12
Superelastic and cyclic response of NiTi SMA at various strain rates and temperatures Sia Nemat-Nasser a, * , Wei-Guo Guo b a Center of Excellence for Advanced Materials, Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, CA 92093-0416, USA b Department of Aircraft Engineering (118#), Northwestern Polytechnical University, Xi’an City, Shaanxi Province, 710072, PR China Received 28 October 2004; received in revised form 19 July 2005 Abstract To characterize the thermomechanical response, especially the superelastic behavior of NiTi shape-memory alloys (SMAs) at various temperatures and strain rates, we have performed a series of both quasi-static and dynamic uniaxial compression tests on cylindrical samples, using an Instron servohydraulic testing machine and UCSD’s enhanced Hopkin- son technique. Strain rates from 10 3 /s to about 4200/s are achieved, at initial temperatures in the range of 77–400 K. The influence of the annealing temperature on the fatigue response is also examined. A few noteworthy conclusions are as fol- lows: (1) the transformation stress and the dissipated energy of NiTi SMAs depend on the annealing temperature; (2) in cyclic loading, the dissipated energy over a cycle tends to a minimum stable value, and cyclic loading leads to a stable superelastic behavior of the alloy; (3) repeated dynamic tests of the alloy produce smaller changes in the shape of the super- elastic loop and in the dissipated energy than do the quasi-static cyclic tests; and (4) the superelastic behavior of this mate- rial has stronger sensitivity to temperature than to strain rate; at very high loading rates, NiTi SMAs show properties similar to ordinary steels, as has been established by the first author and coworkers. Ó 2005 Elsevier Ltd. All rights reserved. Keywords: NiTi shape-memory alloys; Superelasticity; Dissipated energy; Strain rate; Fatigue; Temperature 1. Introduction Depending on the temperature, in general, SMAs can exist in two different crystal structures, the mar- tensitic phase (at low temperatures) and the austen- itic phase (at high temperatures). The parent austenitic phase has a cubic lattice (B2) while the martensitic phase is monoclinic (B19 0 ) (Otsuka and Ren, 1999), consisting of only lattice twins (see Fig. 1). When a NiTi SMA in the martensitic phase is heated, it begins to change into the austen- itic phase. This phenomenon starts at a temperature denoted by A s , and is complete at a temperature denoted by A f . When an austenitic NiTi SMA is cooled, it begins to return to its martensitic struc- ture at a temperature denoted by M s , and the process is complete at a temperature denoted by M f . The difference between the transformation 0167-6636/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.mechmat.2005.07.004 * Corresponding author. Tel.: +1 858 534 4914; fax: +1 858 534 2727. E-mail address: [email protected] (S. Nemat-Nasser). Mechanics of Materials 38 (2006) 463–474 www.elsevier.com/locate/mechmat

Superelastic and cyclic response of NiTi SMA at various strain

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Superelastic and cyclic response of NiTi SMA at various strain

Mechanics of Materials 38 (2006) 463–474

www.elsevier.com/locate/mechmat

Superelastic and cyclic response of NiTi SMA at variousstrain rates and temperatures

Sia Nemat-Nasser a,*, Wei-Guo Guo b

a Center of Excellence for Advanced Materials, Department of Mechanical and Aerospace Engineering,

University of California, San Diego, La Jolla, CA 92093-0416, USAb Department of Aircraft Engineering (118#), Northwestern Polytechnical University, Xi’an City, Shaanxi Province,

710072, PR China

Received 28 October 2004; received in revised form 19 July 2005

Abstract

To characterize the thermomechanical response, especially the superelastic behavior of NiTi shape-memory alloys(SMAs) at various temperatures and strain rates, we have performed a series of both quasi-static and dynamic uniaxialcompression tests on cylindrical samples, using an Instron servohydraulic testing machine and UCSD’s enhanced Hopkin-son technique. Strain rates from 10�3/s to about 4200/s are achieved, at initial temperatures in the range of 77–400 K. Theinfluence of the annealing temperature on the fatigue response is also examined. A few noteworthy conclusions are as fol-lows: (1) the transformation stress and the dissipated energy of NiTi SMAs depend on the annealing temperature; (2) incyclic loading, the dissipated energy over a cycle tends to a minimum stable value, and cyclic loading leads to a stablesuperelastic behavior of the alloy; (3) repeated dynamic tests of the alloy produce smaller changes in the shape of the super-elastic loop and in the dissipated energy than do the quasi-static cyclic tests; and (4) the superelastic behavior of this mate-rial has stronger sensitivity to temperature than to strain rate; at very high loading rates, NiTi SMAs show propertiessimilar to ordinary steels, as has been established by the first author and coworkers.� 2005 Elsevier Ltd. All rights reserved.

Keywords: NiTi shape-memory alloys; Superelasticity; Dissipated energy; Strain rate; Fatigue; Temperature

1. Introduction

Depending on the temperature, in general, SMAscan exist in two different crystal structures, the mar-tensitic phase (at low temperatures) and the austen-itic phase (at high temperatures). The parentaustenitic phase has a cubic lattice (B2) while the

0167-6636/$ - see front matter � 2005 Elsevier Ltd. All rights reserved

doi:10.1016/j.mechmat.2005.07.004

* Corresponding author. Tel.: +1 858 534 4914; fax: +1 858 5342727.

