Transcript

Thermal aging behavior in Cu–Al–Ni–xCo shape memory alloys

Safaa N. Saud • E. Hamzah • T. Abubakar •

H. R. Bakhsheshi-Rad

Received: 29 July 2014 / Accepted: 21 October 2014 / Published online: 7 November 2014

� Akademiai Kiado, Budapest, Hungary 2014

Abstract The effects of aging on the phase transforma-

tion temperatures, morphology, and mechanical properties

were investigated in the Cu–Al–Ni–xCo (x = 0.4, 0.7, and

1.0 mass%) SMAs via differential scanning calorimetry,

field emission scanning electron microscopy, X-ray dif-

fraction, tensile, vickers microhardness, and shape memory

effect test. The results show that the phase morphology and

the presence of c2 precipitates in different sizes, volume

fractions, and distributions have an obvious effect on the

phase transformation characteristics and properties. In

addition, a noticeable variation in the transformation tem-

peratures and thermodynamic parameters were occurred

with the aging. In tension, the highest tensile strength and

tensile strain of 750 MPa and 5.5 % were indicated in the

Cu–Al–Ni–1 mass% Co alloy after being aged at 523 K for

24 h and 373 K for 48 h, respectively. However, the results

of the strain recovery by shape memory effect were varied

in accordance with the variation of the c01 and b01 mor-

phology and volume fraction of c2 precipitates. In aged

alloy of 523 K for 48 h, the thickness ofc01 and b01 phases

and the volume fraction of c2 precipitates increase, thus,

the movement of martensitic interfaces is restricted causing

an increase in et by SME.

Keywords Shape memory alloys � Cu–Al–Ni–Co � Aging

treatment � Shape memory effect

Introduction

Shape memory alloys have a remarkable capability to act

as a sensor and an actuator, which makes them a multi-

functional material with the unique ability to precisely fit

the requirements of a particular usage [1, 2]. Among the

shape memory alloys, NiTi SMAs have been used with the

greatest amount of feedback in a numerous applications

including biomedical and industrial applications [3, 4].

However, the low transformation temperatures (-373 to

373 K) and high producing cost of NiTi alloys render their

usage for many applications impractical, especially for

high temperature usage. Therefore, Cu–Al–Ni SMAs are

the most preferred option as successful alternatives to NiTi

due to their high transformation temperatures (-573 to

573 K), low producing cost, and high thermal stability

[5–8]. However, the brittleness [9], low strength, large

elastic anisotropy, and large grain size hinder its practical

applications. As a result, many researchers have tried to

refine the grain size of Cu-based SMAs through the addi-

tion of alloying elements and/or applying different thermal

aging treatment conditions [10–18]. The unique properties

of Cu–Al–Ni SMA are mainly attributed to the thermo-

elastic martensitic transformation that takes place in a

temperature range of (173–473 K) depending on the

composition of alloy. These properties are significantly

affected by the movements of interfaces (twin boundaries,

martensitic variants, and parent/martensite phase bound-

aries) [19]. Some other parameters also exhibited a sig-

nificant influence on the properties of these alloys, for

instance, the formation of vacancies, dislocations, grain

boundaries, and precipitates, along with the way of their

distribution and volume fraction. On the other hand, the

Cu–Al–Ni alloys are liable to aging treatment [20, 21],

which it leads to form a different types, volume fractions,

S. N. Saud � E. Hamzah (&) � T. Abubakar �H. R. Bakhsheshi-Rad

Faculty of Mechanical Engineering, Universiti Teknologi

Malaysia UTM, 81310, Johor Bahru, Johor, Malaysia

e-mail: [email protected]

123

J Therm Anal Calorim (2015) 119:1273–1284

DOI 10.1007/s10973-014-4265-6

and distributions of precipitates in the microstructure,

associated with the successive of martensitic transforma-

tion from b01 to c01 phase [22]. However, it is well estab-

lished that different types of martensitic phases, c01 (2H), b01(18R), and a01 (6R) form in the Cu–Al–Ni SMAs,

depending on the chemical composition, mode of applied

load, test temperature, and crystal orientation [23–25].

Nevertheless, Balo’s research on Cu–Al–Ni alloys dis-

covered that an appropriate aging process may significantly

vary the transformation temperatures [14]. Furthermore, a

proper aging process may produce a favorable combination

of features, such as high strength and a large strain

recovery [11, 26, 27]. As these alloys are susceptible to

post-quench aging at high temperature service conditions,

their transformation temperatures, martensitic phases, and

the mechanical properties can change with the time of

treatment. As a consequence, a large number of studies

were conducted on various aspects of aging in these

alloys and their influence on the shape memory properties

[10, 11, 13, 14, 28, 29]. However, only a limited number of

studies that too confined to a small number of aging con-

ditions were conducted on the mechanical response of the

aged Cu–Al–Ni alloys [10, 14, 28, 29]. The present work

was initiated to conduct a comprehensive study on the phase

transformation temperatures, morphology, and mechanical

properties in the Cu (84.1-x)–11.9 Al–4 Ni–xCo (x = 0.4,

0.7, and 1) (in mass%) SMAs after conducting an aging

treatment at 373, 423, and 523 K for 24 and 48 h.

Experimental

Materials preparation

The alloy was produced by melting high purity metals of

Cu (99.999 %), Al (99.999 %), Ni (99.995 %), and Co

(99.95 %) using an induction furnace. These metals were

melted in a silicon carbide crucible at a temperature about

1,573 K with continuous stirring and then poured into a

cast iron mold with dimensions of 270 9 50 9 20 mm3.

Three ingots were produced with different percentages of

Co (0.4, 0.7, and 1.0 mass%) and each ingot was cut into

six samples using Electrical discharge machining wire and

then homogenized at the 1,173 K for 30 min, and then

quenched in water which led to the formation of martensite.

