Skip to main content
Log in

On the grain boundary network characteristics in a martensitic Ti–6Al–4V alloy

  • Metals & corrosion
  • Published:
Journal of Materials Science Aims and scope Submit manuscript

Abstract

The characteristics of the intervariant boundary network that resulted from the \(\beta \to \alpha^{\prime}\) martensitic phase transformation in a Ti–6Al–4V alloy were studied using the crystallographic theories of displacive transformations, five-parameter grain boundary analysis and triple junction analysis. The microstructure of Ti–6Al–4V martensite consisted of fine laths containing dislocations and fine twins. The misorientation angle distribution revealed four distinct peaks consistent with the intervariant boundaries expected from the Burgers orientation relationship. The phenomenological theory of martensite predicted four-variant clustering to have the lowest transformation strain among different variant clustering combinations. This configuration was consistent with the observed Ti–6Al–4V martensitic microstructure, where four-variant clusters consisted of two pairs of distinct V-shape variants. The \(63.26^\circ /[\overline{10}\, 5\, 5\, \overline{3}]_{{\alpha^{\prime}}}\) and \(60^\circ /[1\, 1\, \overline{2}\, 0]_{{\alpha^{\prime}}}\) intervariant boundaries accounted for ~ 38% and 33% of the total population, respectively. The five-parameter boundary analysis showed that the former had a twist character, being terminated on the \((\overline{3}\, 2\, 1\, 0)_{{\alpha^{\prime}}}\) plane, and the latter revealed a symmetric tilt \((1\, 0\, \overline{1}\, 1)_{{\alpha^{\prime}}}\) boundary plane. The \(63.26^\circ /[\overline{10}\, 5\, 5\, \overline{3}]_{{\alpha^{\prime}}}\) and \(60^\circ /[1\, 1\, \overline{2}\, 0]_{{\alpha^{\prime}}}\) had the highest connectivity at triple junctions among other intervariant boundaries. Interestingly, the boundary network in Ti–6Al–4V martensite was significantly different from the commercially pure Ti martensite, where only \(60^\circ /[1\, 1\, \overline{2}\, 0]_{{\alpha^{\prime}}}\) intervariant boundaries largely were found at triple junctions due to the formation of three-variant clustering to minimize the transformation strain. This difference is thought to result from a change in the martensitic transformation mechanism (slip vs twinning) caused by the alloy composition.

Graphic abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10

Similar content being viewed by others

References

  1. Lütjering G, Williams JC (2007) Titanium, 2nd edn. Springer, Berlin

    Google Scholar 

  2. Lütjering G (2002) Influence of processing on microstructure and mechanical properties of (α+β) titanium alloys. Mater Sci Eng A 243:32–45. https://doi.org/10.1016/s0921-5093(97)00778-8

    Article  Google Scholar 

  3. Ghonem H (2010) Microstructure and fatigue crack growth mechanisms in high temperature titanium alloys. Int J Fatigue 32:1448–1460. https://doi.org/10.1016/j.ijfatigue.2010.02.001

    Article  CAS  Google Scholar 

  4. Serra A, Bacon DJ, Pond RC (2002) Twins as barriers to basal slip in hexagonal-close-packed metals. Met Mater Trans A 33:809–812. https://doi.org/10.1007/s11661-002-1012-6

    Article  Google Scholar 

  5. Wang J, Liu L, Tomé CN et al (2013) Twinning and de-twinning via glide and climb of twinning dislocations along serrated coherent twin boundaries in hexagonal-close-packed metals. Mater Res Lett 1:81–88. https://doi.org/10.1080/21663831.2013.779601

    Article  CAS  Google Scholar 

  6. Skrotzki W, Wendt H, Carter CB, Kohlstedt DL (1988) The relation between the structure and mechanical properties of a Σ= 51 tilt boundary in germanium. Acta Metall 36:983–994

    Article  CAS  Google Scholar 

  7. Guo Y, Britton TB, Wilkinson AJ (2014) Slip band-grain boundary interactions in commercial-purity titanium. Acta Mater 76:1–12. https://doi.org/10.1016/j.actamat.2014.05.015

    Article  CAS  Google Scholar 

  8. Beladi H, Cizek P, Taylor AS et al (2017) Static softening in a Ni-30Fe austenitic model alloy after hot deformation: microstructure and texture evolution. Metall Mater Trans A Phys Metall Mater Sci 48:855–867. https://doi.org/10.1007/s11661-016-3880-1

    Article  CAS  Google Scholar 

  9. Kobayashi S, Kobayashi R, Watanabe T (2016) Control of grain boundary connectivity based on fractal analysis for improvement of intergranular corrosion resistance in SUS316L austenitic stainless steel. Acta Mater 102:397–405. https://doi.org/10.1016/j.actamat.2015.08.075

