Grain size dependence of fracture toughness and crack-growth resistance

5
Grain size dependence of fracture toughness and crack-growth resistance of superelastic NiTi Aslan Ahadi a,b , Qingping Sun a, a Department of Mechanical and Aerospace Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Hong Kong b School of Civil Engineering, State Key Lab of Water Resources and Hydropower Engineering, Wuhan University, China abstract article info Article history: Received 12 August 2015 Received in revised form 10 October 2015 Accepted 22 October 2015 Available online xxxx Keywords: Shape memory alloys Martensitic phase transformation Microstructure Toughness Intrinsic and extrinsic toughening mechanisms The grain size (GS) dependence of fracture toughness (K IC ) and static crack-growth resistance (K R ) of superelastic NiTi with average GS from 10 to 1500 nm are investigated. The measurements of strain and temperature elds at the crack-tip region are synchronized with the forcedisplacement curves under mode-I crack opening tests. It is found that with GS reduction down to nanoscale, the K IC and the size of crack-tip phase transformation zone monotonically decrease and the K R changes from a rising to a at R-curve. The roles of intrinsic and extrinsic toughening mechanisms in the GS dependence of the fracture process are discussed. © 2016 Elsevier Ltd. All rights reserved. Due to a reversible martensitic phase transformation (PT), superelastic (SE) NiTi shape memory alloys (SMAs) are widely used in a variety of applications. However, the narrow temperature window of superelasticity [1], poor cyclic stability [2], and poor fatigue life [3], signicantly limit the applications of SE NiTi. In recent years, signicant steps have been taken to break the above limitations. Introducing de- fects via doping [4,5], addition of Nb nanowires to NiTi [6], and GS re- duction down to nanocrystalline (nc) regime [710] are unique ways to achieve novel properties. It has been shown that with GS reduction to nc regime, SE NiTi indeed exhibits novel properties such as linear superelasticity with vanishing hysteresis [11], broadening of the tem- perature window of superelasticity down to 60 °C [7], and improved thermomechanical cyclic stability [8]. The fracture toughness (K IC ) and crack-growth resistance (K R ) are essential properties for successful application of metals and alloys [12]. It is well-known that with GS reduction to nc regime, metals exhib- it lower K IC due to limited ductility [13]. Owing to a stress-induced mar- tensitic PT, the most distinctive feature in the fracture process of SMAs is the formation of a kidney-like crack tip PT zone which can transform back at the crack wake [14]. In this regard, several experimental, analyt- ical, and numerical researches have been published [12,1520]. Howev- er, the GS effects on the fracture behavior and intrinsic/extrinsic toughening mechanisms of SE NiTi remain to be unveiled by systematic experiments. In this paper, we have studied the K IC and K R behaviors of SE NiTi with average GS from 10 to 1500 nm. The strain and tempera- ture elds at the crack-tip region and the crack wake are synchronized with the forcedisplacement curves by in-situ digital image correlation (DIC) and in-situ infrared (IR) thermography under the mode-I crack opening tests. SE NiTi (Ti-50.9 at.% Ni) sheets with thickness of 1.7 mm were sandwiched in stainless steel sheets and were cold-rolled to 42% of thickness reduction. The as-rolled sheets were heat-treated at tempera- tures of 250 °C for 45 min, 520 °C for 2 min, 520 °C for 3 min, 520 °C for 6 min, and 600 °C for 45 min followed by quenching in water [7,8,11]. The average GS of the cold-rolled and heat-treated specimens was mea- sured with TEM (see Fig. S1) and WilliamsonHall method. Dog-bone tensile specimens with thickness of 1 mm, width of 2 mm, and gauge length of 30 mm were cut along the rolling direction and tested to ob- tain stressstrain curves (see Fig. S2b) at a strain rate of 1 × 10 4 s 1 using an MTS (UTM-RT/10) machine. Transformation temperatures were determined using differential scanning calorimeter (DSC-TA Q1000) with heating/cooling rate of 10 K/min. Notched compact tension (CT) specimens with the dimension shown in Fig. S3a were EDM cut from the cold-rolled and heat-treated sheets. The highly polished CT specimens were pre-cracked (see Fig. S3b) with a MTS-858 table top machine at a frequency of 5 Hz with sinusoidal wave function under decreasing load amplitudes (ΔP n b b ΔP 1 ) as suggested by ASTM E 561-10 (ΔP= P max P min ). The load ratio of R = P min /P max = 0.1 was used (see inset in Fig. S3c). Different lengths of pre-cracks with 7.6 b a b 10.29 mm (0.4470 b a/W b 0.6053 mm) were obtained. The crack length was monitored ex-situ using a Leica mi- croscope. The strain and temperature elds at the vicinity of the crack tip Scripta Materialia 113 (2016) 171175 Corresponding author. E-mail address: [email protected] (Q. Sun). http://dx.doi.org/10.1016/j.scriptamat.2015.10.036 1359-6462/© 2016 Elsevier Ltd. All rights reserved. Contents lists available at ScienceDirect Scripta Materialia journal homepage: www.elsevier.com/locate/scriptamat