E-mail address: [email protected] (S. Nemat-Nasser).

martensitic phase is monoclinic (B19 0) (Otsukaand Ren, 1999), consisting of only lattice twins(see Fig. 1). When a NiTi SMA in the martensiticphase is heated, it begins to change into the austen-itic phase. This phenomenon starts at a temperaturedenoted by As, and is complete at a temperaturedenoted by Af. When an austenitic NiTi SMA iscooled, it begins to return to its martensitic struc-ture at a temperature denoted by Ms, and theprocess is complete at a temperature denoted byMf. The difference between the transformation

.

Page 2: Superelastic and cyclic response of NiTi SMA at various strain

Temperature

Aus

teni

te (

%)

100

Mf

Ms

As

Af Md

HB19’

B2

Fig. 1. Phase transformation in NiTi SMAs.

0

200

400

600

800

1000

1200

1400

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07Strain

Stre

ss (

MP

a)

NiTi Specimen

A M

Loading

Unloading

σtr

σre

NiTi, 10-3/s, T0 = 296K

A E D

C

B

I H GF J

A M

Fig. 2. Superelastic effect of NiTi SMA after 42% cold workfollowed by 30 min annealing 823 K.

464 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

temperatures in heating and cooling is called thehysteresis temperature. In practice, the hysteresistemperature is generally defined as the differencebetween the temperatures at which the material is50% transformed to the austenite upon heatingand 50% transformed back to the martensite uponcooling; this is denoted by H in Fig. 1.

The chemical composition and the metallurgicaltreatment have a significant effect on the transfor-mation temperatures. When a NiTi SMA is stressedat a temperature close to Af, it can display super-elastic (or pseudoelasticity) behavior which is oftendefined in terms of the ability of the material toreturn to its original shape upon unloading, aftera substantial deformation. This stems from thestress-induced martensite formation, since stresscan produce the martensitic phase at a temperaturehigher than Ms, where macroscopic deformation isaccommodated by the formation of martensites.When the applied stress is released, the martensiticphase transforms back into the austenitic phaseand the specimen returns back to its original shape.Fig. 2 represents a typical experimental result. Inthis figure, a superelastic NiTi SMA is strained upto 6%, which is several times greater than the elasticlimit of ordinary metal alloys, and then unloaded,showing no permanent deformations. This is, how-ever, only observed over a specific temperatureand strain range. The heavy black bar in Fig. 1 sug-gests the temperature range on which superelasticbehavior can occur. The highest temperature atwhich martensitic transformation can no longer beinduced by an applied stress, is denoted by Md.Above the Md temperature, NiTi SMAs behave asordinary metals. Below the As temperature, theydeform while in the martensitic phase, and usuallydo not recover their original shape upon unloading.

However, if they are then heated to a certain tem-perature, the NiTi SMAs can recover and attaintheir original shape. The stress-induced superelasticbehavior appears in a temperature range from nearAf up to Md, as is suggested in Fig. 1. When a NiTiSMA is tested at a temperature below Af or aboveMd, the measured surface temperature distributionof the specimen is uniform, indicating the homoge-neity of the process of phase transformation. Defor-mation at a temperature T, with Af < T < Md, leadsto a heterogeneous surface temperature, pointingout that the development of martensitic transforma-tion is non-homogeneous (Gadaj et al., 2002) in thistemperature range.

2. Superelastic behavior of NiTi SMA

A sample of NiTi that has been annealedat 823 K for 30 min is deformed in uniaxial com-pression at a strain rate of 10�3/s and an initialtemperature 296 K. The resulting engineering stress–engineering strain curve is displayed in Fig. 2. Thesample at room temperature, 296 K, is in an austen-itic phase. The initial AB segment of the stress–strain curve in Fig. 2 is the elastic deformation ofthe austenitic phase (parent phase). Close to pointB, microscopic martensites are generated preferen-tially within the parent phase, because of stressing.In the initial BC segment, some plastic deformationsmay accompany the martensitic phase transforma-tion. This plastic deformation is produced in orderto accommodate locally the resulting martensiticphase (Hosogi et al., 2002). Upon further deforma-tion, extensive stress-induced transformation takesplace in the BC segment, at essentially a constantstress. At point C, the maximum stress-induced

Page 3: Superelastic and cyclic response of NiTi SMA at various strain

S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474 465

phase-transformation strain is attained. The CDsegment corresponds to the elastic unloading ofthe martensitic phase. The reverse transformationstarts around point D and is complete around pointE. The reverse transformation on the DE segmentoccurs at almost a constant stress. The EF segmentis the elastic unloading of the austenites. At F, asmall permanent residual strain remains, represent-ing the plastic deformation in the martensitic phase.This residual plastic strain has actually occurred inthe BC segment. Its magnitude mainly depends onthe maximum total strain and the deformation tem-perature. Basically, the deformation at a tempera-ture close to or slightly greater than Af could leadto a small residual strain. Also, the treatment condi-tions and the chemical composition of the materialaffect this irrecoverable strain. The annealing tem-perature is found to have a great effect on the super-elastic behavior of NiTi SMAs. In particular, themaximum plateau strain (the forward stress-inducedtransformation) increases with increasing annealingtemperature, being 9% for specimens annealed at873 K (Huang and Liu, 2001). Miller and Lagoudas(2001) point out that the level of cold work and theannealing temperature affect the transformationcharacteristics of the thermally induced phasetransformation under a constant applied stress,specifically the transformation strain and transfor-mation-induced plastic strain.

