Exotic metal alloys showing simultaneously steel-like ultrahigh strength and polymer-like ultrahigh flexibility have long been desired for many emerging technologies such as morphing aircrafts1,2 and superman-type artificial muscles3. Such an unconventional “strong yet flexible” property combination (red regime of Fig. 1a) will enable a large recoverable shape change under a small driving power and simultaneously possess strong resistance against fracture or yielding failure under large load. As a result, such alloys would enable a morphing wing in an aircraft or superstrong artificial muscle in a humanoid robot, to name just a few possibilities.
However, achieving such unconventional alloys has remained challenging due to the trade-off relation between strength and flexibility (the latter being measured by elastic compliance or inverse of elastic modulus)5,8,10, as manifested by the grey band (known as Ashby plot4) of Fig. 1a. It reveals a common observation4,10,11 that steels can be made very strong (with a high yield strength σy >1 GPa) but they are stiff (with a high Young’s modulus E ~200 GPa), whereas organic materials such as fiber-reinforced polymers (FRP) have the opposite property of being flexible (with a typical low E ~10 GPa) but they are weak (σy < 0.3 GPa). The strength-flexibility trade-off stems from the fact that the two properties are oppositely correlated to the bonding strength of a material, which is represented roughly by elastic modulus5,8,10. Therefore, this inevitable trade-off relation precludes a 1 GPa-class steel-like high-strength alloy from being of 10 GPa-class polymer-like low modulus.
Over the past decades, significant efforts have been made to seek metal alloys with simultaneously high strength and low modulus4-9,12-17; but an alloy showing both steel-like high yield strength (σy >1 GPa) and polymer-like low modulus (E ~10 GPa) still remains unattainable. So far, several alloys based on shape memory alloys (SMA)5,16,17 are reported to reveal 1 GPa-class high strength and a moderately low modulus E ~30 GPa, and a conventional Mg-Sc strain glass alloy7 has recently been shown to possess a lower modulus E ~20 GPa but with a lower strength σy ~0.3 GPa. Despite these efforts, existing alloys still fall into the conventional grey band of Fig. 1a, and the desired polymer-like ultrahigh-strength property (red regime of Fig. 1a) still remains elusive.
Here we show that a Ti-50.8 at.% Ni “dual-seed strain glass” (DS-STG) alloy can overcome the strength-flexibility trade-off and demonstrates simultaneously a steel-like ultrahigh yield strength σy ~1.8 GPa and a polymer-like ultralow Young’s modulus E ~10.5 GPa (corresponding to an ultrahigh elastic compliance S=1/E ~0.1 GPa-1) (Fig. 1a), together with a superlarge rubber-like J-shaped elastic strain of ~8% (Fig. 1b). As a result, it exhibits an unprecedented flexibility figure of merit σy/E ~0.17 which far exceeds that of existing structural materials (Fig. 1c), and the alloy behaves like a superstrong polymer (Fig. 1d). In the following, we shall first demonstrate its unique mechanical properties, then show its 3-step thermomechanical fabrication route towards its unique DS-STG microstructure, and finally discuss the origin of mechanical properties in relation to the DS-STG microstructure.
Unique mechanical properties of DS-STG alloy
Figure 1a shows that the DS-STG alloy overcomes the well-observed strength-flexibility trade-off relation of existing structural materials (shown as the grey band of Fig. 1a) and exhibits an unprecedented combination of ultrahigh-strength-steel class yield strength of σy ~1.8 GPa and a polymer-like ultralow Young’s modulus E ~10.5 GPa (corresponding to ultrahigh elastic compliance of S=1/E ~0.1 GPa-1). Such a property combination places the DS-STG alloy into a hitherto inaccessible “strong yet flexible” regime (red regime) in Fig. 1a.