Transcript of Grain size dependence of fracture toughness and crack-growth resistance

Page 1: Grain size dependence of fracture toughness and crack-growth resistance

Scripta Materialia 113 (2016) 171–175

Contents lists available at ScienceDirect

Scripta Materialia

j ourna l homepage: www.e lsev ie r .com/ locate /scr ip tamat

Grain size dependence of fracture toughness and crack-growth resistanceof superelastic NiTi

Aslan Ahadi a,b, Qingping Sun a,⁎a Department of Mechanical and Aerospace Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Hong Kongb School of Civil Engineering, State Key Lab of Water Resources and Hydropower Engineering, Wuhan University, China

⁎ Corresponding author.E-mail address: [email protected] (Q. Sun).

http://dx.doi.org/10.1016/j.scriptamat.2015.10.0361359-6462/© 2016 Elsevier Ltd. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 12 August 2015Received in revised form 10 October 2015Accepted 22 October 2015Available online xxxx

Keywords:Shape memory alloysMartensitic phase transformationMicrostructureToughnessIntrinsic and extrinsic toughening mechanisms

The grain size (GS) dependence of fracture toughness (KIC) and static crack-growth resistance (KR) of superelasticNiTi with average GS from 10 to 1500 nmare investigated. Themeasurements of strain and temperature fields atthe crack-tip region are synchronizedwith the force–displacement curves under mode-I crack opening tests. It isfound that with GS reduction down to nanoscale, the KIC and the size of crack-tip phase transformation zonemonotonically decrease and the KR changes from a rising to a flat R-curve. The roles of intrinsic and extrinsictoughening mechanisms in the GS dependence of the fracture process are discussed.

© 2016 Elsevier Ltd. All rights reserved.

Due to a reversible martensitic phase transformation (PT),superelastic (SE) NiTi shape memory alloys (SMAs) are widely used ina variety of applications. However, the narrow temperature windowof superelasticity [1], poor cyclic stability [2], and poor fatigue life [3],significantly limit the applications of SE NiTi. In recent years, significantsteps have been taken to break the above limitations. Introducing de-fects via doping [4,5], addition of Nb nanowires to NiTi [6], and GS re-duction down to nanocrystalline (nc) regime [7–10] are unique waysto achieve novel properties. It has been shown that with GS reductionto nc regime, SE NiTi indeed exhibits novel properties such as linearsuperelasticity with vanishing hysteresis [11], broadening of the tem-perature window of superelasticity down to −60 °C [7], and improvedthermomechanical cyclic stability [8].

