Next Article in Journal
MPCCN: A Symmetry-Based Multi-Scale Position-Aware Cyclic Convolutional Network for Retinal Vessel Segmentation
Previous Article in Journal
Rethinking Electron Statistics Rules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Shape-Memory Effect and the Topology of Minimal Surfaces

The Blackett Laboratory, Imperial College London, London SW7 2AZ, UK
*
Author to whom correspondence should be addressed.
Symmetry 2024, 16(9), 1187; https://doi.org/10.3390/sym16091187
Submission received: 2 August 2024 / Revised: 19 August 2024 / Accepted: 29 August 2024 / Published: 10 September 2024
(This article belongs to the Section Physics)

Abstract

:
Martensitic transformations, viewed as continuous mappings between triply periodic minimal surfaces (TPMSs), as suggested by Hyde and Andersson (Z. Kristallogr. 1986, 174, 225–236), are extended to include paths between the initial and final phases. Reversible transformations, which correspond to shape-memory materials, occur only if lattice points remain at flat points on a TPMS throughout a continuous transformation. For the shape-memory material NiTi, the density functional calculations by Hatcher et al. [Phys. Rev. B 2009, 80, 144203] yield irreversible and reversible paths with and without energy barriers, respectively, in agreement with our theory. Although there are TPMSs for face-centered and body-centered cubic crystals for iron, the deformation between them is not reversible, in agreement with the non-vanishing energy barriers obtained from the density functional calculations of Zhang et al. (RSC Advances 2021, 11, 3043–3048).

1. Introduction

Martensitic transformations are diffusionless solid-to-solid transformations triggered by the rapid collective movement of atoms across distances smaller than a lattice spacing [1,2,3]. This transformation is abrupt, showing a structural discontinuity at a particular temperature, and it is often accompanied by changes in physical properties, which can be exploited for applications. Materials exhibiting martensitic transformations include metals and alloys [2], ceramics [4], and biological systems [5].
The shape-memory effect (SME) [3,6], where a material returns to its original shape when heated after a plastic deformation, accompanies some martensitic transformations. Shape-memory materials have been known since the 1930s [7], but studies of this effect began in earnest only in the 1960s, with the discovery of NiTi-based materials [8,9]. Ni-Ti alloys are the most widely used shape-memory and superelastic alloys, combining a pronounced SME with superelasticity, corrosion resistance, biocompatibility, and superior engineering properties. Shape-memory alloys have various practical applications [10]. For example, the shape-memory alloy NiTi is used in pipe coupling [6], orthodontic treatment devices [11], in the fabrication of self-expandable vascular stents [12,13], and various venous filters [14].
Here, we study martensitic transformations from a geometric and topological perspective. The martensitic transformation is interpreted as a continuous deformation between minimal surfaces, and it includes information about the relevant symmetry and topological properties of the two end states. We build on the approach of Hyde and Andersson [15], to develop a theory of martensitic transformations that encompasses not just the end points, as in their original work, but also the path between the initial and transformed states. Transformation paths that not only maintain the topological properties of the end states, but remain as triply periodic minimal surfaces over the path, are reversible and are, therefore, expected to exhibit the shape-memory effect. Our approach is consistent with available density functional calculations along martensitic transformation paths and provides a topological distinction between reversible and irreversible transformations.

2. Minimal Surfaces

Surfaces that minimize their area for given boundary conditions are called minimal. Determining the surface with the least area that spans a boundary curve was first formulated by Lagrange as a variational problem. He showed that the equation for a minimal surface must satisfy what is now known as the Euler–Lagrange equation. In geometrical terms, the local shape near a point of a three-dimensional smooth surface is determined by the principal curvatures  k 1  and  k 2  measuring the maximum and minimum bending near that point. The surface is characterized by the mean curvature  M = 1 2 ( k 1 + k 2 )  and the Gaussian curvature  K = k 1 k 2  [16,17]. A surface with  K = 0  everywhere is flat, e.g., a plane. A surface with  M = 0  everywhere is locally area-minimizing and is called minimal. The catenoid (the surface of revolution of a catenary), the (circular) helicoid (whose boundary is a circular helix), and the plane are standard examples of minimal surfaces.
The Bonnet transformation [18] establishes a mapping between minimal surfaces. This transformation is continuous, preserves the mean and Gaussian curvatures at every point on the surface, and is an isometry of the surface: that is, all distances within the surface are preserved, as there is no stretching or wrinkling. The Bonnet transformation is the weighted sum of two minimal surfaces parametrized by the Bonnet rotation angle. As the rotation angle increases, a continuous family of minimal surfaces is generated.
Minimal surfaces with repeating units in three independent directions are called triply periodic minimal surfaces (TPMSs). Minimal surfaces can have points where  K = 0 ; these are called flat points, and they will play a central role in what follows. The flat points on the minimal surfaces considered here are isolated. The symmetry of each TPMS is a space group, just as for ordinary crystals [19]. The first TPMSs were discovered by Schwarz [20], including the primitive (P), diamond (D), and hexagonal (H) surfaces (Figure 1). Schoen [21,22] discovered the gyroid (G), which is obtained when deforming P to D surfaces along the Bonnet path and is locally isometric to both the P and D surfaces [21,23]. Chen and Weber [24] discovered that the P surface can deform into the H surface within a space of TPMSs; the intermediate surfaces are in the  o P a  and  o P b  families. Surfaces in the  o P a  family can be obtained by stretching or compressing the P surface along its three necks [Figure 2a,b], while those in the  o P b  family involve displacements of necks and holes on the surface [Figure 2c,d]:
Since the point group of a lattice is a (proper or improper) subgroup of the point group of its TPMS [25], we require the point group of the crystal to be a subgroup (proper or improper) of that of the TPMS for the two end states. There is also a group–subgroup relation between the point group of the crystal and its corresponding Bravais lattice. Thus, the TPMS can be understood as the Bravais lattice that gives the basic symmetry and structural information of this crystal. Now, a topological transformation between two surfaces is bicontinuous (continuous in both directions), one-to-one, and reversible. Some properties are preserved under topological transformations and others are not. The genus of a surface is its topological type. Roughly speaking, the genus is the number of holes in a surface. Thus, surfaces with the same genus are topologically equivalent to each other, but surfaces with a different genus are not topologically equivalent.
The transformation between body-centered cubic (BCC) and face-centered cubic (FCC) lattices is not a topologically permitted martensitic phase transformation, because the genus of the corresponding TPMS is 4 for BCC but is 6 for FCC [26]. Hence, there cannot be a topological transformation between these lattices. However, the hexagonal close-packed (HCP)–BCC, HCP–hexagonal transformation in Zr, the BCC–HCP transformation in Na, and the B2–B19′ transformation in NiTi are topologically permitted to be martensitic.

