Vous êtes sur la page 1sur 13

Available online at www.sciencedirect.

com

Acta Materialia 57 (2009) 35493561 www.elsevier.com/locate/actamat

Transformation-induced plasticity during pseudoelastic deformation in NiTi microcrystals


D.M. Noreet a, P.M. Sarosi b, S. Manchiraju c, M.F.-X. Wagner d, M.D. Uchic e, P.M. Anderson c, M.J. Mills c,*
a Engineering Systems Inc., 3851 Exchange Ave., Aurora, IL 60504, USA General Motors, R&D Tech Center, 30500 Mound Road, Warren, MI 48090, USA c Department of Materials Science and Engineering, The Ohio State University, 2041 College Road, Columbus, OH 43201, USA d Lehrstuhl Werkstowissenschaft, Institut fu r Werkstoe, Ruhr-Universita t Bochum, Universita tsstr. 150, D-44801 Bochum, Germany e Air Force Research Labs RXLM, Wright-Patterson AFB, OH 45433, USA b

Received 20 February 2009; received in revised form 4 April 2009; accepted 7 April 2009 Available online 10 May 2009

Abstract [1 1 0]-oriented microcrystals of solutionized 50.7 at.% NiTi were prepared by focused ion beam machining and then tested in compression to investigate the stress-induced B2-to-B190 transformation in the pseudoelastic regime. The compression results indicate a sharp onset of the transformation, consistent with little prior plasticity. Post-mortem scanning transmission electron microscopy reveals no apparent retained martensite but rather a macroscopic band of dislocation activity within which are planar arrays of $100 nm dislocation loops involving a single ah0 1 0i{1 0 1} slip system. Micromechanics analyses show that the angle of the band is consistent with activation of a favored martensite plate. Further, the stress from the individual variants within the plate is shown to favor activation of the observed slip system. The work done by the applied stress during the B2-to-B190 transformation is estimated to be $34 MJ m3 at ambient temperature. 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Shape memory alloy; NiTi; Pseudoelastic response; Micromechanics; Stressstrain response

1. Introduction Shape memory alloys (SMAs) have remarkable properties that stem from a martensitic transformation that can be induced by an applied stress or cooling, and reversed upon unloading or heating. Several well-known alloys exhibit this response and they have enabled a variety of applications from valves and actuators to surgical devices [13]. The dominant transformation in NiTi SMAs is from the B2 crystal structure to a martensite variant. In a solutionized form, the transformation is directly from the B2-to-B190 structures [4]. In the pseudoelastic mode, a NiTi SMA with a B2 structure is stressed suciently to induce deformation via the
*

Corresponding author. Tel.: +1 614 292 1537. E-mail address: mills.108@osu.edu (M.J. Mills).

transformation to the B190 structure. Upon unloading, reverse deformation occurs via the reverse transformation. The crystallographic aspects of the martensitic transformation, including the orientation relationship with the ordered B2 matrix, have been calculated and veried in several systems [59]. However, some fundamental aspects of shape memory and pseudoelastic behavior are not yet understood, notably how the matrix accommodates the large strain associated with the transformation. Theoretically, accommodation may be achieved either by matrix plasticity or by inducing other transformation variants. As elaborated in Section 4.1, there are few direct experimental observations using transmission electron microscopy (TEM) that directly explore these accommodation mechanisms [10,11], and no complementary modeling that attempts to rationalize these mechanisms, even for the

1359-6454/$36.00 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actamat.2009.04.009

3550

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

archetype SMA system, NiTi. This fundamental question is technologically important since the useful service life of components under cyclic loading or loading/heating (i.e. functional and structural fatigue [12,13]) is determined by the evolution of the microstructure and defect substructure. Much attention has been focused on the eect of sample size on dislocation plasticity, using compression samples with diameters less than 10 lm. A tremendous strength increase can be achieved in pure nickel and pure gold samples when the diameter is decreased from 20 lm to submicron dimensions [1419]. The hypothesis is that small sample dimensions can impede or interact with dislocation motion, generation and plasticity. For NiTi, there is evidence that dislocation plasticity degrades the pseudoelastic eect by causing irreversible or residual strain [20]. Thus, enhanced pseudoelastic eects might be achieved if small sample sizes can inhibit plasticity in NiTi alloys, thereby raising the yield strength relative to the phase transformation stress. Similar to Frick et al. [21,22], the present investigation employs focused ion beam machining to fabricate singlecrystal cylinders that are subsequently compressed and unloaded. However, the sample diameters in this investigation are larger (5 and 20 lm), and, in addition, signicant improvements are made in the machining and loading geometries to minimize artifacts associated with sample misalignment and nonuniform sample cross-sectional area. Relative to the NiTi literature as a whole, a second significant advance is to couple the microcrystal experiments with post-mortem, ex situ, scanning transmission electron microscopy (STEM). An important nding is that such microcrystalline tests can activate particular martensite plates, enabling more precise measurements of the critical stress for forward/backward transformation in the absence of large-scale interaction between competing martensite plates. Further, thin foils are prepared that encompass the entire slip plane or entire gage length of the sample. The results reveal detailed dislocation congurations associated with the phase transformation. A nal advance is to