The fracture toughness (KIC) and crack-growth resistance (KR) areessential properties for successful application of metals and alloys[12]. It iswell-known thatwithGS reduction to nc regime,metals exhib-it lower KIC due to limited ductility [13]. Owing to a stress-inducedmar-tensitic PT, themost distinctive feature in the fracture process of SMAs isthe formation of a kidney-like crack tip PT zone which can transformback at the crackwake [14]. In this regard, several experimental, analyt-ical, and numerical researches have been published [12,15–20]. Howev-er, the GS effects on the fracture behavior and intrinsic/extrinsictougheningmechanisms of SE NiTi remain to be unveiled by systematicexperiments. In this paper, we have studied the KIC and KR behaviors of

SE NiTi with average GS from 10 to 1500 nm. The strain and tempera-ture fields at the crack-tip region and the crack wake are synchronizedwith the force–displacement curves by in-situ digital image correlation(DIC) and in-situ infrared (IR) thermography under the mode-I crackopening tests.

SE NiTi (Ti-50.9 at.% Ni) sheets with thickness of 1.7 mm weresandwiched in stainless steel sheets and were cold-rolled to 42% ofthickness reduction. The as-rolled sheets were heat-treated at tempera-tures of 250 °C for 45 min, 520 °C for 2 min, 520 °C for 3 min, 520 °C for6 min, and 600 °C for 45 min followed by quenching in water [7,8,11].The average GS of the cold-rolled and heat-treated specimenswasmea-sured with TEM (see Fig. S1) and Williamson–Hall method. Dog-bonetensile specimens with thickness of 1 mm, width of 2 mm, and gaugelength of 30 mmwere cut along the rolling direction and tested to ob-tain stress–strain curves (see Fig. S2b) at a strain rate of 1 × 10−4 s−1

using an MTS (UTM-RT/10) machine. Transformation temperatureswere determined using differential scanning calorimeter (DSC-TAQ1000) with heating/cooling rate of 10 K/min.

Notched compact tension (CT) specimenswith thedimension shownin Fig. S3a were EDM cut from the cold-rolled and heat-treated sheets.The highly polished CT specimens were pre-cracked (see Fig. S3b)with aMTS-858 table topmachine at a frequency of 5 Hzwith sinusoidalwave function under decreasing load amplitudes (ΔPn b … b ΔP1) assuggested by ASTM E 561-10 (ΔP = Pmax − Pmin). The load ratio ofR = Pmin/Pmax = 0.1 was used (see inset in Fig. S3c). Different lengthsof pre-cracks with 7.6 b a b 10.29 mm (0.4470 b a/W b 0.6053 mm)were obtained. The crack lengthwasmonitored ex-situ using a Leicami-croscope. The strain and temperaturefields at the vicinity of the crack tip

Page 2: Grain size dependence of fracture toughness and crack-growth resistance

172 A. Ahadi, Q. Sun / Scripta Materialia 113 (2016) 171–175

were synchronized using in-situ DIC and IR thermography. The experi-mental setup for mode-I crack opening tests is shown in Fig. S3c. To cre-ate a fine and random pattern needed for strain field measurementsusing DIC, silicon micro-particles (with average size of 1 μm) werespeckled to the specimen surface (Fig. S4). Ncorr open source 2D DICwas used for DIC calculations [21]. A fast multi-detector IR thermal cam-era (FLIR SC7700BB) equipped with a close-up lens with working dis-tance of 30 cm and field of view of 9.6 × 7.2 mm2 was used to recordthe temperature field at the crack-tip region. To achieve a uniformblack body, one side of the specimens' surface was covered with candlefume. The mode-I crack opening tests were performed at the loadingrate of 5 mm/min in order to capture the temperature variations at thecrack-tip region. The fractography studies were performed in a JEOL-JSM 6390 SEM.

