Vous êtes sur la page 1sur 29

Deformation Behaviour of Steels Having Different Stacking Fault Energies

MM694: Seminar Report Submitted in partial fulfilment of the requirements of the degree of Master of Technology by Karthik V V Subramaniyan Iyer (Roll No. 123114004)

Supervisor Prof. Prita Pant

Department of Metallurgical Engineering and Materials Science INDIAN INSTITUTE OF TECHNOLOGY BOMBAY (October 2012)

ABSTRACT
There is a ubiquitous need to increase tensile strength while simultaneously reducing yield strength and maintain high ductility for better formability of steels. To address this challenge it is essential to understand the various strengthening & deformation mechanisms at work and the factors that influence them. Stacking fault energy (SFE) is an important parameter that significantly affects the mechanical properties in steels. In the present work a comprehensive review of literature published on the relation between deformation mechanism & stacking fault energy is presented. A general description of stacking fault energy, the various approaches to determine it is also discussed. The influence of alloy chemical composition & temperature on SFE from numerous references is tabulated. It was found that there is consensus in the published literature about influence of SFE on deformation mechanism/ products, however there is a scatter in the value at which these mechanisms start or are active. The results on influence of alloying elements on SFE are also widely scattered and depend on the approached (experimental or theoretical) followed with each approach having its own set of limitations. It is crucial to further develop experimental methods and modelling tools to enhance the knowledge relating to SFE & deformation.

CONTENTS
Abstract ...................................................................................................................................... i Contents ..................................................................................................................................... ii List of Figures ..........................................................................................................................iii List of Tables ............................................................................................................................iii Chapter 1 Introduction ............................................................................................................ 1 1.1 Aim of Study .................................................................................................................... 1 Chapter 2 Literature Survey ................................................................................................... 3 2.1 Stacking Fault Energy (SFE) ............................................................................................ 3 2.1.1 Effect of Alloying Elements on Stacking Fault Energy 7 2.1.2 Effect of Temperature on Stacking Fault Energy 8 2.2 Mechanisms of Plastic Deformation in Steels ................................................................ 12 2.2.1 -martensite 12 2.2.2 `-martensite 13 2.2.3 Deformation Twins 13 2.3 Correlation between Deformation Mechanism & Stacking-Fault Energy ..................... 17 References................................................................................................................................ 22 Acknowledgement ................................................................................................................... 25

ii

LIST OF FIGURES
Figure 2.1 Schematic representation of slip in a (1 1 1) plane of a fcc crystal ......................... 5 Figure 2.2 Dark-field TEM micrographs of the deformation microstructure of a Fe22Mn 0.6C steel at 77 K (-martensite platelets) and at room temperature (mechanical twins). ...................................................................................................................... 9 Figure 2.3 Schematic representation of the formation of twinning in FCC ............................ 15

LIST OF TABLES
Table 2.1 Effect of alloying elements on stacking fault energy ............................................... 10 Table 2.2 Twelve possible twinning systems in FCC materials ............................................... 15 Table 2.3 Stacking fault energy & deformation behaviour ...................................................... 19

iii

Chapter 1

Introduction

1.1 Aim of Study


Steels as engineering materials are of utmost importance to society because of their wide range of properties which are used in variety of applications. Theoretically, steels are interstitial alloys of carbon in iron where carbon is 0.008 to 2% by weight. The need to further enhance their properties has led to the development of variety of steels spanning various generations. Each new generation of steels is a result of advancement in understanding the physical metallurgy leading to a better insight into various parameter property relationships.

There is a ubiquitous need to increase tensile strength while simultaneously reducing yield strength and maintain high ductility for better formability especially in the automobile sector. To address this challenge it is essential to understand the various strengthening & deformation mechanisms at work and the factors that influence them. Stacking fault energy (SFE) is an important parameter that significantly affects the mechanical properties in steels. 1

The aim of the present work was to carry out a survey of existing literature and understand the deformation behaviour of steels with different SFE. Steels having BCC structure were not studied for the reasons discussed as follows. The stacking sequence in BCC metals is ABABAB. Stacking faults associated with BCC structure may lead to two A or B layers coming in contact with one another. This leads to very high theoretical stacking fault energy (~200 mj/m2) and hence stacking faults in BCC structure have not been observed directly [1,2]. Even though the SFE of BCC iron has been simulated [3], to the best of my knowledge no experimental information exists about its SFE & its effect on deformation.

Chapter 2

Literature Survey

2.1 Stacking Fault Energy (SFE)


Stacking faults are extremely significant feature in a crystal structure. They arise because there is little to choose electrostatically between the stacking sequence of the close packed planes in the FCC metals ABCABC and that in the HCP metals ABABAB. It is possible that atoms in a part of one of the close-packed layers may fall into the wrong position relative to the atoms of the layers above and below, so that a mistake in the stacking sequence occurs (e.g.ABCBCABC ). Even though such an arrangement will be more stable; work needs to be performed to produce it.