interpret the stressstrain and STEM data via micromechanics analyses of the most preferred martensitic plates to form, as well as the resolved shear stress on slip planes in the vicinity of martensitic plates. 2. Materials and methods 2.1. Experimental Bulk polycrystalline NiTi bar stock containing 50.7 at.% Ni was heat-treated in an argon atmosphere at 1000 C for 20 h, in order to increase the average grain size to >300 lm. The samples were then water quenched to prevent the formation of metastable precipitates (e.g. Ni4Ti3). They were cut to button-sized pieces and mechanically parallel-polished using an Allied High Tech MultiPrep System. Colloidal silica was used in the nal polishing step to achieve a surface nish of 0.05 lm. Grain orientations were determined from electron backscatter diraction (EBSD) on a FEI XL-30 environmental scanning electron microscope. Fig. 1a shows a grain orientation image map (OIM). The large grain indicated by the arrow in the OIM was selected for subsequent machining into cylindrical compression samples. Fig. 1b shows that the surface normal of the large grain is parallel to within 4 of the [1 1 0] direction, with a rotation toward [1  1 0]. Prior work by Gall et al. [20] has shown that, compared to other crystal orientations, bulk [1 1 0] single-crystal compression samples exhibit a moderate strain driving the B2-to-B190 transformation during loading, and upon unloading, intermediate values of remnant strain result. Micromachining was performed using a FEI DB 235 dual-beam focused ion beam (FIB) and the procedure introduced by Uchic et al. [14]. Two 5 lm and one 20 lm diameter micropillars were fabricated from the large grain, each having an aspect ratio of about 2.5:1. Fig. 2 shows an SEM micrograph of a 5 lm pillar. The circular, ducial mark on the top surface aids the automated fabrication

Fig. 1. (a) OIM of polycrystalline 50.7 at.% NiTi sample. The large central grain (arrow) was chosen for pillar fabrication. (b) h1 1 0i pole gure showing the near [1 1 0] orientation of this grain, together with the crystal indexing used in this paper. The grain normal (pillar compression axis) is at the center of the pole gure, rotated by about 4 away from [1 1 0] toward [1 1 0], corresponding to an approximate orientation of [7 8 0]).

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

3551

STEM detector operated at spot size = 1, camera length = 700 mm and objective aperture = 40 microns. The brighteld STEM imaging enables diraction-contrast over a relatively large eld of view while minimizing bend contour contrast which typically obscures diraction-contrast imaging [18,19]. An appropriate convergence angle (in this case 9 mRad) was used to minimize bending contour contrast yet maintain dislocation contrast, enabling g b analysis to furnish dislocation Burgers vectors when a two-beam condition is used [18,19]. 2.2. Modeling
Fig. 2. (a) Pillar specimen (5 lm) resulting from the lathe machining FIB processing. Note the parallel and vertical sidewalls that can be achieved. (b) Schematic showing the crystallographic orientation of the TEM foils extracted from the pillars. The foil plane includes the primary [0 1 0] and [1 0 0] Burgers vectors for this crystal orientation.

process. A precise pillar shape results, with a constant cross-section and at contact surface. This is in contrast to less-rened techniques that produce rounded cylinder tops and signicant tapers (e.g. $3) in the cylinder walls, especially at submicron sample sizes [21,22]. The samples were compressed uniaxially at ambient laboratory temperature under constant displacement rate, giving a nominal strain rate of 104 s1. A MTS Nano Indenter XP was used with a cleaved diamond indenter tip, creating a at platen to conform with the top cylinder surface [14]. An integrated goniometer enabled small adjustments to the sample orientation during repeated loading and unloading, so that the initial, nonlinear toein portion of the loaddisplacement trace was reduced to a minimum. TEM samples were prepared with a foil plane that contains the [1 0 0] and [0 1 0] Burgers vectors (Fig. 2b). They were extracted with an in situ plucking device (OmniProbe), and thinned by FIB to achieve an electron-transparent thickness (<250 nm). The microscopy observations used a 200 kV FEI/Phillips Tecnai TF20 with a eld-emission electron source and the high angle annular dark eld

Two types of modeling were employed to identify: (i) the most favored types of martensite plates to form under h1 1 0i compressive loading; and (ii) the most stressed slip systems in the vicinity of these favored plates. In particular, the phenomenological theory of martensitic transformations [58] can be applied to capture the essential features of the B2-to-B190 transformation. Fig. 3 shows a schematic of a martensitic plate (twinning mode), which is a weighted combination of two martensite variants, k and m. They are arranged in alternating parallel laths that are typically a few nanometers in thicknesssubstantially smaller than the plate dimensions, 2c ( 2a % 2b. The crystallographic normal to the plate is m(k,m) and the transformation induces an average displacement vector b(k,m) of a material point in the plate, relative to another material point positioned a unit distance along the m(k,m) direction. The term average is used in the sense of smearing out the individual martensite variants into a continuum with a uniform transformation strain. In reality, large local stress magnitudes are generated near individual B2martensite variant interfaces. The twin interface between laths has a crystallographic normal n(k,m) and displacement vector a(k,m), dened analogously to b(k,m). For the B2-to-B190 transformation, there are 12 possible martensite variants and 192 possible plates specied by m(k,m), b(k,m) (see Table 7.3 in Ref. [8]) The rst goal is addressed by ranking the transformation strain along the axis of the pillar for each of the 192

Fig. 3. Martensite plate geometry formed by martensite variants k and m. The plate is modeled as a cuboid with dimensions 2a, 2b, 2c, where a = b = 20c. x3 = c are invariant planes with a crystallographic normal m(k,m), x2 is the orthogonalized shear direction given by b(k,m) (b(k,m) m(k,m))m(k,m), where b(k,m) is the transformation displacement vector. x1 is orthogonal to x2 and x3. The interface between variants k and m has the crystallographic normal n(k,m) and displacement vector a(k,m).