The transformation temperatures measured from DSC (see Fig. S2a)are summarized in Table 1. It is seen that all the transformation temper-atures Af (austenite finish temperature), Ms (martensite start tempera-ture), and Mf (martensite finish temperature) decrease with GSreduction. The effects of GS on the isothermal stress–strain tensile prop-erties are shown in supplementarymaterials (Fig. S2b). It is seen that allthe specimens are SE as they can recover large strains of 5%.With GS re-duction to nc regime (GS b 60 nm) the stress–strain response changesfrom a plateau-type superelasticity to hardening SE response withmuch reduced hysteresis loop area [9]. The critical transformation stress(σtr), determined from the cross-cutting of the tangent lines in the elas-tic and PT regions (see Fig. S2b), increases from 381 MPa for GS =1500 nm to 655 MPa for GS = 10 nm. The gradual increase of σtr withGS reduction is due to the fact that the reduction of GS makes the PTmore difficult [11,22]. The plastic yield's stress of martensite (σY

M), de-termined from the cross-cutting of the tangent lines in the martensiteelastic deformation and plastic deformation regions, also increaseswith GS reduction since dislocation activities are significantly sup-pressed at nano-scale. Moreover, the strain to failure (εf) decreaseswith GS reduction to nc indicating a gradual reduction in the ductility.

Fig. 1a shows the force–displacement curves of the CT specimenswith different average GS (a/W=0.4501). It is seen that for all the spec-imens, after an initial linear deformation (up to ~415 N), the force–displacement curve deviates from linearity due to stress-inducedmartensitic PT at the crack-tip region [23]. With further increase ofthe displacement the force first reaches a maximum Pmax and thendecreases either gradually (for GS = 1500, 80, 18, and 10 nm) due torelative stable crack growth or rapidly (for GS = 64 and 42 nm) dueto fast (burst-like) crack growth. It is evident that the amount ofdeviation from linearity and the value of Pmax decrease with GSreduction. This GS dependencies of the fracture behavior forms a starkcontrast with the uniform tensile stress–strain curves where specimenswith smaller GS can sustain larger loads. According to the linear elasticfracture mechanics (LEFM), the critical stress intensity factor (KIC) formode-I loading can be calculated as:

KIC ¼ Pmax

BffiffiffiffiffiffiW

p � 2þ a=Wð Þ1−a=Wð Þ3=2

� f a=Wð Þ ð1Þ

Table 1The GS dependence of phase transformation properties of SE NiTi.

Heat treatment Average GS (nm) As/Af (°C) Ms/Mf (°C)

Cold-rolled 10 N/A N/A250 °C — 45 min 18 N/A N/A520 °C — 2 min 42 −12.7/8 −60/−86520 °C — 3 min 64 −12.5/6 −55/−78520 °C — 6 min 80 −4.8/10 −50/−70600 °C — 45 min 1500 −4.6/3 −30/−33

where Pmax is the applied maximum force, B is the specimen thick-ness, W is the specimen width measured from the load line, a is thephysical crack length (see Fig. S3a), and f(a/W) can be expressed as(see ASTM E 561-08):

f a=Wð Þ ¼ 0:886þ 4:64 a=Wð Þ−13:32 a=Wð Þ2 þ 14:72 a=Wð Þ3−5:6 a=Wð Þ4h i

:

ð2Þ

Whenusing Eq. (1) onehas tonote that this equation is only valid forplane strain condition where the size of the PT zone is much smallerthan the specimen thickness. Looking at the size of PT zone in Fig. 2, itis seen that this condition is barely met in our experiments. Therefore,the values of KIC and KR reported here are nominal and only serve asfirst approximations and are used for comparative purpose. Moreexact calculations such as using finite element method is needed to de-termine the K values. For each GS, themeasured KIC averaged over eightdifferent values of a/W are shown in Fig. 1b. For the GS = 1500 and80 nm, the KIC values are 46.3 MPa

ffiffiffiffiffim

pand 42.4 MPa

ffiffiffiffiffim

p, respectively.

These values are consistent with the reported values for SE NiTi of thesimilar microstructure and stress–strain curves [16,23]. It is seen thatwhen GS is reduced below 80 nm, the value of KIC monotonically de-creases and reaches 25.4 MPa

ffiffiffiffiffim

pfor GS = 10 nm. Fig. 1c shows the

GS dependence of crack-growth resistance KR (R-curve) as a functionof crack extension (a− a0). For GS= 1500 and 80 nm, the KR keeps in-creasing (known as rising R-curve) with crack extension and eventuallysaturates. However, for GS=18 and 10nm, the R-curves becomeflat. Inother words, more energy is dissipated during crack propagation for SENiTi with GS = 1500 and 80 nm while for the GS = 18 and 10 nm thefracture behavior resembles those of brittle materials.

SEM images of the fracture surfaces of the CT specimens are shown inFig. 1d. The fracture surface of the specimen with GS= 1500 nmmainlyshows dimples indicating a dominant ductile fracture. For GS = 80 nmthe fracture surface is characterized by amixture of dimples and cleavageplanes.With further GS reduction, the tendency of the cleavage increasesin the fracture process as it is shown for GS = 42 and 10 nm. Therefore,based on the consistent observations of (1) stress–strain behavior(2) crack-tip inelastic zone size (3) KR curves and (4) the SEM observa-tions, it is suggested that the failuremode transitions fromductile to brit-tle fracture when GS is reduced down to nanoscale.

The effects ofGSon the strain (εyy) distribution around the crack-tip re-gion taken at Pmax are shown in Fig. 2. It is seen that the strain field at thecrack-tip vicinity is in the shapeof two lobes. For theGS=1500and80nmthe lobes are inclined to an angle of about 48 and 51° from the crack line,respectively. The angle of inclination gradually increases to 63° for GS =10 nm. Moreover, the DIC images indicate that with GS reduction thesize of the lobe at Pmax immediately prior to the crack growth decreases.From the plot of εyy as a function of the distance from the crack tip,as shown in Fig. 2, one can determine the size of the PT zone along thex-axis rA → M by referring to the stress–strain curves [16,32]. For GS =1500 nm the 0.25 ≤ rA→ M ≤ 1.7mmat the crack tip region andwith GS re-duction the rA → Mmonotonically decreases to 0.23 ≤ rA → M ≤ 0.81 mm forGS= 80 nm and 0.19 ≤ rA → M ≤ 0.48 mm for GS= 64 nm. The measured

σtr (MPa) σYM (MPa) EAust (GPa) H5% (MPa)

655 N/A 50.5 1.05502 1465 45.8 3.4472 1452 49.8 6.9461 1457 46.3 9.51425 1411 47.4 11.25381 527 45 13.11

Page 3: Grain size dependence of fracture toughness and crack-growth resistance

Fig. 1. (a) GS effects on the force–displacement curves duringmode-Imonotonic crack opening fracture tests, (b) GS effects on the K=KIC (K at Pmax) for specimens of different initial cracklengths (with 0.4470 b a/W b 0.6053 mm), (c) GS effects on the crack-growth resistance (KR), and (d) top-view SEM images of the fracture surface of CT specimens.

173A. Ahadi, Q. Sun / Scripta Materialia 113 (2016) 171–175

values of rA→ M are consistentwith the published data on nano-grained SENiTi [16]. The GS dependence of rA → M and stress distribution can be read-ily explainedby the theoreticalmodels of stress-induced PT at the crack tipregion [18,24].When the size of PT zone is small comparedwith the size ofthe specimen (small scale PT), the size and the shape of PT zone for eachGS can be approximated as [25]:

rA→M θð Þ ¼ 12π

Kapp

σ tr

� �2cos2

θ2

1þ 3 sin2 θ2

� �ð3Þ

where σtr is the critical transformation stress and depends on the GS andKapp is the applied stress intensity factor at the crack tip. For a stationarycrack = KIC. According to this equation, rA → M scales with (Kapp/σtr)2.Therefore, for a given Kapp a gradual reduction of rA → M with GS reductionis expected since σtr increases significantly with GS reduction. Another ef-fect of GS reduction is that for a given Kapp the stress levels in the PT zoneand plasticity zone are higher due to higher σtr and σY

M as schematicallyshown in Fig. S5. A direct consequence of such higher stress levels at thecrack tip region is that the fracture behaves more like brittle materialsand has features of brittle fracture such as a flat R-curve.

The synchronized IR thermographic images of the front surface of thespecimens at Pmax (loading rate of 5 mm/min) are shown in Fig. 3a. It is

Fig. 2. GS effects on the variation of εyy with the distance (x) from the crack tip (in the xdirection). The inset shows the DIC images taken at Pmax with a field of view of6.34 × 4.75 mm2. A reduction of PT zone size with GS reduction is observed.

seen that the crack-tip region is warmer than the rest of the specimendue to release of PT latent heat and heat transfer via conduction and con-vection [2,26,27]. The figure shows two important features. First, with GSreduction the size of the hot zone decreases which is consistent with theDIC measurements of rA → M (Fig. 2a). Second, under the given loadingrate of 5 mm/min, the magnitude of the temperature increase in thePT zone decreases with GS. For example, the magnitude of the temper-ature rise at the crack tip is 5.83 °C for GS = 1500 nm and it monoton-ically decreases to 1.02 °C for GS=18nm, and 0.03 °C for a GS=10nm.This is due to the reduction of latent heat with GS especially for GS =18 nm (see Fig. S2b) [7–9]. Fig. 3b shows the thermographic imagesduring crack propagation. Comparing Fig. 3b with 3a, one can note a re-markable increase in the size of the crack-tip hot zone for GS = 1500and 80 nm. This is consistent with the rising R-curve characteristic ofthe two specimens. Furthermore, it is clearly seen that once the crackpropagates it leaves two cold regions behind with temperatures lowerthan the room temperature (black arrows). For the GS = 1500 nm theminimum temperature in the crack wake is 15.5 °C which is 4.5 °Cbelow the room temperature. This cooling effect is due to a rapidabsorption of latent heat and provides a compelling evidence to the re-verse PT when the crack passed by [26]. For the same reasons, thecooling in the wake becomes also weaker with GS reduction.

To elucidate the effects of GS on the controlling mechanisms of KIC

and KR, a qualitative discussion of the intrinsic and extrinsic tougheningmechanisms at the crack-tip region and crack wake [28] are given inFig. 4a and b, respectively. For the intrinsic mechanisms, it is well-known that the most dominant toughening mechanism in SE NiTi isthe formation of stress-induced martensite at the crack-tip regionwhich act as a shielding mechanism. According to Baxevanis et al. [12,15], the degree of the toughening due to the crack-tip PT dependsmain-ly on the dimensionless parameter α= σtrεtr/[σtr

2/EA] (ratio of dissipat-ed to stored energy) where σtr is the transformation stress, εtr is thetransformation strain, and EA is the Young's modulus of austenite. Thehigher value of α is beneficial for toughness enhancement [25]. Assuch, the gradual decrease of α (due to a gradual increase of σtr) is in-deed the main reason for a gradual reduction of KIC [29]. Another factorthat needs to be considered is the GS dependence of the plastic yield'sstress of themartensite (σY

M). As can be seen in Table 1, theσYM increases

with GS reduction to nc. In the same way as that of the elastic–plastic

Page 4: Grain size dependence of fracture toughness and crack-growth resistance

Fig. 3. (a) IR thermographic images showing the GS effects on the temperature field near the crack-tip region at K= KIC, (b) GS effects on the temperaturefield in the crack-growth regime(a = a0 + Δa) showing heating/cooling due to forward/reverse PT in the crack-tip/wake.