3. Deformation Paths of Martensitic Transformations

Explanations for BCC–FCC (and reverse) transformation include the celebrated Bain tetragonal distortion [27], the Kurdjumow–Sachs (KS) [28] and Nishiyama–Wassermann (NW) [29,30] shear models, together with the Bogers–Burgers (BB) [31] and the Olson–Cohen (OC) [32] hard-sphere models. The geometries for these paths are shown in Figure 3. The Bain path [Figure 3a,b] is based on a deformation along  001 , which yields a body-centered tetragonal (BCT) structure. The BCT unit cell is compressed by about 21% along the z-direction and expanded by about 12% along the x- and y-directions [33].
Despite its simplicity and intuitive appeal, the orientation relationship between the parent (FCC) and daughter (BCC) phases of the Bain path has not been observed in experiments. The NW path [Figure 3c,d] begins with a shear that produces a monoclinic unit cell. Along this path, the unit cell changes from orthogonal to monoclinic and back to orthogonal. Computational studies [33,34] indicate this path is favored over the Bain path, though there is still a barrier for the transformation. The KS shear model [Figure 3e,f], similar to the NW model, starts with a shear that produces a monoclinic cell, but has different orientational relationships. The BB (OC) hard-sphere model indicates the FCC–BCC lattice transformation is realized by a shear of type  112 { 111 }  [31,32].
In non-magnetic materials, these deformation paths are not topologically continuous. Take the Bain deformation as an example [Figure 4a]. Each lattice point in a BCC lattice has 8 nearest neighbors while in an FCC lattice each lattice point has 12 nearest neighbors. Here, the lattice points are equivalent points in a Bravais lattice. The crystal structure of an elemental or compound material is obtained by assigning a basis of one or more atoms to each lattice point. For instance, each lattice point in BCC Fe is conventionally chosen to be occupied by a Fe atom. However, for BCC NiTi, the basis of each lattice point consists of one Ni atom and one Ti atom. In an FCC lattice, the length of the green bond line ( l g ) and the length of the cyan bond line ( l c ) are equal, while in a BCC lattice they are not. When lattice parameters satisfy  a = b > c , we have  l g > l c  and some bond lines connecting the lattice point A will “break”, which leads to a decrease in the number of bond lines connecting one lattice point to its nearest neighbors from 12 (FCC) to 8 (BCC). Similar arguments can be made for the KS, NW, and BB (OC) paths [Figure 4b–d].

4. Density Functional Calculations of Martensitic Transformation Paths

Table 1 compiles energy barriers calculated with the density functional theory (DFT) of materials undergoing martensitic transformations along several paths. The calculations for Fe did not account for magnetism. We can see that paths involving Bain/NW/BB deformations, which are not topologically continuous, have a non-vanishing energy barrier. Only shape-memory alloys have barrierless paths that reside in a space of TPMSs. However, the austenite (B2) phase of NiTi has the space group  P m 3 ¯ m . After a rapid quench, NiTi transforms to martensite. We can locate the Bravais lattice of B2 NiTi on a P surface, as shown in Figure 5.
DFT calculations [35] also show that energy barriers vary with the paths of the B2→B19′ NiTi martensitic transformation, with the energy barrier going to zero when a path meets our geometrical criteria. The paths involving only the monolayer (resp., bilayer)  100 { 011 }  shear/shuffle have an energy barrier up to 37 meV/atom (resp., 22 meV/atom), while the path that admits the intermediate phase 109°–B19′, allowing both shear and structural relaxation, has no barrier, with the energy per atom decreasing throughout the transformation [35]. As shown in Figure 6b, a single-layer  100 { 011 }  shear requires  l 1 / l 2 = 1.40  and  γ = 98 .
We will report elsewhere that the end phases of a martensitic phase transformation yield Bravais lattices whose lattice sites are located at flat points on a TPMS, such that (i) the lattice structure and periodicity are preserved, (ii) the point group of the crystal is a (proper or improper) subgroup of the point group of the TPMS. The lattice parameters used to find the TPMS corresponding to NiTi in different phases are given in Appendix B. The TPMS corresponding to the phase under a single-layer  100 { 011 }  shear closest to the requirements given by Figure 5b is shown in Figure 6d, in which  l 1 / l 2 = 1.40  and  γ = 98 . However, in Figure 6b,  γ 2 = 141 . 89  and  γ 3 = 98 , while in Figure 6d,  γ 2 = 125 . 09  and  γ 3 = 112 . 86 . Similarly, for the bilayer  100 { 011 }  shear, as shown in Figure 6c,  l 1 / l 2 = 0.70  and  g a m m a = 98 . The TPMS closest to this requirement is given in Figure 6e, with  l 1 / l 2 = 0.70  and  γ = γ 3 = 98 . However,  γ 2 = 112 . 86 , which disagrees with the requirement given by Figure 6d that  γ 2 = 109 . 47 . Thus, the TPMS given in Figure 6e cannot describe the crystal lattice shown in Figure 6d. For another choice of lattice points, as shown in Figure 6f, the conditions  l 1 / l 2 = 0.7  and  γ = 98  cannot be satisfied simultaneously. Therefore, none of the single-layer or bilayer  100 { 011 }  shears can reside on a TPMS.
However, for the path that allows the occurrence of both the bilayer and single-layer  100 { 011 }  shear with an intermediate 109°–B19′ phase, as shown in Figure 7a, we can find a TPMS belonging to the oPb family [Figure 7c] on which this phase can reside. Both the 109–B19′ phase and the B19′ phase [Figure 7b,d] can reside on a surface belonging to the  o P b  family, and the shear mechanism of this path is compatible with the continuous deformation within a space of TPMSs formed by surfaces belonging to the  o P b  family. According to Table 1, this differs from the two paths mentioned above, so this path is barrierless.
We see similar arguments can be made for the B2–R phase transformation via the  [ 111 ] B 2  elongation. As shown in Figure 8, we can find the Bravais lattice of the R phase of NiTi described by an H surface. We can see the  [ 111 ] B 2  elongation can be viewed as a combination as an extension of the lattice along  [ 110 ] B 2  and an extension along  [ 001 ] B 2 . This deformation conforms to the deformation from the P surface to the H surface within a space of the  o P b  family. This path is also barrierless. The B2–B33 transformation via an alternate bilayer  100 [ 011 ]  shear and the B2–B19 transformation via a  1 ¯ 10 { 110 }  basal shear are shown in Figure 9 and Figure 10. It can be seen that the B2–B33 path conforms to the surface deformation within the  o P a  family while the B2–B19 path does not, and we see the B2–B33 path is barrierless while the B2–B19 path is not.
The details of the atomic-scale mechanisms responsible for martensitic transformations are important for a quantitative understanding of the transformation processes, including the calculation of energy barriers. Our approach, based on the transformations between minimal surfaces, provides a conceptually simple description based on geometrical and topological principles. As with transformation paths, such as proposed by Bain and NW, and according to the group–subgroup relation [51], no attempt is made to model the temperature-driven nucleation process, just to establish a relationship between the initial and final phases and, in our case, the intermediate states as well. We are looking at the change of crystal structures during the stress-driven diffusionless transformations that are determined by the energy barrier calculated by DFT.