    Article  CAS  Google Scholar 

  10. Papillon F, Rohrer GS, Wynblatt P (2009) Effect of segregating impurities on the grain-boundary character distribution of magnesium oxide. J Am Ceram Soc 92:3044–3051. https://doi.org/10.1111/j.1551-2916.2009.03327.x

    Article  CAS  Google Scholar 

  11. Pang Y, Wynblatt P (2006) Effects of Nb doping and segregation on the grain boundary plane distribution in TiO 2. J Am Ceram Soc 89:666–671. https://doi.org/10.1111/j.1551-2916.2005.00759.x

    Article  CAS  Google Scholar 

  12. Baram M, Chatain D, Kaplan WD (2011) Nanometer-thick equilibrium films: the interface between thermodynamics and atomistics. Science 80(332):206–209. https://doi.org/10.1126/science.1201596

    Article  CAS  Google Scholar 

  13. Dillon SJ, Tang M, Carter WC, Harmer MP (2007) Complexion: a new concept for kinetic engineering in materials science. Acta Mater 55:6208–6218. https://doi.org/10.1016/j.actamat.2007.07.029

    Article  CAS  Google Scholar 

  14. Beladi H, Nuhfer NT, Rohrer GS (2014) The five-parameter grain boundary character and energy distributions of a fully austenitic high-manganese steel using three dimensional data. Acta Mater 70:281–289. https://doi.org/10.1016/j.actamat.2014.02.038

    Article  CAS  Google Scholar 

  15. Beladi H, Rohrer GS (2013) The relative grain boundary area and energy distributions in a ferritic steel determined from three-dimensional electron backscatter diffraction maps. Acta Mater 61:1404–1412. https://doi.org/10.1016/j.actamat.2012.11.017

    Article  CAS  Google Scholar 

  16. Haghdadi N, Cizek P, Hodgson PD, He Y, Sun B, Jonas JJ, Rohrer GS, Beladi H (2020) New insights into the interface characteristics of a duplex stainless steel subjected to accelerated ferrite-to-austenite transformation. J Mater Sci 55(12):5322–5339

    Article  CAS  Google Scholar 

  17. Kumar M, Schwartz AJ, King WE (2002) Microstructural evolution during grain boundary engineering of low to medium stacking fault energy fcc materials. Acta Mater 50:2599–2612. https://doi.org/10.1016/S1359-6454(02)00090-3

    Article  CAS  Google Scholar 

  18. Palumbo G, Lehockey EM, Lin P (1998) Applications for grain boundary engineered materials. Jom 50:40–43. https://doi.org/10.1007/s11837-998-0248-z

    Article  CAS  Google Scholar 

  19. Schwartz AJ (1998) The potential engineering of grain boundaries through thermomechanical processing. JOM 50:50–55. https://doi.org/10.1007/s11837-998-0250-5

    Article  CAS  Google Scholar 

  20. Shimada M, Kokawa H, Wang ZJ et al (2002) Optimization of grain boundary character distribution for intergranular corrosion resistant 304 stainless steel by twin- induced grain boundary engineering. Acta Mater 50:2331–2341

    Article  CAS  Google Scholar 

  21. Watanabe T (2011) Grain boundary engineering: historical perspective and future prospects. J Mater Sci 46:4095–4115. https://doi.org/10.1007/s10853-011-5393-z

    Article  CAS  Google Scholar 

  22. Michiuchi M, Kokawa H, Wang ZJ et al (2006) Twin-induced grain boundary engineering for 316 austenitic stainless steel. Acta Mater 54:5179–5184. https://doi.org/10.1016/j.actamat.2006.06.030

    Article  CAS  Google Scholar 

  23. Haghdadi N, Zarei-Hanzaki A, Farabi E, Cizek P, Beladi H, Hodgson PD (2017) Strain rate dependence of ferrite dynamic restoration mechanism in a duplex low-density steel. Mater Des 132:360–366

    Article  CAS  Google Scholar 

  24. Beladi H, Chao Q, Rohrer GS (2014) Variant selection and intervariant crystallographic planes distribution in martensite in a Ti–6Al–4V alloy. Acta Mater 80:478–489. https://doi.org/10.1016/j.actamat.2014.06.064

    Article  CAS  Google Scholar 

  25. Beladi H, Rohrer GS (2017) The role of thermomechanical routes on the distribution of grain boundary and interface plane orientations in transformed microstructures. Metall Mater Trans A 48:2781–2790. https://doi.org/10.1007/s11661-016-3630-4

    Article  CAS  Google Scholar 

  26. Randle V (1999) Mechanism of twinning-induced grain boundary engineering in low stacking-fault energy materials. Acta Mater 47:4187–4196. https://doi.org/10.1016/S1359-6454(99)00277-3

