Open access
Technical Papers
Jan 23, 2021

Orientation Dependence of Plasticity and Fracture in Single-Crystal Superelastic Cu-Al-Mn SMA Bars

Publication: Journal of Materials in Civil Engineering
Volume 33, Issue 4

Abstract

The objective of this paper is to examine the orientation dependence of plasticity and fracture in single-crystal superelastic Cu-Al-Mn shape memory alloy (SMA) bars. For this purpose, we prepared 14 single-crystal superelastic Cu-Al-Mn SMA bars of 15-mm diameter and 140-mm length. The crystal orientation was measured using electron backscatter diffraction. This work involved cyclic tension tests up to 10% strain and consecutive monotonic tension tests up to fracture. From these results, strong orientation dependence was observed in plasticity and fracture. Ductility was poor or moderate when the orientation of the specimen was close to the <101> direction. On the other hand, a highly ductile response was observed when the orientation was close to the <112>, <113>, or <001> direction. In these specimens, the fracture strain ranged from 47% to 92%. The large rotation of crystal lattice and the propagation of slip band along the long distance in the longitudinal direction are the reasons for the highly ductile response. Such a highly ductile response is desirable in structural materials, especially in seismic applications for civil structures like buildings and bridges.

Introduction

Since the advent of Ni-Ti shape memory alloys (SMAs) in 1963, tremendous progress has been made in the development and characterization of SMAs. Among many types of SMAs, Ni-Ti SMAs have been the most dominant in practical applications (Otsuka and Wayman 1998; Kainuma 2018). The difficulty in cold work of Ni-Ti SMAs, however, limits their use only to simple forms like wires and tubes. On the other hand, Cu-Al-Mn SMAs, discovered based on phase diagram study in the late 1990s, have excellent cold workability (Kainuma et al. 1996, 1998). The lower material cost and higher workability make Cu-Al-Mn SMAs strong candidates to pioneer new practical applications, such as seismic applications (Ozbulut et al. 2011; Chang and Araki 2016), which are the focus of this work. SMAs demonstrate a shape recovery property upon heating or unloading. The former is called shape memory effect and the later superelasticity. In seismic applications, superelasticity of SMAs is often exploited to minimize residual deformations in civil structures like buildings and bridges after earthquakes.
Due to the strong crystallographic anisotropy in transformation strain of Cu-Al-Mn SMAs—which is different from Ni-Ti SMAs—the mismatch of grain orientations at grain boundaries easily introduces dislocations in polycrystalline Cu-Al-Mn SMAs (Sutou et al. 2005, 2013). The accumulation of such dislocations degrades superelasticity and leads to intergranular fracture. One solution to this problem is texture control, where uniform grain orientation is obtained by cold work prior to heat treatment (Sutou et al. 2002). Another solution is single crystallization. It is well known that Cu-Al-Mn SMAs, as well as other Cu-based SMAs like Cu-Al-Ni (Otsuka et al. 1976) and Cu-Al-Be (Qiu and Zhu 2014), have excellent superelasticity in the form of single crystals. However, the cost of producing single-crystal Cu-based SMAs has been much higher than that of polycrystalline Ni-Ti SMAs. The Czochralski method and the Bridgman method are popular techniques for producing single crystals (Reed 2006); however, these methods have very slow crystal growth, which is a major disadvantage in commercialization of single-crystal SMAs. When SMAs are used in seismic applications, the use of large-diameter bars and/or thick plates is required. Commercialization of such large single-crystal SMAs has been considered extremely difficult, if not impossible.
Under such circumstances, a technique called cyclic heat treatment was developed, wherein only a repetition of heat treatments is necessary to accelerate the grain growth in Cu-Al-Mn SMA bars (Omori et al. 2013). This technique enables production of bamboo-like microstructures, wherein the size of grains is larger than the bar diameter. Although excellent superelasticity can be obtained by realizing the bamboo-like microstructure even in large-diameter bars, the risk of intergranular fracture remains if the mismatch of grain orientations is significant. To overcome this difficulty, the process of cyclic heat treatment was improved to further accelerate grain growth (Kusama et al. 2017). The improved process demonstrates that it is possible to produce large single-crystal superelastic bars of 15-mm diameter and 700-mm length within a reasonable time frame in the Cu-Al-Mn SMA. As such, improved cyclic heat treatment provides an avenue for commercializing large single-crystal superelastic Cu-Al-Mn SMAs in seismic application, which has been considered virtually impossible until now.
Of course, some issues remain to be addressed before commercialization of single-crystal Cu-Al-Mn SMAs. As mentioned above, strong anisotropy exists in mechanical properties of superelastic Cu-Al-Mn SMAs. Sutou et al. (2002, 2005, 2013) extensively studied the orientation dependence of superelasticity. It has been reported that the recoverable strain, due to transformation strain in the L21/6M(18R) martensitic transformation, changes in the range of 2% and 10% in tension, depending on the loading direction (Sutou et al. 2002). The temperature dependence of critical stress for martensitic transformation (apparent yield stress) is inversely proportional to the transformation strain. Therefore, critical stress becomes high when the transformation strain is low at a fixed temperature. These mechanical properties related to superelasticity have been experimentally and theoretically investigated. Nevertheless, the orientation dependence of plastic and fracture response—or the response after reaching the maximum recoverable strain up to fracture—is still not well known. Because the ductility of materials plays a critical role in seismic applications, characterizing its orientation dependence is an essential task toward commercialization.
The objective of this paper is to study the orientation dependence of plasticity and fracture in single-crystal superelastic Cu-Al-Mn SMA bars. The key components of the present study are: (1) systematic investigation focusing on plasticity and fracture; and (2) use of large single-crystal SMA bars as test specimens. The study performed cyclic tension tests up to 10% strain and consecutive monotonic tension tests up to fracture. The test specimens were 14 single-crystal superelastic Cu-Al-Mn SMA bars of 15-mm diameter and 140-mm length, with a variety of grain orientations. The microstructure and the fracture surface of the SMA bars were analyzed to consider the possible reasons for the large difference observed in the plastic and fracture responses.