3552

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

possible plates. In particular, the martensitic transformation induces an average deformation gradient F I b  m; or F ij dij bi mj 1

3. Results 3.1. NiTi microcrystal compression results Fig. 4 displays the resulting uniaxial compression stress strain curves from three NiTi microcrystals: two 5 lm diameter samples and one 20 lm diameter sample. These samples were strained to a total compressive value $2.53% and unloaded (Fig. 4a). Superimposed is the response of a bulk [1 1 0]-oriented single crystal from Gall et al. [20]. Fig. 4b details the response of the 5 lm samples, one of which has two cycles. Fig. 4c shows the one- and two-cycle response of the 20 lm sample, and Fig. 4d shows the response for all four cycles of the 20 lm sample. All of the micron-scale cases exhibit a ag-shaped hysteresis loop, indicative of pseudoelastic behavior. As exemplied in Fig. 4c, the typical loading response consists of (i) an initial linear-elastic portion, (ii) a small upper yield point, (iii) a near-plateau region, and (iv) signicant hardening, especially in the 20 lm case. The unloading response is (v) initially linear-elastic, giving way to (vi) a nonlinear region where a large fraction of total strain is recovered, followed by (vii) linear-elastic unloading. Several comparative aspects are noted. The very small (e $ 0.1%) toe-in region at the beginning of each test is evidence of the good sample alignment that can be achieved using the goniometer stage. This is in contrast to the large (>1%) toe-in regions in Frick et al. [21,22]. Despite the reduced toe-in region in the present tests, the elastic modulus is consistently smaller than Eh110i = 148 GPa based on published anisotropic elastic moduli [29]. The discrepancy is likely due to error in the nominal strain as well as machine and substrate compliance. The pillars show an abrupt transition to the plateau in region (iii), occurring at a compressive stress of $610 625 MPa. The abrupt nature of the transition suggests nucleation of a single or limited number of dominant plates in a homogeneously stressed sample, with negligible prior plasticity. In contrast, the bulk single crystal has a smooth, gradual transition, beginning with a noticeable departure from linear-elastic behavior at $380 MPa (Fig. 4a) and rising to a plateau at $600 MPa. The large ($75%) increase in stress cannot be rationalized in terms of the activation of a single plate. Rather, the internal stress state must be suciently inhomogeneous to nucleate plates over a large range of stress. The prospect of nucleating a single or limited number of plates under such conditions is small. Bulk single crystals with other orientations have similar gradual transitions [20]. Thus, micron-scale tests are unique and oer the capacity to precisely determine the onset of transformation. An additional consideration is that the bulk sample has a concentration of 50.9 at.% Ni compared to 50.7 at.% Ni for the micron-scale samples. The forward transformation stress for the 50.9 at.% Ni bulk case is therefore decreased by $60 MPa to replicate a 50.7 at.% Ni bulk case. This correction is based on a ClausiusClapeyron construction

within the plate. The identity matrix is I and the Kronecker delta function dij = 1 when i and j have the same value (1, 2 or 3) and = 0 otherwise. The transformation causes a vector v in the B2 phase to stretch and rotate to a vector v0 in the martensite plate, where v0 F v; or v0i
3 X j1

F ij vj

The resulting axial strain along the compression axis induced by the transformation is e v0 v 1 vT F v 1 1 3

where v = ec is a unit vector along the compression axis. The second goal is addressed by solving for the stress eld around an inclusion of volume X within which there is a transformation strain 1 e F T F I 2 4

where F is dened in Eq. (1). The goal is to nd a displacement eld u(x) satisfying [23]
3 X 3 X 3 X k 1 l1 j1

C ijkl

3 X 3 X 3 X o2 uk oe C ijkl kl ox l ox j ox j k 1 l1 j1

The components Cijkl of the elastic moduli are approximated to be the same for both the B2 and martensite phases. Recent rst-principles calculations, although strictly valid at 0 K, show the elastic moduli of B2 and B190 to be similar in magnitude [24,25]. Closed-form solutions to Eq. (5) are provided for particular inclusion shapes [23,26,27]. The stress eld from transformation regions of arbitrary shape is obtained by introducing a scalar function g(x) that = 1 inside the transformed region and = 0 otherwise. This order parameter can be expressed as a Fourier series X gn1 ; n2 ; n3 ein1 x1 ;n2 x2 ;n3 x3 6 gx
n1 ;n2 ;n3

where the Fourier coecients g(x1, x2, x3) are determined by the shape of the transformation region and the dimensions (L1, L2, L3) of the periodic cell containing the transformation region. Eq. (5) is solved for u(x) using the fast Fourier transform (FFT) method with strain eij(x) = 0.5(oui/ oxj + ouj/oxi) and stress rij(x) = Cijklekl(x), where ekl(x) is the elastic strain eld. Two approaches are taken to model the transformation. The plate approach takes X to be a plate with a uniform, volumeaverage transformation strain (Eq. (4)). The variant approach uses several X transformation volumes, each corresponding to an individual martensite variant in the plate.