In the past decade, the thermomechanicalresponse and superelastic behavior of NiTi SMAsunder different conditions have been extensivelystudied (Shaw and Kyriakides, 1995; Gall et al.,1999; Otsuka and Wayman, 1998). The dynamicresponse of NiTi SMAs has also been studied to acertain extent. Tobushi et al. (1998) carried outdynamic tensile tests, and found that, for_e P 10%=min, the stress and temperature for themartensitic transformation increase with an increasein _e in the loading process, but the reverse trans-formation stress and temperature decrease in theunloading process. The compressive stress–strainbehavior of SMAs depends on the strain rate, andunder dynamic loading conditions, it displays anopen hysteresis loop because of the resultingresidual strain which however, may disappear atroom temperature, and, after a few seconds to sev-eral hours, the recovery may be complete (Chenet al., 2001). More recently, Nemat-Nasser et al.(2005a,b) have experimentally established the strain-rate sensitivity of the transition stress for stress-induced martensite formation, and the existence of

a critical strain rate which determines the deforma-tion mechanism of austenitic SMAs. This is in con-trast to Liu et al. (1999a,b) who point out that,when the strain rate changes from 3 · 10�4 to3000/s, the characteristics of the stress–strain curvefor martensitic NiTi are insensitive to the strainrate, suggesting that the deformation mechanismdoes not change with the strain rate. Recently, aone-dimensional polycrystalline SMA bar was stud-ied numerically and experimentally using a Hopkin-son bar by Lagoudas et al. (2003). The energydissipation calculation for both detwinning of themartensite and stress-induced phase transformationshowed that the energy can be reduced by 80–90%,suggesting that SMAs can be used effectively inshock-absorbing devices.

Over the past few years, an experimental pro-gram to study the dynamic response of NiTi SMAshas been initiated by Nemat-Nasser and coworkersin an effort to understand the dynamic behavior ofthese alloys. The present work has been an integralpart of that experimental study, focusing on thesuperelastic behavior of NiTi SMAs at variousloading rates and temperatures, and heat treatmentconditions. In addition, cyclic quasi-static andrepeated dynamic loadings are used to examinethe fatigue effect on the superelastic behavior of thismaterial.

3. Dissipated energy in superelastic loop

When a sample of a NiTi SMA is subjected to acycle of deformation within its superelastic strainrange, it dissipates a certain amount of energy with-out a permanent strain, as is illustrated in Fig. 2which shows a typical loading and unloadingresponse of a sample in compression. In this figure,the area SABCG is the total energy input per unit ini-tial volume during the loading, and the area ofSFEDCG is the energy release per unit volume duringunloading. Thus, the dissipated energy, per unit ini-tial volume, is given by

Edisp ¼ SABCG � SFEDCG ¼ISloop

rde

where r is the engineering stress, e is the engineering(nominal) strain, and the integration is taken alongthe closed loop. The dissipated energy is due tophase transformation, from austenitic to martens-itic, during loading and the reverse transformationback to austenitic in unloading, resulting in a netrelease of heat energy. This energy dissipation

Page 4: Superelastic and cyclic response of NiTi SMA at various strain

0

400

800

1200

1600

2000

0.00 0.25Strain

Stre

ss (

MP

a)

296K 473K 573K

723K 823K 873K

923K

NiTi, 10-3/s, T0 = 296K Annealing Temperature (K) :Duration: 30 Minutes

Strain = 0.04

Fig. 3. Stress–strain curves at a strain rate of 10�3/s and 296 Ktemperature, for indicated annealing temperatures and 30 minduration; each test starts at zero strain.

0

200

400

600

800

1000

200 300 400 500 600 700 800 900 1000Annealing Temperature (K)

Tra

nsfo

rmat

ion

Stre

ss (

MP

a)

σtr

σre

Annealing Duration: 30 MinutesNiTi, 10-3/s, T0 = 296K

Fig. 4. Transformation stress as a function of annealingtemperature.

466 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

property of SMAs may be used in shock and vibra-tion mitigation.

In Fig. 2, rtr denotes the austenite to martensitetransformation stress, defined by the intersectionof the lines that are tangent to the initial elasticand the upper plateau of the loading stress–straincurve, and rre denotes the reverse martensite to aus-tenite transformation stress.

4. Experiments

4.1. Material and specimens

The NiTi SMA material is obtained from NitinolDevices & Components (NDC). The as-receivedwires are 4.75 mm in diameter, with initial conditionof 42% cold work. Transformation temperature, Af,of this NiTi material is about �10 �C when com-pletely annealed according to the NDC suggestion.The chemical composition of the material is givenin Table 1.

All samples are cut from the same NiTi alloywire. The wire is machined into cylinders of4.5 mm nominal diameter and 5 mm height. Toreduce the end friction on the samples during thedeformation, the sample ends are first polishedusing waterproof silicon carbide paper, 1200 and4000 grit, then the samples are annealed at differentdesired temperatures, and finally, the ends aregreased before the tests.

4.2. Annealing temperature effect on superelastic

behavior

Different specimens made from the above-described NiTi SMA are annealed at differenttemperatures and durations in a furnace with anaccuracy of ±2 �C. Then, the samples are com-pressed at a strain rate of 10�3/s at room tempera-ture, to a maximum strain of about 5.5%, using anInstron servohydraulic testing machine. The result-ing stress–strain curves are shown in Fig. 3 for theindicated annealing temperatures; note that eachtest starts at zero strain, but the various curves areshifted horizontally for clarity in the display.Fig. 4 displays the forward and the reverse transfor-

Table 1The chemical composition of NiTi (wt.%)

