Skip to main content
Log in

A review of some elements for the history of mechanical twinning centred on its German origins until Otto Mügge’s K 1 and K 2 invariant plane notation

  • Interfaces and Intergranular Boundaries
  • Published:
Journal of Materials Science Aims and scope Submit manuscript

Abstract

There is a specific nomenclature for the so-called mechanical twins or twins obtained by mechanical deformation. This nomenclature, (K 1, η 1, K 2, η 2), is due to Otto Mügge in 1889. The German mineralogist and crystallographer Mügge (1858–1932) rationalised twinning observations made by himself and by other German scientists since 1859 such as Friedrich Pfaff, Heinrich Wilhelm Dove, Eduard Reusch and Heinrich Baumhauer, not to forget a formal contribution by William Thomson and Peter Guthrie Tait noticed by Theodor Liebisch who informed Mügge about it, and further classifications by Arrien Johnsen, one of Mügge’s pupils. The presentation of this scientific development also mentions the medieval alchemist Pseudo-Geber and the Renaissance metallurgist Vannoccio Biringuccio who ‘heard’ the cry of the tin now known to be due to deformation twinning, the first models of progressive twinning involving some local atomic rearrangement, with Woldemar Voigt in 1898, Yacov Il’ich Frenkel and Tatyana Kontorova in 1938 and the first drawings of a twin(ning) dislocation by Vladimirskii in 1947 and by Charles Frank and Jan van der Merwe two years later.

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

Similar content being viewed by others

Notes

  1.  Before these relatively modern scientific observations, one can also quote the much older, auditory, observations by Pseudo-Geber in the thirteenth century and Vannoccio Biringuccio in the sixteenth century. See Appendix 1.

  2.  Calcite’s chemical formula is CaCO3. It is trigonal, with a rhombohedric lattice (as α-Al2O3, see [20]). About the history of calcite, see [21]. Calcium carbonite CaCO3 can also exist in the aragonite, orthorhombic, form, stable over 107–108 years at ambient pressure and temperature, and also in the (less stable) vaterite, hexagonal, form.

  3.  (Alexius Burkhard Immanuel) Friedrich Pfaff (1825–1886), nephew of the better known pure mathematician Johann Friedrich Pfaff (1765–1825). Differential expressions used in thermodynamics (first and second laws) are Pfaffian forms. I gathered in Table 1 the names and dates of most of the people mentioned in this article.

  4.  Besides his own apparatus for a controlled pressing of his crystal samples, Pfaff writes that he used a Nörremberg’s polariscope. Before making his polariscope, probably a little after 1833, Nörrenberg (sic) (1787–1862) had devised the vacuum coffee pot (Beschreibung einer Kaffehmaschine, Zeitschrift für Physik und Mathematik (1827) 3:267–271).

  5.  Gustave Rose (1793–1875), together with his brother Heinrich Rose (1795–1864), was a famous mineralogist from a celebrated scientific Rose dynasty which could have inspired Gertrude Stein A Rose is a Rose is a Rose, had she known about them.

  6.  One should be aware that ‘twin crystal’ may mean either a group of two crystals or just one element of such a group, a ‘twinned’ part, as here. See Appendix 2.

  7.  Gustave Rose, who had not answered Reusch’s letter, published in 1868 an interesting paper where he interpreted hollow canals optically observed in calcite and previously named tubes by Brewster as being due to multiple twinning, formed where two lamellae parallel to different planes intersect [26]. These hollow canals are now known as Rose channels and have also been observed very recently in natural diamonds [27].

  8.  Auguste Bravais already spoke very similarly before the Société philomathique de Paris in 1850 [28], a Society which still exists today and whose current general treasurer is the author. Also see [1]. The word hemitropy was introduced by René Just Haüy in his 1801 Treatise of minerals: "hemitropy which means half-turned… I call hemitropic crystals the crystals which undergoe these inversions". See the Appendix of [29].

  9.  William Thomson (1824–1907) was knighted in 1866, as a reward for his participation in laying the Atlantic telegraph cable and service to science and received the title of Lord Kelvin, baron of Largs in 1892.