    Article  CAS  Google Scholar 

  27. Haghdadi N, Cizek P, Hodgson PD et al (2018) Effect of ferrite-to-austenite phase transformation path on the interface crystallographic character distributions in a duplex stainless steel. Acta Mater 145:196–209. https://doi.org/10.1016/j.actamat.2017.11.057

    Article  CAS  Google Scholar 

  28. Haghdadi N, Cizek P, Hodgson PD et al (2018) Five-parameter crystallographic characteristics of the interfaces formed during ferrite to austenite transformation in a duplex stainless steel. Philos Mag 98:1284–1306. https://doi.org/10.1080/14786435.2018.1434321

    Article  CAS  Google Scholar 

  29. Beladi H, Rohrer GS (2013) The distribution of grain boundary planes in interstitial free steel. Metall Mater Trans A Phys Metall Mater Sci 44:115–124. https://doi.org/10.1007/s11661-012-1393-0

    Article  CAS  Google Scholar 

  30. Beladi H, Tari V, Timokhina IB et al (2017) On the crystallographic characteristics of nanobainitic steel. Acta Mater 127:426–437. https://doi.org/10.1016/j.actamat.2017.01.058

    Article  CAS  Google Scholar 

  31. Beladi H, Rohrer GS, Rollett AD et al (2014) The distribution of intervariant crystallographic planes in a lath martensite using five macroscopic parameters. Acta Mater 63:86–98. https://doi.org/10.1016/j.actamat.2013.10.010

    Article  CAS  Google Scholar 

  32. Hutchinson B, Komenda J, Rohrer GS, Beladi H (2015) Heat affected zone microstructures and their influence on toughness in two microalloyed HSLA steels. Acta Mater 97:380–391. https://doi.org/10.1016/j.actamat.2015.05.055

    Article  CAS  Google Scholar 

  33. Wang SC, Aindow M, Starink MJ (2003) Effect of self-accommodation on α/α boundary populations in pure titanium. Acta Mater 51:2485–2503. https://doi.org/10.1016/S1359-6454(03)00035-1

    Article  CAS  Google Scholar 

  34. Farabi E, Hodgson PD, Rohrer GS, Beladi H (2018) Five-parameter intervariant boundary characterization of martensite in commercially pure titanium. Acta Mater 154:147–160. https://doi.org/10.1016/j.actamat.2018.05.023

    Article  CAS  Google Scholar 

  35. Burgers WGG (1934) On the process of transition of the cubic-body-centered modification into the hexagonal-close-packed modification of zirconium. Physica 1:561–586. https://doi.org/10.1016/S0031-8914(34)80244-3

    Article  CAS  Google Scholar 

  36. Qiu D, Shi R, Zhao P et al (2016) Effect of low-angle grain boundaries on morphology and variant selection of grain boundary allotriomorphs and Widmanstätten side-plates. Acta Mater 112:347–360. https://doi.org/10.1016/j.actamat.2016.04.033

    Article  CAS  Google Scholar 

  37. Shi R, Dixit V, Fraser HL, Wang Y (2014) Variant selection of grain boundary α by special prior β grain boundaries in titanium alloys. Acta Mater 75:156–166. https://doi.org/10.1016/j.actamat.2014.05.003

    Article  CAS  Google Scholar 

  38. Qiu D, Shi R, Zhang D et al (2015) Variant selection by dislocations during α precipitation in α/β titanium alloys. Acta Mater 88:218–231. https://doi.org/10.1016/j.actamat.2014.12.044

    Article  CAS  Google Scholar 

  39. Shi R, Dixit V, Viswanathan GB et al (2016) Experimental assessment of variant selection rules for grain boundary α in titanium alloys. Acta Mater 102:197–211. https://doi.org/10.1016/j.actamat.2015.09.021

    Article  CAS  Google Scholar 

  40. Banerjee S, Mukhopadhyay P (2007) Martensitic transformations. In: Phase transformations: examples from titanium and zirconium Alloys, first ed. Elsevier, pp 259–376

  41. Balachandran S, Kashiwar A, Choudhury A et al (2016) On variant distribution and coarsening behavior of the α phase in a metastable β titanium alloy. Acta Mater 106:374–387. https://doi.org/10.1016/j.actamat.2016.01.023

    Article  CAS  Google Scholar 

  42. Williams AJ, Cahn RW, Barrett CS (1954) The crystallography of the β-α transformation in titanium. Acta Metall 2:117–128. https://doi.org/10.1016/0001-6160(54)90101-7

    Article  CAS  Google Scholar 

  43. Hua K, Zhang Y, Gan W et al (2018) Correlation between imposed deformation and transformation lattice strain on α variant selection in a metastable β-Ti alloy under isothermal compression. Acta Mater 161:150–160. https://doi.org/10.1016/j.actamat.2018.09.022