Materials and Methods

Materials Preparation

An ingot of Cu 17 atomic % of AL and 11.4 atomic % of Mn alloy was prepared from 99.99% Cu, 99.99% Al, and 99% Mn by melting in a high-frequency vacuum induction furnace. Then, the ingot was hot forged down to 40-mm diameter at 800°C. A long bar of 15-mm diameter was obtained by repeating cold rolling and drawing at the cold working rate of 30% with an intermediate annealing at 520°C for 60  min. The long bar was cut into bars of 300-mm length and then heat treated under air to produce single-crystal bars. In the heat treatment, the bars were first cooled down from the β phase at 900°C to the α+β phase at 500°C at the rate of 1°C/min. After keeping the temperature at 500°C for 1 h, the temperature was heated up to the β phase at 900°C at the rate of 1°C/min. This process was repeated 10 times and quenched. The bars were etched using an acid ferric chloride solution (100  g Fe3Cl, 250  mL HCl, and 1,000  mL water) to examine the existence of grain boundaries by naked eye. Fig. 1 shows the obtained bars after etching. Although the manufacturing numbers are tentatively shown here, the sample number will be specified after the electron backscatter diffraction (EBSD) measurement. Note that marking line was used to emphasize the grain boundaries in the photograph. It was confirmed that 13 single-crystal bars were obtained from 16 bars heat treated except the 4th, 11th, and 12th bars from the top. In the 4th bar, the grain boundary was observed near the right end. These 13 single-crystal bars, along with the 4th bar, were used as test specimens in the following study.
Fig. 1. Photographs of the bars after heat treatment and etching.
From each bar, three bars of 10-mm length and one bar of 140-mm length were obtained by cutting. The three 10-mm-long bars were used to measure grain orientation, hardness, and transformation temperatures. The bar of 140-mm was machined to the shape shown in Fig. 2 to be used in tension tests. The diameters of center and end portions were 8 and 12 mm, respectively. The lengths of the central and end portions were 56 and 30 mm, respectively. The curvature radius of the transition zone between the center and end portions was 20 mm. All the bars after machining were aged at 130°C for 30  min to stabilize the superelastic property.
Fig. 2. Schematic drawing and photographs of tension test specimens.