The shortest lattice vector in the FCC structure joins a cube corner atom to a neighbouring face centre atom and defines the observed slip direction; one such slip vector a/2[1 0 ] is shown as b1 in Figure 2.1a, which is for glide in the (1 1 1) plane. However, an atom which sits in a B position on top of the A plane would move most easily initially towards a C 3

position and, consequently, to produce a macroscopical slip movement along [1 0 ] the atoms might be expected to take a zigzag path of the type BCB following the vectors b 2 =a/6[2 ] and b3 =a/6[1 1 ] alternately.

It will be evident, of course, that during the initial part of the slip process when the atoms change from B positions to C positions, a stacking fault in the (1 1 1) layers is produced and the stacking sequence changes from ABCABC . . . to ABCACABC . . .. During the second part of the slip process the correct stacking sequence is restored.

To describe the atoms movement during slip, discussed above, Heidenreich and Shockley have pointed out that the unit dislocation must dissociate into two half-dislocations, which for the case of glide in the (1 1 1) plane would be according to the reaction:

Such a dissociation process is algebraically correct as well energetically favourable since it reduces the energy of the system (the sum of value of the partial dislocations is less the single dislocation). These half dislocations, or Shockley partial dislocations, repel each other by a force and separate, as shown in Figure 2.1b. A sheet of stacking faults is then formed in the slip plane between the partials, and it is the creation of this faulted region, which has a higher energy than the normal lattice, that prevents the partials from separating too far. [1,4]

Both partials are glissile on the (111) plane. If the stacking sequence of the (111) planes changes from the regular stacking to the faulted stacking by removing part of a {111} plane, i.e., ABCABCABCABCACABCA, the fault is called an intrinsic stacking fault. In the case of an extrinsic stacking fault, an additional part of a (111) plane is inserted in the crystal and the stacking sequence consequently changes for instance as ABCABCABCABCACBCAB, hence containing an excess plane with C stacking. In some cases it can be more convenient to think of an extrinsic stacking fault as the overlapping of two intrinsic stacking faults on successive (111) planes.

Figure 2.1 Schematic representation of slip in a (1 1 1) plane of a fcc crystal [1]

As mentioned earlier there is net repulsion force associated with the formation of the Shockley partials. With increasing separation r of the partials there is also an increase in energy ESF of the stacking fault associated with it given by the relation,

where L is the length of the dislocation and SFE the stacking fault energy (per unit area) & r is the shift of the dislocation line perpendicular to itself, so that a force opposing the repulsion of the partials is

The equilibrium separation of the dissociated dislocation is hence controlled by the minimum of the total stored energy in the stacking fault, corresponding to the force balance at the partial dislocations: 5

Based on this the separation width d of the stacking fault is given by the relation [5]

where G is the shear modulus and bp the absolute value of the Burgers vector of the Shockley partials. From the equation it is apparent that the width d of the stacking fault is essentially controlled by the stacking fault energy SFE. d in turn controls the ease of cross slip of screw dislocations and is thus largely responsible for the prevailing deformation mechanism or deformation type, i.e. planar glide, wavy glide, deformation twinning, or martensitic transformation [6,7,8].

Different methods to determine SFE has been reported in literature. They fall broadly under two categories: experimental and theoretical calculations. One of the methods based on observation of extended dislocation nodes is used to measure SFE experimentally from electron diffraction measurements in transmission electron microscopy [9,10]. This method is direct & considered to be especially accurate for the materials with the low SFE [11,12]. The line profile analysis of XRD spectrum is another experimental technique which is used indirectly to determine SFE [11,13,14]. Neutron diffraction has also been used to determine SFE & as reported by the authors is better than the earlier mentioned experimental methods [15].

Experimental measurements of the SFE are particularly subtle and are subject to many sources of bias. They are not widely reported in literature and are often sources of controversy. For a given steel composition and temperature, the values reported can vary significantly from one author to another [16,17].

Most authors have thus chosen an indirect approach trying to correlate calculations of SFE on a thermodynamic basis with direct observations of deformation mechanisms by TEM. These calculations are based on a relationship that exists between the SFE and the driving force for -martensite formation first established by Hirth [18] and popularized by Olson and Cohen [19]. The approach considers that an intrinsic SF is in fact equivalent to a platelet of martensite of a thickness of only two atomic layers creating two new that: interfaces. It follows

SFEintrinsic = 2G + 2/

Where is the molar surface density along {1 1 1} planes and / is the surface energy of the interface /. The molar surface density can be calculated using the lattice parameter a and Avogadros constant N as

The estimation of the Gibbs energy G for the bulk transformation is critical for calculation of SFE [16,20,21]. An alternate approach for calculating SFE from quantum-mechanical first principles by using exact muffin-tin orbital method (EMTO) is also reported [17,22]. Also Peierls-Nabarro model fitted to generalized stacking fault energies is also used to determine SFE [23]. In order to optimize the mechanical properties of austenitic steels as desired, SFE has to be adjusted to an appropriate value [24]. The predominant variables; chemical composition & temperature which affect SFE are discussed below.