The chemical composition analysis for the Cu–Al–Ni–Co

SMAs was investigated using inductively coupled plasma

mass spectrometry (ICP–MS). The aging treatments were

carried out at 373, 423, and 523 K for 24 and 48 h in the

normal atmosphere. The optimization of the aging times

and temperatures has been selected in accordance with

extensive studies done by the authors.

Material characterization

Flat specimens were cut from the aged samples with

dimensions of 10 9 10 9 2 mm3 for the microstructural

and X-ray diffraction (XRD) characteristics. Filings of the

alloys removed of about 2–6 mg were taken for the differ-

ential scanning calorimetry measurements using a Mettler

Toledo DSC 822e, where the scanning rate was 10 K min-1

in the 323–673 K range. The phase identifications and

crystal structure determinations were carried out using a

D5000 Siemens X-Ray diffractometer fitted with CuKaX-ray source with a locked couple mode, 2h range between

30–80�, and 0.05� sec-1, is the scanning step. The quenched

samples were ground and polished, and then etched in a

solution containing 2.5 g ferric chloride (FeCl3.6H2O) and

48 mL methanol (CH3OH) in 10 mL HCl for 4 min.

Mechanical tests

Tensile test

The tensile test was performed using an Instron 5982-type

universal testing machine operated at a constant strain rate

of 0.1 mm min-1. The tests were carried out at room

temperature until failure occurred, and then the fracture

stress–strain levels were determined under the tensile load.

Shape memory effect test

The shape memory effect test was carried out using a spe-

cially designed machine. The specially designed contents

were analyzed using an Instron 5982-type universal testing

machine operated with special program parameters

according to the shape memory test which was connected to

a heater tape and digital thermocouple in order to control the

applied temperature, and an external extensometer to mea-

sure the shape extension and recovery. The tests were car-

ried out at a temperature below Mf, which was about 373 K,

where the alloys would be able to obtain shape recovery.

Then the deformed sample that still had an unrecoverable

shape was subsequently heated above the austenite finish

temperature (Af ? 333 K) for 10 min followed by a water

quench to recover the residual strain (er). The recovered

shape was attributed to the transformation of the detwinned

martensite to the austenite phase, which had been termed as

a transformation strain (et). After the cooling process, the

martensite again formed in a self-accommodated structure.

1274 S. N. Saud et al.

123

Results and discussion

Microstructural analysis

The typical microstructures of the Cu–Al–Ni–xCo

(x = 0.4, 0.7, and 1.0 mass%) under different aging con-

ditions are shown in Fig. 1a–r. In the aged alloys of

0.4 mass% Co at 373 K for 24 and 48 h, a fine needle-like

and plate-like of b01 and c01 are formed in different orien-

tations, volume fractions, and sizes along with some pre-

cipitates/intermetallic compounds are formed, as indicated

by the red dot arrow in Fig. 1a, j. Perhaps it seems that with

the Co addition, there is certainly an innovative new phase

formed which begin to grow up into the matrix, and

523

423

Agi

ng te

mpe

ratu

re/K

Cobalt concentration/mass%

Agi

ng te

mpe

ratu

re/K

373

0.4 0.7 1.0

Cobalt concentration/mass%0.4

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

(j) (k) (l)

(m) (n) (o)

(p) (q) (r)

0.7 1.0

γ2

γ2

γ 1

β 1

γ2

γ 1

β 1

523

48 hr

423

373

24 hr

Fig. 1 FESEM micrographs of

the microstructures of Cu–Al–

Ni–xCo SMA (x = 0.4, 0.7, and

1.0 mass%) under different

aging conditions

Thermal aging behavior in Cu–Al–Ni–xCo 1275

123

consequently above the needle-like and plate-like of b01phases, which is typically known as c2 phase. In accor-

dance with an EDS analysis of a spot scanned for the c2

phase area, it was found that these precipitates are Co-rich,

which are an amalgamation of Co, Ni, and Al in compound

of Al75Co22Ni3, as is being pointed out in Fig. 2. With

increasing aging time and temperature to 423 K for 24 and

48 h, the volume fraction and size of c2 phase will increase.

On the other hand, this increment can lead to form and

accumulates the c2 phase at the grain boundaries, thus, the

thickness of the grain boundaries has increased as shown in

Fig. 1d, m. Generally speaking, with an increase in aging

time and temperature, the driving force of the precipitation

hardening is increasing over the subsequent aging [11, 28].

With further increases in the aging time and temperature to

523 K for 24 and 48 h, the c2 precipitates are starting to

form and diffuse into the matrix associated with the

decreases of their sizes as shown in Fig. 1g, p. From

another point of view, it was noticed that with the aging

treatment, the volume fraction of b01 phase increased, which

is consequently associated with the increases of c01 thick-

ness, even though both phases are formed in the matrix and

orientated randomly in accordance to the aging treatment

condition.

At 0.7 mass% Co-aged alloys, the microstructures have

exhibited the same variations in the volume fraction, size,

and morphology of the c01 phase, b01 phase, and c2 precip-

itate as shown in Fig. 1. These variations occur according

to the principles of the aging treatment condition. It can be

seen that within 373 K aging temperature for 24 h, the

volume fraction and size of c01 phase are increased, while

the volume fraction of c2 precipitates is shown much less

and look like they have been penetrated into the micro-

structure as shown in Fig. 1b. With increasing the aging

time to 48 h at 373 K, the volume fraction of b01 phase and

c2 precipitates will be increased, and they have formed

randomly in the matrix as observed in Fig. 1k. However, at

the aging temperature of 423 K and time of 24 h, the c2

precipitates are starting to create a coarse boundary, and

therefore, clear and thick boundaries have occurred along

with a thick and sharp plate-like of c01 phase are formed,

while the remaining area is full of the needle-like of b01phase as shown in Fig. 1e. The c2 precipitates are formed

randomly in the microstructure of the 423 K for 48 h aged

alloy and the volume fraction of b01 phase has increased that

is associated with the decrement of plate thickness of c01phase as shown in Fig. 1n. With a further increase in the

aging time and temperature to 523 K for 24 and 48 h, the

plate thickness of c01 phase has increased and started to

form in a crisscross with each other obtaining a V shape

crossed with the needle-like of b01 phase, whereas, the c2

precipitates have shown the same decrement of the previ-

ous alloy as shown in Fig. 1h, q.