Test Methods

Grain orientation, hardness, and transformation temperatures were measured as follows. After polishing and buffing along the longitudinal direction, the grain orientation of the 10-mm-long bar was measured using EBSD. Obtained results were plotted as the inverse pole figure using an orientation imaging microscopy analysis software. After similar polishing and buffing, the Vickers hardness was measured at room temperature (20°C) using a micro-Vickers hardness tester. Although there are a variety of methods for evaluating hardness, the Vickers hardness test was selected to obtain accurate results from the small samples. The applied load was 1.96  N and the loading time was 15  s. The transformation temperatures were measured using a differential scanning calorimeter. The mass of the specimen was between 20 and 50  mg, and the heating and cooling rate was 0.17°C s1.
The mechanical properties were measured by tension tests at room temperature using a precision tensile tester with load capacity of 50 kN and a noncontact digital video extensometer. The gauge length was 40 mm in the strain measurements. The photograph of the setup of the specimen in the tension test equipment is shown in Fig. 3. Cyclic tension tests up to 10% strain were performed first. Then, monotonic tension tests were performed up to fracture. Throughout the loading process, the strain rate was controlled to be 3×104  s1. In the cyclic tension tests, the target strain amplitudes were increased from 1% to 10% with an increment of 1%. Only one tension cycle was applied for each target strain amplitude. The forced displacement was applied in the loading direction until the target strain was reached. Then, the forced displacement was reversed and applied in the unloading direction until the stress reached zero. This way, only tension load was applied during the whole cyclic loading process. After the last cycle, monotonic tension loading was applied up to the fracture.
Fig. 3. Setup of the tension test.

Results

The black circles in Fig. 4(a) show the grain orientation of each bar along the longitudinal direction. As seen from the inverse pole figure, the grain orientations are distributed randomly around the orientations <101>, <112>, <113>, and <001>. Fig. 4(b) shows the numbering of the samples on the contour curves of the transformation strains [8, 9]. For later discussions, the numbering of the bar specimens was performed in increasing order of the estimated transformation strains. The specimens can be classified into four groups in accordance with the orientations. The first group, G1, is from No. 1 to No. 6, whose orientation is around <101>. The second, G2, is from No. 7 to No. 10, around <112>. The third, G3, is from No. 11 and No. 12, around <113>. And the fourth, G4, is No. 13 and No. 14, around <001>. Fig. 5 shows the stress–strain curves obtained from the tension tests. Table 1 shows the values of Young’s modulus E, the transformation stress σt, the fracture stress σf, the transformation strain εt, and the fracture strain εf obtained from the stress–stress curves shown in Fig. 5. The definitions of these variables are schematically shown in Fig. 6. The measured values of the Vickers hardness HV, the martensite start temperature Ms, and the austenite finish temperature Af are also shown in Table 1. The orientation dependence of Ms and Af is negligibly small and within the range of the measurement error. The variations in HV are also minor, and no clear orientation dependence can be seen. On the other hand, strong orientation dependence exits in the other variables, which are discussed in more detail in the following sections.
Fig. 4. (a) Grain orientation along longitudinal direction of single-crystal samples; and (b) numbering of the samples on the contour map of isotransformation strain.
Fig. 5. Results of the tension tests.
Table 1. Vickers hardness HV, Young’s modulus E, transformation stress σt, maximum stress σmax, the fracture stress σf, transformation strain εt, fracture strain εf, martensite start temperature Ms, and austenite finish temperature Af obtained from the tests
SampleHVE (GPa)σt (MPa)σmax (MPa)σf (MPa)εt (%)εf (%)Ms (K)Af (K)Reference (Product no.)
1242622458358106.516.022223592-2-5
2253802559309086.910.622023192-2-15
3241702507607.17.122123292-2-7
423670220>1,000>1,0007.121923192-2-1
5237652295304107.621.822323992-2-13
6237602317904487.420.122423392-2-13
7235402283202807.972.022723892-2-10
8235422002952398.367.622223592-2-8
9234432103102518.247.322423792-2-16
10229442002952409.070.523224292-2-6
11214361692401319.892.222624092-2-4
12230371612481299.364.722123292-2-2
132302218026019010.287.222823992-2-14
142262317925018110.379.822623992-2-9
Fig. 6. Schematic illustration showing the definitions of specific values for characterization of mechanical and superelastic properties.