Ni O H C Co Cu Fe Ti55.9 <0.0314 <0.002 0.0027 <0.005 <0.005 0.007 Balance

mation stresses as functions of the annealing tem-perature. From this figure, it is seen that thetransformation stresses, rtr and rre, of the NiTiSMA decrease with the increasing annealing tem-perature. The dissipated energy, Edisp, is plotted inFig. 5 as a function of the annealing temperature.The dissipated energy, given by the area of thehysteresis loop, also changes with the annealingtemperature. It attains a minimum at a 600 Kannealing temperature, and a maximum at both296 K (room temperature) and 873 K. With thesetwo latter temperatures, a residual strain (RS) ofabout 0.85% is observed. If the annealing temp-erature is increased beyond 900 K, the NiTi SMAshows substantial irrecoverable strains. The stress–strain curve at an annealing temperature of 923 Kactually is similar to that of an ordinary austenite

Page 5: Superelastic and cyclic response of NiTi SMA at various strain

0

5

10

15

20

25

200 300 400 500 600 700 800 900 1000Annealing Temperature (K)

Dis

sipa

ted

Ene

rgy

Ed

(J/m

3 )

NiTi, 10-3/s, T0 = 296K, εmax = 5.4%

(R. S.) (R. S.)

Fig. 5. Dissipated energy as a function of annealing temperature.

S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474 467

steel. To obtain the results shown in Fig. 6, theannealing duration has been changed, keepingthe annealing temperature at 473 K. It is seen thatthe annealing duration does not seem to affect thesuperelastic behavior of the material at this lowtemperature.

Based on the experimental observations, we maymake the following observations:

(1) The annealing temperature has a significanteffect on the superelastic behavior of NiTiSMAs. The transformation stresses, rtr andrre, and the dissipated energy, Edisp, changewith the annealing temperature.

(2) The dissipated energy, Edisp, attains a mini-mum when the annealing temperature reachesabout 600 K; it can thus be maximized byadjusting the annealing temperature.

(3) Although the transformation stress, rtr, ofNiTi SMAs increases with the decreasing

0

200

400

600

800

1000

1200

1400

0.00 0.01 0.02 0.03 0.04 0.05 0.06Strain

Stre

ss (

MP

a)

NiTi, 10-3/s, T0 = 296K; Annealing Temperature 473K

Solid Dots : Annealing duration 5 MinutesSolidCurves : Annealing duration 30 Minutes

Fig. 6. Effect of annealing duration on superelastic behavior.

annealing temperature, it also induces agreater residual strain.

(4) When the annealing temperature is about923 K, the material actually is no longersuperelastic.

(5) The annealing duration (at a temperature of473 K) has a weak effect on the superelasticbehavior as compared with that of the anneal-ing temperature.

4.3. Irrecoverable strain in NiTi SMAs

Examining the results in Figs. 3–5, it is seen thatthe transformation stress, rtr, and the dissipatedenergy, Edisp, both attain maximum values for theannealing temperature of about 473 K. The valuesof rtr and Edisp are closely related to the superelasticstrain of the material.

To measure the irrecoverable strain, samples areannealed at 473 K for 30 min, and are compressedat a strain rate of 10�3/s and 296 K. The resultsare shown in Fig. 7. For a maximum strain of about6%, the residual strain is about 0.5%. Fig. 8 showsthe results of compression tests on three other sam-ples, with the maximum strain of about 6%, attainedat a strain rate of 10�3/s. It is seen that when the testtemperature is 296 K, the residual strain is 0.5%,whereas when the test temperature is 273 K, theresidual strain is 0.25%, and when the test tempera-ture is further reduced to 253 K, the residual strainbecomes 0.5% again. Therefore, when the annealingtemperature is 473 K, a good superelastic property(smaller residual strain) is found at a test tempera-ture of 273 K for a strain rate of 10�3/s. The resultsin Fig. 9 show that, if samples are compressed to a

0

500

1000

1500

2000

2500

3000

3500

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

Strain

Stre

ss (

MP

a)

NiTi, 10-3/s, T0 = 296K, Annealing Temperature 473K, 30 Minutes

E = 43GPa

Fig. 7. Stress–strain curve at indicated annealing temperatureand duration.

Page 6: Superelastic and cyclic response of NiTi SMA at various strain

0

500

1000

1500

2000

2500

3000

0.00 0.02 0.04 0.06 0.08 0.10 0.12

Strain

Stre

ss (

MP

a)

NiTi, 10-3/s, Annealing Temperature 473K, 30 Minutes

T0 = 296K

T0 = 253KT0 = 273K

Fig. 9. Residual strains at indicated test temperatures.

0

500

1000

1500

2000

2500

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08

Strain

Stre

ss (

MP

a)

NiTi, 10-3/s, Annealing Temperature 473K, 30 Minutes

T0 = 296K(R.S.=0.5%)

T0 = 273K(R.S.=0.25%)

T0 = 253K(R.S.=0.5%)

Fig. 8. Stress–strain curves at a strain rate of 10�3/s andindicated test temperatures.

468 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

strain of about 8.4% at a strain rate of 10�3/s, thenthe irrecoverable strain for test temperatures of 253and 273 K is about 3%. As is discussed later on, for30 min annealing at 823 K, three cycles of strainingto 5.8% maximum strain result in zero residualstrain (see Fig. 10), but a fourth cycle to 8% maxi-

0

400

800

1200

1600

2000

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10Strain

Stre

ss (

MP

a)

NiTi, 10-3/s, T0 = 297K

Annealing Temperature 823K, 30 Minutes

N = 1 2 3 4

Fig. 10. Change in superelasticity due to repeated loading.

mum strain results in a 4% residual strain. A similarirrecoverable strain is observed for 30 min annealingat 473 K. Based on these results, we infer that thisNiTi SMA could have a recoverable strain up toabout 6% maximum strain.