5. Summary and Conclusions

We have studied the martensitic transformation and the shape-memory effect from the perspective of the topology and differential geometry of TPMSs. We propose that martensitic transformations can be understood as continuous deformations between TPMSs, with the Bravais lattice positions of two end phases being at flat points. If a martensitic transformation continues within a space of TPMSs and all the lattice points remain at flat points throughout the process then we expect the occurrence of the SME. Our method does not account for the nucleation process, nor for the effects of temperature in such a stress-driven diffusionless process. Nevertheless, our method is relatively simple, general, and far less time-consuming in comparison to DFT and, thus, can provide a fast screening method for the discovery of more novel shape-memory alloys.

Author Contributions

Conceptualization, M.Y. and D.D.V.; methodology, M.Y. and D.D.V.; software, M.Y.; validation, M.Y. and D.D.V.; formal analysis, M.Y.; investigation, M.Y.; resources, D.D.V.; data curation, M.Y.; writing—original draft preparation, M.Y.; writing—review and editing, D.D.V.; visualization, M.Y. and D.D.V.; supervision, D.D.V.; project administration, D.D.V.; funding acquisition, D.D.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
SMEShape-memory effect
TPMSTriply periodic minimal surface
BCCBody-centered cubic
FCCFace-centered cubic
HCPHexagonal close-packed
BCTBody-centered tetragonal
FCTFace-centered–tetragonal
KSKurdjumow–Sachs
NWNishiyama–Wassermann
BBBogers–Burgers
OCOlson–Cohen
DFTDensity functional theory

Appendix A

To describe minimal surfaces, Weierstrass and Enneper [16,17,52,53] used complex analysis to construct what has become known as the Weierstrass representation of these surfaces. This construction is based on the images obtained by first mapping minimal surfaces according to the directions of their normal vectors onto a unit sphere (also known as the Gauss map), then taking a stereographic projection of this “sphere” on to the complex plane. Generally, the Weierstrass representation  ( G s , d h )  of a minimal surface consists of two elements: its Gauss map  G s  onto the sphere in  R 3  and its height differential  d h , such that  G s d h , d h / G s , d h  are analytic in  C  [54]. A minimal surface S can be described by a map  σ : U R 2 S R 3 , which is an integral on  C  that reverses the mapping mentioned above [16,17], and any point on the surface can be written as (in Cartesian coordinates):  ( x , y , z ) = σ ( u , v ) , ( u , v ) U . To be specific, in Cartesian coordinates, points on this surface can be expressed as [54]
x = Re w 1 2 ( ϕ 2 ϕ 1 ) d w , y = Re w i 2 ( ϕ 2 + ϕ 1 ) d w , z = Re w d h d w ,
where
ϕ 1 = G s d h , ϕ 2 = d h G s ,
and  w , w C . More details about  ( G s , d h )  of different minimal surfaces are listed in Table A1.
All relevant Gauss map expressions in Table A1 are listed below [24,54]:
G s ( z ) θ 1 ( z b s , τ ) θ 1 ( z + b s , τ ) θ 1 ( z c s , τ ) θ 1 ( z + c s , τ ) θ 1 ( z a s , τ ) θ 1 ( z + a s , τ ) θ 1 ( z d s , τ ) θ 1 ( z + d s , τ ) ,
where  θ 1 ( z , τ )  is the Jacobi’s theta function [55],
b s = 1 2 a s , c s = a s + 1 2 τ , d s = b s 1 2 τ ,
where  a s R  and  τ  is an imaginary number determined by  a s  [24,54].
G s ( z ) θ 1 ( z a s , τ ) θ 1 ( z + a s , τ ) θ 1 ( z c s , τ ) θ 1 ( z + c s , τ ) θ 1 ( z b s , τ ) θ 1 ( z + b s , τ ) θ 1 ( z d s , τ ) θ 1 ( z + d s , τ ) ,
where
c s = 1 2 ( 1 τ ) a s , d s = 1 2 ( 1 + τ ) b s ,
where  a s , b s R , and  τ  is determined by  a s , b s  [24].
Table A1. Weierstrass representations  ( G s , d h )  of different triply periodic minimal surfaces in  R 3 . The column labelled Parameters contains tuneable parameters whose variation produces a continuous surface deformation ensuring all associated surfaces are minimal surfaces.
Table A1. Weierstrass representations  ( G s , d h )  of different triply periodic minimal surfaces in  R 3 . The column labelled Parameters contains tuneable parameters whose variation produces a continuous surface deformation ensuring all associated surfaces are minimal surfaces.
PhaseSurface G s dh Parameters
B2Schwarz PEquation (A3) [24,54]1     a s
B19′ o P b  FamilyEquation (A5) [24,54]1     a s , b s
B19 o P a  FamilyEquation (A3) [24]1     a s
B33 o P a  FamilyEquation (A3) [24,54]1     a s
B32Schwarz DEquation (A3) [15] e 1 2 i π      a s
R o P b  FamilyEquation (A5) [24,54]1     a s , b s
109°–B19′ o P b  FamilyEquation (A5) [24,54]1     a s , b s