Discussions

Elasticity and Superelasticity

This section discusses the orientation dependence of response up to the transformation strain. Fig. 7 illustrates the stress–strain curves up to 10% strain obtained from the tension tests. Fig. 8 shows the measured values of the transformation strain εt and those of the transformation stress σt, plotted on the isotransformation strain contour curves. From Fig. 7, excellent superelasticity can be seen in each orientation. After aging heat treatment, Cu-Al-Mn alloy has the L21 ordered structures. It is known that the martensite induced from the L21 ordered structures (Kainuma et al. 1998) has the 6M long period stacking ordered structures (Kainuma et al. 1996). It is therefore considered that the stress-induced β(L21)β(6M) transformation took place in the superelastic response. Fig. 8(a) shows agreement between the measured and calculated transformation strains in each orientation. This result supports the validity of the transformation strains between L21 and 6M calculated by Sutou et al. (2002, 2005). It is found from the following Clausius–Clapeyron Eq. (1) that the transformation stress is inversely proportional to the transformation strain
dσtdT=ΔSεt
(1)
σt1εtMsTΔSdT
(2)
where T = temperature; and ΔS = entropy change per volume (Wollants et al. 1979). Fig. 8(b) shows the calculated transformation stress using Eq. (2) and transformation strain shown in Fig. 7(a), Here, the transformation stress in <001> is assumed to be 170 MPa. In accordance with the calculation, from Figs. 8(a and b), it can be observed that the transformation stress tends to become lower as the transformation strain becomes larger.
Fig. 7. Orientation dependence of the stress–strain curves up to 10% strain.
Fig. 8. Measured values on the isotransformation strain contour curves: (a) transformation strain εt (%); and (b) transformation stress σt (MPa).
Fig. 9 shows the values of Young’s modulus obtained from the tension tests plotted on the inverse pole figure. The iso-Young’s modulus contour curves were calculated from the elastic constants (Xu et al. 2020). These values can be classified into four groups: G1 60 to 80 GPa; G2 40 to 44 GPa; G3 36 and 37 GPa; G4 22 and 23 GPa. The observed orientation dependence in Young’s modulus of the present alloy is similar to that of the Cu-Al-Ni alloy reported by Horikawa et al. (1988).
Fig. 9. Measured values of Young’s modulus E.

