1972 mechanism straininducednucleationmartensitictrans

Upload: nigar-akcay

Post on 07-Jul-2018

222 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    1/12

    Journal of the Less-Common Metals

    Elsevier Sequoia S.A.. Lausanne - Printed in The Netherlands

    1 7

    A MECHANISM FOR THE STRAIN-INDUCED NUCLEATION OF

    MARTENSITIC TRANSFORMATIONS*

    G. B. OLSON and MORRIS COHEN

    Department of Metallurgy and Materials Science, Massachusetts Institute qf Technology, Cumbridge,

    Massachusetts U.S.A)

    (Received January 18th, 1972)

    SUMMARY

    The previous work of Professor W. G. Burgers and Dr. A. J. Bogers is used to

    develop a mechanism of strain-induced martensitic nucleation, involving two inter-

    secting shear systems. We distinguish between strain-induced nucleation and stress-

    assisted nucleation, the latter involving the same sites and embryos as does the regular

    spontaneous transformation. The strain-induced nucleation, on the other hand,

    depends on the creation of new sites and embryos by plastic deformation; this

    phenomenon may also contribute in a major way to autocatalytic nucleation during

    the course of martensitic transformation.

    For the case of strain-induced nucleation, it is possible to focus on specific

    intersecting-shear systems when the austenitic stacking-fault energy is low and

    E (h.c.p.) martensite can form as a part of the shear displacements. It then becomes

    feasible to extend the intersecting-shear mechanism from this special case to alloys of

    higher stacking-fault energy, where E is no longer stable relative to the austenite.

    It should be noted, however, that these events are very early stages in the for-

    mation of martensitic plates and relate primarily to the genesis of embryos; the actual

    growth start-ups which determine the operational (measured) nucleation rates may be

    controlled by subsequent processes.

    INTRODUCTION

    We are privileged to take part in this international tribute to the distinguished

    Professor W. G. Burgers, and take this opportunity to present some ideas about a

    subject which has fascinated him for many years, namely the nucleation of martensitic

    transformations. Professor Burgers’ interest in this phenomenon is typified by his

    classic paper withA. J. Bogers on “Partial dislocations on the {1 o} planes in the b.c.c.

    lattice and the transition of the f.c.c. into the b.c.c. lattice”‘. This remarkable piece of

    Dedicated to Professor Dr. W. G. Burgers in celebration of his 75th birthday.

    This paper is based on research at the Massachusetts Institute of Technology sponsored by the

    Office of Naval Research under Contract No. N00014-67-A-0204-0027 and by the National Science

    Foundation under Grant No. GK-26631.

    J. Less-Common Metals, 28 1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    2/12

    108

    G. B. OLSON, M. COHEN

    work has stimulated us to build on it, and we now wish to extend the Bogers-Burgers

    model to strain-induced nucleation.

    Although most investigations on the kinetics of martensitic reactions have

    dealt with the spontaneous transformation unassisted by externally applied stresses,

    the fact remains that virtually the total portion of such transformations (at least in

    iron-base alloys) involves autocatalytic nucleationzP4. Because of the macroscopic

    displacements which accompany martensitic transformations, the surrounding

    parent phase is plastically strained, and it seems reasonable to assume that strain-

    induced nucleation contributes to the autocatalysis.

    Still another important aspect of strain-induced nucleation arises from the

    circumstance that, when this type of nucleation is systematically studied through

    intentionally-imposed plastic deformation, the nucleation sites and corresponding

    embryos can be identified’ - ‘,

    at least in special cases, and channeled into specific

    models. It should be emphasized, however, that we are concerned here with nucleation

    events which do not necessarily determine the observed nucleation rates3. The latter

    quantities are operational in the sense that the measurements reflect only the frequency

    of growth start-ups which produce martensitic units of a visible size’. The rate-con-

    trolling steps for triggering-off such growth are treated elsewhereg*“. In the present

    paper, on the other hand, we consider only the genesis of nucleation sites and embryos

    by plastic deformation, this being a precursor to whatever may control the later

    growth start-up process.

    STRESS-ASSISTED

    VS.