From the microstructures of 1.0 mass% Co alloys aged

at 373 K for 24 and 48 h, it was seen that the plate sizes of

b01 and c01 phases became more visible, thus corresponding

with the increase in the thickness of the coarse variation of

the c01 phase. In addition, these variants started to obtain a

discontinuous growth with the increase of the distance

between them, as shown in Fig. 1c, l. However, after

increasing the aging temperature to 423 K, the coarse

variants showed a complete discontinuous growth along

with obtaining random orientations. Moreover, the size of

the c2 precipitates was reduced due to the dissolution of

these particles into the microstructure, as shown in Fig. 1f

and o. When the aging temperature increased to 523 K for

24 and 48 h, it was found that a completed growth of the c01phase with a lamella structure was occurred, and the plate-

10 μm Electron image 1

1

AlCo

Ni

Cu

Co

Elements Wt.% At.%

Al 9.58 19.89Ni 0.96 0.91Co 4.71 4.49Cu 84.75 74.71

Spectrum 2

Ni

Full Scale 6308 cts Cursor: 3.202 (56 cts) keV2 3 4 5 6 7 8 9 10

1Full Scale 2783 cts Cursor: 5.338 (22 cts) keV

2 3 4 5 6 7 8 9 10

CoNi Al Elements Wt.% At.%

Al 56.34 73.8Co 38.52 23.1Ni 5.14 3.1

Spectrum 1

(a)

(b)

(c)

Fig. 2 A spot scanned of the EDS analysis of the aged Cu–Al–Ni–0.4 mass% Co a Micrograph of scanned area; b Spectrum 1; c Spectrum 2

1276 S. N. Saud et al.

123

like groups of the b01 phase become thicker, as shown in

Fig. 1i, r, along with an increase in the size of the c2

precipitates.

Figure 3(a–c) shows the XRD diffraction patterns of the

aged Cu–Al–Ni–xCo (x is 0.4, 0.7, and 1.0 mass%).

According to the influence of the aging treatment, the

obtained peaks vary in terms of presence, 2h value, and

intensity, which are completely reflected on the structure

parameters, such as miller induced or crystallite size. These

variations are attributed to the morphology and orientation

of the martensitic phase, and to the volume fraction and size

of the precipitates. Generally, the pattern peaks of the XRD

diffraction represent the existing phases into the micro-

structure, in which the pattern peaks of the peaks of (200)

and (202) represent the c01 phase, and (122), (0018), (128),

(1210), (2010), (1123), (208), (320), (040), and (311) rep-

resent the b01 phase. The pattern peaks of (-421), (712), and

(314) represent the c2 precipitates. According to the

matching standard of this alloy, it was found that these

peaks are related to the (Al75Co22Ni3) phase that we refer to

as c2. Furthermore, it was observed that the c2 precipitate

pattern peaks vary in intensity values and shift according to

the volume fraction of these precipitates in the micro-

structure, which is in complete agreement with the micro-

structural variations. At 0.4 mass% Co addition-aged

alloys, the results show that peak patterns varied in terms of

presence and intensities due to changes in aging time and

temperature. It was also found that the highest intensities

30

Inte

nsity

/a.u

.

(200

)

(122

)(2

02) (0

018)

(128

)

Inte

nsity

/a.u

.

Inte

nsity

/a.u

.

Inte

nsity

/a.u

.

35 40 45 50 552θ/°

60 65 70 75 80 30 35 40 45 50 552θ/°

2θ/°

60 65 70 75 80

30 35 40 45 50 552θ/°

60 65 70 75 80

(121

0)

(040

)

(311

)

(421

)

(112

3)(2

08)

(712

)

(314

)

(200

)

(122

)

(200

)

(122

)(2

02)

(128

)(1

210)

(201

0)

52 54 56 58 60 62 64 66 68 70

(001

8)

(112

3)

(208

)

(712

)

(314

)

(202

) (001

8)(1

28)

(121

0)(2

010)

(320

)(0

40)

(311

)

(112

3)

(208

)

(712

)

(314

)

(421

)

(421

)

(320

) (040

)

(311

)

373 K-24 hr

373 K-48 hr423 K-24 hr

423 K-48 hr523 K-24 hr

523 K-48 hr

JCPDS XRDAl75Co22Ni3(49-1278)

Inte

nsity

/a.u

.

Inte

nsity

/a.u

.

2θ/°

2θ/°

5550 60 65 70

48 50 52 54 56 58 60 62 64 66 68 70

(a) (b)

(c)

Fig. 3 X–ray diffraction patterns of a Cu–Al–Ni–0.4 Co SMA, b Cu–Al–Ni–0.7 Co SMA, c Cu–Al–Ni–1.0 Co SMA

Thermal aging behavior in Cu–Al–Ni–xCo 1277

123

were observed with the 423 K aging temperatures for 24 h.

This may be attributed to higher volume fraction of c2

precipitates into the matrix and at the grain boundaries of

this alloy. Almost the same behavior has been repeated for

the 0.7 and 1.0 mass% aged alloys as shown in Fig. 3b, c.

The lattice parameters and crystallite size of aged alloys

of Cu–Al–Ni–Co SMA were determined from the XRD

patterns and are recorded in Table 1. The lattice parameters

were evaluated in accordance to the orthorhombic 18R

structure, which was proven by the XRD indexing patterns.