4.4. Dissipated energy, Edisp in cyclic loading

SMAs possess high damping capacity both inthe austenite state as a result of the stress-inducedmartensitic transformation, and in the martensitestate as a result of the stress-induced martensiticvariant reorientation (Liu et al., 1999a,b). Follow-ing Fig. 2, in near-equiatomic Ti–Ni alloys thethermoelastic martensitic transformation from theparent phase with a B2 structure to a monoclinicB19 0 structure occurs upon loading, then the ther-moelastic martensitic transformation from theB19 0 to the B2 structure occurs upon unloading.By deformation cycling in the superelastic tempera-ture range (above Af), the shape of the recoverysuperelastic loop in the stress–strain curve changesgradually; the yield stress and the hysteresis widthof the loop decrease, and the permanent elongationincreases. Although the change in the shape of thesuperelastic loop is significant in the early cycles,it becomes insignificant after 100 cycles (Saburi,1998). The cycle of superelastic deformation affectsthe critical stress to start the stress-induced trans-formation (Kato et al., 1999). Cycling at higherstrain rates has been found to increase the residualelongation that remains after the specimen isunloaded, and to cause a more rapid declineof the critical stress for martensite formation(Strnadel et al., 1995).

Several NiTi SMA samples are first annealed at773 K for 30 min, then they are compressed repeat-edly to a common maximum stress at various tem-peratures. The curves in Fig. 11 are at a 296 K testtemperature. Their corresponding energy and resid-ual strain vs. the number of cycles are shown inFig. 12. In this figure, the dissipated energy, Edisp

decreases and tends to a stable value with anincreasing number of cycles. The accumulatedresidual strain also increases gradually and attainsa stable value, as the number of cycles increases.The change in the superelastic loop is significantin several of the early cycles, as also reported bySaburi (1998). The stress–strain curves in Fig. 13are at a 326 K test temperature. These show thatinitially the shape of the superelastic loops isalmost the same, but it begins to change after the

Page 7: Superelastic and cyclic response of NiTi SMA at various strain

0

200

400

600

800

1000

1200

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07Strain

Stre

ss (

MP

a)

N = 1 N = 2 N = 3 N = 4 N = 5 N = 10 N = 20 N = 30 N = 50 N = 80 N = 100 N = 150 N = 200 N = 250

NiTi, f = 0.05Hz, T0 = 296KAnnealing Temperature 773K, 30 Minutes

Fig. 11. Stress–strain curves in cyclic loading of 0.05 Hz at296 K.

0

2

4

6

8

10

12

14

16

18

20

0 50 100 150 200 250 300

Number of Cycles (N)

Dis

sipa

ted

Ene

rgy

Ed

(J/m

3 )

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Accum

ulated Residual Strain (%

)

NiTi, f = 0.05Hz, T0 = 296K

Accumulated Residual Strain

Annealing Temperature 773K, 30 Minutes

Fig. 12. Dissipated energy, and accumulated residual strain vsnumber of cycles.

0

200

400

600

800

1000

1200

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08

Strain

Stre

ss (

MP

a)

N = 1N = 2N = 3N = 4N = 5N = 10N = 15N = 20N = 30N = 50N = 80

NiTi, f = 0.05Hz, T0 = 326KAnnealing Temperature 773K, 30 Minutes

Fig. 13. Stress–strain curves in cyclic loading of 0.05 Hzfrequency.

0

2

4

6

8

10

12

14

16

18

20

0 10 20 30 40 50 60 70 80 90 100

Number of Cycles (N)

Dis

sipa

ted

Ene

rgy

Ed

(J/m

3 )

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Accum

ulated Residual Strain (%

)

Accumulated Residual Strain

NiTi, f = 0.05Hz, T0 = 326KAnnealing Temperature 773K, 30 Minutes

Fig. 14. Dissipated energy, and accumulated residual strain vsnumber of cycles.

S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474 469

fifth cycle. The corresponding energy and the resid-ual strain are shown in Fig. 14 as functions of the

number of cycles. Comparing curves in Fig. 14 withthose in Fig. 12, it is seen that the shape of thesuperelastic loop, the dissipated energy, Edisp, andthe accumulated residual strain tend to their stablevalues much sooner at 326 K than at 296 K. InFigs. 11 and 13, the stress–strain curves shift tothe right due to the accumulated residual strain.According to Figs. 11–14, the dissipated energy,first quickly decreases, and then Edisp, graduallytends to its stable value, while the accumulatedresidual strain first increases rapidly and then grad-ually tends to its stable value, as the number offatigue cycles increases. In practice, SMAs arerepeatedly deformed to obtain stable and wellbehaved superelastic behavior. This is usuallycalled the ‘‘training process.’’ In general, repeatedtraining results in smaller dissipated energy forSMAs. In the present tests, the dissipated energyis greater at a 326 K test temperature than at296 K. In Fig. 15, the sample is annealed at a tem-perature of 473 K for 30 min, then tested. As canbe seen, the residual strain in the first cycle is quitelarge, but in the 100th cycle, the increment of resid-ual strain for that cycle basically disappears. Thedissipated energy also decreases rather rapidly asthe number of cycles increase.

Based on the above results, it is seen that (1)the energy dissipated in each cycle of loading–unloading of NiTi SMAs decreases, and even disap-pears as the number of cycles increases; similarly,the accumulated residual strain tends to a stablevalue with an increasing number of fatigue cycles;and (2) the residual strain depends on the test tem-perature, the fatigue stress amplitude, and theannealing temperature.