    STRAIN-INDUCED NUCLEATION

    The complex interrelationships among applied stress, plastic strain, and mar-

    tensitic transformations have been studied extensively by Bolling and Richman” - l4

    in Fe-Nix and some Fe-Ni-Cr-C alloys. They were able to define a temperature,

    M,” (lying above M,), below which “yielding” under applied stress is initiated by the

    onset of martensite formation, and above which yielding under stress is initiated by

    regular slip processes in the parent phase. Accordingly, the temperature dependence

    of the yield stress is negative (normal) above M,“, but is positive below M,“. In the latter

    regime, the stress to start the martensitic transformation (and hence yielding) ap-

    proaches zero as the MS temperature is approached.

    These relationships are illustrated schematically in Fig. 1. At temperatures

    under Mp, the yielding which accompanies the martensitic transformation occurs

    below the regular yield stress (oY) of the austenite (as extrapolated from higher to

    lower temperatures). Here, the observed yield stress follows a temperature dependence

    consistent with the dependence of MS on applied elastic stress16. We regard the

    nucleation under these conditions as being

    stress-assisted.

    Above M,“, however, the applied stress must reach or exceed @Yn order to

    initiate the martensitic transformation. As indicated by Fig. 1, there is evidence

    (serrated stress-strain curves)14,15

    that the stress at which the transformation is

    initiated at temperatures just above

    Mf

    follows the regular yield strength o,,. At still

    higher temperatures the initiating stress rises above (TV ntil a temperature limit is

    reached at Md*. It is important to note that, in the temperature range above M,“, the

    * The symbol M, has its usual meaning here, being the temperature above which martensite is not

    produced by plastic deformation.

    J. Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    3/12

    NUCLEATION OF MARTENSITIC TRANSFORMATIONS

    109

    _-

    Md

    Temperature --+

    Fig. 1. Schematic representation of interrelationships between stress-assisted (below Mp) and strain-

    induced(aboveM,“)nucleationofa’martensiteinFe-Ni-Calloys.After resultsofBolIingandRichman”- 14,

    stress at which transformation is initiated lies significantly below the relatively-steep

    stress-assisted transformation line extrapolated up from below

    M “.

    Therefore, we

    postulate that the nucleation occurring above Mf is different in kind from the stress-

    assisted nucleation manifested below M “. This distinction is easily rationalized on the

    supposition that plastic deformation at 0,. and above introduces new, highly-potent

    sites which allow nucleation at appreciably lower stresses than in the case of stress-

    assisted nucleation. Hence, we consider the nucleation above M ” to be (plastic) strain-

    induced.

    We can now summarize this overall situation. At the usual MS the chemical

    driving force is sufficiently large for the pre-existing nucleation sites or embryos in the

    parent austenite to become operative without the application of stress. At temperatures

    between MS and M “ such nucleation can still occur, but only with the aid of applied

    stress. This is the stress-assisted transformation regime. Here, the required initiating

    stress is in the elastic range (below gY), but increases with increasing temperature

    because of the concommitant decrease in chemical driving force. At M “ the initiating

    stress for nucleation, based on the original sites and embryos, reaches o,, and plastic

    straining enters the picture. Evidently, the resulting strain-induced nucleation can be

    activated at lower stresses than the stress-assisted nucleation, and so the initiating

    stress tends to follow gY ust above M “. At still higher temperatures, however, the

    further reduction in chemical driving force necessitates additional plastic straining in

    order to produce detectable amounts of transformation; this requires the initiating

    stress to rise above aY.

    The foregoing description applies particularly to paramagnetic austenites.

    Bolling and Richman13 have shown that, below M “ the stress-assisted transformation

    characteristics are rather similar for both paramagnetic and ferromagnetic austenites.

    J.

    Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    4/12

    110

    G. B. OLSON, M. COHEN

    On the other hand, above M,” he strain-induced transformation tends to be suppressed

    if the austenite is ferromagnetic. Although this curious behavior requires further

    elucidation, it appears that the ferromagnetic state has a strong inhibiting effect on

    strain-induced nucleation in contrast to its negligible effect on stress-assisted nucle-

    ation; this can be taken as additional evidence for the separate nature of the two types

    of nucleation.

    Neutron irradiation experiments”

    on austenitic stainless steel demonstrate

    convincingly that the strain-induced transformation is primarily dependent on plastic

    strain rather than on the acting stress. The amount of martensitic transformation as a

    function of plastic strain was found to be the same in both the annealed and irradiated

    materials, even though the latter was much stronger (due to radiation damage) than

    the former.

    SPECIAL SITES FOR STRAIN-INDUCED NUCLEATION

    The formation of E (h.c.p.) martensite in y’(f.c.c.) austenite of low stacking-fault

    energy is well-established5. There are also some instances in which

    E

    has been observed

    to provide favorable nucleation sites for the formation of a’ (b.c.c. or b.c.t.) martensite.