Appendix B

In this appendix, we provide the parameters for the TPMS corresponding to phases along martensitic transformation paths of NiTi. The Bravais lattice given by the TPMS converges to that found either in experiments or by DFT calculations. The source of the DFT calculations is given next to each phase.
Table A2. Parameters for the TPMS corresponding to phases along martensitic transformation paths of NiTi. For each phase, the references for computed and/or measured values of the ratios  a / c  and  b / c  of lattice parameters and values  γ  of the non-right angles are indicated. The corresponding values of these quantities obtained from our minimal surface theory are indicated by (T). The parameters  a s b s , and  ψ  determine each minimal surface from expressions given in Table A1.
Table A2. Parameters for the TPMS corresponding to phases along martensitic transformation paths of NiTi. For each phase, the references for computed and/or measured values of the ratios  a / c  and  b / c  of lattice parameters and values  γ  of the non-right angles are indicated. The corresponding values of these quantities obtained from our minimal surface theory are indicated by (T). The parameters  a s b s , and  ψ  determine each minimal surface from expressions given in Table A1.
Phase a / c a / b a / c  (T) a / b  (T) γ γ  (T)Parameters
B2 [35,56,57]1111 a s = 0.5  ,  ψ = 0
B19′ [35,58]1.131.601.131.6098°97.82° a s 0.096 b s 0.198
B19′ [35,59]1.131.601.131.6097.78°97.70° a s 0.096 b s 0.198
B19′ [35]1.151.601.151.6098°97.69° a s 0.098 b s 0.197
B19′ [56]1.161.601.161.6098.97°97.92° a s 0.099 b s 0.195
B19 [35,60]1.561.071.561.07 a s 0.151
B19 [35]1.621.111.621.11 a s 0.158
B19 [56]1.751.081.751.08 a s 0.175
B33 [39]1.241.671.241.67107.4°107.4° a s 0.153
B33 [39]3.192.363.192.36 a s 0.153
B32 [56]1111 a s = 0.5 ψ = 1 2 π
R1 [35,61]1.3911.391120°120° a s 0.113 , b s 0.220
R1 [35]1.3611.361120°120° a s 0.111 , b s 0.223
R2 [37]1.3711.371120°120° a s 0.112 , b s 0.222
R2 [35]1.3911.391120°120° a s 0.113 , b s 0.221
R3 [35]1.3511.351120°120° a s 0.110 , b s 0.224
109°–B19′ [37]1.131.611.131.61108.88°109.9° a s 0.070 b s 0.192