Figure 1b shows that the DS-STG alloy demonstrate a J-shaped stress-strain curve with a superlarge pseudoelastic strain εre ~8% before reaching its elasticity limit at σy~1.8 GPa for wire samples and σy~1.3 GPa for plate samples. The 1.8 GPa yield strength of DS-STG alloy even exceeds that of most ultrahigh-strength steels like quenched spring steel, and the 8% pseudoelastic strain value not only far exceeds the small elastic strain εre ~0.2-1% for typical metal alloys18-20 but also is larger than that (εre ~1-5%) of many organic structural materials like wood, bamboo, bones, polymer, and fiber-reinforced polymers21-24. It is noted that the large J-shaped elasticity behavior associated with a small stress hysteresis of ~15% resembles that of ultra-flexible materials like rubbers25, and it is very different from the superelasticity of conventional shape memory alloys, which is characterized by a stress plateau with a much larger initial modulus of E~30-75 GPa and a huge hysteresis (~50-90%)26-28. The strongly hysteretic behavior of shape memory alloys is known to be undesirable for many applications28-30.
The combination of ultrahigh strength and ultralow modulus endows the DS-STG alloy with a superhigh flexibility figure of merit σy/E ~0.17, far exceeding that of existing structural materials including typical metal alloys and organic materials, as shown in Fig. 1c. Flexibility figure of merit σy/E is a measure of ultimate material flexibility before yielding failure10, and existing structural materials usually have a much lower σy/E ~0.01-0.06.
Figure 1d and Supplementary Movie 1 show visual evidence for the unconventional “strong yet flexible” (i.e., polymer-like ultrahigh-strength) property of the DS-STG alloy, which contrasts with the conventional behavior of two reference materials: a “strong & stiff” spring steel (with high modulus E ~190 GPa and high strength σy ~1.36 GPa), and a “weak & flexible” FRP (with low modulus E ~11 GPa and low strength σy ~0.13 GPa). Thus, the DS-STG alloy combines the high strength of a steel with the high flexibility of a polymer, and it behaves like a superstrong organic material, which has long been desired by many emerging technologies.
A 3-step thermomechanical fabrication route towards DS-STG alloy and the corresponding microstructure feature after each step
The DS-STG alloy was fabricated through a 3-step thermomechanical processing route from a starting solution-treated Ti-50.8Ni alloy, as shown in Fig. 2a. The simplicity of this processing method makes it scalable to industry lines. The corresponding microstructure after each step is described below.
Step-1 processing: formation of deformation-stabilized martensite
The starting solution-treated Ti-50.8Ni alloy (Fig. 2a(i)) is in its B2 parent state (strictly speaking an unfrozen strain glass state31-33 (Extended Data Fig. 1a)) at room temperature since it has a sub-zero martensitic transformation temperature of Ms ~250 K. Such a B2 state exhibits a high elastic modulus E~71 GPa and an incomplete superelasticity with a large remanent strain and hysteresis (Extended Data Fig. 1b) during loading/unloading even at a moderate stress of ~0.3 GPa. This is a well-observed behavior of Ti-Ni alloys when dislocation slip occurs during a stress-induced martensitic transformation14. Thus, the starting B2 alloy is neither a strong alloy nor a polymer-like alloy.
In Step-1 processing, the B2 alloy is severely deformed by 50% tensile elongation (Fig. 2(ii)). This cold-working process results in a deformation-stabilized B19’ martensite which is preserved at room temperature (Extended Data Fig. 2a). This cold-working-strengthened B19’ martensite exhibits a quasi-linear elastic behavior with a high yield stress of ~1.3 GPa and a moderate elastic modulus of E~37 GPa, as shown in Extended Data Fig. 2b. Such mechanical properties are consistent with previous reports on tensile-deformed Ti-Ni alloys12,34; thus the deformation-stabilized B19’ martensite alloy is a high-strength alloy but not an ultralow modulus alloy.