    Figure 2, taken from a study by Venables’, shows ~1’ormed at the intersection of two

    E plates in an austenitic stainless steel (Type 304) deformed at 78 K. This is clearly a

    case where a nucleating site has been created by plastic straining. The u.’ hus generated

    may be construed as a strain-induced embryo; it has not triggered-off into a countable

    martensitic plate in the sense of operational nucleation, as discussed earlier.

    Lagneborg6 has reported that the intersections of active slip systems with E

    plates may also produce favorable sites for a’ nucleation. Manganon and Thomas’

    found that a’ can be nucleated by the intersection of two E plates or by the intersection

    of an E plate with a twin or grain boundary in the austenite.

    Notwithstanding these instances in which E intersections obviously play a

    definite role in the strain-induced nucleation of a’ martensite, it can be argued that the

    Fig. 2. Nucleation of

    a’

    (b.c.c.) martensite at the intersection of two E (h.c.p.) plates in austenitic (f.c.c.)

    stainless steel. Venables’.

    J. Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    5/12

    NUCLEATION OF MARTENSITIC TRANSFORMATIONS

    111

    E phase is not necessary for this purpose, even in those alloys where E can form.

    Lagneborg6 has noted, later confirmed by Goodchild et al.‘*, that a’ can be generated

    in austenitic grains which are so oriented relative to the tensile-straining axis that E

    formation is suppressed. Similarly, Breedis and Kaufman” have also concluded that

    CI’ an nucleate independently of the E, particularly in those cases where the E is not

    even thermodynamically stable with respect to either the y or a’

    There seems to be a real correlation in Fe-Ni-Cr alloys between decreasing

    stacking-fault energy of the f.c.c. austenite and increasing ease of CI’ ormation with

    plastic strain

    ‘OJ~ In such experiments, E is found in the alloys of lowest stacking-

    fault energy, but there is no accompanying discontinuity in the observed trend of ci

    formation with stacking-fault energy. Thus, although strain-induced nucleation is

    favored by low stacking-fault energy, the formation of E does not appear to be a

    necessary condition.

    Despite these limitations, the E intersection mode of strain-induced IX’ ucle-

    ation is worthy of detailed attention because the site and associated crystallography

    have been well determined. It is also conceivable that, even in instances where E is not

    detected, E or faulted E may actually participate in some transient way, and so the E

    case may offer tangible clues concerning other possible intersection mechanisms.

    Moreover, there is a dynamic aspect to such intersections which should not be over-

    looked. When austenitic stainless steel is plastically strained (about 15 ) at 60°C

    some E is formed, and then considerably more appears on cooling under load to

    - 30°C but the resulting conversion to CI’s much smaller than when the austenite is

    plastically strained to 15 directly at - 30°C6. Clearly, strain-induced nucleation to

    c(’ akes place more readily during plastic deformation than under applied stress after

    the deformation

    INTERSECTING-SHEAR MECHANISMS FOR AN F.C.C.-B.C.C. TRANSFORMATION

    The observations of CI’ ucleation at the intersections of E plates with other E

    plates, with active slip systems, or with twin boundaries, all have a common feature.

    The region of intersection defines a lath- or rod-shaped volume which has been doubly

    sheared, and further, the elements of both shears are of the type (111\(112),,,. We

    shall now examine this double-shear feature as a mechanism for transforming austenite

    to martensite. Bogers and Burgers’ have already built an excellent foundation for this

    problem with their ingenious hard-sphere model, reproduced here in Figs. 3 and 4.

    Figure 3(a) and (d)* illustrates an f.c.c. packing of spheres lying in { 11 }rcc

    planes. Before dealing with the transition to a b.c.c. structure, it is instructive to

    visualize the configuration of spheres after a regular f.c.c. twinning shear, shown in

    Fig. 3(c) and (f), h

    erein successive layers parallel to the close-packed plane PVQ

    have shifted by a,,,/6( 112) in the direction QT perpendicular to PV. The result of

    this twinning shear is that the shear plane PVQ and its conjugate plane PVS remain

    as close-packed

    [

    111 ]rcc planes, whereas the other two close-packed planes QSV and

    QSP become { 1001_rcc

    lanes. Thus, the 60” angle denoted in Fig. 3(a) is enlarged to

    90” in Fig. 3(c).