Plasticity and Fracture

Classification

From Fig. 5, the response after the transformation strain can be roughly classified into the following two cases: Case 1, with strong hardening and poor or moderate ductility, seen in the specimens in Group G1 (Specimens 1 to 6); and Case 2, with weak hardening and superior ductility, observed in the specimens in Groups G2, G3, and G4 (Specimens 7 to 14). The response in Group G1 can be further divided into the following two types:
Case 1a, Specimens 1 to 4
Plastic deformation after hardening can be seen in Specimens 1 and 2, although the fracture strain was relatively small. Brittle fracture took place during hardening in Specimen 3 without any plastic deformation. Testing was terminated before reaching fracture in Specimen 4 because the tensile strength reached the loading capacity of the tensile tester.
Case 1b, Specimens 5 and 6
Superelasticity can be observed after hardening. These specimens had moderate ductility with the fracture strain of over 20%.
The superior ductility without strong hardening observed in Groups G2 to G4 are noteworthy from the viewpoint of seismic applications, where ductility is critical. The small or moderate hardening is important in avoiding brittle fracture at connections (Kise et al. 2018). To the authors’ knowledge, such superior ductility of single-crystal SMAs—especially in the form of bars of 8-mm diameter and 40-mm gauge length—has not been reported in the literature; however, the orientation dependence of superelasticity (Horikawa et al. 1988) and multistage superelasticity (Otsuka et al. 1979) were reported for single-crystal superelastic Cu-Al-Ni SMAs. It has been reported that the 6M martensite is transformed to 2M by stress in Cu-Al-Zn-Mn (Šittner et al. 1998), and multistage superelasticity is possibly caused by a similar transformation, although such transformation has not been reported in this alloy, probably due to lower ductility in the polycrystalline form.
To analyze the fracture modes mentioned above, we examine the specimens after the tension tests. Fig. 10 shows the photograph of all the specimens after the tension tests. Fig. 11 shows the close-up view of the specimens, where each specimen is rotated 0°, 90°, 180°, and 270° with respect to the circumferential direction. Fig. 12 shows typical deformation patterns in Cases 1a, 1b, and 2 of classifications mentioned above. Table 2 shows the values of the maximum and minimum diameters at the location 10-mm apart from the fracture section, to show the change of the cross-section shape due to the propagation of the slip band. Table 2 also shows the values of the fracture strain and the reduction of area (RA) at the fracture section, for reference, where RA was calculated assuming ellipsoidal shapes.
Fig. 10. Photograph of whole bars for all the specimens.
Fig. 11. Photographs of each specimen around the fractured portion (rotated with 90° increments).
Fig. 12. Schematic illustration showing deformation of specimens corresponding to stress–strain properties: Case 1a, Specimen 3; Case 1b, Specimen 5; and Case 2, Specimen 12.
Table 2. Maximum and minimum diameters at the location 10-mm apart from the fractured section, and fracture strain εf and reduction of area at the fracture section
SampleMaximum diameter (mm)Minimum diameter (mm)εf (%)Reduction of area (%)
18.027.5616.0
28.037.9310.6
38.038.027.1
48.048.04
58.048.0421.852.5
68.037.9920.144.6
77.215.2172.057.5
87.255.3167.647.0
97.305.2147.353.7
107.274.8070.562.5
117.325.1392.287.1
127.214.6164.782.1
137.065.1087.262.2
147.054.9979.864.4
From these figures and tables, strong orientation dependence can be observed also in the fracture modes. We can summarize the orientation dependence of the fracture mode in accordance with the classification of the response after the transformation strain.
Case 1a, Specimens 1 to 4
Fracture surface is about π/4 radian from the longitudinal axis of each bar. This implies that shear fracture took place in these specimens. In Specimens 1 and 2, periodic striped patterns parallel to the fracture surface can be seen in the reduced section portion. At the location 10-mm apart from the fractured section, small reduction of the diameter can be seen in Specimens 1 and 2, while little reduction can be seen in Specimens 3 and 4. This suggests that the slip band propagated along the longitudinal direction in Specimens 1 and 2, but did not propagate in Specimen 3. This agrees with the observation that the fracture strains of Specimens 1 and 2 were significantly larger than that in Specimens 3 and 4.
Case 1b, Specimens 5 and 6
Deformation is localized around the fracture section. Significant reduction of sectional area can be seen at the fracture section. On the other hand, almost no reduction can be seen at the location 10-mm apart from the fracture section, as shown in Table 2. Such a fracture mode is similar to that typically observed in tension tests of low-carbon steel.
Case 2, Specimens 7 to 14
In each specimen, a large reduction of sectional area can be seen in a wide region along the longitudinal direction of the specimen. Both ends of this region are indicated by the white arrows in Fig. 10. Similar to Specimens 1 and 2, a periodic striped pattern can be seen clearly in the region. This also suggests that the slip band propagated over a long distance along the longitudinal direction. The key difference from Specimens 1 and 2 is that these specimens have much larger rotation angles as shown in Fig. 10. It can be clearly observed from Figs. 4 and 10 that both the length of the slip band propagation and the rotation angle of the crystal lattice have strong correlations with the fracture strain.
The Schmid factor is often used to explain the reasons for orientation dependence in plasticity and fracture from the viewpoint of crystal structure. From the results discussed above, it is inferred that the value of the Schmid factor is low in Case 1 and high in Case 2. However, when we assume that the slip system is {111}110 for the face-centered-cubic (fcc) martensite, the significant difference in the yield stress cannot be explained by the Schmid factor. From the stress–strain curves, the two-step martensitic transformation from 6M to detwinned 2M (fcc) seems to occur in Case 1, as observed in other Cu-based alloys (Otsuka et al. 1979), but does not occur in Case 2. This difference in the twin structure in martensite might be related to the large difference in stress. To determine the effective Schmid factor, information on the crystal orientation prior to plastic deformation and on the slip system is necessary. Although this topic is beyond the scope of this paper, the authors are currently working on it and plan to report the results in future work.