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

3553

Fig. 4. Near [1 1 0] compressive stressstrain response of solutionized 50.7 at.% NiTi microcrystalline pillars with 5 or 20 lm diameters (present work) and a solutionized 50.9 at.% NiTi bulk crystalline sample [18] for (a) all samples; (b) 5 lm diameter samples; (c) 20 lm diameter sample; and (d) 20 lm diameter and bulk samples.

using data for NiTi [4] that Ms is about 20 C higher for 50.7 at.% Ni. Thus, the plateau stress for the micron-scale pillars is $14% larger than that for a bulk sample of similar composition. After the abrupt transition (iii), the hardening out to $3% strain is a function of specimen size. For example, the average hardening rate in this strain range varies from $1 to 3 GPa for the 5 lm case to $6 GPa for the 20 lm case to $12 GPa for the bulk case. The bulk sample has a maximal hardening rate at small strain with a monotonic decay to nearly zero at 3% strain. In contrast, the 5 lm samples have a negligible hardening rate initially but this increases with strain (Fig. 4b). Beyond 3% strain, the bulk sample has negligible hardening but the 5 lm samples show continued hardening. The micron-scale samples in the present work are consistently stronger in the forward (loading) direction than the bulk sample case. The dierence is most pronounced early in the nonlinear deformation process. At large strain, the

ow stress for the 20 lm case (Fig. 4d) far exceeds the bulk case ($1.2 GPa vs. $600 MPa at 12% strain). Thus, the plastic ow stress at large strain appears to be enhanced in the micron-scale case. The unloading behavior is also a function of sample size as well as cycle number. During the rst cycle, both 5 lm cases display two transitionsone at $470 MPa (point B, Fig. 4b) that is consistent with the onset of the transformation back to B2 and a very sharp transition at $360 MPa (point C, Fig. 4b) that is consistent with the completion of the back-transformation. These transitions are not as sharp on the second cycle nor are they as sharp for the 20 lm or bulk cases. These two transition points occur at smaller stress as sample size increases, at least for the rst unloading cycle. The trend suggests that during the rst unloading cycle, smaller samples have less substructure to constrain the back-transformation. The limited data presented here does not show a strong sample size eect on the magnitudes of remnant strain and

3554

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

reversible strain upon unloading. In particular, the 5 and 20 lm cases have a comparable remnant strain ($0.2%) after a similar total imposed strain ($3%) on the rst cycle. Second, the 20 lm (fourth cycle) and bulk (rst cycle) cases show a comparable reversible strain (4.5% vs. 5.0%) after a large total imposed strain of $1012% (Fig. 4d). The substructure generated by the rst cycle produces characteristic features in subsequent cycles. The second cycle loading curves (5 and 20 lm cases) transition abruptly when the rst cycle plateau is reached. For the 5 lm case, the stress drop is not present in the second cycle, suggesting that the substructure aids the forward transformation upon reloading. The nature of the substructure is not sensitively dependent on the total strain in the rst cycle. In Fig. 4b, the two rst-cycle cases complete the back-transformation at $360 MPa and have $0.2% remnant strain, even though the total strain is dierent ($3% vs. $3.4%). In addition, the two unloading curves from point A nearly overlap initially, suggesting that the substructure does not interfere with the rst regions to backtransform, but interferes with the later regions to backtransform. 3.2. STEM results for the one- and two-cycle 5 lm pillars The bright-eld STEM images of Fig. 5a and b show a band of remnant dislocation structure in each of the one-

and two-cycle 5 lm samples, respectively. The bands are most likely regions that transformed to martensite during loading, and then transformed back to austenite upon unloading. Diraction patterns from various locations in each of the foils show no indication of residual martensite. The dislocation density is much greater in the top portion of the band, suggesting that the transformation probably initiated at the upper surface. This is not surprising since even a slight misalignment between the indenter and pillar produces stress concentrations at the corners. Fig. 6 shows images used for conventional dislocation analysis. All dislocations inside and outside the transformation zone satisfy an invisibility condition using g = [ 1 0 0], indicating that they all have a Burgers vector of a[0 1 0]. This is most clearly demonstrated in the lower dislocation density, one-cycle sample shown in Fig. 6b. The relatively strong residual contrast is discussed in more detail below. The densely arrayed dislocation congurations inside the former transformation zones exhibit similar attributes for both the one- and two-cycle samples. At lower magnication (Fig. 6b and c), linear features are apparent that run nearly parallel to the [1 1 0] loading axis (rotated several degrees clockwise toward [ 1 1 0]). Fig. 7 shows that these linear features are arrays of elongated dislocation loops that appear to lie on parallel (1 0  1) planes. A series of images at dierent tilt angles shows that the dislocations

Fig. 5. Diraction-contrast STEM images using g = [0 1 0] for (a) one-cycle and (b) two-cycle 5 lm pillars. The foil normal is [0 0 1].

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

3555

Fig. 6. Higher-magnication diraction-contrast STEM images. Within the former transformation zone for the 5 lm one-cycle sample using (a) g[0 1 0] and (b) g[1 0 0]. At the periphery of the transformation zone for the two-cycle sample using (c) g[0 1 0] and (d) g[100]. The foil normal is [0 0 1]. Arrows indicate the g-vectors for each image.

outside the apparent, former transformation zone are elongated along the nonorthogonal directions [1 4 1] and [ 14 1] in the parent phase, giving them a distinctive arrowhead shape. Thus, the small loops within the transformation zone and the arrowhead congurations at the periphery of the transformation zone both lie on (1 0  1) planes. Fig. 6b and c have residual contrast since g (b u) is non-zero. For instance, the arrowhead congurations have |g (b u)| = 1 and generate a typical weak, doublepeak contrast. Frequent intersections of the dislocation loop features with the foil surfaces may also contribute to the strong residual contrast. The result that much of the a[0 1 0] dislocation content lies on (1 0  1) planes is rather surprising since the h1 0 0i/ {1 1 0} family has a lower Schmid factor (0.35) compared

to the h1 0 0i/{1 0 0} family (0.50). Hence, the observed dislocations are not on a slip system with the highest Schmid factor, with respect to the macroscopic applied stress. The dramatic increase in dislocation content between the one- and two-cycle samples, both of which have been deformed to the same maximum compressive strain, clearly suggests a direct interplay between the transformation and the observed dislocation structures. 3.3. Analytic modeling of the preferred martensite plate The procedure outlined in Section 2.2 was used to identify the most likely twinning modes (plates) to form, based on: (i) the macroscopic forward transformation strain; (ii) a ranking of the modes that generate the largest compressive