    *

    Figure 3(dHf) are sections through the hard-sphere model in Fig. 3(aHc), corresponding to

    1 lOif,, planes. Such planes are normal to the shear plane PVQ and contain both the shear direction QT

    and the dilatational direction.

    J. Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    6/12

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    7/12

    NUCLEATION OF MARTENSITIC TRANSFORMATIONS

    113

    If we now trace the paths taken by the hard spheres during the transition from

    Fig. 3 (a) and (d) to 3 (c) and (f), each sphere must ride up to a saddle-point position

    between the initial and final states. Hence a dilatational component normal to the

    shear plane is involved; this expansion amounts to 5.4 at

    one-third of the twinning

    sheur, a,,,/18(112), as shown in Fig. 3(b) and (e). Figure 3(b) further indicates that

    the shear plane PVQ and its conjugate plane PVS are still dimensionally unchanged,

    but the planes QSV and QSP are distorted in such a way that the 60” angle in Fig. 3 (a)

    is enlarged to 70” 32’ in Fig. 3 (b), thus attaining the geometry of a 1 10)t,CC lane.

    Although the stacking sequence of these planes is not correct for a true b.c.c. structure,

    the latter can be obtained if successive planes (parallel to QSV or QSP) are sheared

    according to Fig. 4’. Referred to the b.c.c. structure, this shear corresponds to a

    displacement of a,,,/8( 110) on each plane, whereas referred to the f.c.c. structure,

    this corresponds to a displacement of arJl2( 112) or just one-half of a twinning shear.

    Again, because of the hard-sphere packing there is a dilatation normal to the shear

    plane. Since each of the above shears entails a dilatational component, they are more

    accurately described as invariant-plane strains.

    According to the Bogers-Burgers model, then, a b.c.c. structure can be

    generated from an f.c.c. structure by two invariant-plane strains (either successively or

    simultaneously), which can be thought of as one-third and one-half f.c.c. twinning

    shears. For convenience, we shall call these the T/3 and T/2 shears.

    Bogers and Burgers’ point out that their first shear (T/3) must involve dis-

    placements on each (ill],,, p

    lane of ar,,/18( 112), the latter being one-third the

    Burgers vector of a Shockley partial dislocation, ur,,/6( 112). They also suggest the

    possibility that such partial displacements might occur by the “spreading” of a

    Shockley partial dislocation over a number of successive {ill),,, planes. This idea,

    although seemingly strange at first thought, is not difficult to imagine. In the con-

    ventional glide motion of an u,,,/6( 112) partial dislocation, the atoms pass through

    appropriate positions for the T/3 shear proposed in the Bogers-Burgers model. Under

    conditions of sufficient chemical driving force, the atoms may tend to “stick in the

    b.c.c. positions” corresponding to the ar,,/lS( 112) displacement (Fig. 3(b) and (e))

    rather than continue through to the full Burgers-vector displacement of a,,,/6( 112)

    (Fig. 3(c) and(f)). H

    owever, the full Burgers vector can be conserved if the atoms in an

    adjacent plane are concurrently “dragged along” to the proper “b.c.c. positions“. In

    this way, the T/3 shear of the Bogers-Burgers model can be achieved by an array of

    u,,,/6( 112) partial dislocations, averaging one on every third (111 ).rccplane.

    A problem arises in connection with the second (T/2) shear. Bogers and Burgers

    suggest that this shear may involve b.c.c. partial dislocations with ut,,,/8( 110)

    vectors since these have been previously proposed to explain b.c.c. twinning”.

    However, the existence of such partial dislocations is not generally accepted. We can

    circumvent this difficulty by considering the second shear (T/2) relative to the f.c.c.

    instead of the b.c.c. structure. This comes about naturally in the shear-intersection

    mechanism to be described below.