Fractography

In this section, we perform failure analysis using fractography. Figs. 1316 show typical examples of the scanning electron microscope (SEM) views of the fracture surface. Specimens 3 and 5 represent Cases 1a and 1b, respectively. The fracture surface of Type 2 response can be classified into two groups: Case 2a, Specimens 11 and 12; and Case 2b, Specimens 7–10, 13, and 14. Specimens 12 and 13 represent Cases 2a and 2b, respectively. The observations made from these specimens are described as follows.
Fig. 13. SEM views of the fracture surface of Specimen 3: (a) whole section view; and (b) close-up view.
Fig. 14. SEM views of the fracture surface of Specimen 5: (a) whole section view; and (b) close-up view.
Fig. 15. SEM views of the fracture surface of Specimen 12: (a) whole section view; and (b) close-up view.
Fig. 16. SEM views of the fracture surface of Specimen 13: (a) whole section view; and (b) close-up view.
Case 1a, Specimen 3
It is observed from Fig. 5 that brittle fracture took place during the sudden hardening right after the transformation strain. From Figs. 11 and 12(a), shear fracture can be observed, where the angle of fracture surface is close to π/4 radian (Callister and Rethwisch 2011; Hosford 2010).
Fig. 13 shows the SEM views of the fracture surface. Fig. 13(a) shows the whole view. The observed patterns can be classified into three regions separated by the white dashed lines. The three photographs in Fig. 13(b) show the close-up view of each region. In Region R1, both equiaxed and elongated dimples exist, which are observed under tension and shear stresses, respectively. In Region R2, elongated dimples and river patterns can be observed at the outer and central portions, respectively. In Region R3, river patterns scatter on a flat surface with a central ridge.
The mechanism of the fracture can be considered as follows (Becker and Lampman 2002). First, fracture took place in Region R1 due to microvoid coalescence. Then, fracture propagated into Region R2, where shear and cleavage fractures coexist. Finally, cleavage fracture occurred in Region R3. The vertical light gray lines, shown in Region R3 in Fig. 13(a), indicate the parallel ridges resulted from brittle fracture. The inclined stripe pattern arises from the contrast between the smooth surface with low dislocation density and the rough surface with river patterns, where dislocation density is high. Note that the fracture surfaces of Specimens 1 and 2 look similar to that in Region R3 of Specimen 3. This suggests that cleavage fracture also took place in Specimens 1 and 2.
Case 1b, Specimen 5
From the stress–strain curves shown in Figs. 5 and 12(b), it is inferred that plastic deformation took place from about 8% strain where the stress–strain curve becomes downward convex. It is also inferred that necking took place at about 12% strain, after which stress continuously decreases. From Figs. 10, 11, and 12(b), it is observed that necking led to significant reduction of the cross-section area.
Fig. 14(a) shows the whole fracture section. Fig. 14(b) shows the close-up view of the rectangular portion whose boundary is indicated by the white dashed line. From the whole view in Fig. 14(a), it is observed that the circular cross section was reduced to an ellipsoidal one, and that the reduction of the sectional area was localized around the fracture section. The fracture section can be roughly classified into the following two portions: (1) central fibrous portion and (2) outer smoother portion. The close-up view in Fig. 14(b) shows that the central fibrous potion has equiaxed dimples and the outer smoother portion has shallow elongated dimples. These dimple patterns indicate that fibrous fracture took place at the central portion under tension stress, and then shear fracture took place in the outer portion (Becker and Lampman 2002).
Case 2a, Specimen 12
From the stress–strain curves shown in Figs. 5 and 12(c), it is inferred that plastic deformation took place from about 17.5% strain where the stress–strain curve becomes downward convex, and no hardening can be seen from about 20% strain up to fracture. From Figs. 10 and 11, propagation of the slip band can be clearly observed. The large difference between the maximum and minimum diameters in Table 2 indicates that a large rotation of crystal lattice took place in the slip band portion.
From the whole section view shown in Fig. 15(a), the stairlike fracture section can be observed. The close-up view in Fig. 15(b) shows the ridges in the stairlike pattern. On the left (dark) side of the ridges, there are ripple patterns, which indicate the separation of the slip plane (Koterazawa 1974; Beachem and Meyn 1968). On the right (light) side of the ridge, elongated dimples can be seen. The ripple patterns and the elongated dimples can be seen alternatively. Let f, s, and m, schematically illustrated in the figure, indicate the normal of the fracture surface, the normal of the slip plane, and the member axis. Note here that s is perpendicular to the surface on which ripple patterns were observed. It is observed that shear fracture took place so that the plane determined by f and m is almost orthogonal to the plane determined by s and m.
Case 2b, Specimen 13
From Fig. 16, it is observed that the response of Specimen 13 is similar to that of Specimen 12, while Specimen 13 has larger fracture strain than Specimen 12. Similar to Specimen 12, Specimen 13 exhibits a significant difference between the maximum and minimum diameters in the slip band portion, which also indicates that a large rotation of crystal lattice took place. On the other hand, it is observed from Figs. 10, 11, and 12 that the slip band propagated along almost the full length of the reduced-section portion in Specimen 13. The longer propagation length in Specimen 13 is the reason for the larger fracture strain.
Fig. 16(a) shows the whole view of the fracture section. Fig. 16(b) shows the close-up view of the region indicated by the white dashed line in the whole view. As shown in Fig. 16(a), the fracture section has a band region at its center, indicated by the arrows. Serpentine glide can be seen at the central portion in Fig. 16(b), which indicates the existence of slip plane at the band region in Fig. 16(a). Elongated dimples can be seen around the serpentine glide. From these observations, it is inferred that shear fracture took place so that all f, s, and m lie on almost the same plane, and that the angle between s and m is larger than that between f and m.