2.1.1 Effect of Alloying Elements on Stacking Fault Energy

The alloying effects on SFE are quite complicated and sometimes contradict each other from different experimental measurements. Additionally, as mentioned earlier the experimental SFE values are not accurate and are usually associated with large error bars. Based on the existing databases, several empirical relationships between SFE and chemical compositions 7

have been proposed [13,25]. However, the application of these empirical relationships is limited and in most cases they are unable to reproduce the complex nonlinear dependence of SFE on the composition [17]. Moreover, these empirical relationships hardly address the interactions between the different alloying elements. A review of stacking fault energy maps based on subregular solution model in high manganese steel is presented here [26].Table 2.1 provides the tabulated data of alloying element effect on SFE at a given temperature for a particular element from various referenced sources. The approach used to find SFE is also mentioned.

2.1.2 Effect of Temperature on Stacking Fault Energy

For a given steel composition, the SFE increases as a function of the temperature according to the following relationship [27,28,16]:

where a is the lattice parameter, N the Avagadro number &

the change in entropy

during the transformation. This equation comes from the derivative of the equation presented above for intrinsic SFE. If the temperature is higher than the -martensite start temperature Es, is negative. Thus, the SFE increases with temperature above Es.

For a given steel composition, at a temperature that is high compared to Es, the SFE is high, and the energetic cost of dissociation of perfect dislocation is unfavourable. The only possible deformation mechanism is dislocation glide. The high temperature and high SFE enable crossslip events and activation of multiple slip systems. At lower temperatures, dislocation glide becomes more and more planar, and below a certain value, large dissociation of dislocations becomes favourable. This enables the twinning process to occur as a competitive mechanism to dislocation glide since the energetic cost of twinning becomes sufficiently low. At even lower temperatures, -martensite formation replaces deformation twinning. The transition between mechanical twinning and -martensite formation is sustained by a shift of the 8

character of the stacking fault (SF) available to serve as a nucleus for these processes (intrinsic or extrinsic for twin & martensite respectively) [16,29]. Figure 2.2 shows TEM micrographs of the deformation microstructure of a Fe22Mn0.6C steel at two different temperature. It can be seen that at lower temperature (77 K) -martensite platelets are visible whereas at higher temperature (300 K) mechanical twins can been seen.

Figure 2.2 Dark-field TEM micrographs of the deformation microstructure of a Fe22Mn0.6C steel at 77 K (-martensite platelets) and at room temperature (mechanical twins) [16].

Table 2.1 Effect of alloying elements on stacking fault energy


Element Temp. (K) 300 Mn 300 Fe20Cr(16-20)Ni(0-8)Mn Fe22Mn0.6C 300 Al 300 Fe22Mn0.6C3Al Fe22Mn0.6C6Al Fe22Mn0.6C Fe22Mn0.6C3.5Al Fe22Mn0.6C4.5Al Fe31Mn0.25Si0.77C 300 Fe31Mn2.03Si0.77C Fe31Mn5.31Si0.77C Fe31Mn8.67Si0.77C Fe31Mn0.25Si0.77C Si 300 Fe31Mn2.03Si0.77C Fe31Mn5.31Si0.77C Fe31Mn8.67Si0.77C Fe22Mn0.6C 300 Fe22Mn0.6C3Si Fe22Mn0.6C8Si mass % at% at% mass % Base Composition (69.5-74.5)Fe13.5Cr12Ni(0-5)Mn (65.5-70.5)Fe17.5Cr12Ni(0-5)Mn Fe20Cr(8-16)Ni(0-8)Mn at% at% at% at% wt% SFE (mJ/m2) 44.6-32.1 28.6-17.6 Varied range Varied range 21.5 36.5 50.7 ~22.5 ~38 ~45 17.39 14.69 10.51 6.33 17.39 ~19 ~22.5 ~25 ~22.5 ~25.5 ~23 Si SFE & then Si SFE T Si SFE T Both non-magnetic & magnetic effect on SFE. Anomaly in shear modulus due to antiferromagnetic order. Inflection point observed at 3.5% Si [31] [32] Si SFE XRD SFE (mJ/m2)=17.53-1.30 (Si at%) [11] Al SFE T [31] Al SFE T [30] Effect on SFE Mn SFE Mn SFE Mn SFE Mn SFE EMTO [22] EMTO Method Remarks Ref.

Cr SFE

[17]

10

Element

Temp. (K) 300

Base Composition Fe(13-25)Cr14Ni Fe(13-25)Cr16Ni at% at% mass %

SFE (mJ/m2) ~45-15 ~47-25 ~22.5 ~18.5 ~17 at% at% at% ~9-40 ~0-35 ~0-27.5 ~27.5-25 at% at% 44.6-43.2 28.6-40.6

Effect on SFE Cr SFE Cr SFE Cr SFE Ni SFE Ni SFE Ni SFE & then Ni SFE Nb SFE Nb SFE Nb SFE Nb SFE

Method

Remarks Ni SFE

Ref.