    A schematic intersection of partial-dislocation arrays for the strain-induced

    formation of a b.c.c. embryo in f.c.c. austenite is illustrated in Fig. 5. The plane of this

    diagram is parallel to (1 10)rcc,and perpendicular to VQ in Fig. 3. Then, planes (ill) rcc

    and (lil),,, correspond to PVQ and QSV, respectively. The T/2 shear (the second of

    the Bogers-Burgers shears) can be accomplished (with spreading) by an array of

    J. Less-Common Metals,

    28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    8/12

      4

    G. B. OLSON, M. COHEN

    Fig. 5. Schematic illustration of intersecting shears due to two arrays of n&6( 112) partial dislocations in

    austenite. One array (T/3) has partial dislocations on every third {11 }rcc plane and averages one-third of a

    twinning shear, while the other array (T/2) has partial dislocations on every second { ll}r,, plane and

    averages one-half of a twinning shear. The resulting doubly-faulted intersection has an exact b.c.c. structure.

    a rCC/6 21i] partial dislocations on every second (lil),,, plane, just as the T/3 shear

    (the first of the Bogers-Burgers shears) can be accomplished (with spreading) by an

    array of

    afcc/6

    [211] partial dislocations on every third (ill) plane. Each of these

    Burgers vectors has a component out of the plane of the diagram. The partial dis-

    locations are shown in their positions (I) after the intersection process, and the stacking

    faults left behind outside the intersected region are denoted by heavy lines. (Of course,

    these faults will be bounded at their other ends by other partial dislocations.)

    The Bogers-Burgers dislocation-spreading hypothesis is most likely to be

    valid within the intersected volume and during the intersection event. Under these

    conditions, the atoms can attain their true b.c.c. positions which, as we have seen,

    happen to lie intermediate between the full displacements involved in the regular

    gliding of a,,,/6( 112) partial dislocations. When referred to the f.c.c. structure, the

    J.

    Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    9/12

    NUCLEATION OF MARTENSITIC TRANSFORMATIONS

    115

    intersected volume is “doubly faulted” (lighter solid lines in Fig. 5) but this region

    now has a true b.c.c. structure.

    It is more realistic to picture the intersection event, not as a collision of two

    fronts of moving partial dislocations, but as the intersection of one moving array with

    a stationary packet of stacking faults left behind by the prior passage of another array

    of partial dislocations. This situation is readily visualized from the fact that the

    set of stacking faults left behind by the T/2 dislocation array produces exactly an

    E (h.c.p.) plate; in other words, the h.c.p. structure can be modeled as a shear of

    ar,,/6( 112) on every second 111 I.rccplane.

    We now examine the intersection of a T/3 array with the coherent interface of

    an

    E

    plate. One can well imagine that a barrier of this type will cause the T/3 set of

    partial dislocations to pile up at the interface. A likely plane in the E for the passage of

    these dislocations would be (1Oil). Every second basal plane in the h.c.p. structure is

    in the proper position for a T/2

    shear

    (except for the dilatation), and all that is necessary

    to generate the exact configuration of a T/2 shear is a “shuffle” on every second plane

    by a displacement of a ,,,/12 [21i]. Moreover, if the atoms behave as hard spheres,

    these shuflIes will produce the afore-mentioned dilatation. It can be shown that the

    conversion of the E structure to an exact T/2 shear configuration transforms the

    (loil),,, plane into a uniformly distorted (ill),,, plane. Returning now to the inter-

    secting T/3 array, we find that the ar,,/6 [211] partial dislocations in the Tj3 packet

    can glide on the uniformly distorted (il l)rcc planes. If these T/3 partials can spread on

    entering the intersected region, as discussed earlier, their resulting passage through

    the intersection will generate b.c.c. martensite. Put in another way, the formation of

    the b.c.c. structure allows the blocked dislocations to pass through the intersected

    volume.

    In accordance with theoretical estimates for the critical size of a nucleus, it is

    improbable that a small number of intersecting dislocations could enter the E plate,

    since the volume of martensite would then be subcritical. Consequently, we should

    expect the blocked dislocations to continue to pile up at the E interface during the

    plastic straining until the T/3 shear packet is thick enough to create a supercritical

    volume in the intersected region under the conditions at hand. A more general postu-

    late regarding strain-induced nucleation might then be: if the motion of a large enough

    array of appropriate partial dislocations in the parent phase is impeded by intersecting

    a set of stacking faults (not necessarily E) in which enough of the atomic positions

    coincide with the “other shear” of the Bogers-Burgers model, the required shuffling of

    the atomic planes within the set of stacking faults will take place concurrently with the

    spreading of the entering partial dislocations, and then the passage of these dislocations

    through the intersected region can occur by the formation of b.c.c. martensite.

    From an experimental standpoint, the most commonly reported strain-induced

    nucleation site is the intersection of two E plates. It is already evident that an E plate

    can accomplish the T/2 shear with the aid of shuffles. However, in order for the other e

    plate to provide the equivalent of a T/3 shear, it would have to be highly faulted.