Conclusions

We examined the orientation dependence of mechanical properties of single-crystal superelastic Cu-Al-Mn SMA bars, focusing on plasticity and fracture. The following conclusions can be drawn from the study. (1) It was confirmed that the orientation dependence in elastic and superelastic response can be predicted quantitatively by calculations based on the theories presented in the literature. (2) Strong orientation dependence was observed in plasticity and fracture. To the authors’ knowledge, such orientation dependence has not been reported in the literature. Ductility was poor or moderate when the orientation of the specimen was close to the <101> direction. On the other hand, a highly ductile response, with fracture strain up to 92%, was observed when the orientation was close to the <112>, <113>, or <001> direction. The large rotation of the crystal lattice and the propagation of the slip band along the long distance in the longitudinal direction are the main reasons for the highly ductile response. Such a highly ductile response is desirable in a structural material, especially for seismic applications.

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

This research was supported by Grant-in Aids for Challenging Research (Exploratory) #17K18825 and for the Promotion of Joint International Research #16KK0149 provided by the Japan Society for the Promotion of Science (JSPS). The authors thank Prof. Y. Sutou of Tohoku University and Prof. M. Nishida and Prof. em. K. Tsuzaki of Kyushu University for their helpful discussions. The authors are grateful to Mr. Y. Oizumi for his assistance in performing the tension tests.