References

  1. Olson, G.B.; Cohen, M. A perspective on martensitic nucleation. Ann. Rev. Mater. Sci. 1981, 11, 1–30. [Google Scholar] [CrossRef]
  2. Christian, J.W. The Theory of Transformations in Metals and Alloys, 3rd ed.; Pergamon: Amsterdam, The Netherlands, 2002. [Google Scholar]
  3. Battacharya, K. Microstructure of Martensite: Why It Forms, and How It Gives Rise to the Shape-Memory Effect; Oxford University Press: New York, NY, USA, 2003. [Google Scholar]
  4. Kelly, P.J.; Rose, L.R.F. The martensitic transformation in ceramics—Its role in transformation toughening. Prog. Mater. Sci. 2002, 47, 463–557. [Google Scholar] [CrossRef]
  5. Olson, G.B. New directions in martensite theory. Mater. Sci. Eng. A 1999, 273–275, 11–20. [Google Scholar] [CrossRef]
  6. Tadaki, K.; Otsuka, K.; Shimuzu, K. Shape memory alloys. Ann. Rev. Mater. Sci. 1988, 18, 25–45. [Google Scholar] [CrossRef]
  7. Ölander, A. An electrochemical investigation of solid cadmium-gold alloys. J. Am. Chem. Soc. 1932, 54, 3819–3833. [Google Scholar] [CrossRef]
  8. Buehler, W.J.; Gilfrich, J.V.; Wiley, R.C. Effect of low-temperature phase changes on the mechanical properties of alloys near composition TiNi. J. Appl. Phys. 1963, 34, 1475–1477. [Google Scholar] [CrossRef]
  9. Kauffman, G.B.; Mayo, I. The story of Nitinol: The serendipitous discovery of the memory metal and its applications. Chem. Educ. 1997, 2, 1–21. [Google Scholar] [CrossRef]
  10. Balasubramanian, M.; Srimath, R.; Vignesh, L.; Rajesh, S. Application of shape memory alloys in engineering—A review. J. Phys. Conf. Ser. 2021, 2054, 012078. [Google Scholar] [CrossRef]
  11. Andreasen, G.F.; Barrett, R.D. An evaluation of cobalt-substituted nitinol wire in orthodontics. Am. J. Orthod. 1973, 63, 462–470. [Google Scholar] [CrossRef]
  12. Simon, M.; Kaplow, R.; Salzman, E.; Freiman, D. A vena cava filter using thermal shape memory alloy. Radiology 1977, 125, 89–94. [Google Scholar] [CrossRef]
  13. Lipscomb, I.P.; Nokes, L.D.M. The Application of Shape Memory Alloys in Medicine; Wiley–Blackwell: Hoboken, NJ, USA, 1996. [Google Scholar]
  14. Yang, C.; Abanteriba, S.; Becker, A. A review of shape memory alloy based filtration devices. AIP Adv. 2020, 10, 060701. [Google Scholar] [CrossRef]
  15. Hyde, S.T.; Andersson, S. The martensite transition and differential geometry. Z. Kristallogr. 1986, 174, 225–236. [Google Scholar] [CrossRef]
  16. Gray, A. Modern Differential Geometry of Curves and Surfaces with Mathematica, 2nd ed.; CRC Press: BocaRaton, FL, USA, 1997. [Google Scholar]
  17. O’Neill, B. Elementary Differential Geometry; Revised 2nd edn; Elsevier: Amsterdam, The Netherlands, 2006. [Google Scholar]
  18. Lidin, S. Some results of the Bonnet transformation. J. Phys. Colloq. 1990, 51, C7-237–C7-242. [Google Scholar] [CrossRef]
  19. Koch, E.; Fischer, W. A crystallographic approach to 3-periodic minimal surfaces. In Statistical Thermodynamics and Differential Geometry of Microstructured Materials; Davis, H.T., Nitsche, J.C.C., Eds.; Springer: New York, NY, USA, 1993; pp. 14–48. [Google Scholar]
  20. Schwarz, H.A. Gesammelte Mathematische Abhandlungen; Springer: Berlin, Germany, 1890. [Google Scholar]
  21. Schoen, A.H. Infinite Periodic Minimal Surfaces without Self-Intersections; Tech. Rep. TN D-5541; NASA: Washington, DC, USA, 1970. [Google Scholar]
  22. Schoen, A.H. Reflections concerning triply-periodic minimal surfaces. Interface Focus 2012, 2, 658–668. [Google Scholar] [CrossRef] [PubMed]
  23. Karcher, H. Embedded minimal surfaces derived from Scherk’s examples. Manuscripta Math. 1989, 64, 291–357. [Google Scholar] [CrossRef]
  24. Chen, H.; Weber, M. An orthorhombic deformation family of Schwartz’ H surfaces. Trans. Am. Math. Soc. 2021, 374, 2057–2078. [Google Scholar] [CrossRef]
  25. Kocian, P.; Schenk, K.; Chapuis, G. Differential geometry: A natural tool for describing symmetry operations. Acta Crystallogr. 2009, A65, 329–341. [Google Scholar] [CrossRef]
  26. Nakahara, M. Geometry, Topology, and Physics, 2nd ed.; Institute of Physics: Bristol, UK, 2003. [Google Scholar]
  27. Bain, E.C.; Dunkirk, N.Y. The nature of martensite. Trans. AIME 1924, 70, 25–47. [Google Scholar]
  28. Kurdjumow, G.; Sachs, G. Über den mechanismus der stahlhärtung. Z. Physik 1930, 64, 325–343. [Google Scholar] [CrossRef]
  29. Nishiyama, Z. Martensitic Transformation; Academic: New York, NY, USA, 1978. [Google Scholar]
  30. Wassermann, G. Einfluß der α-γ-Umwandlung eines irreversiblen Nickelstahls auf Kristallorientierung und Zugfestigkeit. Archiv. Eisenhüttenwesen 1933, 6, 347–351. [Google Scholar] [CrossRef]
  31. Bogers, A.J.; Burgers, W.G. Partial dislocations on the {110} planes in the B.C.C. lattice and the transition of the F.C.C. into the B.C.C. lattice. Acta Metall. 1964, 12, 255–261. [Google Scholar] [CrossRef]
  32. Olson, G.; Cohen, M. A mechanism for the strain-induced nucleation of martensitic transformations. J. Less-Common Met. 1972, 107–118. [Google Scholar] [CrossRef]
  33. Sandoval, L.; Urbassek, H.M.; Entel, P. The Bain versus Nishiyama–Wassermann path in the martensitic transformation of Fe. New J. Phys. 2009, 11, 103027. [Google Scholar] [CrossRef]