3556

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

Fig. 7. Tilt series of STEM images for the 5 lm one-cycle sample using g[0 1 0] for all images. The images are tilted about a [1 1 0] rotation axis. (a) 26 tilt from [0 0 1] zone toward [1 1 0]. (b) Near the [0 0 1] zone. (c) 26 tilt from [0 0 1] zone toward [1 1 0]. (df) Corresponding schematic illustrations of the dislocation loop arrays observed in the former transformation zone. Arrows indicate the g-vectors for each image.

strain; and (iii) the orientations of the band of residual dislocation content and planar arrays of dislocations. The rst aspect was addressed by noting that for the two-cycle sample, the rst forward cycle imparts a macroscopic compressive transformation strain of $1.65% (Fig. 4b). This strain could be achieved if the transformation strain in the band were $5%, since STEM images show the volume fraction of the band to be $1/3. The second aspect is pursued by ranking the compressive axial strain produced by each of the 192 possible twinning modes, using Eq. (3). Table 1 lists the 32 twinning modes that produce an axial [1 1 0] compressive strain of >5%. These are arranged into eight classes, T1T8, each associated with a mode (A, B or C) and type (I or II) following the convention in Table 7.3 of Bhattacharya [8]. Each class is comprised of four crystallographic permuations of m(k,m) and b(k,m), with axial compressive transformation strains ranging from 5.2% to 5.6%. The consistency with the macroscopic forward transformation strain suggests activation of these highly favored twinning modes. The nal aspect is pursued by superimposing the predictions for the T4 class and a STEM image from the 5 lm two-cycle sample, as shown in Fig. 8. The assumption here is that the observed dislocation content is intimately related to the activation of a particular martensite plate. The projected value h0 = 31 ts remarkably well.

In principal, the h0 = 30 prediction for the T1 class also ts well. However, the T4 class has a twin interface orientation (k0 = 9) that aligns well with the linear, nearly vertical lines in Fig. 8, which are traces of the dislocation loop arrays in Fig. 6. In contrast, the T1 class has poor alignment (k0 = 38). Thus, the T4 class is viewed as most consistent. The slight misorientation (see Section 2) puts the compression axis close to [7 8 0] rather than [1 1 0]. In this case, the T4 combinations (k, m) = (5, 8) and (8, 5) produce the maximum axial strain, relative to all 192 possible martensite plate types. 3.4. Analytic modeling of stressed slip systems near a martensite plate The stress state around the T4 class of plates may explain why dislocations are observed on a lower Schmid factor (0.35) system, (1 0  1)/[0 1 0]. The FFT method (Section 2.2) is used with the package FFTW [28]. The periodic cell dimensions L1 = 4a = L2 = 4b and L3 = 4c isolate the stress elds of individual plates, and sucient accuracy is obtained with a three-dimensional grid of 256 256 512 points. The thin plate geometry in Fig. 3 is used, with the invariant planes at x3 = c, where c = a/20 = b/ 20. The Cartesian coordinate system has x3 || m(k,m) and x2 || b(k,m)-(b(k,m) m(k,m))m(k,m).

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561 Table 1 Analysis of potential twinning modes and martensitic variants.a Caseb T1 (B,I) T2 (B,I) T3 (B,II) T4 (B,II)h eaxialc (%) 5.2 5.6 5.6 5.3 (k, m)d (1,4), (4,1) (5,8), (8,5) (1,4), (4,1) (5,8), (8,5) (1,4), (4,1) (5,8), (8,5) (5,8) (8,5) (1,4) (4,1) (1,5), (4,8) (5,1), (8,4) (1,5), (4,8) (5,1), (8,4) (1,5), (4,8) (5,1), (8,4) (1,5), (4,8) (5,1), (8,4) m(k,m) (0.91, 0.25, 0.33) (0.25, 0.91, 0.33) (0.35, 0.83, 0.43) (0.83, 0.35, 0.43) (0.38, 0.77, 0.51) (0.77 0.38, 0.51) (0.22, 0.89, 0.40) (0.22, 0.89, 0.40) (0.89, 0.22, 0.40) (0.89, 0.22, 0.40) (0.0, 0.81, 0.58) (0.81, 0.0, 0.58) (0.86, 0.28, 0.43) (0.28, 0.86, 0.43) (0.91, 0.20, 0.36) (0.20, 0.91, 0.36) (0.04, 0.89, 0.46) (0.89, 0.04, 0.46) b(k,m) (0.05, 0.11, 0.05) (0.11, 0.05, 0.05) (0.12, 0.03, 0.04) (0.03, 0.12, 0.04) (0.12, 0.02, 0.05) (0.02, 0.12, 0.05) (0.10, 0.06, 0.06) (0.10, 0.06, 0.06) (0.06, 0.10, 0.06) (0.06, 0.10, 0.06) (0.09, 0.04, 0.04) (0.04, 0.09, 0.04) (0.0, 0.09, 0.06) (0.09, 0.0, 0.06) (0.0, 0.1, 0.05) (0.01, 0.0, 0.05) (0.1, 0.03, 0.04) (0.03, 0.10, 0.04) h0 e 30 68 71 31 31 31 31 45 +63 +57 42 k0 f +38 38 9 +9 +9 +9 +9 0 0 90 +90
i j j i