References

Beachem, C. D., and D. A. Meyn. 1968. “Sources of fracture toughness: The relation between KIc and the ordinary tensile properties of metals.” In Applications related phenomena in titanium alloys, edited by H. Conrad, R. Jaffee, H. Kessler, and W. Minkler, 5–32. West Conshohocken, PA: ASTM.
Becker, W. T., and S. Lampman. 2002. Fracture appearance and mechanisms of deformation and fracture. 11th ed. Materials Park, OH: American Society for Metals.
Callister, W. D., and D. G. Rethwisch. 2011. Materials science and engineering. 9th ed. New York: Wiley.
Chang, W. S., and Y. Araki. 2016. “Use of shape-memory alloys in construction: A critical review.” Proc. Inst. Civ. Eng. Civ. Eng. 169 (2): 87–95. https://doi.org/10.1680/jcien.15.00010.
Horikawa, H., S. Ichinose, K. Morii, S. Miyazaki, and K. Otsuka. 1988. “Orientation dependence of β1β1’ stress-induced martensitic transformation in a Cu-Al-Ni alloy.” Metall. Mater. Trans. A 19 (4): 915–923. https://doi.org/10.1007/BF02628376.
Hosford, W. F. 2010. Mechanical behavior of materials. Cambridge, UK: Cambridge University Press.
Kainuma, R. 2018. “Recent progress in shape memory alloys.” Mater. Trans. 59 (3): 327–331. https://doi.org/10.2320/matertrans.M2017340.
Kainuma, R., N. Sutou, X. J. Liu, I. Ohnuma, and K. Ishida. 1998. “Phase equilibria and Heusler phase stability in the Cu-rich portion of the Cu–Al–Mn system.” J. Alloys Compd. 266 (1–2): 191–200. https://doi.org/10.1016/S0925-8388(97)00425-8.
Kainuma, R., S. Takahashi, and K. Ishida. 1996. “Thermoelastic martensite and shape memory effect in ductile Cu-Al-Mn alloys.” Metall. Mater. Trans. A. 27 (8): 2187–2195. https://doi.org/10.1007/BF02651873.
Kise, S., A. Mohebbi, S. M. Saiidi, T. Omori, R. Kainuma, K. C. Shrestha, and Y. Araki. 2018. “Mechanical splicing of super elastic Cu-Al-Mn alloy bars with headed ends.” Smart Mater. Struct. 27 (6): 065025. https://doi.org/10.1088/1361-665X/aabf0d.
Koterazawa, R. 1974. “Fractography.” J. Soc. Mater. Sci. Jpn. 23 (248): 412–421. https://doi.org/10.2472/jsms.23.412.
Kusama, T., T. Omori, T. Saito, S. Kise, Y. Tanaka, T. Araki, and R. Kainuma. 2017. “Ultra-large single crystals by abnormal grain growth.” Nat. Commun. 8 (1): 1–9. https://doi.org/10.1038/s41467-017-00383-0.
Omori, T., T. Kusama, S. Kawata, I. Ohnuma, Y. Sutou, Y. Araki, K. Ishida, and R. Kainuma. 2013. “Abnormal grain growth induced by cyclic heat treatment.” Science 341 (6153): 1500–1502. https://doi.org/10.1126/science.1238017.
Otsuka, K., H. Sakamoto, and K. Shimizu. 1979. “Successive stress-induced Martensitic transformation pseudoelasticity in Cu-Al-Ni alloys.” Acta Metall. 27 (4): 585–601. https://doi.org/10.1016/0001-6160(79)90011-7.
Otsuka, K., and C. M. Wayman. 1998. Shape memory materials. Cambridge, UK: Cambridge University Press.
Otsuka, K., C. M. Wayman, K. Nakai, H. Sakamoto, and K. Shimizu. 1976. “Superelasticity effect and stress-induced martensitic transformation in Cu-Al-Ni alloys.” Acta Metall. 24 (3): 207–226. https://doi.org/10.1016/0001-6160(76)90071-7.
Ozbulut, O., S. Hurlebaus, and R. DesRoches. 2011. “Seismic response control using shape memory alloys: A review.” J. Intell. Mater. Syst. Struct. 22 (14): 1531–1549. https://doi.org/10.1177/1045389X11411220.
Qiu, C., and S. Zhu. 2014. “Characterization of cyclic properties of superelastic monocrystalline Cu–Al–Be SMA wires for seismic applications.” Constr. Build. Mater. 72 (Dec): 219–230. https://doi.org/10.1016/j.conbuildmat.2014.08.065.
Reed, R. C. 2006. The superalloys: Fundamentals and applications. Cambridge, UK: Cambridge University Press.
Šittner, P., V. Novak, and N. Zarubova. 1998. “Martensitic transformations in [001] CuAlZnMn single crystals.” Acta Mater. 46 (4): 1265–1281. https://doi.org/10.1016/S1359-6454(97)00323-6.
Sutou, Y., T. Omori, R. Kainuma, and K. Ishida. 2013. “Grain size dependence of pseudoelasticity in polycrystalline Cu–Al–Mn-based shape memory sheets.” Acta Mater. 61 (10): 3842–3850. https://doi.org/10.1016/j.actamat.2013.03.022.
Sutou, Y., T. Omori, R. Kainuma, N. Ono, and K. Ishida. 2002. “Enhancement of superelasticity in Cu-Al-Mn-Ni shape-memory alloys by texture control.” Metall. Mater. Trans. A 33 (9): 2817–2824. https://doi.org/10.1007/s11661-002-0267-2.
Sutou, Y., T. Omori, K. Yamauchi, N. Ono, R. Kainuma, and K. Ishida. 2005. “Effect of grain size and texture on pseudoelasticity of Cu-Al-Mn-based shape memory wire.” Acta Mater. 53 (15): 4121–4133. https://doi.org/10.1016/j.actamat.2005.05.013.
Wollants, P., M. D. Bonte, and J. R. Roos. 1979. “Thermodynamic analysis of the stress-induced martensitic-transformation in single-crystal.” Z. Metallk. 70 (2): 113–117.
Xu, S., R. Tsukuda, M. Zhao, X. Xu, T. Omori, K. Yoshimi, and R. Kainuma. 2020. “Negative Poisson’s ratio in Heusler-type Cu–Al–Mn shape memory alloy.” Scr. Mater. 177 (Mar): 74–78. https://doi.org/10.1016/j.scriptamat.2019.10.015.

Information & Authors

Information

Published In

Go to Journal of Materials in Civil Engineering
Journal of Materials in Civil Engineering
Volume 33Issue 4April 2021

History

Received: Apr 29, 2020
Accepted: Jul 20, 2020
Published online: Jan 23, 2021
Published in print: Apr 1, 2021
Discussion open until: Jun 23, 2021

Authors

Affiliations

Manager, Technology Development Dept., Special Metals Div., Furukawa Techno Material Co., Ltd., 5-1-8, Higashi-Yawata, Hiratsuka 2540016, Japan. Email: [email protected]
Professor, Graduate School of Environmental Studies, Nagoya Univ., Chikusa, Nagoya 4648603, Japan (corresponding author). ORCID: https://orcid.org/0000-0001-9569-1753. Email: [email protected]
Toshihiro Omori [email protected]
Associate Professor, Dept. of Materials Science, Graduate School of Engineering, Tohoku Univ., Aoba, Sendai 9808579, Japan. Email: [email protected]
Ryosuke Kainuma [email protected]
Professor, Dept. of Materials Science, Graduate School of Engineering, Tohoku Univ., Aoba, Sendai 9808579, Japan. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share