174 A. Ahadi, Q. Sun / Scripta Materialia 113 (2016) 171–175

materials, a higher yield's stress indicates a smaller size of the plastic de-formation zone and higher stress levels (see Fig. S5) at the crack tipleading to less shielding effect and toughness enhancement. However,it is important to note that the effect of plastic deformation zone atthe crack tip on the shielding and dissipation is believed to be smallcompared with the dissipation due to PT [12]. In addition, with GS re-duction, the tortuous crack-path configuration (see Fig. 4a1) becomesstraight (see Fig. 4a2). This causes a change in the mode of K from mixof mode-I and mode-II in the former, to near mode-I in the latter. Thisleads to a decrease of the threshold (critical) driving force for crackpropagation that in turn leads to reduction of KIC [30].

The effects of GS on the extrinsic mechanisms are schematicallyshown in Fig. 4b1 (for GS ≥ 80 nm) and Fig. 4b2 (for nc NiTi). As it isseen in Fig. 4b, during crack propagation the crack wake undergoes acomplete cycle of forward and reverse PT (see also Fig. 3b for the ther-mographic evidence of the reverse PT at the crackwake). The toughnessenhancement ΔK = Ka − KIC for steady-state crack propagation due tosuch forward/reverse PT or energy dissipation can be described by thefollowing equation [31]:

K2a ¼ K2

IC þ2E

1−υ2ð Þ� �Z H

0Ud yð Þdy ð4Þ

where E is the elastic modulus, υ is the Poisson's ratio, H is the height ofthe crackwake, and Ud (= ∮ σdε) is the specificmechanical energy dis-sipated in thewake and is represented by theσ− ε hysteresis loop area.From Eq. (4) it is clearly seen that the gradual change of KR with GS

Fig. 4. Schematics of the GS effects on (a) the intrinsic toughening

reduction (from a rising R-curve to a flat R-curve) is not only causedby a reduction of H due to the increase of σtr, but also caused by a grad-ual reduction of the stress hysteresis (Ud → 0) of the material [7].

In summary, the grain size (GS) dependence of fracture toughness(KIC) and static crack-growth resistance (KR) of SE NiTi wereinvestigated by synchronized measurements of the force–displacementcurves and the crack-tip strain and temperature fields. It is shown thatwith GS reduction to nanoscale, the KIC gradually decreases and KR tran-sitions from rising to flat R-curve. The in-situ measurements showed amonotonic reduction in the size of the phase transformation zone atthe crack-tip region. Using the controlling parameters of the fractureprocess of NiTi SMAs and their grain size dependencies, the intrinsicand extrinsic toughening mechanisms of the fracture behavior of theNiTi were analyzed and discussed. The results provide a better under-standing of the fracture behavior of nanocrystalline superelastic NiTiwhich possesses exceptional thermomechanical properties.

The authors are grateful for the financial support of this work fromthe Hong Kong Research Grant Council (RGC) through the GRF grant(project no. 16214215) and the 973 Program of China (project no.2014 CB046902). We also thank Maziar Jamishidi for DIC data analysisand Hamed Mokhtari for helping with thermographic experiments.

Appendix A. Supplementary data

Supplementary data to this article can be found online at http://dx.doi.org/10.1016/j.scriptamat.2015.10.036.

mechanisms and (b) the extrinsic toughening mechanisms.

Page 5: Grain size dependence of fracture toughness and crack-growth resistance

175A. Ahadi, Q. Sun / Scripta Materialia 113 (2016) 171–175

References

[1] T. Omori, K. Ando, M. Okano, X. Xu, Y. Tanaka, I. Ohnuma, R. Kainuma, K. Ishida, Sci-ence 333 (2011) 68 (80-.).

[2] H. Yin, Y. He, Q. Sun, J. Mech. Phys. Solids 67 (2014) 100.[3] A.L. McKelvey, R.O. Ritchie, Metall. Mater. Trans. A 32 (2001) 731.[4] D. Wang, Z. Zhang, J. Zhang, Y. Zhou, Y. Wang, X. Ding, Y. Wang, X. Ren, Acta Mater.