    Article  CAS  Google Scholar 

  44. Khundzhua AG, Ptitsyn AG, Brovkina EA, Chzhen S (2012) Self-accommodation of crystals of martensitic phases in titanium and zirconium based alloys. Phys Met Metallogr 113:1035–1040. https://doi.org/10.1134/s0031918x12110117

    Article  Google Scholar 

  45. Nishida M, Nishiura T, Kawano H, Inamura T (2012) Self-accommodation of B19 ′ martensite in Ti–Ni shape memory alloys—Part I. Morphological and crystallographic studies of the variant selection rule. Philos Mag 92:2215–2233. https://doi.org/10.1080/14786435.2012.669858

    Article  CAS  Google Scholar 

  46. Miyazaki S, Otsuka K, Wayman CM (1989) The shape memory mechanism associated with the martensitic transformation in TiNi alloys-I. Self-Accomm Acta Metall 37:1873–1884. https://doi.org/10.1016/0001-6160(89)90072-2

    Article  CAS  Google Scholar 

  47. Srivastava D, Madangopal K, Banerjee S, Ranganathan S (1993) Self accommodation morphology of martensite variants inZr-2. 5wt % Nb alloy. Acta Metall Mater 41:3445–3454

    Article  CAS  Google Scholar 

  48. Germain L, Gey N, Humbert M et al (2005) Analysis of sharp microtexture heterogeneities in a bimodal IMI 834 billet. Acta Mater 53:3535–3543. https://doi.org/10.1016/j.actamat.2005.03.043

    Article  CAS  Google Scholar 

  49. Humbert M, Germain L, Gey N et al (2006) Study of the variant selection in sharp textured regions of bimodal IMI 834 billet. Mater Sci Eng A 430:157–164. https://doi.org/10.1016/j.msea.2006.05.047

    Article  CAS  Google Scholar 

  50. Divinski SV, Dnieprenko VN, Ivasishin OM (1998) Effect of phase transformation on texture formation in Ti-base alloys. Mater Sci Eng A 243:201–205

    Article  Google Scholar 

  51. Qiu D, Zhao P, Shi R et al (2016) Effect of autocatalysis on variant selection of α precipitates during phase transformation in Ti–6Al–4V alloy. Comput Mater Sci 124:282–289. https://doi.org/10.1016/j.commatsci.2016.07.032

    Article  CAS  Google Scholar 

  52. Lieberman DS (1958) Martensitic transformations and determination of the inhomogeneous deformation. Acta Metall 6:680–693. https://doi.org/10.1016/0001-6160(58)90059-2

    Article  CAS  Google Scholar 

  53. Lieberman DS, Wechsler MS, Read TA (1955) Cubic to orthorhombic diffusionless phase change—experimental and theoretical studies of AuCd. J Appl Phys 26:473–484

    Article  CAS  Google Scholar 

  54. Wayman CM (1964) Introduction to the crystallography of martensitic transformations. Macmillan, Newyork

    Google Scholar 

  55. Saylor DM, El-Dasher BS, Adams BL, Rohrer GS (2004) Measuring the five-parameter grain-boundary distribution from observations of planar sections. Metall Mater Trans A 35:1981–1989. https://doi.org/10.1007/s11661-004-0147-z

    Article  Google Scholar 

  56. Rohrer GS, Saylor DM, El Dasher B et al (2004) The distribution of internal interfaces in polycrystals. Zeitschrift für Met 95:197–214. https://doi.org/10.3139/146.017934

    Article  CAS  Google Scholar 

  57. Ratanaphan S, Yoon Y, Rohrer GS (2014) The five parameter grain boundary character distribution of polycrystalline silicon. J Mater Sci 49:4938–4945. https://doi.org/10.1007/s10853-014-8195-2

    Article  CAS  Google Scholar 

  58. Gey N, Humbert M (2002) Characterization of the variant selection occurring during the α to β phase transformations of a cold rolled titanium sheet. Acta Mater 50:277–287. https://doi.org/10.1016/S1359-6454(01)00351-2

    Article  CAS  Google Scholar 

  59. Wang J, Beyerlein IJ (2012) Atomic structures of symmetric tilt grain boundaries in Hexagonal Close-Packed (hcp) crystals. Metall Mater Trans A Phys Metall Mater Sci 20:1–22. https://doi.org/10.1088/0965-0393/20/2/024002

    Article  CAS  Google Scholar 

  60. Glowinski K, Morawiec A (2012) A toolbox for geometric grain boundary characterization. In: Proceedings of the 1st international conference on 3d materials science. Springer, Cham, pp 119–124

  61. Otsuka K, Wayman CM (1999) Shape memory materials. Cambridge University Press, Cambridge

    Google Scholar 

  62. Banerjee S, Mukhopadhyay P (2010) Phase transformations: examples from titanium and zirconium alloys. Elsevier, Amsterdam