Thus, the lattice parameters were determined using the

following relation [30]:

1

d2¼ 1

a2

h2

sin2b

� �þ k2

b2þ 1

c2þ l2

sin2b

� �� 2hl cosb

ac sin2b: ð1Þ

The crystallite size was determined by a Scherrer

equation [31, 32] for the highest intensity of the two peaks

of (0018) and (128), as follows:

Crystallite size dð Þ ¼ 0:9 � ðkÞB � cos h

; ð2Þ

where k is the XRD wavelength, b is the full width at half

maximum, and h is the Bragg’s angle. It was found that the

lattice parameters and crystallite size are varied with the

variation of aging time and temperature. However, the

highest crystallite sizes were observed 19.4, 17.4, and

24.8 nm with the 373 K–24 h for 0.4, 0.7, and 1 mass%

Co-aged alloys, respectively. The maximum value was

obtained with the aged 1.0 mass% Co alloys.

Transformation temperatures

Figure 4 shows the endothermic and exothermic curves of

the Cu–Al–Ni–1.0 mass% Co SMA aged at 523 K for

24 h. After applying the aging treatment at different times

and temperatures, the transformation temperatures are

slightly increased. The determined data from the DSC

curves for the aged and unaged alloys are presented in

Table 2. The transformation temperature increased gradu-

ally with an increase in the percentage of Co addition from

0.4 to 1 mass%, on the other hand, as a comparison with

the aged samples, it was also found that the transformation

temperatures decreased with increase in the aging times

and temperatures. This may attributed to the fact that the

aging treatment led to increase in the dislocation density,

and thus decrease the transformation temperatures [33].

The transformation temperatures of the aged alloys are

varied in according to the variations of the structures and

morphologies as well as the variations of volume fraction

of precipitates/intermetallic compounds that are associated

during the aging treatment. In other words, the phenomena

behind the variations in transformation temperatures are

related to the nucleation and growth of c2 precipitates that

may induce a stress field by the coherent boundaries

between the existing phases of c01 and b01 and precipitates of

c2 [34]. Remarkably, during the aging process, this stress

field can affect the transformation temperatures. Further

growth of the c2 precipitates can lead to loss of coherent

interface, and therefore, its influence on the transformation

Table 1 Lattice parameters and crystallite size of Cu–Al–Ni–XCo SMA under different aging conditions

Alloys Conditions a/A´

b/A´

c/A´ b Crystallite size/A

´

Cu–Al–Ni–0.4 mass% Co 373 K–24 h 3.711 5.79 43.985 96.364 194

373 K–48 h 3.877 5.125 39.285 74.467 168

423 K–24 h 4.132 5.268 38.406 91.924 167

423 K–48 h 3.892 5.145 39.573 73.219 180

523 K–24 h 3.914 5.211 39.605 73.419 157

523 K–48 h 4.0913 5.289 39.449 94.687 185

Cu–Al–Ni–0.7 mass% Co 373 K–24 h 4.138 5.234 38.186 87.570 174

373 K–48 h 4.138 5.238 38.189 87.569 172

423 K–24 h 4.108 5.254 38.625 94.091 143

423 K–48 h 4.144 5.234 38.206 87.757 169

523 K–24 h 3.747 5.287 38.523 98.532 172

523 K–48 h 3.821 5.183 38.569 79.557 161

Cu–Al–Ni–1 mass% Co 373 K–24 h 4.139 5.250 38.086 87.912 248

373 K–48 h 4.002 5.155 38.869 75.639 204

423 K–24 h 4.131 5.237 38.051 88.771 198

423 K–48 h 3.988 5.153 38.938 75.351 226

523 K–24 h 3.753 5.267 38.592 94.777 187

523 K–48 h 3.753 5.268 38.580 95.016 204

1278 S. N. Saud et al.

123

temperatures. A number of the aged alloys have shown

multiple endothermic/exothermic peaks in their transfor-

mation curves; their multiple peaks are attributed to the

interface transformations. These intermartensitic transfor-

mations are a first-order phase transformation between

martensites with different structures at temperatures below

the Ms. So far, several intermartensitic phases have been

found in the shape memory alloys, which have modulated a

lattice with different periodicity of stacking sequences [35].

However, the structures and transformation temperatures of

these intermartensitic phases depend on the levels of

applied stress and the chemical composition. On the other

hand, these interphases are very sensitive to the internal

stress of the alloy [36].