  34. Zhang, L.H.; Cheng, M.J.; Shi, X.H.; Shuai, J.W.; Zhu, Z.Z. Bain and Nishiyama–Wassermann transition path separation in the martensitic transitions of Fe. RSC Adv. 2021, 11, 3043–3048. [Google Scholar] [CrossRef] [PubMed]
  35. Hatcher, N.; Kontsevoi, O.Y.; Freeman, A.J. Role of elastic and shear stabilities in the martensitic transformation path of NiTi. Phys. Rev. B 2009, 80, 144203. [Google Scholar] [CrossRef]
  36. Niu, J.G.; Geng, W.T. Anti-precursor effect of Fe on martensitic transformation in TiNi alloys. Acta Mater. 2016, 104, 18–24. [Google Scholar] [CrossRef]
  37. Hatcher, N.; Kontsevoi, O.Y.; Feemen, A.J. Martensitic transformation path of NiTi. Phys. Rev. B 2009, 79, 020202(R). [Google Scholar] [CrossRef]
  38. Morris, J.R.; Ye, Y.; Krcmar, M.; Fu, C.L. The Role of Phase Stability in Ductile, Ordered B2 Intermetallics. MRS Online Proc. Libr. 2006, 980, 610. [Google Scholar] [CrossRef]
  39. Krcmar, M.; Samolyuk, G.D.; Morris, J.R. Stacking faults and alternate crystal structures for the shape-memory alloy NiTi. Phys. Rev. Mater. 2020, 4, 103606. [Google Scholar] [CrossRef]
  40. Kibey, S.; Sehitoglu, H.; Johnson, D.D. Energy landscape for martensitic phase transformation in shape memory NiTi. Acta Mater. 2009, 57, 1624–1629. [Google Scholar] [CrossRef]
  41. Trinkle, D.R.; Hatch, D.M.; Stokes, H.T.; Hennig, R.G.; Albers, R.C. Systematic pathway generation and sorting in martensitic transformations: Titanium α to ω. Phys, Rev. B 2005, 72, 014105. [Google Scholar] [CrossRef]
  42. Hennig, R.G.; Trinkle, D.R.; Bouchet, J.; Srinivasan, S.G.; Albers, R.C.; Wilkins, J.W. Impurities block the α to ω martensitic transformation in titanium. Nat. Mater. 2005, 4, 129–133. [Google Scholar] [CrossRef]
  43. Ghosh, P.S.; Arya, A.; Tewari, R.; Dey, G.K. Alpha to omega martensitic phase transformation pathways in pure Zr. J. Alloys Compd. 2014, 586, 693–698. [Google Scholar] [CrossRef]
  44. Suzuki, T.; Ledbetter, H.M. Barrier energy for the b c.c.-f.c.c. martensitic transition in sodium. Phil. Mag. 1983, 48, 83–94. [Google Scholar] [CrossRef]
  45. Okatov, S.V.; Kuznetsov, A.R.; Gornostyrev, Y.N.; Urtsev, V.N.; Katsnelson, M.I. Effect of magnetic state on the γ-α transition in iron: First-principles calculations of the Bain transformation path. Phys. Rev. B 2009, 79, 094111. [Google Scholar] [CrossRef]
  46. Wang, T.L.; Du, J.L.; Wei, S.Z.; Liu, F. Ab-initio investigation for the microscopic thermodynamics and kinetics of martensitic transformation. Prog. Nat. Sci. Mater. Int. 2021, 31, 121–128. [Google Scholar] [CrossRef]
  47. Wang, L.G.; Šob, M. Structural stability of higher-energy phases and its relation to the atomic configurations of extended defects: The example of Cu. Phys. Rev. B 1999, 60, 844–850. [Google Scholar] [CrossRef]
  48. Jeong, S. Structural properties of bulk copper: Pseudopotential plane-wave-basis study. Phys. Rev. B 1996, 53, 13973–13976. [Google Scholar] [CrossRef]
  49. Mishin, Y.; Mehl, M.J.; Papaconstantopoulos, D.A.; Voter, A.F.; Kress, J.D. Structural stability and lattice defects in copper: Ab initio, tight-binding, and embedded-atom calculations. Phys. Rev. B 2001, 63, 224106. [Google Scholar] [CrossRef]
  50. Mishin, Y.; Farkas, D.; Mehl, M.J.; Papaconstantopoulos, D.A. Interatomic potentials for monoatomic metals from experimental data and ab initio calculations. Phys. Rev. B 1999, 59, 3393–3407. [Google Scholar] [CrossRef]
  51. Bhattacharya, K.; Conti, S.; Zanzotto, G.; Zimmer, J. Crystal symmetry and the reversibility of martensitic transformations. Nature 2004, 428, 55–59. [Google Scholar] [CrossRef] [PubMed]
  52. Weierstrass, K. Mathematische Werke; Mayer & Müller: Berlin, Germany, 1903. [Google Scholar]
  53. Enneper, A. Analytisch-Geometrische Untersuchung. Z. Math Phys. 1864, 9, 96–125. [Google Scholar]
  54. Weyhaupt, A.G. New Families of Embedded Triply Periodic Minimal Surfaces of Genus Three in Euclidean Space. Ph.D. Thesis, Indiana University, Bloomington, IN, USA, 2006. [Google Scholar]
  55. Stein, E.M.; Shakarchi, R. Complex Analysis; Princeton University Press: Princeton, NJ, USA, 2010. [Google Scholar]
  56. Liu, G.T.; Liu, H.Y.; Feng, X.L.; Redfern, S.A.T. High-pressure phase transitions of nitinol NiTi to a semiconductor with an unusual topological structure. Phys. Rev. B 2018, 97, 140104. [Google Scholar] [CrossRef]
  57. Sittner, P.; Lukás, P.; Neov, D.; Novák, V.; Toebbens, D.M. The martensite transition and differential geometry. J. Phys. IV (Fr.) 2003, 112, 709–711. [Google Scholar] [CrossRef]
  58. Prokoshkin, S.D.; Korotitskiy, A.V.; Brailovski, V.; Turenne, S.; Khmelevskaya, I.Y.; Trubitsyna, I.B. On the lattice parameters of phases in binary Ti–Ni shape memory alloys. Acta Mater. 2004, 52, 4479–4492. [Google Scholar] [CrossRef]
  59. Kudoh, Y.; Tokonami, M.; Miyazaki, S.; Otsuka, K. Crystal structure of the martensite in Ti-49.2 at. %Ni alloy analyzed by the single crystal X-ray diffraction method. Acta Metall. 1985, 33, 2049–2056. [Google Scholar] [CrossRef]