    Google Scholar 

  63. Zhang Y, Li Z, Esling C et al (2010) A general method to determine twinning elements. J Appl Crystallogr 43:1426–1430. https://doi.org/10.1107/S0021889810037180

    Article  CAS  Google Scholar 

  64. Bilby BA, Crocker AG (1965) The theory of the crystallography of deformation twinning. Proc R Soc A Math Phys Eng Sci 288:240–255

    CAS  Google Scholar 

  65. Hammond C, Kelly P (1969) The crystallography of titanium alloy martensites. Acta Metall 17:869–882. https://doi.org/10.1016/0001-6160(69)90107-2

    Article  CAS  Google Scholar 

  66. Nishiyama Z, Oka M, Nakagawa H (1965) 10–11 transforamtion twins in titanium. Jpn Inst Met 7:1–4

    Google Scholar 

  67. Grabovetskaya GP, Melnikova EN, Kolobov YR et al (2006) Evolution of the structural and phase states of a Ti–6Al–4V alloy in forming submicrocrystalline structure with the use of temporary hydrogenation. Russ Phys J 49:442–447. https://doi.org/10.1007/s11182-006-0123-8

    Article  CAS  Google Scholar 

  68. Swarnakar AK, Van Der Biest O, Baufeld B (2011) Thermal expansion and lattice parameters of shaped metal deposited Ti–6Al–4V. J Alloys Compd 509:2723–2728. https://doi.org/10.1016/j.jallcom.2010.12.014

    Article  CAS  Google Scholar 

  69. Shibata M, Ono K (1977) On the minimization of strain energy in the martensitic transformation of titanium. Acta Metall 25:35–42. https://doi.org/10.1016/0001-6160(77)90243-7

    Article  CAS  Google Scholar 

  70. Morawiec A (2012) On “interface-plane scheme” and symmetric grain boundaries. Zeitschrift fur Krist 227:199–206. https://doi.org/10.1524/zkri.2012.1475

    Article  CAS  Google Scholar 

  71. Randle V (2010) Role of grain boundary plane in grain boundary engineering. Mater Sci Technol 26:774–780. https://doi.org/10.1179/026708309X12567268926641

    Article  CAS  Google Scholar 

  72. Randle V (2010) Grain boundary engineering: an overview after 25 years. Mater Sci Technol 26:253–261. https://doi.org/10.1179/026708309X12601952777747

    Article  CAS  Google Scholar 

  73. Randle V (2004) Twinning-related grain boundary engineering. Acta Mater 52:4067–4081. https://doi.org/10.1016/j.actamat.2004.05.031

    Article  CAS  Google Scholar 

  74. Gruber J, George DC, Kuprat AP et al (2005) Effect of anisotropic grain boundary properties on grain boundary plane distributions during grain growth. Scr Mater 53:351–355. https://doi.org/10.1016/j.scriptamat.2005.04.004

    Article  CAS  Google Scholar 

  75. Gruber J, Miller HM, Hoffmann TD et al (2009) Misorientation texture development during grain growth. Part I: Simulation and experiment. Acta Mater 57:6102–6112. https://doi.org/10.1016/j.actamat.2009.08.036

    Article  CAS  Google Scholar 

  76. Ratanaphan S, Olmsted DL, Bulatov VV et al (2015) Grain boundary energies in body-centered cubic metals. Acta Mater 88:346–354. https://doi.org/10.1016/j.actamat.2015.01.069

    Article  CAS  Google Scholar 

  77. Dillon SJ, Rohrer GS (2009) Mechanism for the development of anisotropic grain boundary character distributions during normal grain growth. Acta Mater 57:1–7. https://doi.org/10.1016/j.actamat.2008.08.062

    Article  CAS  Google Scholar 

  78. Li J, Dillon SJ, Rohrer GS (2009) Relative grain boundary area and energy distributions in nickel. Acta Mater 57:4304–4311. https://doi.org/10.1016/j.actamat.2009.06.004

    Article  CAS  Google Scholar 

  79. Rohrer GS, Li J, Lee S et al (2010) Deriving grain boundary character distributions and relative grain boundary energies from three-dimensional EBSD data. Mater Sci Technol 26:661–669. https://doi.org/10.1179/026708309X12468927349370

    Article  CAS  Google Scholar 

  80. Sutton AP (1991) An analytic model for grain-boundary expansions and cleavage energies. Philos Mag A 63:793–818

    Article  Google Scholar 

  81. Wolf D (1990) Correlation between structure, energy, and ideal cleavage fracture for symmetrical grain boundaries in fcc metals. J Mater Res 5:1708–1730. https://doi.org/10.1016/0036-9748(89)90482-1

    Article  CAS  Google Scholar 

  82. Fan Q (2012) A new method of calculating interplanar spacing: The position-factor method. J Appl Crystallogr 45:1303–1308. https://doi.org/10.1107/S0021889812037764

    Article  CAS  Google Scholar 

  83. Wang J, Beyerlein IJ (2012) Atomic structures of symmetric tilt grain boundaries in hexagonal close packed (hcp) crystals. Model Simul Mater Sci Eng 20:24002. https://doi.org/10.1088/0965-0393/20/2/024002

    Article  CAS  Google Scholar 

  84. Greninger AB, Troiano AR (1940) Crystallography of austenite decomposition. Trans AIME 140:307–336

    Google Scholar 

  85. Mackenzie J, Bowles J (1954) The crystallography of martensite transformations II. Acta Metall 2:138–147. https://doi.org/10.1016/0001-6160(54)90103-0

    Article  CAS  Google Scholar 

  86. Bowles JS, Mackenzie JK (1954) The crystallography of martensite transformations I. Acta Mater 2:129–137. https://doi.org/10.1016/0001-6160(54)90102-9

    Article  CAS  Google Scholar 

  87. Bhadeshia HKDH (2001) Worked examples in the Geometry of Crystals, 2nd edn. The Institute of Metals, London

    Google Scholar 

Download references

Acknowledgements

Extensive discussions with Prof. Bevis Hutchinson are greatly acknowledged. Deakin University's Advanced Characterization Facility is acknowledged for use of the EBSD instruments and assistance from Dr. Mark Nave. Financial support provided by the Australian Research Council is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hossein Beladi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Appendices

Appendix

Phenomenological theory of martensite transformation

The phenomenological theory of martensite transformation, which is based on the double shear mechanism introduced by Greninger-Troiano [84], has been developed by Bowles and Mackenzie (B–M theory [85, 86]) and by Wechsler, Lieberman, and Read (W–L–R theory [52, 53]). This implies that the crystallographic changes during the martensite transformation produce a linear strain resulting in existence of an interface plane with no rotation and distortion (invariant plane) at the vicinity of the martensite lath and the matrix. Thus, the strain (i.e. deformation) on this plane is called invariant plane strain. However, for an employed homogeneous deformation resulting from the crystallographic changes of the parent and product lattices, known as the Bain deformation, an invariant plane condition cannot be met. This can be generally observed through the imposing of the Bain distortion on sphere parent crystal. The deformation resulting in an ellipsoid shape shows that no lines or planes in the parent sphere crystal (shown by letters in Fig. 

Figure 11
figure 11

(a) The effect of Bain strain (dashed ellipsoid) on the parent phase (solid circle), (b) after combination with a rigid body rotation. The invariant line “cd” can be observed in (b) [40, 87]

11a) can be unextended or undistorted, meaning that the Bain distortion is not an invariant plane strain in nature. In this regard, there should be an additional shear, which does not impose a macroscopic shape, although having a microscopically inhomogeneous character. This lattice invariant shear (LIS), when superimposed on the Bain strain, results in an invariant plane within the shear transformation. Since the LIS cannot produce any crystal changes, it should have an inhomogeneous character such as deformation by dislocation slip or twinning. This can be illustrated in Fig. 7 where the lattice strain changes the initial crystal into a deformed one (Fig. 7a), and thus, the magnitude of A′B′ vector can be brought back to AB through a lattice invariant shear either by slip (Fig. 7d) or twinning (Fig. 7c) [40, 87]. Moreover, there should be an additional rotation within the matrix to coincide the A′B′ vector with the original AB (Fig. 7b).

Therefore, the martensite transformation should consist of a lattice deformation (Bain distortion), a lattice invariant shear and a lattice rotation, which can be represented in a matrix form of B, P and R, respectively. The total shape deformation can be represented, as \(P_{1} = RBP\). \(P_{1}\) can be described by assuming the lattice invariant shear P. Based on the prior descriptions and the fact that the LIS can be ascertained by slip or twinning, the total shape deformation in martensite transformation is described through Wechsler, Lieberman and Read theories in the following.

Wechsler–Lieberman–Read theory

Here, the shape deformation is presented as the P1, which can be described through the production of three matrices as follows:

$$P_{1} = RBP$$
(5)

where the R, B and P are known as the rigid body rotation, lattice deformation and a simple shear, respectively. It is now necessary to determine the plane and direction for the lattice invariant shear. The rigid body rotation does not change the length of any vectors; therefore, the vector v must remain unchanged in magnitude as a result of the combination of P and B. Meaning,

$$v^{\prime}P^{\prime}B^{\prime}BPz = v^{\prime}v$$
(6)

Here, the prime mark denotes the transpose of the matrix. Through this equation, the habit plane is determined as an undistorted plane.

If the combination matrix of BP can be defined as F, the relation changes into:

$$v^{\prime}F^{\prime}Fv = v^{\prime}v$$
(7)

Now, it is more convenient to express F as the product of an orthogonal matrix R3 and a symmetric matrix Fs. Therefore, the Fs can be diagonalized by means of orthogonal transformation:

$$z^{\prime} = R_{4} z$$
(8)

In this regard, in the basis where the Fs is diagonal, the P1 can be described as:

$$P_{1} = RR_{3} R_{4} F_{d} R^{\prime}_{4}$$
(9)

where Fd is a diagonal matrix described as:

$$F_{d} = \left[ {\begin{array}{*{20}c} {\lambda_{1} } & 0 & 0 \\ 0 & {\lambda_{2} } & 0 \\ 0 & 0 & {\lambda_{3} } \\ \end{array} } \right]$$
(10)

Now, based on the this theorem, where the condition for Eq. (6) is the presence of an undistorted plane, one of the eigenvalues of Fd must remain unity and the eigenvector must remain in the undistorted plane, which gives,

$$F^{\prime}F = R_{4} F_{d}^{2} R^{\prime}_{4}$$
(11)

Therefore, a characteristic equation:

$${\text{Det}} \left( {F^{\prime}F - \lambda^{2} I} \right) = 0$$
(12)

which can be solved. Here, I is the identity matrix.

Now, an orthonormal basis, which the lattice invariant shear takes from, needs to be introduced. This basis is defined by R5 as a 3 by 3 matrix, where each column defines the unit shear direction (i.e. slip direction, d2), the unit vector parallel to the shear plane normal (i.e. slip plane normal, p2) and t through the cross product of d2 and P2.

In this regard, the P in the orthonormal basis (P0) can be defined as:

$$P_{o} = \left[ {\begin{array}{*{20}c} 1 & g & 0 \\ 0 & 1 & 0 \\ 0 & 0 & 1 \\ \end{array} } \right]$$
(13)

Now, Bain strain can be transformed into the orthonormal basis through,

$$B_{o} = R^{\prime}_{5} BR_{5}$$
(14)

Then, F can be defined easily in the orthonormal basis,

$$F_{o} = B_{o} P_{o}$$
(15)

In this regard, Eq. (12) takes a quadratic form and can be solved. Concurrently, the corresponding eigenvectors define the orthogonal matrix R4, which as indicated diagonalizes the \(F^{\prime}F\) (i.e. Eq. 11).

Now, the g value and the undistorted plane in Eqs. 6 and 13 can be solved, respectively. The R can be also obtained using the Euler theorem. Here, two vectors in the habit plane can be considered in the orthonormal basis, naming \(\sigma\) and \(\vartheta\). Then, the vectors should be defined in the orthogonal matrix R4, as \(\overline{\sigma }\) and \(\overline{\vartheta }\). The magnitude of u (i.e. vector parallel to the desired axis of rotation) and the tangent half angle of the rotation can be obtained:

$$\frac{{\left[ {\overline{\vartheta } - \vartheta \left] \times \right[\overline{\sigma } - \sigma } \right]}}{{\left[ {\overline{\sigma } - \sigma } \right] \cdot \left[ {\overline{\vartheta } + \vartheta } \right]}} = u\tan \left( {\frac{\theta }{2}} \right)$$
(16)

Then, u and θ can be utilized to calculate the rigid body rotation (R). The R and P1 values can be obtained using the following equation and Eq. 6, respectively:

$$R = \left[ {\begin{array}{ccc} {\alpha _{{11}}^{2} \beta + \cos \theta } & {\alpha _{{11}} \alpha _{{12}} \beta - \alpha _{{31}} \sin \theta } & {\alpha _{{11}} \alpha _{{31}} \beta + \alpha _{{21}} \sin \theta } \\ {\alpha _{{21}} \alpha _{{11}} \beta + \alpha _{{31}} \sin \theta } & {\alpha _{{21}}^{2} \beta + \cos \theta } & {\alpha _{{21}} \alpha _{{31}} \beta - \alpha _{{11}} \sin \theta } \\ {\alpha _{{31}} \alpha _{{11}} \beta + \alpha _{{21}} \sin \theta } & {\alpha _{{31}} \alpha _{{21}} \beta + \alpha _{{11}} \sin \theta } & {\alpha _{{31}}^{2} \beta + \cos \theta } \\ \end{array} } \right]$$
(17)

where the \(\alpha_{ij}\) are the indices for the orthonormal matrix R5.

For the accommodation of lattice invariant strain through twinning, the parent phase is evolved through two twin-related orientations in the martensitic phase. Therefore, two equivalent crystallographic strains need to be operated to produce two twin-related variants in the product phase. This is expected to form the symmetry configuration point of view, as the transformation process has a general tendency to restore the reduction in symmetry elements through creating a number of crystallographic variants. An illustration of such is depicted in Fig 12. It is observed that the parent lattice indicated by the notation ABCD produces two twin-related equivalent lattices. Therefore, the parent lattice loss of mirror symmetry is compensated through forming the twinned product crystals.

Figure 12
figure 12

Schematic representation of the symmetry correspondence between two crystallographic variants [40]

For such transformation, the lattice deformation (Bain strain, B1) for the two twin-related orientations can be represented in their principal axes systems, as follows:

$$\begin{gathered} B^{\prime}_{1} = \left[ {\begin{array}{*{20}c} {\eta_{1} } & 0 & 0 \\ 0 & {\eta_{2} } & 0 \\ 0 & 0 & {\eta_{3} } \\ \end{array} } \right] \hfill \\ B^{\prime}_{2} = \left[ {\begin{array}{*{20}c} {\eta_{1} } & 0 & 0 \\ 0 & {\eta_{3} } & 0 \\ 0 & 0 & {\eta_{2} } \\ \end{array} } \right] \hfill \\ \end{gathered}$$
(18)

However, it is more convenient to represent the lattice strains in the parent crystal coordinate system, meaning that the axis system of the martensite crystals should be rotated into the parent phase. Therefore, B1 and B2 can be obtained, as follows:

$$\begin{gathered} B_{1} = \left[ {\begin{array}{*{20}c} {\frac{{\eta_{1} + \eta_{2} }}{2}} & {\frac{{\eta_{2} - \eta_{1} }}{2}} & 0 \\ {\frac{{\eta_{2} - \eta_{1} }}{2}} & {\frac{{\eta_{1} + \eta_{2} }}{2}} & 0 \\ 0 & 0 & {\eta_{3} } \\ \end{array} } \right] \hfill \\ B_{2} = \left[ {\begin{array}{*{20}c} {\frac{{\eta_{1} + \eta_{2} }}{2}} & 0 & {\frac{{\eta_{2} - \eta_{1} }}{2}} \\ 0 & {\eta_{3} } & 0 \\ {\frac{{\eta_{2} - \eta_{1} }}{2}} & 0 & {\frac{{\eta_{1} + \eta_{2} }}{2}} \\ \end{array} } \right] \hfill \\ \end{gathered}$$
(19)

Similar to the Bowles and Mackenzie theory, B1 and B2 can be brought into twin-related orientations through a rigid body rotation of \(\phi_{1}\) and \(\phi_{2}\), respectively. These rotations are indicated by the 1 and 2 numbers, respectively, as clearly observed in Fig. 12. Therefore, \(\phi_{1}\) and \(\phi_{2}\) describe the rotations of the principal axes of the pure distortions in regions 1 and 2 relative to an axis system fixed in the untransformed parent phase.

It should be considered that the volume fraction of each orientation can be expressed in terms of their thickness ratio, which is determined based on the IPS condition required for the lattice invariant deformation. Therefore, an arbitrary vector such as r in the matrix (Fig. 7) becomes like a twinned martensite crystal shown as a zigzag line where each orientation consumes a thickness of (1 − x) and x, respectively. The transformation of vector r into the r′ can be expressed as the sum of the vectors \(OV = OA + AB + BC + \cdots + UV\) or by the indicated lattice distortions and rigid body rotation, as follows:

$$r^{\prime} = \left[ {\left( {1 - x} \right)\phi_{1} B_{1} + x\phi_{2} B_{2} } \right]r$$
(20)
$$r^{\prime} = Er$$
(21)
$$E = \left[ {\left( {1 - x} \right)\phi_{1} B_{1} + x\phi_{2} B_{2} } \right]$$
(22)

In this regard, the total distortion matrix, E, can transform any vector in the parent matrix into an internally twinned martensite.

To define the rigid body rotations, \(\phi_{1}\) and \(\phi_{2}\) cannot be determined from the available data set. However, a rotation \(\phi\) (\(\phi_{2} = \phi_{1} \phi\)), which gives the relative rotation between \(\phi_{1}\) and \(\phi_{2}\), can be defined and the total macroscopic shear can be expressed, as follows:

$$E = \phi_{1} \left[ {\left( {1 - x} \right)B_{1} + x\phi B_{2} } \right] = F\phi_{1}$$
(23)

where \(F = \left( {1 - x} \right) B_{1} + xB_{2}\).

Therefore, the macroscopic shear tacking place during the martensitic transformation contains three components, naming \(\phi_{1}\)(i.e. rigid body rotation), F (i.e. the fraction of the shear by the twins) and \(B_{1}\) (i.e. the Bain strain). For the certain values of x (magnitude of lattice invariant shear, LIS), the matrix produces an eigenvalue problem for the vectors in the habit plane.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Farabi, E., Tari, V., Hodgson, P.D. et al. On the grain boundary network characteristics in a martensitic Ti–6Al–4V alloy. J Mater Sci 55, 15299–15321 (2020). https://doi.org/10.1007/s10853-020-05075-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10853-020-05075-7

Navigation