    Strain-induced E is, indeed, often observed to be highly faulted, and it is conceivable

    that local regions exist in such

    E

    where the average shear matches that of a T/3 array.An

    alternative for the cast of relatively perfect E plates is that one-third of the dislocations

    attempting to pass into the intersection from the second plate may be left at the inter-

    face. producing a semicoherent interface. Hence, operation of the proposed inter-

    .f. Less-Common Metals, 28 (1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    10/12

    116

    G. 8. OLSON, M. COHEN

    setting-shear mechanism is feasible even when only I: plates are involved

    The proposed intersecting-shear mechanism for E plates offers a good explana-

    tion for Lagneborg’s finding6 that E formed by plastic straining at low temperatures

    leads to much more effective b.c.c. nucleation sites than does the same plastic straining

    at higher temperatures (also generating E) and then cooling under load to the low

    temperature. In the intersection processes described, the b.c.c. structure can only

    nucleate ~~~i~g he intersection event and only if there is a substantial chemical driving

    force for the f.c.c.-b.c.c. transformation at play. At the higher temperature, where this

    driving force is small or nonexistent, the intersection mechanism may actually

    regenerate the f.c.c. phase, as proposed by Sleeswijk23.

    A further deviation from the “ideal” T/3-T/2 shear-intersection case is that of

    an E plate penetrating an f.c.c. twin boundary. Consider the leading a rcJ6 [2li]

    partial dislocations of an advancing E plate (the T/2 array in Fig. 5) colliding with a

    (il l)foc twin boundary (parallel to the plane of the T/3 array in Fig. 5). The twinned

    lattice across the (71 l)fcc boundary can be equivalently described by a twinning shear

    of a,,,/6(112) on every (ill),,, plane in either of the [211],,,, [El],,, or [TlZJ,,,

    directions. To cross into the twinned region, the partial dislocations should glide on

    the planes in the twin that correspond to the glide planes in the untwinned region.

    Because of the three equivalent twinning shears, there are three correspondences, but

    they do not permit the dislocations to cut straight across the twin. Nevertheless, this

    could become feasible for one of the correspondences if the appropriate shuttles

    of (illhcc planes in the twin could occur to produce the T/3 shear configuration

    during the deformation. However, inasmuch as there are relatively few positions in

    the twin which are already T/3 shear positions, it may be anticipated that the inter-

    section of E plates with f.c.c. twin boundaries will be a less potent mechanism for b.c.c.

    nucleation than the case of two intersecting E plates and, moreover, a higher chemical

    driving force will be required. This is in line with the fact that b.c.c. nucleation at E-

    plate/twin-interfa~ intersections is less common than at E/E ntersections6,

    In alloys of slightly higher stacking-fault energy where the E phase is not stable

    relative to the austenite, packets of stacking faults in the f.c.c. structure (with no

    evidence of an h.c.p. phase) are sometimes associated with nucleation of b.c.c. marten-

    sitez4. Here, it may be presumed that the nucleation takes place when the intersection

    of two such packets locally approximates the conditions represented by Fig. 5; for

    example, one packet with a stacking-fault probability of l/3 intersecting a second

    packet with a stacking-fault probability of l/2. Inasmuch as the length of these packets

    (distance between pairs of partial dislocations) will decrease with increasing stacking-

    fault energy, and so might the thickness and number of these packets, the probability

    of packet intersections will also decrease with increasing stacking-fault energy. This

    relationship could account for the adverse effect of stalking-fault energy on the

    propensity to strain-induced nucleation.

    Plausibility arguments have now been presented here for mechanisms by

    which the two shears of the Bogers-Burgers model can bring about the observed

    nucleation sites. l-lowever, an important aspect not yet considered is the problem of

    coherency strains. In the idealized shear intersections which can be analyzed in

    detail, the b.c.c. phase thus nucleated is fully coherent with its surroundings. The

    coherency strains have to be accommodated by a dilatation normal to the stacking-

    fault arrays outside the intersected region. In the case of E intersections, the accom-

    J. Less-Common Met 28 1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    11/12

    NUCLEATION OF MARTENSITIC TRANSFORMATIONS

    117

    modation would distort the h.c.p. structure to a non-ideal c/a ratio. The coherency

    strains are further increased because the real b.c.c. structure in iron has a smaller

    atomic diameter than provided by the hard-sphere model of f.c.c. iron. It may be

    expected, therefore, that at some point in its formation, the b.c.c. region will undergo

    plastic deformation to relieve the coherency strains and so create a semicoherent

    interface. This deformation process is also likely to cause a rigid-body rotation of the

    intersected region, all of which will influence the orientation relationships and inter-

    face planes of the embryo thus formed.