  60. Nam, T.H.; Saburi, T.; Nakata, Y.; Shimizu, K.I. Shape Memory Characteristics and Lattice Deformation in Ti–Ni–Cu Alloys. Mater. Trans. JIM 1990, 31, 1050–1056. [Google Scholar] [CrossRef]
  61. Hara, T.; Ohba, T.; Okinushi, E.; Otsuka, K. Structural study of R-phase in Ti-50.23 at. %Ni and Ti-47.75 at. %Ni-1.50 at. %Fe Alloys. Mater. Trans. JIM 1997, 38, 11–17. [Google Scholar] [CrossRef]
Figure 1. Periodic units that repeat along three independent directions, to form the corresponding TPMS: (a) P surface, (b) I- W P  surface, (c) F- R D  surface, and (d) H surface.
Figure 1. Periodic units that repeat along three independent directions, to form the corresponding TPMS: (a) P surface, (b) I- W P  surface, (c) F- R D  surface, and (d) H surface.
Symmetry 16 01187 g001
Figure 2. Periodic units of (a) P surface deformed to (b) a surface belonging to the  o P a  family, and (c) the top view of the P surface deformed to (d) a surface belonging to the  o P b  family.
Figure 2. Periodic units of (a) P surface deformed to (b) a surface belonging to the  o P a  family, and (c) the top view of the P surface deformed to (d) a surface belonging to the  o P b  family.
Symmetry 16 01187 g002
Figure 3. Atomic arrangements of (a) the Bain transformation path from the FCC to the BCC phase, with an intermediate BCT unit cell (shaded), (c) the NW path, with an intermediate body-centered triclinic unit cell (shaded), (e) the KS transformation path, with an intermediate body-centered triclinic cell, and (g) the BB (OC) path, with an intermediate body-centered triclinic cell depicted in (h). The unit cells are isolated in (b), (d), (f), and (h), respectively. The red and blue spheres indicate atoms in the parent and transformed phases, respectively.
Figure 3. Atomic arrangements of (a) the Bain transformation path from the FCC to the BCC phase, with an intermediate BCT unit cell (shaded), (c) the NW path, with an intermediate body-centered triclinic unit cell (shaded), (e) the KS transformation path, with an intermediate body-centered triclinic cell, and (g) the BB (OC) path, with an intermediate body-centered triclinic cell depicted in (h). The unit cells are isolated in (b), (d), (f), and (h), respectively. The red and blue spheres indicate atoms in the parent and transformed phases, respectively.
Symmetry 16 01187 g003
Figure 4. The (a) Bain deformation, (b) NW deformation, (c) KS deformation, and (d) BB (OC) deformation. The blue ad spheres signify lattice points in the BCC and FCC cells, respectively. The green and cyan bond lines, which connect the nearest neighbors in the FCC lattice, connect nearest and second-nearest neighbors in the BCC lattice. Therefore, none of these paths is continuous.
Figure 4. The (a) Bain deformation, (b) NW deformation, (c) KS deformation, and (d) BB (OC) deformation. The blue ad spheres signify lattice points in the BCC and FCC cells, respectively. The green and cyan bond lines, which connect the nearest neighbors in the FCC lattice, connect nearest and second-nearest neighbors in the BCC lattice. Therefore, none of these paths is continuous.
Symmetry 16 01187 g004
Figure 5. The NiTi lattice against the P surface. The red and blue spheres are the Ni and Ti atoms, respectively. The magenta spheres are Bravais lattice points. The entire Bravais lattice is shown in (a), with the detail of the positions of the lattice points in (b).
Figure 5. The NiTi lattice against the P surface. The red and blue spheres are the Ni and Ti atoms, respectively. The magenta spheres are Bravais lattice points. The entire Bravais lattice is shown in (a), with the detail of the positions of the lattice points in (b).
Symmetry 16 01187 g005
Figure 6. (a) B2 NiTi under (b) a single layer  100 { 011 }  shear and (c) a bilayer  100 { 011 }  shear. The orange plane in (a) denotes the  { 011 }  invariant plane; the red and blue spheres denote Ni and Ti atoms, respectively; (b,c) are viewed from the  [ 0 1 ¯ 1 ]  direction. In (b,c),  l 1 , l 2  are the lengths of cell edges outlined by black dashed lines in the crystal lattice. In (b),  l 1 / l 2 = 1.40 , and in (c l 1 / l 2 = 0.70 . In (bf),  γ = 98 . In (df),  l 1 , l 2  are the lengths of the cell edges outlined by black dashed lines on the surface. In (d),  l 1 / l 2 = 1.40 ; in (e),  l 1 / l 2 = 1.40 ; and in (f),  l 1 / l 2 = 0.62 .
Figure 6. (a) B2 NiTi under (b) a single layer  100 { 011 }  shear and (c) a bilayer  100 { 011 }  shear. The orange plane in (a) denotes the  { 011 }  invariant plane; the red and blue spheres denote Ni and Ti atoms, respectively; (b,c) are viewed from the  [ 0 1 ¯ 1 ]  direction. In (b,c),  l 1 , l 2  are the lengths of cell edges outlined by black dashed lines in the crystal lattice. In (b),  l 1 / l 2 = 1.40 , and in (c l 1 / l 2 = 0.70 . In (bf),  γ = 98 . In (df),  l 1 , l 2  are the lengths of the cell edges outlined by black dashed lines on the surface. In (d),  l 1 / l 2 = 1.40 ; in (e),  l 1 / l 2 = 1.40 ; and in (f),  l 1 / l 2 = 0.62 .
Symmetry 16 01187 g006