  10.  In the twentieth of his Baltimore Lectures on Molecular Dynamics and the Wave Theory of Light, Thomson wrote: "I never satisfy myself until I can make a mechanical model of a thing. If I can make a mechanical model I can understand it. As long as I cannot make a mechanical model all the way through I cannot understand" (in the 1884 edition of the Baltimore Lectures). The following excerpt from his Lecture IX is also worth of quotation: We shall then be able to put our formula into numbers; and I feel that I understand it much better when I have an example of it in numbers than when it is merely in a symbolic form". Accurate measures of angles in crystallography correspond to this demand to check hypotheses, since Haüy. Whereas Heinrich Baumhauer simply used Naumann’s symbol −(1/2)R to specify the twinning plane of his rhombohedron of calcite in 1879, Otto Mügge will say that Naumann’s symbols being unsuitable for calculations, he will use our modern Miller indices (Whewell-Miller and Bravais’s four indices for hexagonal systems, more balanced than Weiss’ four indices). The −(1/2)R twin is the classical ‘rhombohedral twin’ whose Bravais indices are (01–12). (The rhombohedral lattice is trigonal, not hexagonal, but the Bravais hexagonal indices are very easy to use and sufficient to compute angles).

  11.  After twinning, one half of the crystal has been sheared. The interfacial plane is obviously one plane of no distortion. All planes parallel to that plane are also planes of no distortion, in a truly simple shear deformation, i.e. forgetting about atomic shuffles which we now know about. These planes are moved, simply translated, by the shear, but not distorted. There is a second set of parallel planes (half-planes) which can be said to be ‘of no distortion’ in the sheared half. They are not distorted, simply rotated.

  12.  Neues Jahrbuch für Mineralogie, Geologie und Palaeontologie, usually abbreviated in NJM. Palaeontologie until 1904, Paläontologie afterwards. In the same spirit, that journal wrote Krystall until 1904, Kristall afterwards. A German spelling ‘reform’ (German Orthographic Conference) occured in 1901.

  13.  O. Mügge wrote the third part of the Kristallographie chapter of the German Encyclopædia of (applied) mathematical sciences in 1905, while Th. Liebisch wrote the first part and Arthur Moritz Schönflies the second part.

  14.  The English ‘shear’ was translated into Schiebung. Mügge used the indexed Greek letter σi for the vectorial directions (and ‘s’ for the shear plane).

  15.  It was Robert Cahn who, in 1953, coined the term ‘compound twin’ for a twin whose four elements are rational [39]. Georges Friedel termed ‘macles correspondantes’ the corresponding twins between reflexion and rotation twins, in 1904 (+ 1911 and 1926, see [1]), considering twin laws and not mechanical twinning.

  16.  Both Mügge and Friedel kept on working on twins until their death, in 1932 for Mügge, in 1933 for Friedel. I try in Appendix 4 to compare Mügge’s description of mechanical twins and Friedel’s classification of twins.

  17.  To the best of my knowledge, the method was provided by Hiroshi Kihô in [42], and again in 1965 by Bruce Bilby and Alan Crocker [43]. Neither Kihô nor Bilby and Crocker mention Johnsen.

  18.  Except for some possible extensions proposed by Crocker in 1962, considering homogeneous shears with orientational changes [46].

  19.  Oddly enough, Egon Orowan proposed a homogeneous twin nucleation model in 1954, which, of course, requires unreasonnable high stresses [50].

  20.  Sixteen years later, however, T. Kontorova wrote a very critical paper with Marina Victorovna Klassen-Neklyudova about the importance of dislocations for plasticity when they reported, in Russian, about Alan Cottrell’s first review paper on that subject, see [52].

  21.  Vladimirskii’s paper in Russian has never been translated. Its title, see [55] in the Reference list, is ‘on the duplication [twinning] of calcite’. Dva means two (duo, Zwei in German). Russian has another word for daily life twins, namely bliznets. The French, respectively, use macle or mâcle for twin crystals and jumeaux for human twins. See footnotes 9 and 10 in [1].

  22.  At room temperature, above 13 °C, tin is white, metallic, yet tetragonal ; hence dislocations cannot move easily and twinning is dominant. Below 13 °C, or much below that temperature because of hysteresis, tin turns grey, cubic diamond, brittle. Zinc is also known to ‘cry’. Zinc is hexagonal close-packed with a quite large c/a ratio, ~1.856 versus 1.633 (√(8/3)). Martensitic transformations in steel can also emit creaks. To hear the cry of tin, please watch https://www.youtube.com/watch?v=7rWIHR4pB9s (by Jessica Gwynne from the University of Cambridge).