EMTO

[17]

Cr 300

Fe22Mn0.6C Fe22Mn0.6C4Cr Fe22Mn0.6C6Cr 300 Fe17Cr(8-20)Ni Fe19Cr(8-20)Ni Fe20Cr(8-16)Ni Fe20Cr(16-20)Ni (69.5-74.5)Fe13.5Cr12Ni(0-5)Nb 300 (65.5-70.5)Fe17.5Cr12Ni(0-5)Nb Fe20Cr(<16)Ni(0-3)Nb Fe20Cr(8-16)Ni(3-8)Nb Fe22Mn0.6C

[31]

EMTO

Cr SFE

[17]

Ni 300

EMTO

Inflection point observed at 16% Ni (Nb&Cr &Fe) SFE

[22]

EMTO

Opposite effect due to interaction

[17]

Nb 300

at%

EMTO

[22]

mass %

~22.5 ~26 ~30 Cu SFE Co SFE N SFE N SFE N SFE N SFE T Inflection point observed at 0.3 % Ni [33] T [31]

Cu

300

Fe22Mn0.6C3.5Cu Fe22Mn0.6C7.5Cu

Co

300

Fe20Cr(8-20)Ni(0-8)Co Fe18Cr10Ni(0-0.3)N

at% wt%

Varied range ~23-24 ~24-8 ~27.5-28.5 ~28.5-13.5

EMTO

[22]

300

Fe18Cr10Ni(0.3-0.6)N Fe18Cr10Ni8Mn(0-0.3)N Fe18Cr10Ni8Mn(0.3-0.6)N

EMTO - Exact Muffin-Tin Orbital Method, T- Thermodynamic Calculations, XRD- X-Ray Diffraction Line Profile Analysis

11

2.2 Mechanisms of Plastic Deformation in Steels


Generally plastic deformation occurs by dislocation glide in steels. However, especially in the case of metastable austenite, either as retained austenite or fully austenitic structure; mechanical twinning & various phase transformation can take place sometimes being the dominant deformation mechanism.

Dislocation glide is the primary mechanism responsible for the change in shape during plastic deformation of most metals; both twinning and phase transformations can produce additional obstacles to dislocation glide and thus reduce the free mean path of dislocations. A reduction in the free mean path of dislocations, in turn, gives rise to higher strain hardening rates. In TRIP and TWIP steels advantage is taken of this phenomenon where martensitic phase transformation and twinning, respectively, occurs during the progress of straining. A certain amount of stress/strain is required to initiate the phase transformation or twinning of crystal planes [6]. The products of alternative deformation mechanisms are discussed briefly below.

2.2.1 -martensite

During the

phase transformation, the close-packed planes and directions in the

two structures are parallel, i.e., {111}||{0001} and <110>||<11 0>, and -martensite forms a hexagonal close-packed (hcp) crystal structure. In terms of {111} planes of the cubic crystal structure, the stacking sequence in the -martensite can be expressed as CACA. Therefore, each single intrinsic stacking fault contains a thin layer of -martensite phase. As mentioned earlier this consideration is used in the SFE calculation approach proposed by Olson and Cohen [19]. Single stacking faults can be regarded as -martensite nuclei, while the growth of perfect -martensite occurs by overlapping of the intrinsic stacking faults on every second {111} plane or extrinsic stacking successive parallel {111} planes [16]. The distinction between single stacking faults, bundles of overlapping stacking faults, and faulted or perfect -martensite is therefore not quite unambiguous. Numerous authors have even observed that both processes could be detected within the same shear bands. [34]

12

2.2.2 `-martensite

Through comparison of the close packed planes of the fcc and hcp lattice, i.e., the {111}- and the {0001}-plane, respectively, with the most closely packed plane {110} of the bcc lattice, it is evident that the close packed planes of the fcc and hcp lattices can be transformed into the {110} bcc-plane through little distortion. The lattice transformation can, e.g., be described through the Bain transformation, which transforms the fcc lattice through buckling and straining into the bcc lattice. The crystallographic orientation relationship between the and the phase obeys the Bain transformation relationship described by {111}|| {001} and <100>|| <110> . Another commonly observed transformation orientation relationship is the Kurdjumov-Sachs transformation, where close-packed planes of the fcc structure transform into the most closely packed planes of the bcc crystal structure, and close-packed directions into close-packed directions, i.e., {111}|| {110} and <110>|| <111>.

Many researchers have reported that the nucleation of -martensite takes place at the intersections of shear bands, while others have reported nucleation within a single shear band. Since in many cases shear bands consist of -martensite, it is often considered an intermediate phase in the phase transformation, which can then be written as

. Whether the crystal structure of the -martensite should be considered bcc or slightly tetragonally distorted bcc (bct), depends on the carbon content [6].

2.2.3 Deformation Twins

The classical definition of twinning requires that the lattices of the twin and the matrix are related either by a reflection in some plane, called the twin plane, or by rotation of 180 about some axis, called the twin axis [2]. In crystals of high symmetry, such as fcc, these orientations are equivalent. Deformation twins in fcc crystals form, at least in principle, by a homogeneous shear of the parent lattice along the {111} plane in the <112> direction. Frank proposed a way of creating a twin by stacking or overlapping of intrinsic stacking faults, i.e., by the glide of Shockley partials with identical Burgers vectors on successive {111} planes. The case of two intrinsic stacking faults overlapping on successive {111} planes is referred to 13

as an extrinsic stacking fault, which can also be considered as the special case of a twin, two atom-layers in thickness. If the overlapping of intrinsic stacking faults proceeds on successive {111} planes, the twin grows in thickness. The thickness of single twins is therefore only of nanometer scale, but stacks of nano-twins can add up to twin bands micrometers in thickness. The actual substructure in twin bands is lamellae like, where arrays of twins and matrix alternate [6].