58 (2010) 6206.[5] Y. Nii, T. Arima, H.Y. Kim, S. Miyazaki, Phys. Rev. B 82 (2010) 214104.[6] S. Hao, L. Cui, F. Guo, Y. Liu, X. Shi, D. Jiang, D.E. Brown, Y. Ren, Sci. Rep. 5 (2015)

8892.[7] A. Ahadi, Q.P. Sun, Appl. Phys. Lett. 103 (2013) 021902.[8] A. Ahadi, Q.P. Sun, Acta Mater. 76 (2014) 186.[9] A. Ahadi, Q.P. Sun, Acta Mater. 90 (2015) 272.

[10] K. Tsuchiya, M. Inuzuka, D. Tomus, A. Hosokawa, H. Nakayama, K. Morii, Y. Todaka,M. Umemoto, Mater. Sci. Eng. A 438-440 (2006) 643.

[11] Q.P. Sun, A. Ahadi, M. Li, M. Chen, Sci. China Technol. Sci. 57 (2014) 671.[12] T. Baxevanis, C.M. Landis, D.C. Lagoudas, J. Appl. Mech. 81 (2013) 041005.[13] M. Dao, L. Lu, R. Asaro, J. Dehosson, E. Ma, Acta Mater. 55 (2007) 4041.[14] A. Boulbitch, A.L. Korzhenevskii, Phys. Rev. Lett. 107 (2011) 085505.[15] T. Baxevanis, A.F. Parrinello, D.C. Lagoudas, Int. J. Plast. 50 (2013) 158.[16] S. Daly, A. Miller, G. Ravichandran, K. Bhattacharya, Acta Mater. 55 (2007) 6322.[17] M.R. Daymond, M.L. Young, J.D. Almer, D.C. Dunand, Acta Mater. 55 (2007) 3929.[18] C. Maletta, F. Furgiuele, Acta Mater. 58 (2010) 92.

[19] M.L. Young, S. Gollerthan, A. Baruj, J. Frenzel, W.W. Schmahl, G. Eggeler, Acta Mater.61 (2013) 5800.

[20] T. Baxevanis, D.C. Lagoudas, Int. J. Fract. 191 (2008) 191.[21] J. Blaber, B.S. Adair, A. Antoniou, Rev. Sci. Instrum. 86 (2015) 035111.[22] J.R. Wollants, P. De Bonte, M. Roos, Z. Met. 82 (1979) 113.[23] S. Gollerthan, M.L. Young, A. Baruj, J. Frenzel, W.W. Schmahl, G. Eggeler, Acta Mater.

57 (2009) 1015.[24] C. Maletta, F. Furgiuele, Int. J. Solids Struct. 48 (2011) 1658.[25] Y. Freed, L. Bankssills, J. Mech. Phys. Solids 55 (2007) 2157.[26] S. Gollerthan, M.L. Young, K. Neuking, U. Ramamurty, G. Eggeler, Acta Mater. 57

(2009) 5892.[27] H. Yin, Y. Yan, Y. Huo, Q.P. Sun, Mater. Lett. 109 (2013) 287.[28] R.O. Ritchie, Nat. Mater. 10 (2011) 817.[29] Q.S. Mei, L. Zhang, K. Tsuchiya, H. Gao, T. Ohmura, K. Tsuzaki, Scr. Mater. 63 (2010)

977.[30] B. Lawn, Fracture of Brittle Solids, 2nd ed. Cambridge University Press, Cambridge,

1975.[31] Q.P. Sun, X.J. Xu, Mater. Sci. 30 (1995) 439.[32] W. LePage, S. Daly, The influence of crystallographic texture on fracture of the

shape-memory alloys, in: A. Bajaj, P. Zavattieri, M. Koslowski, T. Siegmund (Eds.), Pro-ceedings of the Society of Engineering Science 51st Annual Technical Meeting, October1-3, 2014, Purdue University Libraries Scholarly Publishing Services, West Lafayette,2014.