3557

Slip systemg

T5 (C,I) T6 (C,I) T7 (C,II) T8 (C,II)


a

5.2 5.6 5.6 5.3

See Fig. 3 for a description of m, b, n, and a. (Mode, type) as indicated in the phenomenological theory of martensite, e.g. see Table 7.3 in Ref. [8]. c As predicted by Eq. (3). d k and m refer to particular martensite variants as dened in Table 4.3 in Ref. [8]. e The counterclockwise angle between the [1 1 0] axis and the invariant plane, projected on the [0 0 1] plane. f The counterclockwise angle between the [1 1 0] axis and the twin interface, projected on the [0 0 1] plane. g The max resolved shear stress, obtained from the micromechanics calculations of Section 3.4. h (5,8) a = (0.02, 0.01, 0.29), a(1,4) = (0.01, 0.02, 0.29), and n(5,8) = (0.81, 0.59, 0), n(1,4) = (0.59, 0.81, 0). The a3 and n3 components are reversed when (k, m) are reversed. i (1 0  1)/[0 1 0] has the maximum stress on x1 = 1.0125a and x2 = 1.0125b; (1 0 1)/[0 1 0] has the maximum stress on x3 = 1.05c. j (1 0 1)/[0 1 0] has the maximum stress on x1 = 1.0125a and x2 = 1.0125b; (1 0  1)/[0 1 0] has the maximum stress on x3 = 1.05c.
b

Table 2 Legend for the austenite slip systems. Label 1 2 3 4 5 6 7 Austenite slip system (0 1 0)/[1 0 0] (0 0 1)/[1 0 0] (0 0 1)/[0 1 0] (0 1 1)/[1 0 0] (0 1  1)/[1 0 0] (1 0 1)/[0 1 0] (1 0  1)/[0 1 0] and and and and and equivalent equivalent equivalent equivalent equivalent (1 0 0)/[0 1 0] (1 0 0)/[0 0 1] (0 1 0)/[0 0 1] (1 1 0)/[0 0 1] ( 110)/[0 0 1]

6 U1 4 e e

c e a d

7 d5 a

3 and

6 U4 4 e a e d

3 7

7 d 5 a

Two representations are employed to study the stress eld. In the plate approach, the transformation strain in the plate has a uniform value 0.5(b  m + m  b). Isotropic elastic moduli (E = 148 GPa, m = 0.3) are also assumed. It was conrmed that the normalized stress distribution r23(x1)/E from the FFT approach agrees well with the analytical solution by Chiu [27] for an isolated cuboidal inclusion with a uniform transformation strain. This validates the FFT method, but the plate approach is unable to capture plasticity at the individual variant scale, as revealed by the STEM images. In the variant approach, the stress eld around a plate composed of an arbitrary number of (k, m) = (4, 1) variant pairs is computed using the FFT approach. The B2-to-B190 transformation produces a strain 0.5(UT i Ui I), where

are referred to the austenite cubic basis, with a = 1.0243, c = 0.9563, d = 0.058 and e = 0.0427 [8]. The austenite elastic moduli are C11 = 130 GPa, C22 = 98 GPa and C44 = 34 GPa [29], with the same values assumed for martensite, as discussed in Section 2.2. A macroscopic compressive transformation stress of 600 MPa is superimposed as observed for the forward transformation stress in the experiments (Fig. 4a). Fig. 9 shows the most stressed austenite slip systems just outside the martensite plate (Fig. 3). Fig. 9ac show results on the cuts x1 = 1.0125a, x2 = 1.05b and x3 = 1.05c, respectively. A type (i) plot indicates the favored austenite {1 0 0}/h0 0 1i or {1 1 0}/h0 0 1i slip system with the largest resolved shear stress. A type (ii) plot shows the corresponding resolved shear stress (in MPa) on the favored system. The stress is clearly localized along contours parallel to the underlying twin interfaces. Fig. 10 shows magnied views of the same three cuts in Fig. 9, except that a lter is applied to show only slip systems for which the resolved shear stress exceeds 1500 MPa. A white background indicates regions where no slip system

3558

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

meets this criterion. Although the 1500 MPa criterion is arbitrary, it serves as an eective lter to identify the most likely spatial regions for slip activity and the particular slip systems involved. A value >1500 MPa would shrink the predicted bands of slip activity and a smaller value would enlarge them, eventually reproducing the type (i) plots in Fig. 10. Three signicant observations are made from Fig. 10. First, the active slip systems in the vicinity of the plate do not include (1 0 0)/[0 1 0], even though it has the largest Schmid factor (0.50). Second, the experimentally observed (1 0  1)/[0 1 0] system is predicted to be dominant on two faces of the plate. Third, the predicted regions for (1 0  1)/ [0 1 0] and (1 0 1)/[0 1 0] activity are aligned to the twin interfaces in the martensite plates, correlating well with the STEM images of planar loop arrays (Fig. 8). Two additional comments pertain to the scope of the variant approach. Figs. 9 and 10 show the results based on the T4(4,1) plate mode in Table 1. However, the rightmost column in Table 1 shows that the other T4 modes always favor slip systems (1 0  1)/[0 1 0] and (1 0 1)/[0 1 0] over the (1 0 0)/[0 1 0] system. Finally, these key results from the variant approach hold for other plate shapes, such as an ellipsoid-shaped inclusion and a cuboid-shaped inclusion with hemispherical ends. 4. Discussion
Fig. 8. Comparison of the crystallographic theory of martensitic transformations for NiTi with TEM observations from the preliminary micropillar testing. Shown are the predictions for the T4 class in Table 1. The line labeled invariant plane shows the predicted intersection of the invariant plane with the [0 0 1] plane of the image. The line labeled twin interface shows the intersection of the predicted martensite martensite interface with the [0 0 1] plane of the image. The compression axis is approximately h= 4 from [1 1 0].