  23.  Yet, Robert Cahn wrote in 1954 that his personal tests "with tin containing substantial amount of added impurity do not bear out the statement" [Cahn54]. Pseudo-Geber was even more astray since he believed, from his tests, that the substance causing the creak in tin was mercury [62]. We, of course, cannot really blame neither Pseudo-Geber nor Biringuccio for these early mistakes. Pseudo in the phrase Pseudo-Geber means: not the true Geber, yet looking like the true Geber. In the same way, a pseudo-symmetry in crystallography looks like a true symmetry although it is not a true symmetry, just an approximate one. Pseudo, from the ancient Greek pseudés (ψευδής) meaning lying or false, does no longer have to have a pejorative meaning which it still has as in pseudo-science for instance.

  24.  Twin crystals were first identified and defined as such by Jean-Baptiste Romé de l’Isle in 1783, for natural crystals analysable with the naked eye and some instrument to measure angles (a goniometer), see the first paragraph of Appendix 3.

  25.  The symbol in the Italian text stands for ‘s’. It was customary in old European texts. One still uses it for the continuous sum, or integral, symbol thanks to Gottfried Wilhelm Leibniz, when Σ is used for discrete sums thanks to Leonhard Euler.

  26.  A group of well-formed crystals which simply happened to grow together is called a crystal cluster.

References

  1. Hardouin Duparc OBM (2011) A review of some elements in the history of grain boundaries centered on Georges Friedel, the coincident ‘site’ lattice and the twin index. J Mater Sci 46:4116–4134. doi:10.1007/s10853-011-5367-1

    Article  Google Scholar 

  2. Mügge O (1889) Ueber homogene Deformationen (einfache Schiebungen) an den triklinen Doppelsalzen BaCdCl4.4aq., Neues Jahrbuch für Mineralogie, Geologie und Palaeontologie Beilage-Band 6:274–304 (+ Taf. IX)

  3. Christian JW (1965, 2002) The theory of transformations in metals and alloys. Elsevier, Oxford

  4. Hirth JP, Lothe J (1982) Theory of dislocations. 1st edn. 1968 McGraw-Hill, 2nd edn. Wiley, New York

    Google Scholar 

  5. Kelly AA, Knowles KM (2012) Crystallography and crystal defects, 2nd edn. Wiley, New York

    Book  Google Scholar 

  6. Niggli P (1920) Lehrbuch der Mineralogie. Gebrüder Borntraeger, Berlin

    Google Scholar 

  7. Friedel G (1926) Leçons de Cristallographie professées à la Faculté des sciences de Strasbourg. Berger-Levrault, Nancy (reprinted as Leçons de Cristallographie in 1964 by A. Blanchard, Paris)

  8. Schmid E, Boas W (1935) Kristallplastizität mit besonderer Berücksichtigung der Metalle. Springer, Berlin; translated, (1950) plasticity of crystals with special references to metals. Hughes and Co., London

    Google Scholar 

  9. Tertsch H (1949) Die Festigkeitserscheinungen der Kristalle. Springer, Vienna

    Book  Google Scholar 

  10. Hall EO (1954) Twinning and diffusionless transformations in metals. Butterworths Scientific Publications, London

    Google Scholar 

  11. Klassen-Neklyudova MV (1964) Mechanical twinning of crystals, 1960 in Russian, revised and expanded for the authorized translation by J.E.S. Bradley. Consultants Bureau, New York

  12. Reed-Hill RE, Hirth JP, Roger HC (eds) (1864) Deformation twinning. Gordon and Breach, New York

    Google Scholar 

  13. Cahn RW (1953) Soviet work on mechanical twinning. Nuovo Cimento 4:350–386

    Article  Google Scholar 

  14. Cahn RW (1954) Twinned crystal. Adv Phys 3:363–445

    Article  Google Scholar 

  15. Partridge PG (1967) The crystallography and deformation modes of hexagonal crystals. Metall Rev 12:169–194

    Google Scholar 

  16. Christian JW, Mahajan S (1995) Deformation twinning. Prog Mater Sci 39:1–157

    Article  Google Scholar 

  17. Beyerlein IJ, Zhang X, Misra A (2014) Growth twins and deformation twins in metals. Annu Rev Mater Res 44:329–363

    Article  Google Scholar 

  18. Pond RC, Hirth JP, Serra A, Bacon DJ (2016) Atomic displacements accompanying deformation twinning: shears and shuffles. Mater Res Lett 4:185–190

    Article  Google Scholar 

  19. Boulliard J-C (2010) Le cristal et ses doubles. CNRS Editions, Paris

    Google Scholar 

  20. Hardouin Duparc OBM (2014) Crystallography, group theory, etymology and ‘pataphysics. Math Intell 36:54–61

    Article  Google Scholar 

  21. Kristjánsson L (2010) Iceland Spar and its influence on the development of science & technology in the period 1780–1930, 3rd edn. Univ, Iceland

    Google Scholar 

  22. Huygens C (1690) Traité de la lumière. Pierre Vander, Leiden

  23. Pfaff Fr (1859) Versuche über den Einfluß des Drucks auf die optischen Eigenschaften doppelbrechender Krystalle. Annalen der Physik und Chemie 107:333–338 (+ a part II, 108:598–601)

  24. Dove HW (1860) Optische Notizen. 1. Ueber Kalkspathzwillinge. Annalen der Physik und Chemie 110:286

    Google Scholar 

  25. Reusch (1869) On the effect of pressure upon rock salt and iceland spar. proceedings of the royal society of edinburgh, session 1866–67 6:134–139 (Because F. Eduard Reusch was not a British Professor and because he wrote in French, the convention was to call him as M. [standing for Monsieur] Reusch, and then give his position, viz. Professor of Natural Philosophy in the University of Tubingen, instead of calling him Professor Reusch). Also see Reusch E (1867) Ueber eine besondere Gattung von Durchgängen in Steinsalz und Kalkspath. Annalen der Physik und Chemie 132:441–451

  26. Rose G (1868) Über die im Kalkspath vorkommenden hohlen Canäle. Abhandlungen der Königlichen Akademie der Wissenschaften, Berlin 55–70

  27. Schoor M, Boulliard JC, Gayou E, Hardouin Duparc O, Estève I, Baptiste B, Rondeau B, Fritsch E (2016) Plastic deformation in natural diamonds: rose channels associated with mechanical twinning. Diam Relat Mater 66:102–106

    Article  Google Scholar 

  28. Bravais A (1851) Études cristallographiques, Bachelier, Paris, 1851. Spec. the third part: Des macles et des hémitropies, first communicated to the Société Philomathique de Paris in 1850

  29. Liang L, Hardouin Duparc O (2016) First-principles study of four mechanical twins and their deformation along the c-axis in pure α-titanium and in titanium in presence of oxygen and hydrogen. Acta Mater 110:258–267

    Article  Google Scholar 

  30. Baumhauer H (1879) Über künstliche Kalkspath-Zwillinge nach–½ R. Zeitschrift für Krystallographie und Mineralogie 3:588–591

    Google Scholar 

  31. Johnsen A (1914) Schiebungen und Translationen in Kristallen. Jahrbuch der Radioaktivität und Elektronik 11:226–262

    Google Scholar 

  32. Thomson W (1890) Molecular Constitution of Matter, Proceedings of the Royal Society of Edinburgh 16:693–724, reprinted in his Mathematical and Physical Papers, vol. 3 (1890) pp. 395–427, spec. p. 420 (58)

  33. Thomson W, Tait PG (1867) Treatise on Natural Philosophy. Cambridge, vol. I (1) 170-171, pp. 105-106 (pp.123-124 in the second edition 1879, §§149-150 in the Elements of Natural Philosophy, a simplified version of the Treatise)

  34. Liebisch Th (1887) Ueber eine besondere Art von homogenen Deformationen krystallisirter Körper. Nachrichten von der Königlichen Gesellschaft der Wisssenschaften und der Georg-Augusts-Universität zu Göttingen 435–448

  35. Liebisch Th (1889) Ueber eine besondere Art von homogenen Deformationen krystallisirter Körper. Neues Jahrbuch für Mineralogie, Geologie und Palaeontologie Beilage-Band 6:105–120 (+ Taf. I)

  36. Mügge O (1883) Beiträge zur Kenntniss der Structurflächen des Kalkspathes und über die Beziehungen derselben untereinander und zur Zwillingsbildung am Kalkspath und einigen anderen Mineralien. Neues Jahrbuch für Mineralogie, Geologie und Palaeontologie 1:32–54

    Google Scholar 

  37. Mügge O (1889) Ueber die Krystallform des Brombaryums BaBr2.2H2O, und verwandter Salze und über Deformationen derselben. Neues Jahrbuch für Mineralogie, Geologie und Palaeontologie 1: 130–178 (+ Taf. II)