Figure 2.3 is a schematic representation of the formation of mechanical twinning. As explained [ earlier the sequential movement of close packed layers along ] on (111) plane forms a twin plate. It is obvious that movement along the opposite

direction of bp1 is unfavourable because the translation required would be -2bp1 which doubles the shear and hence is not favoured from an energy point of view. Thus the shear displacement during twin formation can take place only in a definite direction but cannot be reversed. The direct consequence of this fact is the polarization of twinning direction which leads to: on each {111} twinning plane, only three <112 > directions out of six can be the twinning direction. Consequently there are 12 twinning systems in FCC, as listed in Table 2.2 [4]. The directional character of twins does not manifest itself in isotropic materials. However, in the presence of texture the twinning stresses in compression and tension can be different [6].

14

Figure 2.3 Schematic representation of the formation of twinning in FCC [4]

Table 2.2 Twelve possible twinning systems in FCC materials [4]

Twinning Plane Twinning Directions

[ ] [ ] [ ]

) [ ] [ ] [ ]

) [ ] ] [ [ ]

) [ ] [ ] [ ]

In most cases the critical event in the twinning process is nucleation, while growth can occur at stresses much lower than the nucleation stress. The critical twinning stress can, at least in theory, be formulated in analogy to the critical resolved shear stress CRSS for dislocation glide. However, the experimental evidence supporting this theory suffers from large scatter in measured twinning stresses. Several critical factors affecting the experimental results can be the reason for this large scatter. On the one hand, the incidence of twinning as well as the twinning-slip competition are very sensitive to orientation and therefore require that very accurate and reproducible orientations in the test specimens are guaranteed. On the other hand, impurity levels and defect structures giving rise to local stress concentrations can have a great effect on the twin nucleation in polycrystals, while in single crystals stress concentration sites of different nature can have a similar effect (e.g. surface notches, internal flaws, etc.). 15

Twinning in fcc metals often nucleates at microscopic defect structures and the local stress at which twinning initiates is considerably higher than the externally applied stress [6].

When an appropriate external shear stress is applied, the width d of the stacking fault may increase or decrease, depending on the type of stress. The equilibrium width when it increases is given by [35]

where is the applied shear stress. Therefore, the split distance is a function of the applied stress and the value of SFE. SFE is the intrinsic factor of the materials, while the applied stress represents the corresponding external deformation conditions. It can be seen that the split distance increases with increasing applied shear stress. If the applied stress increases to a critical value, the split distance will approach infinity. A critical value of shear stress clearly exists, above which the propagation of SF is catastrophic. In general, forming a wide SF is the prerequisite step of nucleation of a deformation twin. Therefore, the required twinning stress must be greater than this critical value. This value can be found out since as d; (

Thus from the above equation we get critical stress C as [35];

Thus C increases with increasing SFE. This has proven to be true mostly for fcc metals. Venables found a linear relationship for the twinning stress as a function of the square root of SFE. Narita et al. have confirmed that in accordance with

, and Narita and Takamura found a proportional relationship for Ni-Ge alloys, Cu, Au, and Ag according to , where bS is the Burgers vector of a Shockley partial [6].

16

The effect of grain size on the twinning stress obeys in most cases the Hall-Petch relationship,

where the slope kT is larger than the slope kS for slip, i.e., the twinning stress is more sensitive to grain size than the slip stress. The reason for the difference is still not fully understood, but Armstrong and Worthington attributed the higher grain size sensitivity of the twinning stress to microplasticity, i.e., to localized dislocation activity at stress levels where the metal is globally in the elastic domain, whereas the yield stress is associated with the onset of global plastic deformation [6,36].

2.3 Correlation between Deformation Mechanism & StackingFault Energy


The deformation mechanisms and mechanical properties of fcc metals are strongly related to their stacking fault energy (SFE) SFE. The dimensionless parameter SFE /Gb, where G is the shear modulus and b the slip distance, is a measure for the ease of cross slip of screw dislocations and therefore determines the work hardening behaviour during deformation. To fully exploit the twinning and/or phase transformation mechanisms in austenitic steels, the magnitude of SFE has to be properly adjusted by choosing the right amount of chemical alloying elements. At low SFE, wide dissociation of dislocations into Shockley partials can hinder dislocation glide and thus favour mechanical twinning (T) or martensitic phase transformation ( or ). The ` transformation occurs in

steels whose with SFE 12 mJ/m2 and generally at low temperature [16,37].

There is a wide variation in the SFE values proposed for martensitic transformation from about SFE 16 mJ/m2 to SFE 20 mJ/m2. For stacking fault energies of the order SFE = 25 mJ/m2 twinning becomes the dominant deformation mechanism. At even higher stacking fault energies of approximately SFE 45 mJ/m2, plasticity is controlled solely by dislocation glide [6,38].The data from numerous references relating the deformation mechanism/ deformation 17

microstructure observed for different steels at various SFE values is tabulated in Table 2.3. The deformation mechanism & deformed microstructure are generally characterized by XRD & TEM.