    CONCLUSIONS

    Evidence for the separate nature of stress-assisted and strain-induced nucle-

    ation of martensitic transformations has been presented. The indications are that

    stress-assisted nucleation depends on the same nucleation sites or embryos which are

    responsible for the usual spontaneous transformation, whereas the strain-induced

    nucleation involves the creation of new sites or embryos by plastic deformation. It is

    likely that strain-induced nucleation plays an important role in the autocatalytic

    nucleation observed in regular martensitic transformations.

    Strain-induced nucleation occurs in austenites with a wide range of stacking-

    fault energies at lower stresses than does stress-assisted nucleation at comparable

    temperatures. When the stacking-fault energy is very low and martensite can form,

    the nucleation sites and ci embryos are generated by two intersecting shear systems

    in the austenite with the elements {11 l} (112). The proposed mechanism is consistent

    with the Bogers-Burgers two-shear model, but is an extension thereof to embrace

    plausible dislocation motions.

    The strain-induced a’ embryos generated in this way are initially coherent.

    The formation of semicoherent interfaces and the subsequent growth start-ups which

    enter into the operational (measured) nucleation rates constitute still later stages in

    the formation of the usual martensitic plates.

    ACKNOWLEDGMENTS

    The authors deeply appreciate the stimulating interest and critical guidance

    offered by Professor J. W. Christian during the development of ideas presented here.

    They are also indebted to the Office of Naval Research and the National

    Science Foundation for the support of resea’rch which provided the general back-

    ground for this paper.

    REFERENCES

    1 A. J. Bogers and W. G. Burgers, Acta Met., 12 (1964) 255.

    2 V. Raghavan and A. R. Entwiste, Iron Steel Inst. Spec. Rept. No. 93, 1965, p. 20.

    3 S. R. Pati and M. Cohen, Acta

    Met.,

    17 (1969) 189.

    4 S. R. Pati and M. Cohen, Acta Met., 19 (1971) 1327.

    5 J. A. Venables, Phil. Msg., 7 (1962) 35.

    6 R. Lagneborg, Acta Met., 12 1964) 823.

    7 D. L. Manganon and G. Thomas, Met. Trans., 1 (1970) 1577.

    8 M. Cohen. to be published in Met. Trans. as a part of the E. C. Bain Symposium (October 1971), Detroit,

    Michigan.

    J. Less-Common Metals, 28 1972)

  • 8/19/2019 1972 Mechanism StrainInducedNucleationMartensiticTrans

    12/12

    118

    G. B. OLSON, M. COHEN

    9 V. Raghavan and M. Cohen, to be published in Acta Met. (March 1972).

    10 V. Ragahavan and M. Cohen, submitted to

    Acta kfet

    11 G. F. Boiling and R. H. Richman, Scripta Met., 4 (1970) 539.

    12 G. F. Bolling and R. H. Richman, Phil. Mag., 19 (1969) 247.

    13 G. F. Bolling and R. H. Richman,

    Acta Met.,

    18 (1970) 673.

    14 R. H. Richman and G. F. Bolling, Met. Trans., 2 (1971) 2451.

    15 R. H. Richman, private communication.

    16 J. R. Pate1 and M. Cohen,

    Acta Met.,

    1 (1953) 531.

    17 A. L. Bement, Effects of Residual Elements on Properties of Austenitic Stainless Steels, Am. Sot. Testing

    Mater. STP418, 1967, p. 46.

    18 D. Goodchild, W. T. Roberts and D. V. Wilson,

    Acta Met., 18 1970) 1137.

    19

    J. F. Breedis and L. Kaufman,

    Met. Trans.,

    2 (1971) 2359.

    20 J. S. Dunning, Univ. Calif. (Berkeley) Rept. UCRL-19052, 1969.

    21 F. Lecroisey and A. Pineau, to be published; private communication from W. G. Burgers.

    22 J. B. Cohen, R. Hinton, K. Lay and S. Sass, Acta Met., 10 (1962) 894.

    23 A. W. Sleeswijk, Phil. Mug., 7 (1962) 1597.

    24 P. M. Kelly,

    Acta Met.,

    13 (1965) 635.

    J. Less-Common Metals, 28 1972)