Page 8: Superelastic and cyclic response of NiTi SMA at various strain

0

400

800

1200

1600

2000

2400

2800

3200

0.00 0.22Strain

Stre

ss (

MP

a)

1 cycle 2 cycle 3 cycle 4 cycle 5 cycle 10 cycle 15 cycle 20 cycle 30 cycle 50 cycle 80 cycle 100 cycle

NiTi, f = 0.005,T0= 296K; Annealing Temperature: 473K, 30 Minutes

Strain = 0.04

Fig. 15. Stress–strain curves for the last cycle of indicatedloading frequency; each curve starts at zero strain.

0

400

800

1200

1600

2000

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08Strain

Stre

ss (

MP

a)

NiTi, 1,400/s, T0 = 296K

Edisp = 19 J/m3

Annealing Temperature 773K, 30 Minutes

Fig. 16. A typical dynamic stress–strain curve for NiTi SMA at1400/s loading strain rate and 296 K initial temperature.

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

2.0E-04 2.5E-04 3.0E-04 3.5E-04 4.0E-04 4.5E-04 5.0E-04

Time (s)

Stra

in

1,400/s 400/s

NiTi, T0 = 296KAnnealing Temperature 773K, 30 Minutes

Fig. 17. A typical dynamic strain time history obtained using theHopkinson bar technique.

470 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

5. Strain rate effect

In general, the elastic properties of metals aretemperature dependent, and high strain rate loadingcan lead to a temperature rise in the material. Forthe superelastic deformation of NiTi SMAs, a smalltemperature change can produce a change in thesuperelastic behavior, as in Figs. 8 and 9. In thepresent study, dynamic tests at high-strain rates(greater than 102/s), and initial temperatures in therange of 77–400 K, are performed using UCSD’srecovery Hopkinson bar facility, enhanced forhigh-temperature and high-strain rate recovery tests(Nemat-Nasser et al., 1991; Nemat-Nasser andIsaacs, 1997). With this facility, the reflected tensilepulse in the incident bar is trapped, so that thesample is subjected to a single compression pulseof desired profile, as discussed by Nemat-Nasseret al. (1991).

5.1. Superelastic response

Several specimens are annealed at 773 K for30 min, and then tested at various strain rates, usingUCSD’s recovery Hopkinson technique with apaper pulse shaper. The result of one such test is dis-played in Fig. 16. The transformation stresses, rtrand rre, for this case are 720 MPa and 280 MPa,respectively, and the dissipated energy is 19 J/m3,with an average strain rate of 1400/s, at a 296 K ini-tial temperature. The sample deformation is com-pletely recovered in this test. The unloading occursat a lower strain rate than that of loading. Thestrain time history is plotted in Fig. 17. By calculat-

ing the slope of the loading and unloading curves,the unloading strain rate is about 400/s, whereasthe loading occurs at a 1400/s strain rate. We haveobserved in all our tests that the unloading for NiTiSMAs occurs at a strain rate of the order of 102/s,independently of the loading rates. In general, theunloading rate cannot be easily controlled. There-fore, hereafter we focus attention on the loadingrates only.

5.2. Superelastic behavior in dynamic fatigue tests

Our quasi-static cyclic loadings show that the dis-sipated energy, Edisp, of NiTi SMAs tends to a sta-ble value with an increasing number of cycles. Thisvalue depends on the maximum strain and the stresslevel. To explore the high strain rate superelasticbehavior of NiTi SMAs under repeated loading,dynamic tests are performed using UCSD’s recov-ery Hopkinson bar facility. The experimental pro-

Page 9: Superelastic and cyclic response of NiTi SMA at various strain

0

200

400

600

800

1000

1200

1400

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08Strain

Stre

ss (

MP

a)

N = 1N = 10N = 20N = 30N = 50

R.S. = 0.05% ( N = 10)

R.S. = 0.15% (N = 50)R.S. = 0.10% (N = 30)

NiTi, 1,000/s (unloading rate : 330/s), T0 = 296K

Fig. 18. Stress–strain curves at 1000/s loading strain rate and296 K initial temperature.

0

400

800

1200

1600

2000

2400

2800

0.00 0.32Strain

Stre

ss (

MP

a)

10-3/s

10-1/s

NiTi, T0 = 296K

1,500/s (330/s)

4,200/s

2,100/s (360/s)

440/s (360/s)

Strain = 0.04

Annealing Temperature 773K, 30 Minutes

Fig. 19. Stress–strain curves at indicated loading strain rates and296 K initial temperature; each curve starts at zero strain.

0

400

800

1200

1600

2000

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07Strain

Stre

ss (

MP

a)

0.001/s0.1/s440/s1,500/s2,100/s4,200/s

NiTi, T0 = 296KAnnealing Temperature 773K, 30 Minutes

Fig. 20. Stress-induced martensitic transformation (austenite tomartensite) curves at indicated loading strain rates and 296 Kinitial temperature.

S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474 471

cess is similar to that in subsection 6.1. A sample isrepeatedly loaded for 50 cycles, each to a maximumstrain of about 5.6%, at a 1000/s strain rate and a296 K initial testing temperature; see Fig. 18. Inthe figure, the unloading curves do not completelyreturn to the origin. However, precise dimensionalmeasurement reveals an almost complete recoveryafter each unloading, indicating that full recoveryhas been attained shortly after the stress–strain mea-surement was terminated. For a typical case, after10 cycles, the accumulated residual strain is about0.05%, and after 50 cycles, it is about 0.15%. Thedissipated energy for each cycle is about 22 J/m3.Compared with the results of the quasi-static case,the dissipated energy, Edisp, remains almost a con-stant in the dynamic case.