Figure 7. (a) Crystal lattice of the 109°–B19′ phase; (b) crystal lattice of the B19′ phase; (c) the TPMS on which the Bravais lattice of the 109°–B19′ phase resides; and (d) the TPMS on which the Bravais lattice of the B19′ lattice resides.
Figure 7. (a) Crystal lattice of the 109°–B19′ phase; (b) crystal lattice of the B19′ phase; (c) the TPMS on which the Bravais lattice of the 109°–B19′ phase resides; and (d) the TPMS on which the Bravais lattice of the B19′ lattice resides.
Symmetry 16 01187 g007
Figure 8. (a) Crystal lattice of B2 NiTi, where the red and blue spheres are Ni and Ti atoms, respectively; (b) crystal lattice and Bravais lattice on the P surface viewed from above; (c) deformed crystal lattice of B2 NiTi under the  [ 111 ]  elongation; (d) Bravais lattice of R NiTi against the H surface viewed from above; (e) crystal lattice and Bravais lattice against the P surface; (f) Bravais lattice of R NiTi against the H surface viewed from above. The arrow points along the  [ 111 ]  direction in the B2 phase.
Figure 8. (a) Crystal lattice of B2 NiTi, where the red and blue spheres are Ni and Ti atoms, respectively; (b) crystal lattice and Bravais lattice on the P surface viewed from above; (c) deformed crystal lattice of B2 NiTi under the  [ 111 ]  elongation; (d) Bravais lattice of R NiTi against the H surface viewed from above; (e) crystal lattice and Bravais lattice against the P surface; (f) Bravais lattice of R NiTi against the H surface viewed from above. The arrow points along the  [ 111 ]  direction in the B2 phase.
Symmetry 16 01187 g008
Figure 9. (a) Crystal lattice of the B2 NiTi viewed from the  [ 0 1 ¯ 1 ] B 2 ]  direction; (b) crystal lattice of the B33 NiTi obtained under an alternate bilayer  100 { 011 }  shear; bilayer marked by a red (resp. blue) dotted rectangle moves along direction denoted by the arrow; in (a,c),  γ = 109 . 47  and in (b,d),  γ = 107 . 41 ; (c) plots the crystal lattice (red and blue spheres for Ni and Ti atoms) and Bravais lattice (magenta sphere) of the B2 NiTi against the P surface; (d) plots the Bravais lattice of the B33 NiTi against its corresponding TPMS belonging to the  o P a  family.
Figure 9. (a) Crystal lattice of the B2 NiTi viewed from the  [ 0 1 ¯ 1 ] B 2 ]  direction; (b) crystal lattice of the B33 NiTi obtained under an alternate bilayer  100 { 011 }  shear; bilayer marked by a red (resp. blue) dotted rectangle moves along direction denoted by the arrow; in (a,c),  γ = 109 . 47  and in (b,d),  γ = 107 . 41 ; (c) plots the crystal lattice (red and blue spheres for Ni and Ti atoms) and Bravais lattice (magenta sphere) of the B2 NiTi against the P surface; (d) plots the Bravais lattice of the B33 NiTi against its corresponding TPMS belonging to the  o P a  family.
Symmetry 16 01187 g009
Figure 10. (a) Crystal lattice (the red and blue spheres represent Ni and Ti atoms, respectively) and Bravais lattice (magenta sphere) of the B2 NiTi against the P surface; (b) Bravais lattice of the B19 NiTi.
Figure 10. (a) Crystal lattice (the red and blue spheres represent Ni and Ti atoms, respectively) and Bravais lattice (magenta sphere) of the B2 NiTi against the P surface; (b) Bravais lattice of the B19 NiTi.
Symmetry 16 01187 g010
Table 1. Energy barriers calculated along different transformation paths for the given materials. All shears are given in the coordinate system of the B2 phase of NiTi.  Δ E  is expressed in units of meV/atom. No magnetism but structure transformation was considered in these calculations.
Table 1. Energy barriers calculated along different transformation paths for the given materials. All shears are given in the coordinate system of the B2 phase of NiTi.  Δ E  is expressed in units of meV/atom. No magnetism but structure transformation was considered in these calculations.
MaterialInitialPathFinal Δ E
NiTi [35]B2 100 { 011 }  basal shearB19′∼37
B2 100 { 011 }  bilayer shearB19′∼22
B2109°–B19′B19′0
NiTi [36]B2[111] elongationR0
NiTi [35,37,38,39]B2 100 { 011 }  alternate bilayer shearB330
NiTi [40]B2 1 ¯ 10 { 110 }  basal shearB19∼13
Ti [41]HCP( α ) [ 0001 ] α | | [ 01 2 ¯ 0 ] ω , { 11 2 ¯ 0 } α | | { 0001 } ω Hexagonal ( ω )∼9
Ti [42]HCP( α ) [ 11 2 ¯ 0 ] α | | [ 01 1 ¯ 1 ] ω , ( 0001 ) α | | ( 0 1 ¯ 11 ) ω Hexagonal ( ω )∼9.5
Zr [43]HCP( α ) [ 11 2 ¯ 0 ] α | | [ 01 1 ¯ 1 ] ω , ( 0001 ) α | | ( 0 1 ¯ 11 ) ω Hexagonal ( ω )22
Na [44]BCCBainFCC∼0.79
Fe [34,45] γ -FeBain (FCT) α -Fe∼20
γ –FeBain α –Fe∼45
Fe [34] γ –FeNW (FCT) α –Fe∼37
Fe [46] γ –FeBB α –Fe∼25
Cu [47]BCT(BCC)BainFCC1
Cu [48]BCT(BCC)BainFCC1.3
Cu [49]FCCBBBCC∼700
Al [50]BCT(BCC)BainFCC∼30
Al [50]FCCBBBCC∼100
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yin, M.; Vvedensky, D.D. Shape-Memory Effect and the Topology of Minimal Surfaces. Symmetry 2024, 16, 1187. https://doi.org/10.3390/sym16091187

AMA Style

Yin M, Vvedensky DD. Shape-Memory Effect and the Topology of Minimal Surfaces. Symmetry. 2024; 16(9):1187. https://doi.org/10.3390/sym16091187

Chicago/Turabian Style

Yin, Mengdi, and Dimitri D. Vvedensky. 2024. "Shape-Memory Effect and the Topology of Minimal Surfaces" Symmetry 16, no. 9: 1187. https://doi.org/10.3390/sym16091187

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop