the true strain to

    rupture (er). For all the alloys the three parameters either increased slightly with SR,

    or remained constant. In general the variation of UTS with SR was slightly higher

    than that of r0:2. No measurements of mechanical properties were made in the range

    120 s1. Since the change of the mechanical properties with SR, in the measured

    range was small and smooth, it was assumed that the same should be expected for

    the SR values not covered during the mechanical properties measurements. Table 2

    shows the mechanical property values found at 105 s1.

    2.2. Environments

    All the solutions used were prepared with analytical grade reagents and triply

    distilled water with a minimum resistivity of 18.2 MX cm. AISI 304 was tested in an

    11.8 M LiCl solution at 130 C. In this case the temperature was kept constant within

    1

    C. All the other systems were tested at room temperature. The AgPd alloy wastested in aqueous 1 M KCl and 1 M KI solutions. The AgAu alloys were tested in

    aqueous 1 M HClO4 solution. Brass was tested in 1 M NaNO2 solution and in a

    solution with the following composition: 0.05 M CuSO4 + 1 M (NH4)2SO4, the pH of

    the solution was adjusted to 6.5 with NH4OH. In the text this last solution is referred

    to as Mattsons solution.

    2.3. Polarization curves

    Anodic polarization curves were measured in a conventional three-electrode glass

    cell, with a potential scanning rate of 0.5 mV s1. The solutions were deaerated withprepurified nitrogen [7], in a scrubber connected to the cell. After a 60 min deaer-

    ation, the solution was transferred to the cell. The samples were allowed to reach a

    Table 2

    Mechanical properties of the alloys used, at 105 s1

    Alloy r0:2 (MPa) UTS (MPa) er (%)

    AISI 304 206 885 34.5Ag15Pd 113 321 26.6

    Ag2Au 55 182 18.8

    Ag5Au 55 178 15.9

    Ag15Au 53 180 16.4

    Brass 95 495 40.0

    No significant differences were observed when varying the strain rate.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 593

  • 7/31/2019 Serega Lv 04

    4/22

    steady potential, and then the potential was scanned, beginning at a potential 100

    mV below the corrosion potential. The reference electrodes used were: for AISI 304

    in the LiCl solution a calomel saturated reference electrode, for brass in aqueous

    NaNO2 solution, as well as for AgPd alloy in KI solution and for AgAu alloys inHClO4 solution a mercurous sulphate reference electrode, for AgPd in KCl, a silver

    chloride reference electrode, prepared according to [8]. In the case of brass in

    Mattsons solution a pure copper wire was used as a reference electrode. All the

    potentials are reported in the standard hydrogen electrode (SHE) scale, with the only

    exception of brass in Mattsons solution, where the potential applied in the straining

    experiments was the equilibrium potential of copper in the solution.

    2.4. Stress corrosion tests

    The SCC tests were carried out at constant strain rate. For convenience, the ex-

    periments were divided in three strain rate ranges: slow strain rate tests (SSRT) for

    SR values in the range of 106104 s1, intermediate strain rate tests (ISRT) for SR

    values in the range of 1041 s1, and ultrafast strain rate tests (UFSRT) for SR

    values higher than 1 s1. For SSRT and ISRT the machines used were conventional

    constant crosshead speed straining machines, and the cell used was described in [9].

    The total length of the sample, between grips, was 120 mm, and the part of the

    samples exposed to the corrosive solutions had an initial area of 0.8 mm 2. The

    UFSRT were performed in a drop weight apparatus described in [10], and the cell

    was similar to that for SSRT, but arranged for vertical straining. In the UFSRT thefailure time tr was calculated by detecting the starting point and the end of the test

    with two switches wired to a Tektronix TDS-210 digital oscilloscope. With the fixed

    image in the oscilloscope tr ttotDLr=DLtot was calculated, ttot and DLtot being thetotal time and length of the mobile grip displacement, and DLr the elongation to

    rupture.

    The electrode potentials were measured through a Luggin capillary and were kept

    constant with a LYP M7 potentiostatgalvanostat. The procedure for loading the

    solution into the cell as well as the reference electrodes used were those described for

    the polarization curves. After the samples reached a steady potential, the chosen

    potential was applied. The samples were kept at constant potential for a short time,and then the straining of the samples was started. The exposure at constant potential

    pointed to obtaining a reproducible initial surface, but without allowing for excessive

    surface corrosion. Typical exposure time values, at constant potential, were the

    following: for AISI 304 and AgPd, 10 min, for AgAu, between 15 s and 10 min,

    depending on the current density and for brass in Mattsons solution, 3 min. The

    potentials applied for each experiment, are described in the next section.

    The samples were strained to fracture. Afterwards the fracture surface and lateral

    surfaces of the samples were observed with a Phillips 500 scanning electron mi-

    croscope (SEM). Then the samples were mounted for metallographic sectioning and

    observation. The crack propagation rate was calculated by dividing the crack lengthLf by the failure time tf. The former was found either by measuring the maximum

    radial length of the brittle part on the fracture surface observed by SEM, or by

    594 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    5/22

    measuring the maximum crack length on the mounted section. In those cases where

    both measurements were taken into consideration, the former observations were

    identified as SEM values, and the latter ones as MET values.

    3. Experimental results

    3.1. Polarization curves

    3.1.1. AISI 304 in LiCl

    The polarization curve of AISI 304 in 11.8 M LiCl solution at 130 C shows a

    passive range was observed between the corrosion potential, Ecorr 0:5 VSHE, and

    )0.1 VSHE. The passive current density was of the order of 10

    5

    A/cm2

    . At potentialsabove Ep 0:1 VSHE, pitting corrosion started, and the current density rose untilohmic drop limitations appeared. The results obtained were similar to those reported

    by Wilde [11], and by Duffoo et al. [12].

    3.1.2. AgPd in halides

    Fig. 1 shows the polarization curves for pure silver and for Ag15Pd in a 1 M KCl

    solution. The arrows indicate the potential values used in the present work for the

    stress corrosion tests. The curves found were similar to those reported by Duffoo and

    Galvele [13]. A detailed analysis of the polarization curves was reported by these

    authors. According to Duffoo and Galvele, for Ag15Pd, in 1 M KCl solution, be-tween 0.2 and 0.4 VSHE selective dissolution of silver was taking place. No traces of

    Pd were detected by them in the solution after exposing the alloy for 22 h at 0.4 VSHE.

    On the other hand, above 0.45 VSHE the increase in the current density was

    0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.810

    -3

    10-2

    10-1

    100

    101

    102

    103

    1M KClAg 99.99%Ag - 15Pd

    I(A/m2

    )

    E(V/SHE)

    Fig. 1. Polarization curves of Ag15Pd and Ag 99.99% in 1 M KCl solution. The arrows show the po-

    tentials applied in stress corrosion tests.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 595

  • 7/31/2019 Serega Lv 04

    6/22

    accompanied by simultaneous dissolution of Ag and Pd. At 0.6 VSHE the amount of

    Pd found by them in the solution indicated a quasi-stoichiometric dissolution of the

    alloy.

    3.1.3. AgAu in HClO4The polarization curves for pure silver and for AgxAu (x 2; 5; and 15 at.%)

    alloys in a 1 M HClO4 solution were similar to those reported by Maier et al. [14].

    These authors measured the dissolution rate of the alloys in 1 M HClO4 solution, at

    constant potential, and calculated the rate of formation of dealloyed films on the

    metal surface.

    3.1.4. Brass in Mattsons solution

    No measurements were made for the polarization curve in this system. The SCC

    tests in this system were made at the equilibrium potential of pure copper in the

    corrosive solution. This choice was based on the observations published by Montoto

    et al. [15].

    3.1.5. Brass in NaNO2For the SCC tests in this system the polarization curves published by Alvarez

    et al. [16] were used.

    3.2. Stress corrosion tests

    3.2.1. AISI 304 in LiCl

    Under the experimental conditions chosen, the fracture surface of AISI 304 was

    always transgranular (TG). The experimental conditions were selected to disfavour

    intergranular (IG) cracking. In this way the complications in the interpretation of the

    results due to mixed cracking morphology were avoided. As shown by Galvele et al.

    for hot MgCl2 [17] and LiCl [12] solutions, when increasing the potential the TG

    fraction in the cracks increased. The fracture surfaces of the samples strained in the

    present work showed typical features of TG cracking, such as river patterns and fan-

    shaped marks starting from the initiation points. The longest cracks were usually

    observed to initiate at large pits. For the lowest SR (2.4 106

    s1

    ), usually there wasonly one, passing-through, crack. Sometimes, tiny cracks appeared in the region

    close to the main crack. When SR increased, up to about 5 103 s1, a higher

    number of cracks, with a length similar to the longest one, was found. For higher SR

    the number of cracks decreased again, but this fact should be associated to a limi-

    tation in the detection of cracks rather than to their absence. For example, for the

    highest SR value in the ISRT (0.044 s1) pitting was still effective in inducing

    cracking, but no cracks were detectable in the UFSRT, at 18 s1.

    Scanning electron microscopy (SEM) and metallographic mounting (MET) usu-

    ally gave comparable values of CPR. Fig. 2 shows crack propagation rates versus

    strain rate. In the same figure the lower detection limit of the constant strain ratetechnique is shown. This limit is defined as the length of the smallest detectable crack

    Lfis;min (here taken conservatively as 4 lm) divided by the largest cracking time tr;max.

    596 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    7/22

    As mentioned in a previous publication [3] and confirmed in the present work,

    there is a significant difference in the slope, a, of the loglog plot when intergranular

    SCC is compared with transgranular SCC. The slope a of the loglog plot is defined

    as

    a o logCPR

    o logSR; 1

    and in the case of AISI 304 in LiCl it was 0.31. An extrapolation of this linear re-

    lationship to a SR of 18 s1 showed that the expected crack propagation rate was

    below the detection limit, even when the expected CPR was as high as 2 105 m s1.

    If cleavage cracks, longer than 4 lm were produced, they should have been detect-

    able. Nevertheless, in the present work no such cracks were found for transgranular

    SCC of AISI 304 in LiCl solution.

    3.2.2. AgPd in halides

    This system has been previously studied with SSRT [13]. It was found that the

    CPR generally increased with the potential, from the potential of formation of the

    silver halide up to an overpotential of %300 mV. For higher potentials the CPRreached a plateau [13].

    As reported by Duffoo and Galvele [13], fracture morphology was always IG in

    both KI and KCl solutions, with no traces of TG cracking. The fracture surfaces

    were covered with AgI and AgCl, respectively, as identified by EDAX. For low

    potentials in KCl, the coverage was incipient, and only a small number of crystals

    were observed. On the other hand, for high potentials a dense coverage with manysilver halide crystals was found. Similarly as in AISI 304, for low SR there was only

    one, passing-through, crack. Small number of cracks were also observed for high

    10

    -6

    10

    -5

    10

    -4

    10

    -3

    10

    -2

    10

    -1

    10

    0

    10

    110

    -9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    CPR

    (m/s)

    Strain Rate (s-1)

    Lower Det. Lim.

    NF

    2 tests

    AISI 304 in 11.8M LiCl

    SEM

    MET

    Fig. 2. Effect of strain rate on crack propagation rates of AISI 304 strained in a 11.8 M LiCl solution at

    )0.1 VSHE and 130 C. The lower detection limit expected for the experimental technique used is shown.

    NF indicates that no cracks were detected at the end of the experiment.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 597

  • 7/31/2019 Serega Lv 04

    8/22

    potentials. When SR increased, or the potential decreased, a larger number of cracks

    with similar length appeared. A characteristic feature, quite common in intergran-

    ular stress corrosion fracture surfaces, was the presence of shallow pits or vacancy

    clusters, as previously reported by Maier et al. [14].

    Unlike to what was observed with AISI 304 in LiCl, AgPd alloy in halide so-

    lutions gave measurable cracks for UFSRT. The cracks in a sample strained in 1 MKI at 5 s1 were about 100 lm in size. The samples strained in KCl showed shorter

    cracks, but still above the detection limit of the experimental technique. The slopes a

    in the logarithmic plot of Fig. 3 were clearly steeper than those found for AISI 304 in

    LiCl solution. The values measured for a were between 0.5 and 0.65, and were listed

    in Table 3. One important observation was that, as shown in Fig. 4, in the UFSRT,

    the cracks measured at 5 s1 were always longer than those measured at 18 s1. This

    observation was important because if the cracks produced at 5 s1 were due to a

    single pseudo-cleavage type event, the same crack length should have been expected

    for the experiments at 18 s1.

    Fig. 5 shows the effect of electrode potential on the crack propagation rate atvarious strain rate values, for Ag15Pd in 1 M KCl solution. It was observed that an

    increase in the potential (E) produced an increase in CPR, up to an overpotential of

    about 300 mV, where a plateau was reached. The shape of the curves of CPR in

    function of E was similar for all strain rates tested, but the curves were shifted to

    higher CPR values as the SR increased. The present results could lead to introduce in

    Eq. (1) the effect of the potential:

    log CPRSR;E fE a log SR 2

    suggesting that the effects of the electrochemical variable (E) and the mechanical

    variable (SR) could be independent. The slope a remained fairly independent of E.According to the measurements of Yu and Parkins, Eq. (2) is also valid for brass in

    nitrites [18]. The inverse slopes p of the ascending part of the CPRE curves were

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    101

    10-9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    Lower Det. Lim.

    1M KCl 0.322VSHE 0.522VSHE0.372V

    SHE0.572V

    SHE

    0.422VSHE

    0.622VSHE

    0.472VSHE

    1M KI-0.15V

    SHE

    Ag-15Pd

    CPR(m/s)

    Strain Rate (s-1)

    Fig. 3. Effect of strain rate on crack propagation rates of Ag15Pd in 1 M KI and KCl solutions.

    598 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    9/22

    approximately 75 mV/decade, close to the value of 100 mV/decade reported by Galvele

    et al. for low-gold AgAu alloys in 1 M potassium halide solutions at SSRT [19].

    3.2.3. AgAu in HClO4

    This system has been previously studied by Maier et al. [14] at SSRT, and byDuffoo and Galvele [20] and by Kelly et al. [21] at UFSRT. The experimental results

    from the present work are shown in Figs. 6 and 7.

    Table 3

    Compilation of the slopes in log CPRlog SR plots from the present work

    Alloy Environment Potential (VSHE) Morph. a a0

    AISI 304 11.8 M LiCl,130 C

    )0.10 TG 0.31 0.31

    Ag15Pd 1 M KI 0.15 IG 0.65 0.52

    Ag15Pd 1 M KCl 0.32 IG 0.59

    Ag15Pd 1 M KCl 0.37 IG 0.63 0.58

    Ag15Pd 1 M KCl 0.42 IG 0.53

    Ag15Pd 1 M KCl 0.47 IG 0.53 0.51

    Ag15Pd 1 M KCl 0.52 IG 0.55 0.47

    Ag15Au 1 M HClO4 0.90 IG 0.59 0.50

    Ag15Au 1 M HClO4 1.00 IG 0.57

    Ag15Au 1 M HClO4 1.15 IG 0.62 0.48

    Ag5Au 1 M HClO4 0.90 IG 0.45

    Ag2Au 1 M HClO4 0.90 IG 0.40

    a-brass Mattsons sol.a (0.0 VCu0=Cu ) IG 0.54 0.44

    a-brass 1 M NaNO2 0.20 TG 0.19 0.19

    The a0 values were calculated excluding the measurements for SR higher than 1 s1.a 0.05 M CuSO4 +1 M (NH4)2SO4 + NH4OH adjusted to pH 6.5.

    100 1010

    10

    20

    30

    40

    50

    60

    Lower Det. Lim.

    Strain Rate (s-1)

    Lfis

    (m)

    0.372VSHE0.472VSHE0.522V

    SHE0.572VSHE0.622VSHE

    Ag-15Pd in 1M KCl

    Fig. 4. Crack lengths measured at 5 s1 and at 18 s1 for Ag15Pd in 1 M KCl solutions. By increasing the

    SR, in the UFSRT region, the crack lengths were found to decrease.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 599

  • 7/31/2019 Serega Lv 04

    10/22

    The fracture surfaces found were always intergranular, though small isolated

    transgranular patches were also detected. Vacancy clusters or shallow pits, as de-

    scribed by Maier et al. [14], were also present in this system. The effect of the po-

    tential and strain rate on the number and length of the cracks was the same as for

    Ag15Pd in KCl, i.e. few and long cracks, with one much longer than the others, was

    the situation favoured by high potentials and low SR.In this system a comparison was made on the effect of SR on CPR as a

    function of alloy composition, Fig. 6, and also for one alloy at different potentials,

    300 350 400 450 500 550 600 65010

    -9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    Strain Rate:

    Ag-15Pd in 1M KCl

    2.3 E-61/s8.8 E-61/s

    7.3 E-51/s1.8 E-31/s4.4 E-21/s5.01/s18 1/s

    CPR(m/s)

    E(mV/SHE)

    Fig. 5. Effect of potential on the crack propagation rate at various strain rate values, for Ag15Pd in 1 M

    KCl solution.

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    101

    10-9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    Lower Det. Lim.

    Ag-2AuAg-5AuAg-15Au

    Ag-xAu in 1M HClO4, 0.9VSHE

    CPR(m/s

    )

    Strain Rate (s-1)

    Fig. 6. Effect of strain rate on crack propagation rates of AgAu alloys in 1 M HClO4, at one potential

    0.9 VSHE.

    600 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    11/22

    Fig. 7. The effect of SR was the same as found with AgPd alloy, the slopes a

    were higher than those found with AISI 304. The a values found were between 0.4

    and 0.65.

    Fig. 7 shows the effect of the electrode potential on the SR versus CPR for Ag

    15Au in 1 M HClO4. When the potential was increased it was found that the CPRalso increased. The inverse slopes pwere about 230 mV/decade, for SSRT, ISRT and

    UFSRT. These results were very close to the value of ca. 200 mV/decade for Ag

    2Au, Ag5Au, Ag10Au and Ag15Au at SSRT reported by Galvele et al. [19]. In

    the present system no plateau velocity was reached, possibly due to the very high

    dissolution rate at high potentials. The effect of increasing SR was again to shift the

    CPRE curves upwards, while keeping the shape of the curves. Here again Eq. (2)

    was applicable.

    3.2.4. Brass in Mattsons solutionThis system showed intergranular stress corrosion cracks on the fracture surfaces,

    with occasional transgranular patches. The variation of the number and distribution

    of the length of cracks with strain rate was similar to that described in the previous

    systems, including the decrease in the number of cracks at UFSRT observed in AISI

    304. Fig. 8 shows the logarithmic plot of CPRSR, where a slope a of 0.54 was

    found.

    3.2.5. Brass in 1 M NaNO2 solution

    The fracture surfaces in this system were transgranular, as described by Alvarezet al. [16] and by Rebak et al. [22]. The results of the straining experiments are shown

    in Fig. 9, and the slope for a was 0.19 [3].

    10-6 10-5 10-4 10-3 10-2 10-1 100 10110

    -9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    0.9VSHE

    1VSHE1.1V

    SHE

    1.15VSHE

    Ag-15Au in 1M HClO4

    CPR(m/s)

    Strain Rate (s-1)

    Lower Det. Lim.

    Fig. 7. Effect of the electrode potential on the SR versus CPR for Ag15Au in 1 M HClO4.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 601

  • 7/31/2019 Serega Lv 04

    12/22

    4. Discussion

    4.1. Difference between IG and TG cracking

    The analysis of the experimental results described above indicates that there is aclear and consistent difference in the values of the logarithmic slope a, Eq. (1), when

    intergranular SCC systems are compared with transgranular SCC systems. It was

    10

    -6

    10

    -5

    10

    -4

    10

    -3

    10

    -2

    10

    -1

    10

    0

    10

    110

    -9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    Lower Det. Lim.

    CPR(m/s)

    Strain Rate (s-1)

    Brass in Mattson's solution

    Fig. 8. Effect of strain rate on crack propagation rates of brass in Mattsons solution (0.05 M CuSO4 + 1

    M (NH4)2SO4 + pH 6.5 adjusted with NH4OH) at 0 VCu.

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    101

    10-9

    10-8

    10-7

    10-6

    10-5

    10-4

    10

    -3

    10-2

    10-1

    CPR - UTS

    CPR - 0.2

    Lower Det. Lim.

    CPR(m/s)

    Strain Rate (s-1)

    Brass in 1M NaNO2solution

    Fig. 9. Effect of strain rate on crack propagation rates of brass in 1 M NaNO2 solution at 25 C and 0.2

    VSHE. The meaning of the CPR lines will be considered in Section 4.

    602 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    13/22

    always found that the slope for intergranular SCC was higher than that for trans-

    granular SCC:

    aIG > aTG: 3

    Table 3 shows the values for a found in the present work. Since the values of CPR

    could not be measured in some systems for UFSRT, Table 3 also includes the values

    for a0, defined by the same Eq. (1) but taking into account the measurement for SR

    up to 1 s1 (i.e. neglecting UFSR tests).

    The data in Table 3 shows that Eq. (3), either for a or a0, is satisfied in all the

    systems studied. One objection that could be raised is that different systems were

    compared. In support of the conclusion reached in Eq. (3) the following facts should

    be taken into account. (a) Only fcc alloys were considered. Thus systems with similar

    deformation mechanisms were compared. (b) One same alloy, from the same batch

    (brass), was subjected in the present work to IGSCC (Mattsons solution) and

    TGSCC (1 M NaNO2 solution) and the results were compared. The values found,

    for the same alloy in two different environments, were aIG 0:54 and aTG 0:19,which confirms Eq. (3). (c) Alvarez et al. [4] studied one single alloy system (AgAu

    alloy) in the same environment, and produced IGSCC and TGSCC. For this purpose

    these authors compared AgAu single crystals with AgAu poly-crystals. The studies

    were made in 1 M HClO4 solution and in 1 M KCl solution. As shown in Table 4,

    and discussed below, Alvarez et al. were able to confirm the validity of Eq. (3).

    To further confirm the validity of Eq. (3), values of a published in previous

    publications were collected in Table 4. As mentioned above, a major point in the

    analysis of the present results is provided by the work of Alvarez et al. [4]. Theseauthors strained single crystals of Ag10Au in 1 M HClO4 and 1 M KCl solutions.

    As for the gold contents x, Maier et al. [14] showed that for x between 2.2 and 15

    at.% Au there was no noticeable effect of the alloy composition on CPR, in experi-

    ments made at a slow strain rate. In the present work it was found that this

    observation is correct also when measurements are made with the ISRT. As a

    consequence, the measurements made by Alvarez et al. [4] can be compared with

    those reported in the present work. Alvarez et al. found for TGSCC aTG 0:22,while for IGSCC they found aIG 0:48. In the present work the lowest value for

    Table 4Compilation of the slopes a in logCPRlog SR plots from previous publications

    Alloy Environ-

    ment

    Reference Potential

    (VSHE)

    Morph. a a0

    Ag20Au 1 M HClO4 [20] 1.20 IG 0.60

    Ag20Au 1 M KCl [20] 0.70 IG 0.63 0.69

    Ag15Au 1 M KCl [4] 0.50 IG 0.68

    Ag15Au 1 M HClO4 [4] 0.90 IG 0.48

    Ag10Au

    single crystal

    1 M KCl [4] 0.50 TG 0.30

    Ag10Au

    single crystal

    1 M HClO4 [4] 0.90 TG 0.22

    IG: Intergranular SCC; TG: Transgranular SCC. The a0 values exclude measurements for SR higher than

    1 s1.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 603

  • 7/31/2019 Serega Lv 04

    14/22

    IGSCC (a0 for Ag2Au) was aIG 0:4, all the other a values being higher. When themeasurements were made in KCl solution, the difference between IGSCC and

    TGSCC was maintained. Alvarez et al. measured for TGSCC aTG 0:30 and for

    IGSCC aIG 0:68. Duffoo and Galvele [20] reported for IGSCC of Ag20Au in 1 MKCl aIG 0:63 and in 1 M HClO4 aIG 0:60.

    Finally, it is important to point out that in the present work, in the case of

    TGSCC of stainless steels in hot concentrated chloride solutions, as well as brass in

    nitrite solutions, no cracks were detected after UFSRT. On the other hand, no

    reference was found in the literature of detection of any incipient transgranular

    cracks after an UFSRT. If there were any cracks, they were smaller than the de-

    tection limit of the technique used in the present work.

    4.2. Mechanistic interpretation

    The above results will be analysed within the scope of the surface mobility SCC

    mechanism. When appropriate, references to the film induced cleavage SCC mech-

    anism [5], and the anodic dissolution SCC mechanism [5] will be made. Since the

    above analysis involves the use of the equations developed for the surface mobility

    SCC mechanism, a brief review of those equations will be made.

    The surface mobility mechanism [5,23,24] is based on the assumption that the

    environment acts by increasing the surface mobility of the alloy, and that the cracks

    propagate by capture of vacancies at the tip of the stressed crack. This mechanism

    was used to quantitatively explain numerous cases of SCC [5,6,10,24,25]. The fol-lowing equation was developed for the crack propagation rate [23,24]:

    CPR DS

    Lexp

    rTa3

    kT

    1

    ; 4

    DS being the surface self-diffusion coefficient at the crack walls, L the diffusion dis-

    tance of the vacancies, rT the elastic stress at the tip of the crack, a the atomic size,

    k the Boltzmann constant, and T the temperature in K.

    Since measured DS values, for the conditions of interest for SCC, are very scarce

    two different approaches can be used to calculate the value ofDS. When a solid film,with a known melting point, contaminates the crack surface, an empirical equation,

    based on the work of Gjostein [26] and Rhead [27,28] can be used to calculate DS[23]:

    DS m2 s1 740 104 exp126Tm=RT 0:014 10

    4 exp54Tm=RT;

    5

    R being the molar gas constant (R 8:314 J mol1 K1), T the absolute temperaturein K, and Tm the melting point of the surface film in K.

    On the other hand, when no surface films are formed, but the SCC process isspecifically affected by the concentration of cations of the noble metal of the alloy, as

    in the case of AgCd alloy in AgNO3 solution, Galvele and Duffoo [29] developed an

    604 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    15/22

    equation for the calculation ofDS, based on the alloy composition and the exchange

    current density of the noble metal.

    4.3. Transgranular stress corrosion cracking

    4.3.1. AISI 304 in LiCl

    The crack propagation rate values measured with SSRT for AISI 304 in LiCl

    solution at 130 C, Fig. 2, are very close to those found by Speidel [30] using fracture

    mechanics techniques. This coincidence in the measured CPR values proved that the

    SSRT is a valid technique for measuring crack propagation rates in the present

    system.

    As mentioned by Duffoo et al. [12] for this SCC system, assuming that the surface

    mobility SCC mechanism was operative, the analysis of the temperature dependenceof CPR showed that DS at 130 C was approximately 3 10

    15 m2 s1. For SSRT the

    predicted CPR values, with the stress rT equal to r0:2 (206 MPa, Table 2), agreed

    very well with the present experimental results. When increasing the strain rate, the

    samples fractured with noticeable plastic deformation. Then it is reasonable to as-

    sume that the value of rT will increase, eventually reaching the UTS value (885 MPa,

    Table 2). As shown in Fig. 10, using these two rT values (Table 2) the extreme CPR

    values predicted by the surface mobility mechanism contain the experimentally

    measured values.

    There is no information about the exact values rT will have at different strain

    rates. Nevertheless, in SSRT the samples fracture with very small plastic deforma-tion; consequently it is reasonable to assume that the stress at the tip of the crack will

    be close to r0:2. On the other hand, under high strain rates, where cracks propagate

    while the samples are undergoing strong plastic deformation, the first assumption

    would be that the maximum value reached by the stress at the tip of the crack would

    be that given by the UTS. The results in Fig. 10 suggest that the surface mobility

    SCC mechanism could give a good account for the correlation found between CPR

    and SR for transgranular SCC. No equations were found in the literature for a

    similar analysis from the point of view of the film induced cleavage SCC mechanism,

    or the anodic dissolution SCC mechanism. Consequently, from Fig. 10 no conclu-

    sions can be drawn either supporting or disqualifying these two mechanisms.

    4.3.2. Brass in NaNO2The reactions taking place at the tip of the crack, during SCC of brass in NaNO 2

    solutions are only ambiguously known. Newman and Burstein [31] suggested that

    decomposition of nitrite ions could lead to the formation of ammonia, with subse-

    quent stability of copper ions by complex formation. On the other hand Rebak et al.

    [22] concluded that passivity breakdown was the step previous to SCC of brass in

    NaNO2 solution. For any of these alternatives, the environment at the crack tip

    would be a solution with a high concentration of copper ions, providing a highexchange current density i0 of Cu$Cu

    . In this case, surface diffusion would be

    associated with the exchange current density, as in the model of Galvele and Duffoo

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 605

  • 7/31/2019 Serega Lv 04

    16/22

    [29]. Giordano et al. [32] and Montoto et al. [15] confirmed the validity of this hy-

    pothesis.

    When the value of DS is not known, but experimental CPR values are available

    for a given experimental condition, Eqs. (4) and (5) can be used to predict the CPR

    value at a different condition. This was done, for example, to predict CPR values foralloy 600 at different temperatures [25]. In the present case it will be assumed that in

    a SSRT the value of rT is equal to r0:2 and with the CPR value measured at SSRT,

    plus Eqs. (4) and (5), a DS value fitting these results can be calculated. Assuming that

    this value does not change with the SR, and also assuming, as done above, that the

    maximum stress value at the tip of the crack will be the UTS, the maximum CPR

    expected can be calculated. Fig. 9 shows that, as it was the case with AISI 304 in LiCl

    solution, the CPR values measured at different SR for SCC of brass in NaNO 2solutions fall between the extreme values predicted by the surface mobility SCC

    mechanism.

    4.4. Intergranular stress corrosion cracking

    4.4.1. Brass in Mattsons solution

    Speidel [33], using fracture mechanics techniques, for brass, of a composition

    equal to that used in the present work and exposed to a NH 4OH solution, reported

    CPR values very close to those found in the present work, with SSRT (Fig. 8). As

    mentioned above for AISI 304, this observation gives support to the validity of the

    CPR values measured with SSRT.

    The same procedure used for brass in NaNO2 solutions, was applied to the CPRresults measured for brass in Mattsons solution. As before, the value of DS fitting

    the SSRT results was calculated. Then with the DS value found, the CPR for rT equal

    10-6 10-5 10-4 10-3 10-2 10-1 100 10110

    -9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    Lower Det. Lim.

    NF2 tests

    AISI 304 in 11.8M LiCl

    SEM CPR - 0.2MET CPR -

    UTS

    CPR(m/s)

    Strain Rate (s-1)

    Fig. 10. Crack propagation rates of AISI 304 in 11.8 M LiCl at 130 C. Comparison of measured values

    with the predicted range of the surface mobility SCC mechanism.

    606 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    17/22

    to the UTS was calculated. As shown in Fig. 11, contrary to what was found for

    transgranular SCC, Figs. 9 and 10, the extreme values of CPR predicted by the

    surface mobility SCC mechanism were not the limits for the experimentally found

    CPR for intergranular SCC. As shown in Fig. 11, when the strain rate increased, the

    measured crack propagation rates exceeded the values predicted by surface mobility,if as done before, the maximum stress rT acting at the tip of the crack was considered

    to be the UTS. A calculation of the maximum stress rT required by Eq. (4) to fit the

    CPR measured in UFSRT gave a value of 2.7 GPa, which was about five times larger

    than the UTS. The stress values required by Eq. (4) to fit the experimental CPR

    values are shown in Fig. 12.

    One possible reason for the different effect on SR, between intergranular and

    transgranular SCC, could be the fact that grain boundaries act as barriers for the free

    movement of dislocations, leading to dislocation pileups [34]. At an atomic level, at

    the tip of the crack, this process will lead to an increase in the value of rT well above

    the externally applied stress. As discussed below, the dislocation pileups could ac-count for the required stress values mentioned in Fig. 12. A very crude calculation of

    the formation rate of dislocations, at different strain rates, can be attempted by using

    the equations developed by Van Bueren [35]. Fig. 13 shows the results of such cal-

    culations for copper. It is interesting to point out the similarity between Fig. 13 and

    Figs. 7 and 8. If the role of dislocations was to produce a surface slip step, where

    corrosion would be localised, as it is assumed in the slip-step anodic dissolution SCC

    mechanism [5], an increase in the strain rate should produce an increase in the crack

    propagation rate, as found in the present work. Nevertheless, the anodic dissolution

    SCC mechanism does not predict a difference in behaviour between IGSCC and

    TGSCC. On the other hand, any mechanism based on the elastic stresses at the tip ofthe crack will be able to explain the difference between IGSCC and TGSCC, because

    of the higher stress values induced by the dislocation pileups at the grain boundaries.

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    101

    10-9

    10-8

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    CPR - UTS

    CPR - 0.2

    CPR(m/s)

    Brass in Mattsons solution

    Strain Rate (s-1)

    Fig. 11. Comparison of crack propagation rates of brass in Mattsons solution (0.05 M CuSO4 + 1 M

    (NH4)2SO4 + pH 6.5 adjusted with NH4OH) at 0 VCu, measured and predicted by surface mobility.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 607

  • 7/31/2019 Serega Lv 04

    18/22

    4.4.2. AgPd in halides

    The same procedure used above was applied to the CPR results measured for Ag

    Pd in halide solutions. As before, the values ofDS fitting the SSRT were calculated.

    Then, with the DS values found, the CPR for values rT equal to the UTS were

    calculated.

    For potassium iodide environment, with the calculated value ofDS, results similar

    to those in Fig. 11 were found. The relation of rT=UTS required to fit the measuredCPR had a maximum value of 4.5, very close to the value %5 indicated in Fig. 12.This observation supports the same assumptions made above for IGSCC.

    10-6 10-5 10-4 10-3 10-2 10-1 100 1010

    500

    1000

    1500

    2000

    2500

    3000

    Stress(MPa)

    0.2

    UTS

    required

    Strain Rate (s-1)

    Brass in Mattson's solution

    Fig. 12. Stresses rT of Eq. (4) required to fit the measured crack propagation rates of brass in Mattson s

    solution (0.05 M CuSO4 + 1 M (NH4)2SO4 + pH 6.5 adjusted with NH4OH) at 0 VCu.

    10-6

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    101

    104

    105

    106

    107

    108

    109

    1010

    1011

    1012

    Disloc.

    Density(cm

    s

    )

    -2

    -1

    Strain Rate (s-1)

    Fig. 13. Rough estimate of the rate of production of dislocations, in copper, at various strain rates, based

    on Van Buerens equations [35].

    608 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    19/22

    For potassium chloride environment a range of potentials was applied, and more

    data is available. The polarization curve, Fig. 1, shows a noticeable discontinuity at

    about 0.45 VSHE, related to the change in the mode of dissolution of the alloy.

    Nevertheless, as reported by Duffoo and Galvele for SSRT [13], and confirmed in thepresent work with different SR values, no discontinuities were observed in CPR at

    any SR, Figs. 3 and 5. These observations suggest that, at least for IGSCC of Ag

    15Pd in 1 M KCl solution, the dissolution mode of the alloy had no effect on the

    SCC process.

    As regards Eq. (2), and following the predictions of the surface mobility SCC

    mechanism, the present results suggest the following: apparently, if Eq. (4) is taken

    into account, the effect of SR is concentrated mainly on the mechanical variable rT,

    while the potential Eacts mainly by changing the value ofDS (and possibly to some

    extent L).

    4.4.3. AgAu in HClO4Due to the high solubility of AgClO4 in water, 26.9 moles per litre of water [36], it

    is probable that during dissolution, an acid concentrated AgClO4 solution will be

    present next to the alloy interface, but no surface compounds will be expected. No

    measurements are available at present of the DS value expected in this system. Vela

    et al. [37] reported that the rate of step movement on pure Ag surfaces, in 1 M

    HClO4 solution, was an order of magnitude higher than in vacuum. On the other

    hand, Duffoo and Galvele [20] reported that SCC of Ag20Au alloy was faster in 1 M

    AgClO4 solution than in 1 M HClO4 solution. While the operating mechanism is stillnot clear, it is safe to assume that the DS value in this system is high.

    Using the same approach applied in the above discussed SCC systems, it was

    found again that with the calculated value ofDS, results similar to those in Fig. 11

    were found. The relation rT=UTS required to fit the measured CPR was very close tothat shown in Fig. 12.

    These are two facts that should be pointed out. The first one is that the effect of

    potential and SR on the CPR can be described by Eq. (2), as it was the case with Ag

    Pd in halides. The second one is that Alvarez et al. [4] found that the inverse slope p

    from a log CPREplot is the same for both IGSCC and TGSCC. These observations

    indicate that there is an association between SR and rT, as well as for a with the SCCmorphology and also an association between E and DS, if the surface mobility

    mechanism is considered, Eq. (4). As a further contribution to this point, Alvarez

    et al. [38] have recently found that the activation energy for IGSCC of AgAu alloy

    was equal to that for TGSCC for the same alloy, and that there was a coincidence

    between the activation energy predicted by the surface mobility SCC mechanism and

    the experimental results.

    As for the film induced cleavage (FIC) mechanism, Maier et al. [14] showed that

    there was no relation between the rate of dealloyed film formation and the CPR, as

    should be expected if the FIC mechanism was operative. The supporters of the FIC

    [39] attributed particular importance to the critical potential for selective dissolution(Ec) in alloys like AgAu. Maier et al. [14] found that changes in the nature of the

    selective dissolution process had no effect on the CPR. This observation was

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 609

  • 7/31/2019 Serega Lv 04

    20/22

    particularly evident for Ag15Au, Ag25Au and Ag40Au alloys 1 M HClO4 so-

    lution [14]. No breaks were found in the CPR vs. Ecurve when the potential crossed

    the critical potential value. These results showed that SCC of AgAu alloys in

    HClO4 solution was independent of the Ec value, and that there was no experimentalevidence correlating the SCC with that critical potential value. A similar observation

    regarding Ec was reported by Lichter et al. [40] for CuAu alloys.

    4.5. Mechanistic considerations

    The above observations indicate that the slope in Eq. (1) is higher for IGSCC than

    for TGSCC, Eq. (3). Besides the numerous considerations made above, the in-

    volvement of SCC mechanisms, other than the surface mobility mechanism, will be

    further considered.In the case of the film rupture, or slip step anodic dissolution mechanism, it

    should be expected that, as found in the present work, an increase in the strain rate

    will lead to an increase in the SCC rate. Nevertheless, as mentioned above, the effect

    should be the same for IGSCC and for TGSCC. In the literature of slip step anodic

    dissolution mechanism no explanation was found for the observation indicated by

    Eq. (3).

    As for the FIC SCC mechanism, as mentioned above, numerous contradictions

    were found by various authors between the predictions of this mechanism and the

    experimental results. On the other hand, the literature published on this mechanism

    does not give, so far, any clue for the observation indicated by Eq. (3). Besides, in thepresent work, Fig. 4, longer cracks were found for 5 s1 than for 18 s1. From the

    descriptions given for the FIC mechanism, a single event of crack propagation

    should be expected at both strain rates, and no difference in the crack length should

    be observed. Unfortunately no further quantitative comparisons can be made for

    this mechanism at the present state of development of the FIC model.

    5. Conclusions

    The following conclusions can be drawn from the present work:

    An increase of the strain rate generally accelerates the crack propagation process

    according to the relation (1), or equivalently

    log CPR / alog SR: 6

    The slope a is different for transgranular and intergranular cracking. The values are

    aTG 0:20.3 and aIG 0:40.7.The effect of potential can be incorporated as an additional term fE in Eq. (6),

    not having a noticeable effect on the SR-dependent term.

    No information was found in the anodic slip step dissolution mechanism, thatcould explain the present experimental results. The FIC mechanism requires further

    development to account for the differences found between IGSCC and TGSCC.

    610 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612

  • 7/31/2019 Serega Lv 04

    21/22

    The surface mobility SCC mechanism can account for the dependence of crack

    propagation rate on strain rate. With an increasing strain rate, the stress at the crack

    tip zone will increase due to the formation of dislocation pileups, which is equivalent

    to say that the driving force for the movement of surface vacancies increases.The surface mobility SCC mechanism can account for the separate effect of strain

    rate and potential on crack propagation rates, according to Eq. (2). The strain rate

    affects the stress at the tip of the crack, as indicated above, and the potential affects

    the surface diffusion coefficient.

    The absence of transgranular cracks, when very high strain rates are used, favours

    a continuous crack propagation model and sheds doubts on any mechanism based

    on very fast and discontinuous crack propagation events.

    Acknowledgements

    The present research has been supported by the Consejo Nacional de Investi-

    gaciones Cientficas y Teecnicas, Argentina, and by the FONCYT, Secretara de

    Ciencia y Tecnologa, Argentina.

    References

    [1] J.C. Scully, D.T. Powell, Corros. Sci. 10 (1970) 719.[2] R.H. Jones, R.E. Eicker, in: R.H Jones (Ed.), Stress Corrosion Cracking, ASM International, Ohio,

    1992, p. 4.

    [3] S.A. Serebrinsky, G.S. Duffoo, J.R. Galvele, Corros. Sci. 41 (1999) 191.

    [4] M.G. Alvarez, S.A. Fernaandez, J.R. Galvele, Corros. Sci. 42 (2000) 739.

    [5] J.R. Galvele, in: R.E. White, J.OM. Bockris, B.E. Conway (Eds.), Modern Aspects of Electrochem-

    istry, vol. 27, Plenum Press, New York, 1995, p. 233.

    [6] J.R. Galvele, in: R.H. Jones (Ed.), Chemistry and Electrochemistry of Stress Corrosion Cracking,

    TMS, Warrendale, PA, 2001, p. 27.

    [7] D. Gilroy, J.E.O. Mayne, J. Appl. Chem. 12 (1962) 382.

    [8] G.J. Janz, in: D.J.G. Ives, G.J. Janz (Eds.), Reference electrodesTheory and practice, Academic

    Press, New York, NY, 1961, p. 179.

    [9] J.R. Galvele, S.M. de De Micheli, I.L. Muller, S.B. de Wexler, I.L. Alanis, in: R.W. Staehle, B.F.Brown, J. Kruger, A. Agrawal (Eds.), Localized Corrosion, NACE, Houston, 1974, p. 580.

    [10] J.R. Galvele, in: S.M. Bruemmer, E.I. Meletis, R.H. Jones, W.W. Gerberich, F.P. Ford, R.W. Staehle

    (Eds.), Parkins Symposium on Fundamental Aspects of Stress Corrosion Cracking, The Minerals,

    Metals & Materials Society, Warrendale, PA, 1992, p. 85.

    [11] B.E. Wilde, J. Electrochem. Soc. 118 (1971) 1717.

    [12] G.S. Duffoo, I.A. Maier, J.R. Galvele, Corros. Sci. 28 (1988) 1003.

    [13] G.S. Duffoo, J.R. Galvele, Corros. Sci. 30 (1990) 249.

    [14] I.A. Maier, S.A. Fernaandez, J.R. Galvele, Corros. Sci. 37 (1995) 1.

    [15] M.L. Montoto, G.S. Duffoo, J.R. Galvele, Corros. Sci. 43 (2001) 755.

    [16] M.G. Alvarez, C. Manfredi, M. Giordano, J.R. Galvele, Corros. Sci. 24 (1984) 769.

    [17] C. Manfredi, I.A. Maier, J.R. Galvele, Corros. Sci. 27 (1987) 887.

    [18] J. Yu, R.N. Parkins, Corros. Sci. 27 (1987) 159.

    [19] J.R. Galvele, I.A. Maier, S.A. Fernaandez, Corrosion 52 (1996) 326.

    [20] G.S. Duffoo, J.R. Galvele, Metall. Trans. A 24 (1993) 425.

    S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612 611

  • 7/31/2019 Serega Lv 04

    22/22

    [21] R.G. Kelly, A.J. Frost, T. Shahrabi, R.C. Newman, Metall. Trans. A 22 (1991) 531.

    [22] R.B. Rebak, R.M. Carranza, J.R. Galvele, Corros. Sci. 28 (1988) 1089.

    [23] J.R. Galvele, Corros. Sci. 27 (1987) 1.

    [24] J.R. Galvele, Corros. Sci. 35 (1993) 419.

    [25] J.R. Galvele, J. Nucl. Mat. 229 (1996) 139.

    [26] N.A. Gjostein, in: J.J. Burke, N.L. Reed, V. Weis (Eds.), Surfaces and Interfaces-I, Syracuse

    University Press, 1967, p. 271.

    [27] G.E. Rhead, Surf. Sci. 15 (1969) 353.

    [28] G.E. Rhead, Surf. Sci. 22 (1970) 223.

    [29] J.R. Galvele, G.S. Duffoo, Corros. Sci. 39 (1997) 605.

    [30] M.O. Speidel, Metall. Trans. A 12 (1981) 779.

    [31] R.C. Newman, G.T. Burstein, J. Electrochem. Soc. 127 (1980) 2527.

    [32] C.M. Giordano, G.S. Duffoo, J.R. Galvele, Corros. Sci. 39 (1997) 1915.

    [33] M.O. Speidel, in: M.O. Speidel, A. Atrens (Eds.), Corrosion in Power Generating Equipment, Plenum

    Press, New York, 1984, p. 85.

    [34] J.P. Hirth, J. Lothe, in: Theory of Dislocations, Krieger Publishing Co., Malabar, FL, 1992, p. 764.[35] H.G. Van Bueren, in: Imperfections in Crystals, North-Holland Publishing Co, Amsterdam, 1960,

    p. 156.

    [36] R.C. Weast (Ed.), CRC Handbook on Chemistry and Physics, 54th ed., CRC Press, Cleveland, 1973,

    p. B-134.

    [37] M.E. Vela, G. Andreasen, R.C. Salvarezza, A. Hernaandez-Creus, A.J. Arvia, Phys. Rev. B 53 (1996)

    10217.

    [38] M.G. Alvarez, S.A. Fernaandez, J.R. Galvele, Corros. Sci. 44 (2002) 2831.

    [39] F. Friedersdorf, K. Sieradzki, Corrosion 52 (1996) 331.

    [40] B.D. Lichter, W.F. Flanagan, J.B. Lee, M. Zhu, in: R.P. Gangloff, M.B. Ives (Eds.), Environment-

    Induced Cracking of Metals (NACE-10), NACE, Houston, TX, 1990, p. 251.

    612 S.A. Serebrinsky, J.R. Galvele / Corrosion Science 46 (2004) 591612