5.3. Strain rate effect on superelastic behavior

In general, the NiTi SMAs in the austenite phaseare cubic with high crystal symmetry. Upon stress-ing, the cubic austenite phase transforms into alower symmetry monoclinic phase (martensite).Upon unloading, the martensitic phase transformsback into the austenitic phase, since the martensiticphase is not stable. This is a diffusionless phasetransformation in solids, in which atoms movecooperatively, and often by a shear-like mechanism.This kind of transformation is sometimes referred toas displacive or military transformation. One openproblem is the relation between this transformationand the applied strain rate. In this paper, in order tounderstand the strain rate effect on the superelasticbehavior of the NiTi SMAs, we compare our lowstrain rate results with those of high strain rates

(greater than 102/s), obtained using UCSD’senhanced Hopkinson technique. Six samples areuniaxially compressed at various loading strainrates and a 296 K initial temperature, and theresults are shown in Fig. 19. It is seen that the trans-formation stresses, rtr, and rre increase with theincreasing strain rate. The plateau range (austeniteto martensite transformation region) tends to disap-pear as the strain rate reaches 4200/s. For strainrates in the range of 0.1–2100/s, the loading portionof the stress–strain curves appears to be similar, asin Fig. 20, but the stress levels are quite different,being greater for the higher strain rate; for a system-atic study of the effect of strain rate on the responseof SMAs, see Nemat-Nasser et al. (2005a,b). Asshown by these authors, for very high strain rates,the response of this material is similar to that ofan ordinary metal. The result for a strain rate of

Page 10: Superelastic and cyclic response of NiTi SMA at various strain

472 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

4200/s in Fig. 20, seems to suggest this tendency.Thus, with increasing stress level or strain rate, thediffusionless shear-like mechanism of deformationchanges to dislocation-based slip plasticity (Nemat-Nasser and his coworkers, 2005a,b). Therefore,when the strain rate increases to 4200/s and higher,the NiTi SMA tends to deform essentially similar toan ordinary austenite metal.

5.4. Temperature effect on superelastic behavior

In general, the parent phase (austenite) ofSMAs is stable at temperatures greater than Af.Thus, greater applied stresses are required forstress-induced martensitic transformation at suchtemperatures. Figs. 21 and 22 show the effect oftemperature on the plateau stress for samplesloaded at 10�3/s strain rate. The plateau disappears

0

400

800

1200

1600

2000

2400

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08Strain

Stre

ss (

MP

a)

T0= 296K

T0= 396K

T0= 356K

T0= 326K

NiTi, 10-3/sAnnealing Temperature 773K, 30 Minutes

Fig. 21. Stress–strain curves at 10�3/s strain rate and 296, 326,356, and 396 K initial temperatures.

0

400

800

1200

1600

2000

2400

0.00 0.20Strain

Stre

ss (

MP

a)

T0 = 296K

T0 = 396KT0 = 356K

T0 = 326K

NiTi, 10-3/s

Strain = 0.02

Annealing Temperature 773K, 30 Minutes

Fig. 22. Stress–strain curves at 10�3/s strain rate and indicatedinitial temperatures; each curve starts at zero strain.

for a 396 K initial test temperature. A similarresponse is observed at high strain rates (Fig. 20,4200/s strain rate). This shows that above a criticaltemperature (here higher than 396 K) the parentphase is so stable that stressing cannot induce mar-tensitic transformation. The transformation stres-ses, rtr and rre, of the NiTi SMAs tend to increasewith increasing temperature, as does the residualstrain; see results in Figs. 21 and 22. In Figs. 23and 24, the stress–strain curves are plotted at differ-ent temperatures for 0.1/s and 1400/s strain rates.The trend of these curves is the same as those ofFigs. 21 and 22. The plateau slope of the stress–strain curves increases with increasing temperature.For a temperature of 356 K and a 1400/s strain rate,the stress-induced martensitic transformation isessentially suppressed (see Fig. 24). In Fig. 25, theloading parts of the curves in Fig. 24 are plotted

0

400

800

1200

1600

2000

2400

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08Strain

Stre

ss (

MP

a)

NiTi, 10-1/s

T0 = 396K

T0 = 296KT0 = 326K

Annealing Temperature 773K, 30 Minutes

Fig. 23. Stress–strain curves at 10�1/s strain rate and 296, 326,and 396 K temperatures.

0

500

1000

1500

2000

2500

0.00 0.35Strain

Stre

ss (

MP

a)

356K 326K 296K 273K250K 210K 77K

NiTi, 1,400/s

T0 =

Strain=0.04

Annealing Temperature 773K, 30 Minutes

Fig. 24. Stress–strain curves at 1400/s loading strain rate andindicated initial temperatures; each test starts at zero strain.

Page 11: Superelastic and cyclic response of NiTi SMA at various strain

0

400

800

1200

1600

2000

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08Strain

Stre

ss (

MP

a)

356K326K296K273K250K210K

NiTi, 1,400/s

T0 = 210K

T0 = 356K

Annealing Temperature 773K, 30 Minutes

Fig. 25. Stress-induced martensitic transformation (austenite tomartensite) at 1400/s loading strain rate and indicated initialtemperatures.

S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474 473

and compared. As is evident, the transformationstress, rtr, increases with increasing temperature.