Step-2 processing: formation of a dual-crossover strain glass (DC-STG) state
In Step-2 processing, the deformation-stabilized B19’ martensite alloy is annealed at 573 K for 10 min (Fig. 2(iii)). Annealing at such a temperature not only fully annihilates the metastable B19’ martensite produced in Step-1 processing, but also results in a unique “dual-crossover strain glass” (DC-STG) state. This DC-STG is an unfrozen R strain glass (R-STG) at room temperature and it undergoes a strain glass transition around Tg ~251 K, which is evidenced by the same signatures as those of a normal strain glass transition reported in the literature31-33, including a frequency (w) dependent Tg following Vogel-Fulcher law (inset of Fig. 2b) and the invariance of average B2 structure across Tg, together with the appearance of nano-sized R strain domains. But being different from a conventional R strain glass which keeps R nanodomain down to 0 K, this DC-STG is able to crossover to R and B19’ dual martensites at lower temperatures, as evidenced by in-situ transmission electron microscopy (TEM) and X-ray diffractometry (XRD) (Fig. 2b). It is noted that the crossover transitions are incomplete and significant amount of strain glass remains even at the lowest temperature tested. This unique R strain glass that can crossover into R and B19 dual martensites is thus named “dual-crossover strain glass” or DC-STG.
It should be noted that this DC-STG state, albeit a forerunner state to the ultralow-modulus DS-STG state, is not a low modulus state by itself. It shows a non-linear and hysteretic superelastic behavior with a moderate initial elastic modulus of ~33 GPa and a high yield strength of ~1 GPa (Extended Data Fig. 3).
Step-3 processing: formation of dual-seed strain glass (DS-STG) state
In Step-3 processing, the DC-STG alloy undergoes a moderate tensile elongation of ~12% (Fig. 2a(iv)). This process introduces a small amount of R and B19’ martensite “dual-seeds” (50-100nm in size) into the room-temperature DC-STG matrix (manifested as nanosized R domains), as revealed by HRTEM image (Fig. 2c, top). The resultant strain glass state containing R and B19’ dual-seeds is thus named “dual-seed strain glass” (DS-STG). Clearly the realization of this unique DS-STG owes to the capability of its forerunner state DC-STG to crossover into two martensites R and B19’, as revealed in Fig. 2b. Such a unique DS-STG state endows this alloy a polymer-like ultralow modulus E ~10.5 GPa, together with a J-shaped superlarge pseudoelastic strain εre ~8% and small hysteresis of ~15% (Fig. 1b). The mechanism of such unconventional elastic behavior will be revealed in Fig. 3.
Low magnification image of the DS-STG alloy (Fig. 2c, bottom) reveals that it contains high density dislocations and a large amount of B2 mechanical twins, which are formed by the severe cold-working during the 3-step processing. These deformation-hardening features are characteristic of severely deformed Ti-Ni alloys16,35-37, and can account for the ultrahigh yield strength of the DS-STG alloy (Fig. 1b).
To confirm the general applicability of the 3-step thermomechanical processing in achieving polymer-like ultrahigh-strength property, we also tested a wire sample of the Ti-50.8Ni alloy by the same 3-step route but the severe deformation of the sample was done with a wire-drawing machine instead of a tensile machine. As can be seen in Fig. 1b, the wire sample exhibits the same ultralow modulus E ~10.5 GPa as that of the plate sample, but with an even higher yield strength σy ~1.8 GPa due to the favorable influence of cold-drawing. Therefore, the present 3-step processing route is applicable to both wire and plate samples with Step-1 deformation being either tensile elongation and cold-drawing.