18

Table 2.3 Stacking fault energy & deformation behaviour


Base Composition Fe Bal. at% 22 28 Bal. at% 27 0.52 25 0.28 0.24 + Mo <0.01 4.1 0.08 0.05 6 1.6 1.6 0.6 0.08 0.08 <0.001 0.05 Mn 22 22 Si Ni Cr Al 0 3 C 0.6 0.6 300 Nb Cu N Temp. (K) SFE (mJ/m ) 21.5 36.5 50.7 27 20.5 T
2

Method

Deformation Mechanism/Microstructure

Remarks

Ref.

Planar glide followed by mechanical twinning Mechanical twinning Mechanical twinning

SFEC Strain for onset of twinning [30]

Below 225K martensite Above 350K dislocation slip Twinning frequency more [39] Twinning frequency less [21]

300

42

Mechanical twinning

21.9 Bal. Mass % 24.6 13 Bal. wt. % Bal. wt. % 22 18 24 22 22 22 21.9

0.45 0.59 0.59 0.3 0.1 0.6 0.7 0.6 0.6 0.6 77 293 693 300 300

15 20 25 ~0 7 19 35 10 19 80 T T T

Mechanical twinning Mechanical twinning Mechanical twinning Deformed & martensite & deformed martensite Mechanical twinning Mechanical twinning martensite Mechanical twinning Dislocation Gliding

[40]

Temp. SFE

[41]

19

Base Composition Fe Mn 1.49 1.71 1.22 Bal. wt. % 6.97 1.23 1.61 1.34 28.0 5.70 9.00 9.73 9.7 Bal. wt. % 9.82 10.19 9.66 Bal. wt. % ~10 ~13 ~16 ~20 0.17 0.25 0.22 17.54 18.02 18.12 0.03 0.15 0.38 0.6 0.6 0.6 0.6 2 2 2 2 0.69 0.42 0.38 9.58 Si 0.43 0.33 1.12 0.35 0.50 0.48 0.51 0.28 0.29 0.40 0.24 0.24 0.26 4.7 1.1 Ni 8.2 8.1 6.4 4.5 6.6 6.6 6.6 Cr 18.2 18.2 16.7 17.6 17.4 17.6 17.4 <0.01 17.3 15.2 17.83 18.06 17.65 1.6 0.001 0.002 0.002 0.005 Al 0.003 C 0.049 0.041 0.093 0.045 0.030 0.019 0.017 0.08 0.047 0.079 0.03 0.03 0.03 2.39 1.68 0.107 0.115 0.39 0.44 0.51 Nb Cu 0.43 0.37 0.25 0.25 0.168 0.22 0.14 N 0.047 0.054 0.074 0.198 0.168 0.094 0.145

Temp. (K)

SFE (mJ/m ) 30 29.2 26.6 24.6


2

Method

Deformation Mechanism/Microstructure Mechanical twinning

Remarks

Ref.

Mechanical twinning In the temperature T range 50T600 K Temp. SFE Mechanical twinning Mechanical twinning mart. & mech. twinning Deformed martensite [42]

300

23.7 23.2 22.6 27 24.9 16.8 10.4 12.2

martensite & mechanical 17.1 300 22.8 20.1 27.6 10 300 12 15 17 T ND twinning Mechanical twinning martensite & mechanical twinning Mechanical twinning Deformed & martensite Deformed & martensite Deformed martensite Deformed martensite [31] [15]

20

Base Composition Fe Mn 20.24 Bal. wt. % 20.72 Bal. wt. % 19.84 19.87 19.87 19 Bal. wt. % 19 19 19 19 19 5 5 5 5 5 5 0 1 2.5 3.5 4 5.5 2.00 2.46 0.01 1.3 1.31 1.31 0.25 0.25 0.25 0.25 0.25 0.25 1.5 3.04 0.022 22.57 21.55 Si 2.44 2.46 3.00 Ni Cr Al 1.95 1.18 0.69 C 0.01 0.01 0.01 Nb Cu N 0 0.011 0.052

Temp. (K)

SFE (mJ/m )
2

Method

Deformation Mechanism/Microstructure Deformed & martensite Deformed Deformed & martensite Mechanical Twinning

Remarks

Ref.

300

XRD

[43]

24.36 300 26.56 28.74 ~9 ~20 300 ~30 ~38 ~45 ~55 XRD T

Mechanical Twinning Mechanical Twinning Mechanical Twinning Deformed & martensite Mechanical twinning Mechanical twinning Mechanical twinning Mechanical twinning Mechanical twinning Al acts as stabilizer at low conc. And as stabilizer at high conc. [45] [44]

T- Thermodynamic Calculations, ND- Neutron Diffraction, XRD- X-Ray Diffraction Line Profile Analysis

21

References
[1] R E Smallman and A H W Ngan, Physical Metallurgy and Advanced Materials, 7th ed.: Elsevier Ltd, 2007. [2] J W Christian and S Mahajan, "Deformation Twinning," Progress in Materials Science, vol. 39, no. 1-2, pp. 1-157, 1995. [3] Ryuji Watanabe, "Generalized stacking fault energy in body cubic iron," Strength, Fracture and Complexity, vol. 5, no. 1, pp. 13-25, 2007. [4] Bo Qin, "Crystallography of TWIP Steel," Graduate Institute of Ferrous Technology, Pohang University of Science and Technology, Pohang, Master's Dissertation 2007. [5] D Hull and D J Bacon, Introduction to Dislocations, 4th ed. London, United Kingdom: Butterworth-Heinemann, 2001. [6] Sven Curtze, "Characterization of the Dynamic Behavior and Microstructure Evolution of High Strength Sheet Steels," Tampere University of Technology, Tampere, Doctoral dissertation ISBN 978-952-15-2288-8, 2009. [7] P Mullner and P J Ferreira, "On the energy of terminated stacking faults," Philosophical Magazine Letters, vol. 73, no. 6, pp. 289-298, June 1996. [8] O Grassel, L Kruger, G Frommeyer, and L W Mayer, "High strength FeMn(Al, Si) TRIP/TWIP steels development properties application," International Journal of Plasticity, vol. 16, no. 10-11, pp. 13911409, 2000. [9] M J Whelan, "Dislocation Interactions in Face-Centred Cubic Metals, with Particular Reference to Stainless Steel," Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, vol. 249, no. 1256, pp. 114-137, January 1959. [10] Jinkyung Kim and B C De Cooman, "On the Stacking Fault Energy of Fe-18 Pct Mn-0.6 Pct C-1.5 Pct Al Twinning-Induced Plasticity Steel," Metallurgical and Materials Transactions A, vol. 42, no. 4, pp. 932-936, April 2011. [11] Xing Tian and Yansheng Zhang, "Effect of Si content on the stacking fault energy in FeMnSiC alloys: PartI. X-ray diffraction line profile analysis," Materials Science and Engineering: A, vol. 516, no. 1-2, pp. 73-77, August 2009. [12] M J Whelan, P B Hirsch, R W Horne, and W Bollman, "Dislocations and Stacking Faults in Stainless Steel," Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, vol. 240, no. 1223, pp. 524-538, July 1957. [13] R E Schramm and R P Reed, "Stacking fault energies of seven commercial austenitic stainless steels," Metallurgical Transactions A, vol. 6, no. 7, pp. 1345-1351, July 1975. [14] Xing Tian, Hong Li, and Yensheng Zhang, "Effect of Al content on stacking fault energy in austenitic FeMnAlC alloys," Journal of Materials Science, vol. 43, no. 18, pp. 6214-6222, September 2008. [15] Tae-Ho Lee, Eunjoo Shin, Chang-Seok Oh, Heon-Young Ha, and Sung-Joon Kim, "Correlation between stacking fault energy and deformation microstructure in highinterstitial-alloyed austenitic steels," Acta Materialia, vol. 58, no. 8, pp. 31733186, May 2010. 22

[16] O Bouaziz, S Allain, C P Scott, P Cugy, and D Barbier, "High manganese austenitic twinning induced plasticity steels: A review of the microstructure properties relationships," Current Opinion in Solid State and Materials Science, vol. 15, no. 4, pp. 141-168, August 2011. [17] L Vitos, J O Nilsson, and B Johansson, "Alloying effects on the stacking fault energy in austenitic stainless steels from first-principles theory," Acta Materialia, vol. 54, no. 14, pp. 3821-3826, August 2006. [18] J P Hirth, "Thermodynamics of Stacking Faults," Metallurgical Transactions, vol. 1, no. 9, pp. 2367-2374, September 1970. [19] G B Olson and Morris Cohen, "A general mechanism of martensitic nucleation: Part I. General concepts and the FCC HCP transformation," Metallurgical Transactions A, vol. 7, no. 12, pp. 1897-1904, December 1976. [20] P J Ferreira and P Mllner, "A thermodynamic model for the stacking-fault energy," Acta Materialia, vol. 46, no. 13, pp. 44794484, August 1998. [21] S Curtze and V T Kuokkala, "Dependence of tensile deformation behavior of TWIP steels on stacking fault energy, temperature and strain rate," Acta Materialia, vol. 58, no. 15, pp. 51295141, September 2010. [22] Song Lu, Qing-Miao Hu, Borje Johansson, and Levente Vitos, "Stacking fault energies of Mn, Co and Nb alloyed austenitic Stainless Steels," Acta Materialia, vol. 59, no. 14, pp. 57285734, August 2011. [23] S Kibey, J B Liu, M J Curtis, D D Johnson, and H Sehitoglu, "Effect of nitrogen on generalized stacking fault energy and stacking fault widths in high nitrogen steels," Acta Materialia, vol. 54, no. 11, pp. 29913001, June 2006. [24] Sangwon Lee, Jinkyung Kim, Seok-Jae Lee, Bruno C, and De Cooman, "Effect of Cu addition on the mechanical behavior of austenitic twinning-induced plasticity steel," Scripta Materialia, vol. 65, no. 12, pp. 10731076, December 2011. [25] P J Brofman and G S Ansell, "On the Effect of Carbon on the Stacking Fault Energy of Austenitic Stainless Steels," Metallurgical Transactions A, vol. 9, no. 6, pp. 879-880, June 1978. [26] A Saeed-Akbari, U Imlau, U Prahl, and W Bleck, "Derivation and Variation in Composition-Dependent Stacking Fault Energy Maps Based on Subregular Solution Model in High-Manganese Steels," Metallurgical and Materials Transactions A, vol. 40, no. 13, pp. 3076-3090, December 2009. [27] L Remy, "Temperature variation of the intrinsic stacking fault energy of a high manganese austenitic steel," Acta Metallurgica, vol. 25, no. 2, pp. 173-179, February 1977. [28] Wan Jianfeng, Chen Shipu, and Xu Zuyao, "The influence of temperature on stacking fault energy in Fe-based alloys," Science in China (Series E), vol. 44, no. 4, pp. 345-352, Aug 2001. [29] H Idrissi, L Ryelandt, M Veron, D Schryvers, and P J Jacques, "Is there a relationship between the stacking fault character and the activated mode of plasticity of FeMn-based austenitic steels?," Scripta Materialia, vol. 60, no. 11, pp. 941-944, June 2009. [30] Kyung-Tae Park et al., "Stacking fault energy and plastic deformation of fully austenitic high manganese steels: Effect of Al addition," Materials Science and Engineering: A, vol. 527, no. 16-17, June 2010. [31] A Dumay, J P Chateau, S Allain, S Migot, and O Bouaziz, "Influence of addition 23

elements on the stacking-fault energy and mechanical properties of an austenitic FeMn C steel," Materials Science and Engineering: A, vol. 483-484, pp. 184-187, June 2008. [32] Xing Tian and Yansheng Zhang, "Effect of Si content on the stacking fault energy in FeMnSiC alloys: Part II. Thermodynamic estimation," Materials Science and Engineering: A, vol. 516, no. 1-2, pp. 78-83, August 2009. [33] I A Yakubtsov, A Ariapour, and D D Perovic, "Effect of nitrogen on stacking fault energy of f.c.c. iron-based alloys," Acta Materialia, vol. 47, no. 4, pp. 1271-1279, March 1998. [34] I Tamura, "Deformation-induced martensitic transformation and transformation-induced plasticity in steels," Metal Science, vol. 16, no. 5, pp. 245-253, May 1982. [35] W Z Han, Z F Zhang, S D Wu, and S X Li, "Combined effects of crystallographic orientation, stacking fault energy and grain size on deformation twinning in fcc crystals," Philosophical Magazine, vol. 88, no. 24, pp. 3011-3029, November 2008. [36] B C De Cooman, O Kwon, and K G Chin, "State-of-the-knowledge on TWIP steel," Materials Science and Technology, vol. 28, no. 5, pp. 513-527, May 2012. [37] G Frommeyer, U Bruex, and P Neumann, "Supra-Ductile and High-Strength ManganeseTRIP/TWIP Steels for High Energy Absorption Purposes," ISIJ Int, vol. 43, no. 3, pp. 438-446, 2003. [38] X D Wang, B X Huang, and Y H Rong, "On the deformation mechanism of twinninginduced plasticity steel," Philosophical Magazine Letters, vol. 88, no. 11, pp. 845-851, October 2008. [39] Shigeo Sato, Eui-Pyo Kwon, Muneyuki Imafuku, Kazuaki Wagatsuma, and Shigeru Suzuki, "Microstructural characterization of high-manganese austenitic steels with different stacking fault energies," Materials Characterization, vol. 62, no. 8, pp. 781 788, August 2011. [40] A Saeed-Akbari, A Schwedt, and W Bleck, "Low stacking fault energy steels in the context of manganese-rich iron-based alloys," Scripta Materialia, vol. 66, no. 12, pp. 10241029, June 2012. [41] S Allain, J P Chateau, O Bouaziz, S Migot, and N Guelton, "Correlations between the calculated stacking fault energy and the plasticity mechanisms in FeMnC alloys," Materials Science and Engineering: A, vol. 387-389, pp. 158162, December 2004. [42] S Curtze, V T Kuokkala, A Oikari, J Talonen, and H Hanninen, "Thermodynamic modeling of the stacking fault energy of austenitic steels," Acta Materialia, vol. 59, no. 3, pp. 10681076, February 2011. [43] B X Huang, X D Wang, L Wang, and Y H Rong, "Effect of Nitrogen on Stacking Fault Formation Probability and Mechanical Properties of Twinning-Induced Plasticity Steels," Metallurgical and Materials Transactions A, vol. 39, pp. 717-724, April 2008. [44] Xian Peng et al., "Stacking fault energy and tensile deformation behavior of high-carbon twinning-induced plasticity steels: Effect of Cu addition," Materials & Design, vol. 45, pp. 518-523, March 2013. [45] B W Oh et al., "Effect of aluminium on deformation mode and mechanical properties of austenitic Fe-Mn-Cr-Al-C alloys," Materials Science and Engineering: A, vol. 197, no. 2, pp. 147156, July 1995.

24

ACKNOWLEDGEMENT

I would like to express my warm gratitude and thanks to Prof. Prita Pant, my seminar supervisor, without whose enterprise and endeavour this seminar report would not have been a possibility. I would also like to express my sincere gratitude to Abhinandan Gangopadhyay and Srijan Sengupta for their support during the course of the present work. Karthik Iyer Roll no. 123114004

25

Vous aimerez peut-être aussi