6. Conclusions

To understand the thermomechanical response,especially the superelastic behavior of NiTi shape-memory alloys (SMAs), a number of uniaxialcompression tests are performed using an Instronservohydraulic testing machine and UCSD’senhanced Hopkinson technique. These tests are per-formed over the range of strain rates from 10�3/s toabout 4200/s, and at various initial temperatures, upto 400 K. The influence of the annealing tempera-ture on the fatigue response is also examined. Sev-eral noteworthy conclusions are as follows:

(1) By changing the annealing temperature, thetransformation stress and the dissipatedenergy of NiTi SMAs can be obviouslychanged, as this changes the transformationtemperatures.

(2) In the quasi-static cyclic fatigue tests, the dis-sipated energy of NiTi SMAs tends to aminimum stable value, and the cyclic loadingcan lead to a stable superelastic behavior forthe SMAs.

(3) Dynamic repeated tests of NiTi SMAs resultin a smaller change in the shape of the super-elastic loop and the dissipated energy than thequasi-static cyclic tests.

(4) The split Hopkinson bar can be effectivelyused for high loading rates, but, the unloadingrate in this system cannot be easily controlled;it has been of the order of 102/s in our tests.

Acknowledgments

This work was supported by ONR grantN00014-02-1-0666 to the University of California,San Diego.

References

Chen, W.W., Wu, Q., Kang, J.H., Winfree, N.A., 2001. Com-pressive superelastic behavior of a NiTi shape memory alloyat strain rates of 0.001–750 s�1. Int. J. Solids Struct. 38, 8989–8998.

Gadaj, S.P., Nowacki, W.K., Pieczyska, E.A., 2002. Temperatureevolution in deformed shape memory alloy. Infrared Phys.Technol. 43, 151–155.

Gall, K., Sehitoglu, H., Chumlyakov, Y.I., Kireeva, I.V., 1999.Tension-compression asymmetry of the stress–strain responsein aged single crystal and polycrystalline NiTi. Acta Mater. 47(4), 1203–1217.

Hosogi, M., Okabe, N., Sakuma, T., Okita, K., 2002. A newconstitutive equation for superelastic deformation and pre-diction of martensite volume fraction in Titanium–Nickel–Copper shape memory alloy. Mater. Trans. JIM 43 (5),822–827.

Huang, X., Liu, Y., 2001. Effect of annealing on the transfor-mation behavior and superelasticity of NiTi shape memoryalloy. Scripta Mater. 45, 153–160.

Kato, H., Ozu, T., Hashimoto, S., Miura, S., 1999. Cyclic stress–strain response of superelastic Cu–Al–Mn alloy single crystal.Mater. Sci. Eng. A A264, 245–253.

Lagoudas, D.C., Ravi-Chandar, K., Sarh, K., Popov, P., 2003.Dynamic loading of polycrystalline shape memory alloy rods.Mech. Mater. 35, 689–716.

Liu, Y., Li, Y., Ramesh, K.T., Humbeeck, J.V., 1999a. Highstrain rate deformation of martensitic NiTi shape memoryalloy. Scripta Mater. 41 (1), 89–95.

Liu, Y., Xie, Z., Humbeeck, J.V., 1999b. Cyclic deformation ofNiTi shape memory alloys. Mater. Sci Eng. A A273–A275,673–678.

Miller, D.A., Lagoudas, D.C., 2001. Influence of cold workand heat treatment on the shape memory effect and plasticstrain development of NiTi. Mater. Sci. Eng. A A308, 161–175.

Nemat-Nasser, S., Isaacs, J.B., 1997. Direct measurement ofisothermal flow stress of metals at elevated temperatures andhigh strain rates with application to Ta and Ta–W alloys.Acta Mater. 45, 907–919.

Nemat-Nasser, S., Isaacs, J.B., Starrett, J.E., 1991. Hopkinsontechniques for dynamic recovery experiments. Proc. R. Soc.Lond. 435 (A), 371–391.

Nemat-Nasser, S., Choi, J.Y., Guo, W.-G., Isaacs, J.B., 2005a.Very high strain-rate response of a NiTi shape-memory alloy.Mech. Mater. 37, 287–298.

Nemat-Nasser, S., Choi, J.Y., 2005b. Strain rate dependence ofdeformation mechanisms in a Ni–Ti–Cr shape-memory alloy.Act Mater. 53, 449–454.

Otsuka, K., Ren, X., 1999. Recent development in the research ofshape memory alloys. Intermetallics 7, 511–528.

Otsuka, K., Wayman, C.M., 1998. Mechanism of shape memoryeffect and superelasticity. In: Otsuka, K., Wayman, C.M.

Page 12: Superelastic and cyclic response of NiTi SMA at various strain

474 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 38 (2006) 463–474

(Eds.), Shape Memory Materials. Cambridge UniversityPress, Cambridge, pp. 27–48.

Saburi, T., 1998. TiNi shape memory alloys. In: Otsuka, K.,Wayman, C.M. (Eds.), Shape Memory Materials. CambridgeUniversity press, Cambridge, pp. 49–96.

Shaw, J.A., Kyriakides, S., 1995. Thermomechanical aspects ofNiTi. J. Mech. Phys. Solids 43 (8), 1243–1281.

Strnadel, B., Ohashi, S., Ohtsuka, H., Miyazaki, S., Ishihara, T.,1995. Effect of mechanical cycling on the pseudoelasticitycharacteristics of Ti–Ni and Ti–Ni–Cu alloys. Mater. Sci.Eng. A A203, 187–196.

Tobushi, H., Shimeno, Y., Hachisuka, T., Tanaka, K., 1998.Influence of strain rate on superelastic properties of NiTishape memory alloy. Mech. Mater. 30, 141–150.