4.1. Comparison with previous TEM studies To our knowledge, the present results are the rst detailed analysis of the specic type and arrangement of dislocations associated with pseudoelastic response in NiTi alloys. This is signicant since dislocation activity and retained martensite are associated with hysteresis loss, a

Fig. 9. Spatial distribution of (i) the most stressed slip system and (ii) corresponding resolved shear stress (in MPa) on three planes (a) x1 = 1.0125a; (b) x2 = 1.0125b and (c) x3 = 1.05c that are just outside the faces of a martensite plate (T4, k = 8, m = 5 in Table 1) as shown in Fig. 10. The numbers in the legend correspond to the slip systems in Table 2.

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

3559

Fig. 10. Spatial distribution of slip systems with a resolved shear stress exceeding 1500 MPa for a stress axis of [7 8 0], on planes (a) x1 = 1.0125a, (b) x2 = 1.0125b and (c) x3 = 1.05c located just outside the faces of a martensite plate. These calculations assume aT4 type plate with (k, m) = (8, 5), as identied in Table 1. Slip system numbers correspond to those listed in Table 2.

decrease in critical transformation stress and an increase in remnant strain [3034]. Previous work [10,32,33] has identied ah1 0 0i dislocations and dislocation activity [11], but without a detailed consideration of the stress generated by martensitic variants and the applied stress. Chumlyakov et al. [33] showed that dislocations within the B2 phase of Ti50.8 at.% Ni0.3 at.% Mo are ah1 0 0i type on {0 0 1} and {0 1 1} planes. Although complete dislocation analyses have not been performed, Hurley et al. [10] and Hamilton et al. [11] have shown dislocation structures with a similar periodicity to the present work, in cyclically deformed NiTi single crystals and thermally cycled, solutionized Ti50.1 at.% Ni, respectively. Coupling between plasticity and transformations has been proposed by Yawny et al. [34], but quantifying this in polycrystals, and even bulk single crystals, is dicult since multiple plates are usually activated. The present work demonstrates that tests of micronscale volumes can overcome these inherent diculties. The plate level modeling identies the likely transformation mode, taking into account the slight misorientation from the [1 1 0] compression axis. From this, the work to induce the B2-to-B190 transformation in an isolated plate at room temperature is $34 MJ m3, based on the product re*, where r $ 600 MPa (Fig. 4) is the compressive transformation stress along [7 8 0] and e* $ 5.7% is the axial strain along [7 8 0] induced by either the (k, m) = (5, 8) or (8, 5) T4 class of plates in Table 1. The variant level modeling (Figs. 9 and 10) identies that the most stressed slip systems near favored martensite plates are (1 0  1)/[0 1 0] and (1 0 1)/[0 1 0], consistent with post-

mortem STEM observations in one- and two-cycle pillars (Figs. 47). Thus, the variant modeling supports the hypothesis that the dislocation loop arrays observed in the STEM studies are driven into the austenite by the local stress eld of these transforming martensite variants. It is not clear what stabilizes these loop structures in spite of the large line tension forces that act to collapse them. The detailed formation of these loop arrays during/after the transformation is likely to require in situ observation techniques. 4.2. Comparison with previous single-crystal and micropillar testing Frick et al. [21,22] reported a loss of pseudoelasticity as the diameter of aged and solutionized single-crystal NiTi pillars is reduced from $1 lm; the loss is especially pronounced in the diameter range $400200 nm. Further, the transitions are not sharp as observed in the present work for [1 1 0]-oriented 5 and 20 lm diameter samples. There are possible sources for these dierences. First, the specimens in Frick et al.s work have rounded top surfaces, nonuniform crosssectional areas due to tapering of sample walls, and are not aligned with a goniometer as in the present study. These issues become more pronounced for submicron samples, inducing localized deformation at the top, bending, deviations from uniaxial loading, and a large, nonlinear toe-in region at the beginning of a test. Second, the compression axes in each study are dierent. However, the loss in pseudoelastic behavior is observed for several orientations, ranging from [1 1 1] compression which favors plasticity, to [1 0 0] which is intermediate, to [2 1 0] which favors transforma-

3560

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561

tion. The [1 1 0] compression axis in the present work is intermediate. Overall, a comparison of the literature and present work suggests that in solutionized form, the plateau value of forward transformation stress may not depend sensitively on sample diameter, at least down to $2 lm. In particular, Frick et al. [22] reported $600 MPa for a $2 lm solutionized sample. This is comparable to the $600 MPa value for the 5 and 20 lm [1 1 0]-oriented pillars in the present work. The bulk [1 1 0] solutionized value is $60 MPa smaller, when adjusted for composition dierences (Section 3.1). The martensite Schmid factor (MSF) % 0.43 in all of these cases [20]. However, Frick et al. did not report a NiTi composition, making exact comparison with their work dicult. Despite the similar plateau values, the present work displays the sharpest transitions in stressstrain response, suggesting that micron-scale samples, when carefully machined and aligned, can be used to study the abrupt onset of the transformation. 5. Conclusions Microcrystal compression specimens (5 and 20 lm) were fabricated from a large [1 1 0] grain in a solutionized 50.7 at.%NiTi polycrystalline sample. Utilizing STEM techniques, the substructure evolution was studied as a function of loading cycle. The observations are:  At the sizes explored, there is no evidence to suggest a complete loss in pseudoelastic behavior, as observed by Frick et al. [21,22] for compression microcrystals in the 200400 nm range.  The compression results indicate a sharp elastic/transformation transition, suggesting that little plasticity occurs prior to the onset of the martensitic transformation.  Type a[0  1 0]/(1 0  1) dislocations are observed both inside and at the periphery of the transformation zone. Those within the transformation zone are arranged as arrays of dislocation loops, while those at the periphery have a distinctive arrowhead morphology.  Four twinning modes that probably produced the remnant bands of dislocations are identied from 192 possible modes, based on a plate-level micromechanics model that employs the phenomenological theory of martensite transformations.  The twin interface predicted for these four twinning modes aligns with the observed planar arrays of dislocation loops, suggesting that loop formation may be driven by local stress elds at the scale of individual variants.  A variant-level micromechanics model reveals that the observed (1 0  1)/[0 1 0] slip system is highly stressed by the four likely twinning modes, in B2 regions that are just outside the martensite plate and parallel to twin interfaces. Further consideration of the small misorien-