Meanwhile, the differences of the austenitic transfor-

mation temperatures (Af–As) are mainly larger than the

46018.5

19

19.5

20

20.5

21

21.5

470

Heating

480 490 500 510

Temperature/K

Hea

t flo

w/m

W

520 530 540 550 560

503 503.5 504 504.5 505Temperature/K

Hea

t flo

w/m

WH

eat f

low

/mW

505.5 506 506.5 507

512

20.94

20.96

20.98

21

21.02

21.04

21.06

18.7

18.75

18.8

18.85

18.9

18.95

513 514 515 516 517 518 519Temperature/K

End

o

Cooling

Fig. 4 DSC diagrams on the

heating and cooling cycle of the

Cu–Al–Ni–1.0 mass% Co SMA

aged at 523 K for 24 h

Table 2 Transformation temperatures of aged Cu–Al–Ni–xCo SMA

Alloys Conditions Transformation temperatures/K

As Af Ms Mf Af–As Ms–Mf To

Cu–Al–Ni–0.4 mass% Co 0 513.4 521.5 510 505.3 8.1 4.7 515.75

373 K–24 h 508.2 515.6 504.5 490.9 7.4 13.6 510.05

373 K–48 h 499.8 514.5 503 490 14.7 13 508.75

423 K–24 h 511.86 516.15 508.1 498.3 4.29 9.8 512.125

423 K–48 h 510 517.56 506 503.2 7.56 2.8 511.78

523 K–24 h 510.1 517.8 506.1 503.7 7.7 2.4 511.95

523 K–48 h 510 517.86 506 503.1 7.86 2.9 511.93

Cu–Al–Ni–0.7 mass% Co 0 515.68 525.2 514.2 507.7 9.52 6.5 519.7

373 K–24 h 508.91 515.66 504.36 490.1 6.75 14.26 510.01

373 K–48 h 500 514.41 503.1 489.8 14.41 13.3 508.755

423 K–24 h 511.8 516.3 508.1 498.6 4.5 9.5 512.2

423 K–48 h 510 517.5 506 503.4 7.5 2.6 511.75

523 K–24 h 510.3 518 506.7 503.4 7.7 3.3 512.35

523 K–48 h 509.9 515 506 503 5.1 3 510.5

Cu–Al–Ni–1 mass% Co 0 522.88 532 521.2 514.4 9.12 6.8 526.6

373 K–24 h 508.4 516 504.5 491.3 7.6 13.2 510.25

373 K–48 h 500.4 514 503 490.6 13.6 12.4 508.5

423 K–24 h 511.4 516.3 510.1 498 4.9 12.1 513.2

423 K–48 h 510.2 518 506.1 503.4 7.8 2.7 512.05

523 K–24 h 513.4 518.2 506.3 503.6 4.8 2.7 512.25

523 K–48 h 510.2 518.1 506.1 503.5 7.9 2.6 512.1

Thermal aging behavior in Cu–Al–Ni–xCo 1279

123

martensitic transformation temperatures (Ms–Mf), except

the aged alloys of 373 K for 24 h of the 0.4 mass% Co

addition, and with 373 and 423 K for 24 h of the

0.7 mass% Co and 1.0 mass% Co addition, which means

the (Ms–Mf) [ (Af–As). On the other hand, a small elastic

strain occurred associated with the occurrence of the plastic

relation and/or lattice softening [37]. However, within the

aging treatment, there will be some aging-induced lattice

softening or maybe some variations in the long-range of the

martensitic structure order.

Mechanical properties

Stress–Strain curve behavior and microhardness test

The typical stress–strain curves of the Cu–Al–Ni–xCo-

aged alloys are presented in Fig. 5a–c. It can be observed

that the tensile properties of the aged alloys are varied in

terms of fracture stress and fracture strain in accordance to

the condition of aging treatment as shown in Table 3.

In the 0.4 mass% Co-aged alloys, the strength is

increased with increasing the aging temperatures from 373

to 423 K, and then dropped at 523 K as shown in Fig. 5a. It

may be attributed to the higher or the more heterogeneous

dispersion of precipitate, which has shown the highest

volume fraction associated with a large size with 423 K for

24 h. Furthermore, it was found that the presence of pre-

cipitate can pin the mobility of dislocations, and therefore

the dislocations are required for a higher stress, either to

shear the precipitate particles, if the particles are large,

which lead to increase the fracture stress or bow around

them resulting in high fracture strain. Figure 5b shows the

stress–strain curve of the 0.7 mass% Co-aged alloy. It

found that the strength of the alloys is increased with

increasing the aging temperatures, where it has observed

the highest strength with the 523 K for 24 h. This may be

00 0.5 1 1.5 2 2.5 3 3.5

100 °C - 24 hr

250 °C-48 hr

100 °C - 48 hr

150 °C - 24 hr150 °C - 48 hr250 °C - 24 hr250 °C - 48 hr

4

50100150200250300350400450

Tens

ile s

tres

s/M

Pa

Tensile strain/% 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5Tensile strain/%

Tens

ile s

tres

s/M

Pa

Tens

ile s

tres

s/M

Pa

500550600650700750

050

100150200250300350400450500550600650700750

050

100150200250300350400450500550600650700750800

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6Tensile strain/%

100 °C - 24 hr

100 °C - 48 hr

150 °C - 24 hr

150 °C - 48 hr

250 °C - 24 hr

250 °C - 48 hr

100 °C - 24 hr

100 °C - 48 hr

150 °C - 24 hr

150 °C - 48 hr

250 °C - 24 hr

250 °C - 48 hr

100

°C-4

8 hr

100

°C-2

4 hr

150 °C-24 hr

150 °C-48 hr

250

°C-2

4 hr

250 °C-48 hr

100

°C-4

8 hr

100

°C-2

4 hr

150 °C-24 hr

150 °C-48 hr

250

°C-2

4 hr

250 °C-48 hr

100

°C-4

8 hr

100

°C-2

4 hr

150 °C-24 hr

150 °C-48 hr

250

°C-2

4 hr

(a) (b)

(c)

Fig. 5 Stress–strain curves obtained from the tensile test performed at room temperature of a Cu–Al–Ni–0.4 mass% Co SMA, b Cu–Al–Ni–

0.7 mass% Co SMA, and c Cu–Al–Ni–1.0 mass% Co SMA

1280 S. N. Saud et al.

123

attributed to the high volume fraction of c01 plate-like that

has affected the dislocation movements, and thus,

increased the strength of the alloy. On the other hand, the

volume fraction, distribution, and size of the c2 precipitates

may also influence the dislocation movements, therefore

influencing the strength and ductility of the alloy [38–40].

However, the highest strain was observed with the 423 K

for 24 h alloy. This may be attributed to the large size of

the c2 precipitates, which force the dislocation to bow

around them, and thus increase the strain of the alloy. The

aged samples of 1.0 mass% Co alloy showed an increase

in the values of fracture strain and fracture stress as shown

in Fig. 5c. The determined data are given in Table 3.

From this point of view, it has been proven that the

brittleness of Cu-based SMA can be slightly reduced

using an aging treatment. These enhancements were

mainly attributed to the microstructure changes, which can

be considered as variations of martensitic phase mor-

phologies and orientations, as well as the behavior and

morphology of the c2 precipitates. It was also seen that the

elongation had reached the maximum value at the aging

condition of 373 K for 48 h, which improved the frac-

tured strain to 5.5 %.

The fracture surface areas of the Cu–Al–Ni–xCo SMAs

are displayed in Fig. 5a–c. It was determined that the

popular fracture features of the aged alloys are varied based

on the changing the time and temperature of aging treat-

ment. Figure 5a reveals the fracture area of the aged Cu–

Al–Ni–0.4 mass% Co SMA. The fracture feature of 373 K

for 24 h alloy shows a mix mode fracture of the inter-

granular and transgranular as shown in Fig. 5a. By

increasing the aging time to 48 h, the fracture feature

transferred from the mix mode to a dimple rupture along

with a little intergranular area exhibited on the left side of

the fracture area as shown in Fig. 5a, whereby the

appearance of these dimples represents a ductile fracture.

The 423 K for 24 h alloy displayed an aggregate mode

fracture between a little area of a quasi-cleavage and a

large area of intergranular feature as shown in Fig. 5a.

With further aging time and temperatures, the alloys show

an intergranular fracture feature, except the alloy 523 K for

24 and 48 h exhibited a little area of quasi-cleavage along

with the intergranular fracture as shown in Fig. 5a. How-

ever, Fig. 5b, c show the fracture features of the aged Cu–

Al–Ni–0.7 mass% Co, and Cu–Al–Ni–1.0 mass% Co

SMAs, respectively. It was indicated that the fracture fea-

tures of both alloys are behaving similar to the 0.4 mass%

Co-aged alloys, whereas the types of fracture are varied

between the transgranular, intergranular, dimple rupture

with a little intergranular, quasi-cleavage, and mix mode.

However, the variations of the fracture modes are attributed

to the amount of the elastic anisotropy at the grain

boundaries and the grain size of the aged alloys. Therefore,

as much as the amount of the elastic anisotropy at the grain

boundaries is small, the dimple rupture is occurring, which

represents the ductile fracture. When the elastic anisotropy

at the grain boundaries is large, a quasi-cleavage fracture is

observed, which represents a brittle fracture mode. This

Table 3 Results obtained from the tensile, microhardness, and shape memory tests on the aged Cu–Al–Ni–xCo SMA at different conditions

Alloys Aging conditions Stress/MPa Strain/% Microhardness/MPa Strain recovery/%

Cu–Al–Ni–0.4 mass% Co 373 K–24 h 620 2.2 331 73

373 K–48 h 600 2.45 317 76

423 K–24 h 680 2.8 345 82

423 K–48 h 710 3.4 358 77

523 K–24 h 670 2.6 340 80

523 K–48 h 700 2.7 351 81

Cu–Al–Ni–0.7 mass% Co 373 K–24 h 600 4.1 340 85

373 K–48 h 600 4.8 342 84

423 K–24 h 700 5 360 87

423 K–48 h 610 3.6 347 80

523 K–24 h 712 3.8 376 82

523 K–48 h 700 3.2 370 80

Cu–Al–Ni–1 mass% Co 373 K–24 h 625 4.5 360 76

373 K–48 h 700 5.5 372 73

423 K–24 h 680 4 367 71

423 K–48 h 650 3.5 356 80

523 K–24 h 750 3.75 387 82

523 K–48 h 740 5 381 100

Thermal aging behavior in Cu–Al–Ni–xCo 1281

123

study can conclude that aging treatment is one of the most

important factors that can influence the amount of an

elastic anisotropy associated with controlling the grain size

of the alloys, and thus, the variations of the aging times and

temperatures are able to vary the fracture features of the

alloys. On the other hand, the fracture surface of the alloys

can also be affected by the structural compounds, for

instance, parent phases, precipitates, and intermetallic

compounds along with their size and distribution into the

matrix [41, 42]. Microhardness of the aged Cu–Al–Ni–xCo

SMA with different aging times and temperatures are

shown in Table 3. The microhardness value varied sub-

stantially through increasing the aging time and tempera-

ture. This study found that the highest microhardness

values were 358, 370, and 387 MPa with the 423 K for

48 h alloy for 0.4 mass% Co, 523 K for 48 h alloy for

0.7 mass% Co, and 523 K for 24 h alloys for 1.0 mass%

Co addition-aged alloys, respectively. The increment of

microhardness after the aging treatment in the shape

memory alloys is essentially linked to dislocation move-

ments [43, 44]. The reason for the microhardness variations

is believed to be related to the c2 precipitates. When the

alloy is aged at a lower temperature such as 373 K, the

nucleation and growth rates of the c2 precipitates are

slower in which this phenomena is greatly depended on the

temperature. As the aging temperature increases to 423 K

and 523, the c2 precipitates will come out gradually, which

led to harden the alloys.

Shape memory effect

The shape memory effects of Cu–Al–Ni–xCo (x is 0.4, 0.7,

and 1.0 mass%) are shown Fig. 6a–c. It is found that the

SMEs are variations in the terms of strain recovery,

according to the condition of aging treatment as shown in

Table 3. For the 0.4 mass% Co, the highest results are

observed with the 423 K for 24 h alloy, which has obtained

an 82 % recovery out of the original shape after being

heated above the Af as shown in Fig. 6a. Generally

speaking, the strain recovery is associated with the mar-

tensite ? austenite transformation, and as well that the

magnitude of the plastic deformation is dependent on the

size and volume fraction of needle-like, plate-like, and

precipitates that occurred around them, and therefore,

00

50

100

150

200

250

300

350

400

450

500

550

600100 °C-24 hr100 °C-48 hr150 °C-24 hr150 °C-48 hr250 °C-24 hr250 °C-48 hr

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3

Tensile strain/%

Tens

ile s

tres

s/M

Pa

Tens

ile s

tres

s/M

Pa

0

50

100

150

200

250

300

350

400

450

500

550

600

Tens

ile s

tres

s/M

Pa

1.4 1.5 1.6 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3

Tensile strain/%1.4 1.5 1.6

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3

Tensile strain/%1.4 1.5 1.6

0

50

100

150

200

250

300

350

400

450

500

100 °C-24 hr100 °C-48 hr150 °C-24 hr150 °C-48 hr250 °C-24 hr250 °C-48 hr

100 °C-24 hr100 °C-48 hr150 °C-24 hr150 °C-48 hr250 °C-24 hr250 °C-48 hr

Loading at 100 °C

Unloading at 100 °C

Recovery at 300 °C

Loading at 100 °C

Unloading at 100 °C

Recovery at 300 °C

Loading at 100 °C

Unloading at 100 °C

Recovery at 300 °C

(a) (b)

(c)

Fig. 6 Shape memory effect curves of a Cu–Al–Ni–0.4 mass% Co, b Cu–Al–Ni–0.7 mass% Co, and c Cu–Al–Ni–1.0 mass% Co

1282 S. N. Saud et al.

123

423 K for 24 h alloys has exhibited the highest strain

recovery. In addition, the flow strength of the austenite

phase may also significantly contribute to the enhancement

of SME that can be obtained as a result of the reordering of

martensite phase. On the other hand, the stiffened parent

phase is capable of accommodating the transformation

strain elastically, which can lead to maintaining the

coherent martensite/austenite interfaces during the direct

and/or reverse transformation with heat treaditions are

shown inted to a certain temperature or by applying a small

stress. However, the precipitation effects may also play an

important role in the enhancement of SME that have been

varied during the aging treatment. This effect can be

explained in terms of: (i) martensite $ transformation

temperature; (ii) the amount of induced stress during the

forward or backward transformation; (iii) the interaction

between the precipitates and dislocations. For the

0.7 mass% Co-aged alloys, the highest shape recovery

occurred with 423 K for 24 h alloy as shown in Fig. 6b.

The reason behind this improvement also relates to the

same phenomena previously mentioned with 0.4 mass%

Co-aged alloys. The 1.0 mass% Co-aged alloys show the

highest performance for the shape recovery with 523 K for

48 h alloy as shown in Fig. 6c, which obtained a completed

shape recovery (100 % of the original shape) after being

heated above the Af without obtaining any residual strain.

This may be attributed to the high volume fraction and size

of c01 phase. As mentioned previously, the increment in the

size and volume fraction of the plate-like can lead to

improving the SME by increasing the strain around the c01plate-like as proven by the micrographs in Fig. 1a–c. On

the other hand, the higher volume fraction and a large size

of c2 precipitates can hinder the movements of dislocations

and martensite variant interfaces associated with a decre-

ment of permanent strain, and thus, increase the shape

recovery.

Conclusions

Aging of the Cu–Al–Ni–Co SMAs at different tempera-

tures and times shows considerable variations in transfor-

mation temperatures and transformation hysteresis due to

successive martensitic transition and precipitate formation.

These variations are suitable factors to select this alloy to

be used for high temperature applications and fluctuating

environmental conditions. Moreover, aging also has a

noticeable influence on the stress–strain properties of the

alloys within the aging temperatures and time considered.

A significant variation in the fracture strain and fracture

stress value with the aging of the martensite alloys was

observed in the phase transition regime. With aging, the

shape memory effects were varied in terms of strain

recovery and this could be as a result of the variation of the

phase morphology and volume fraction of the c2 precipi-

tates. However, the highest strain recovery was observed

with the Cu–Al–Ni–1.0 mass% Co SMA after being aged

at 523 K for 48 h, in which it exhibited a 100 % recovery

of the original shape after being heated above the tem-

perature of Af.

Acknowledgements The author(s) would like to thank the Malay-

sian Ministry of Higher Education (MOHE) and Universiti Teknologi

Malaysia for providing the financial support and facilities for this

research, under Grant No. R.J130000.7824.4F150.

References

1. Gandhi V, Thompson BD. Smart materials and structures.

Springer: New York; 1992.

2. Duerig TW, Melton KN, Stoeckel D, Wayman CM. Engineering

aspects of shape memory alloys. Mater Des. 1990;11(6):301.

3. Van Humbeeck J. Non–medical applications of shape memory

alloys. Mater Sci Eng A. 1999;273–275:134–48.

4. Schetky LM. A review of: ‘‘Engineering aspects of shape mem-

ory alloys’’ by T.W. Duerig, K.N. Melton, D. Stockel, and CM.

Waymanb Butterworth-Heinemann 499 pages, hardcover, 1990.

Mater Manuf Process. 1993;8(4–5):567–9.

5. Recarte V, Hurtado I, Herreros J, No ML, San Juan J. Precipi-

tation of the stable phases in Cu–Al–Ni shape memory alloys.

Script Mater. 1996;34:255–60.

6. Recartea V, Perez–Saeza RB, Bocanegrac EH, Noc ML, San Juan

J. Dependence of the martensitic transformation characteristics

on concentration in Cu–Al–Ni shape memory alloys. Mater Sci

Eng A. 1999;273–275:380–384.

7. Recarte V, Perez–Landazabal JI, No ML, San Juan J. Study by

resonant ultrasound spectroscopy of the elastic constants of the bphase in Cu–Al–Ni shape memory alloys. Mater Sci Eng A.

2004;370(1–2):488–91.

8. Zarubova N, Gemperlova J, Gartnerova V, Gemperle A. Stress–

induced martensitic transformations in a Cu–Al–Ni shape mem-

ory alloy studied by in situ transmission electron microscopy.

Mater Sci Eng A. 2008;481–482:457–61.

9. Otsuka K, Wayman CM. Shape memory materials. Reprint,

illustrated ed. Cambridge University Press: Cambridge; 1999.

10. Aydogdu A, Aydogdu Y, Adiguzel O. The influence of aging pn

the martensite ordering and stabilization in shape memory Cu–

Al–Ni alloys. Mater Res Bull. 1997;32:507–13.

11. Suresh N, Ramamurty U. Aging response and its effect on the

functional properties of Cu–Al–Ni shape memory alloys. J Alloy

Compd. 2008;449(1–2):113–8.

12. Sutou Y, Koeda N, Omori T, Kainuma R, Ishida K. Effects of

aging on stress-induced martensitic transformation in ductile Cu–

Al–Mn–based shape memory alloys. Acta Mater. 2009;57(19):

5759–70.

13. Sari U, Kirindi T, Ozcan F, Dikici M. Effects of aging on the

microstructure of a Cu–Al–Ni–Mn shape memory alloy. Int J

Miner Metall Mater. 2011;18(4):430–6.

14. Balo SN, Neslihan S. Effects of thermal aging on transformation

temperatures and some physical parameters of Cu–13.5 wt% Al–

4 wt% Ni shape memory alloy. Thermochim Acta. 2012;536:

1–5.