Origin of the polymer-like ultrahigh strength of DS-STG alloy
In-situ tensile XRD experiment (Fig. 3) shows that the DS-STG alloy with the above “dual-seeds in a STG matrix” microstructural feature (Fig. 2c) undergoes a unique nucleation-free reversible transition between STG and R and B19’ martensites during a stress loading/unloading cycle up to 1.3 GPa and this process leads to the observed ultralow elastic modulus. Before loading (i.e., σ =0 GPa), the DS-STG, being a DC-STG matrix embedded with a small fraction of R and B19’ martensite seeds, reveals a broadened B2 peak (representing the average structure of the random STG nanodomains38) with a long tail covering R and B19’ positions in its XRD profile. With increasing stress, the R and B19’ martensite peaks grow smoothly without needing a critical stress until the whole sample transforms almost fully into B19’ martensite at σ =1.3 GPa. Clearly this is a result of R and B19’ martensite seeds that bypass the nucleation barrier of the stress-induced transition, consequently leading to the polymer-like ultralow elastic modulus (E ~10.5 GPa). Upon unloading from 1.3 GPa, a smooth reverse transition occurs from B19’ to a mixture of R and B19’, and then to the original state of DS-STG. Thus the original DS-STG state is recovered upon unloading. The large pseudoelastic strain (εre ~8%) is a natural result of this reversible transformation, as the maximum transformation strain to form B19’ martensite in Ti-Ni alloys is about 8-9%26.
Another important consequence of the nucleation-free DS-STG to R and B19’ transition is its J-shape and narrow hysteresis features in the stress-strain curve of the DS-STG alloy, because the R and B19’ martensite seeds can bypass the nucleation barrier during a stress-induced transition. This results in ultralow initial slope in the stress-strain curve, i.e., the “J-shape”, and small hysteresis. The small hysteresis is also reflected by the similarity of the XRD profiles between loading and unloading at the same stress level. The smooth, J-shape, and narrow-hysteretic behavior of DS-STG, contrasts with that of normal stress-induced STG to martensite transition (Extended Data Fig. 3) or a conventional stress-induced martensitic transformation (Extended data Fig. 1b ), where the inevitable nucleation leads to not only a much higher elastic modulus of E ~30-70 GPa but also a strongly nonlinear and hysteretic stress-strain behavior with a stress plateau14,16.
Fig. 4a shows that a DS-STG state is essential to achieve a polymer-like ultralow modulus. Such a state is obtained by a 3-step processing with Step-2 annealing at Ta ~573 K. When the Step-2 annealing temperature Ta deviates from the optimal 573 K, the alloy deviates from a DS-STG state and this results in an increase in elastic modulus and a decrease in recoverable strain.
Fig. 4b explains why the DS-STG state is obtained only when the Step-2 annealing temperature Ta is around 573 K. This is because the DS-STG’s forerunner state DC-STG is achieved only around Ta ~573 K, as shown in the phase diagram of Fig. 4b (established from data from Fig. 2b, Extended Data Fig. 5 and 6). When Ta deviates significantly from 573 K, the forerunner DC-STG will no longer exist and consequently a R+B19’ dual-seed strain glass cannot be achieved after Step-3 processing (i.e., 12% elongation).
The phase diagram (Fig.4b) shows that a significantly lower Ta (e.g., Ta ~473 K) will lead to a mono-crossover strain glass (MC-STG) after Step-2 processing, which is a R strain glass that can crossover (partially) into one martensite B19’ at low temperature. After Step-3 processing (12% elongation) it will change into a mono-seed strain glass (MS-STG, i.e., with B19’ seeds only), which has a higher modulus of ~17 GPa (~1.5 times high than that of DS-STG) and a smaller recoverable strain of ~6%.
At a significantly higher Ta ~773 K, which is a common annealing temperature for TiNi alloys26, strain glass transition is no longer existent and the alloy is characterized by a normal 2-step martensitic transformation into R and B19’ martensite. After Step-3 processing (12% elongation) the alloy will change into an aligned B19’ martensite, which has an even higher modulus of ~30 GPa and an even smaller recoverable strain of ~4%.
The lower elastic modulus of the DS-STG state as compared with non-DS-STG states can be explained by the phase instability of the system in the vicinity of three different phases. This results in a nearly flat free energy landscape which facilitates the transition from one phase to another and consequently the lattice becomes the softest. Such an effect has been observed in many ferroelectrics and relaxor materials associated with multiple ferroelectric phases40.