tation (4) from the [1 1 0] deformation axis suggests that the most highly stressed martensitic plate system is activated. An estimate of the stress work for activation at room temperature is re* $ 34 MJ m3.  Application of the crystallographic theory of martensite transformations to experimental STEM images oers a powerful method to study the coupled interaction between transformation and plasticity. This preliminary theoryexperiment approach suggests that ne-scale plasticity is produced on a scale and orientation commensurate with individual martensite phases.

Acknowledgments D.M.N., P.M.S. and M.J.M. acknowledge the support of the AFOSR Science and Technology Workforce for the 21st Century program. P.M.A. and S.M. acknowledge support from the Ohio Supercomputer Center (Grant PAS0676-5) for computing resources in the micromechanics modeling. References
[1] Schmassmann A, Meyenberger C, Knuchel J, Binek J, Lammer F, Kleiner B, et al. Am J Gastr 1997;92:400. [2] Wagner M, Richter J, Frenzel J, Gro nemeyer D, Eggeler G. Materialwissenschaft und Werkstotechnik 2004;35:320. [3] Krone L, Mentz J, Bram M, Buchkremer HP, Sto ver D, Wagner M, et al. Adv Eng Mater 2005;7:613. [4] Otsuka K, Ren X. Prog Mater Sci 2005;50:511. [5] Weschler MS, Liberman DS, Read TA. Trans AIME 1953;197:1503. [6] Bowles JS, MacKenzie JK. Acta Metall 1954;2:12938. [7] Wayman CM. Introduction of the crystallography of martensitic transformations. New York: Macmillan; 1964. [8] Bhattacharya K. Microstructure of martensite: why it forms and how it gives rise to the shape-memory eect. Oxford: Oxford University Press; 2003. [9] Hane KF, Shield TW. Acta Mater 1999;47:2603. [10] Hurley J, Ortega AM, Lechniak J, Gall K, Maier HJ. Z Metallkde 2003;94:547. [11] Hamilton RF, Sehitoglu H, Chumlyakov Y, Maier HJ. Acta Mater 2004;52:3383. [12] Eggeler G, Hornbogen E, Yawny A, Heckmann A, Wagner M. Mater Sci Eng 2004;A378:24. [13] Wagner M, Nayan N, Ramamurty U. J Physics 2008;D 41:18540814. [14] Uchic MD, Dimiduk DM, Florando JN, Nix WD. Science 2004;305:986. [15] Dimiduk DM, Uchic MD, Parthasarathy TA. Acta Mater 2005;53:4065. [16] Greer JR, Oliver WC, Nix WD. Acta Mater 2005;53:1821. [17] Greer JR, Nix WD. Appl Phys A 2005;A80:1625. [18] Noreet DM, Uchic MD, Dimiduk DM, Mills MJ. Acta Mater 2008;56:2988. [19] Maher DM, Joy DC. Ultramicroscopy 1976;1:239. [20] Gall K, Dunn ML, Liu Y, Labossiere P, Sehitoglu H, Chumlyakov YI. J Eng Mater Technol (Trans ASME) 2002;124:238. [21] Frick CP, Orso S, Artz E. Acta Mater 2007;55:3845. [22] Frick CP, Clark BG, Orso S, Ribic PS, Arzt E. Scripta Mater 2008;59:79. [23] Mura T. Micromechanics of defects in solids. Berlin: Springer Verlag; 1982. [24] Wagner M, Windl W. Acta Mater 2008;56:6232.

D.M. Noreet et al. / Acta Materialia 57 (2009) 35493561 [25] [26] [27] [28] [29] Wagner M, Windl W. Scripta Mater 2009;60:207. Eshelby JD. Proc Roy Soc (London) 1957;241:376. Chiu Y. J Appl Mech Trans ASME 1977;44:587. http://www.tw.org. Brill TM, Mittelbach S, Assmus W, Mu llner M, Lu thi B. J Phys Condens Matter 1991;3:9621. [30] Melton KN, Mercier O. Acta Metall 1979;27:137.

3561

[31] Gall K, Sehitoglu H, Chumlyakov Y, Kireeva IV. Scripta Mater 1999;40:7. [32] Matsumoto O, Miyazaki S, Otsuka K, Tamura H. Acta Mater 1987;35:2137. [33] Chumlyakov YI, Surikova NS, Korotayev AD. Fiz Met Metalloved 1996;82:148. [34] Yawny M, Sade M, Eggeler G. Z Metallkde 2005;96:608.

Vous aimerez peut-être aussi