15. Tang SM, Chung CY, Liu WG. Preparation of Cu–AI–Ni–based

shape memory alloys by mechanical alloying and powder met-

allurgy method. J Mater Process Technol. 1997; 63: 307–132.

Thermal aging behavior in Cu–Al–Ni–xCo 1283

123

16. Saud S, Hamzah E, Abubakar T, Zamri M, Tanemura M. Influ-

ence of Ti additions on the martensitic phase transformation and

mechanical properties of Cu–Al–Ni shape memory alloys.

J Therm Anal Calorim. 2014;118:111–2.

17. Saud S, Hamzah E, Abubakar T, Bakhsheshi-Rad HR, Zamri M,

Tanemura M. Effects of Mn additions on the structure,

mechanical properties, and corrosion behavior of Cu–Al–Ni

shape memory alloys. J Mater Eng Perform. 2014: 1–10.

18. Yildiz K, Kok M. Study of martensite transformation and

microstructural evolution of Cu–Al–Ni–Fe shape memory alloys.

J Therm Anal Calorim. 2014;115(2):1509–14.

19. Van Humbeeck J. Damping capacity of thermoelastic martensite

in shape memory alloys. J Alloy Compd. 2003;355(1):58–64.

20. Zarubova N, Gemperle A, Novak V. Ageing phenornena in a Cu–

Al–Ni Alloy. Le J Phys IV. 1997; 7(C5):C5–281–C5–6.

21. Zarubova N, Gemperle A, Novak V. Initial stages of c2 precipi-

tation in an aged Cu–Al–Ni shape memory alloy. Mater Sci Eng

A. 1997;222(2):166–74.

22. Picornell C, Pons J, Cesari E. Effects of thermal ageing in,

B-phase in Cu–Al–Ni single crystals. J Phys–Colloq. 1997;7(5):

323–8.

23. Karagoz Z, Canbay CA. Relationship between transformation

temperatures and alloying elements in Cu–Al–Ni shape memory

alloys. J Therm Anal Calorim. 2013;114(3):1069–74.

24. Swann P, Warlimont H. The isothermal ordering of a b copper–

aluminum alloy. Acta Metall. 1962;10(9):903–5.

25. Jellison J, Klier E. The cooling transformations in the beta eu-

tectoid alloys of the Cu–Al system. Trans Metall Soc AIME.

1965;233:1694–702.

26. Suresh N, Ramamurty U. The effect of ageing on the damping

properties of Cu–Al–Ni shape memory alloys. Smart Mater

Struct. 2005;14(5):N47.

27. Bouabdallah M, Baguenane-Benalia G, Saadi A, Cheniti H, Ga-

chon J-C, Patoor E. Precipitation sequence during ageing in b1

phase of Cu–Al–Ni shape memory alloy. J Therm Anal Calorim.

2013;112(1):279–83.

28. Suresh N, Ramamurty U. Effect of aging on mechanical behavior

of single crystal Cu–Al–Ni shape memory alloys. Mater Sci Eng

A. 2007;454–455:492–9.

29. Rodriguez P, Guenin G. Thermal aging behaviour and origin of a

Cu–Al–Ni shape memory alloy. Mater Sci Eng A. 1990;129:

273–7.

30. Xuan Q, Bohong J, Hsu TY. The effect of martensite ordering on

shape memory effect in a copper–zinc–aluminium alloy. Mater

Sci Eng. 1987;93:205–11.

31. Cullity BD, Stock SR. Elements of x-ray diffraction. Prentice

Hall: NJ; 2001.

32. Patterson AL. The Scherrer formula for x-ray particle size

determination. Phys Rev. 1939;56(10):978–82.

33. Duerig TW, Melton KN, Stockel D, Wayman CM. Engineering

aspects of shape memory alloys. Books on demand; 1990.

34. Ma J, Subirana JI, Pu ZJ, Wu K–H, editors. Influence of aging

treatment on Ni46Ti44Ta10 shape memory alloy; 1999.

35. Martynov V, Kokorin V. The crystal structure of thermally-and

stress-induced martensites in Ni2MnGa single crystals. J Phys III.

1992;2(5):739–49.

36. Wang W, Liu Z, Zhang J, Chen J, Wu G, Zhan W, et al. Ther-

moelastic intermartensitic transformation and its internal stress

dependency in Ni 52 Mn 24 Ga 24 single crystals. Phys Rev B.

2002;66(5):052411.

37. Otsuka K, Ren X. Martensitic transformations in nonferrous

shape memory alloys. Mater Sci Eng A. 1999;273:89–105.

38. Michler T, Naumann J. Influence of high pressure hydrogen on

the tensile and fatigue properties of a high strength Cu–Al–Ni–Fe

alloy. Int J Hydrogen Energy. 2010;35(20):11373–7.

39. Roh D, Kim J, Cho T, Kim Y. Tensile properties and micro-

structure of microalloyed Cu–Al–Ni–X shape memory alloys.

Mater Sci Eng A. 1991;136:17–23.

40. Sarı U, Kırındı T. Effects of deformation on microstructure and

mechanical properties of a Cu–Al–Ni shape memory alloy. Mater

Charact. 2008;59(7):920–9.

41. Yang QP, Sun L, Feng YZ. Experiment of beam crack repair

using shape memory alloy. Appl Mech Mater. 2011;71–78:

1707–10.

42. Gollerthan S, Young M, Baruj A, Frenzel J, Schmahl W, Eggeler

G. Fracture mechanics and microstructure in NiTi shape memory

alloys. Acta Mater. 2009;57(4):1015–25.

43. Miyazaki S. Development and characterization of shape memory

alloys. Springer: Vienna; 1996.

44. Morawiec H, Gigla M. Effect of ageing on shape recovery in Cu–

Al–Ni alloy with Ti?B additions. Acta Metall Mater. 1994;42(8):

2683–6.

1284 S. N. Saud et al.

123


Recommended