  38. Mügge O (1894) Ueber „reciproke“ einfache Schiebungen an den trikliniken Doppelsalzen K2Cd(SO4)2.H2O, K2Mn(SO4)2.H2O und verwandten. Neues Jahrbuch für Mineralogie Geologie und Palaeontologie 1:106–108

    Google Scholar 

  39. Cahn RW (1953) Plastic deformation of alpha-uranium; twinning and slip. Acta Mater 1:49–70

    Article  Google Scholar 

  40. Johnsen A (1907) Untersuchungen über Kristallzwillinge und deren Zusammenhang mit anderen Erscheinungen. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie Beilage-Band 23:237–344

    Google Scholar 

  41. Wallerant FFA (1904) Des macles secondaires et du polymorphisme. Bulletin de la Société française de minéralogie 27:169–189

    Google Scholar 

  42. Kihô H (1954) The Crystallographic Aspect of the Mechanical Twinning in Metals. J Phys Soc Jpn 9:739–747

    Article  Google Scholar 

  43. Bilby BA, Crocker AG (1965) The theory of the Crystallography of Deformation Twinning. Proceedings of the Royal Society A 288:240–255

    Article  Google Scholar 

  44. Johnsen A (1916) Die wahrscheinlichsten atombewegungen im wismut während einer schiebung. Centralblatt für Mineralogie, Geologie und Paläontologie 385-392

  45. Johnsen A (1917) Die einfachsten Bahnen der Atom während der Schiebungen im Eisenglanz und Korund. Centralblatt für Mineralogie, Geologie und Paläontologie 433–445

  46. Crocker AG (1962) Double twinning. Phil Mag 7:1901–1925

    Article  Google Scholar 

  47. Voigt W (1898) Bemerkung über dir Größe der Spannungen und Deformationen, bei denen Gleitschichten im Kalkspat entstehen. Nachrichten von der Königlichen Gesellschaft der Wissenschaften zu Göttingen. Mathematische-physikalische Klasse. 146–153; reproduced in (1899) Annalen der Physik und Chemie 67: 201–208 (with the Esszet letter ‘ß’ turned back into ‘ss’)

  48. Polanyi M (1921) Über die Natur des Zerreißvorganges. Zeitschrift für Physik 7:323–327

    Article  Google Scholar 

  49. Zwicky F (1923) Die Reissfestigkeit von Steinsalz. Physikalische Zeitschrift 24:131–137

    Google Scholar 

  50. Orowan O (1954) Dislocations and mechanical properties. In: Cohen M (ed) Dislocations in metals. Am Inst Mining Metall Eng New York, 1954, pp. 69–195

  51. Kontorowa TA, Frenkel YI (1938) K teorii plasticheskoï deformatsii i dvoïnikovanija. Zhurnal Eksperimental’noï Teoreticheskoï Fiziki 8:89–95, translated as Frenkel YI, Kontorowa (1938) On the theory of plastic deformation and twinning. Physikalische Zeitschrift der Sowjetunion 13:1–10; J Phys Acad Sci USSR, (1939) 1:137–149

  52. Friedel J, Hardouin Duparc OBM (2013) Alan Cottrell, a fundamental metallurgist. Philos Mag 28–30:3703–3713

    Article  Google Scholar 

  53. Seitz F, Read WT (1941) Theory of the plasticity of solids III. J Appl Phys 12:470–486

    Article  Google Scholar 

  54. Frank FC, van der Merwe JH (1949) One-dimensional dislocations I. Static theory. Proc R Soci A 198:205–216

    Article  Google Scholar 

  55. Vladimirskii KV (1947) O dvoïnikovanie kalcita. Zhurnal Eksperimental’noï teoreticheskoï Fiziki 17:530–536

    Google Scholar 

  56. Khater HA, Serra A, Pond RC (2013) Atomic shearing and shuffling accompanying the motion of twinning disconnections in zirconium. Phil Mag 10–12:1279–1298

    Article  Google Scholar 

  57. Pond RC (1989) Line defects in interfaces. In: Nabarro FRN (ed) Dislocations in solids. North-Holland, Amsterdam, pp 1–66

    Google Scholar 

  58. Pond RC (1989) TEM studies of line defects in interfaces. Ultramicroscopy 30:1–7

    Article  Google Scholar 

  59. Hirth JP (1994) Dislocations, steps and disconnections at interfaces. J Phys Chem Solids 55:985–989

    Article  Google Scholar 

  60. Pond RC, Hirth JP (1994) Defects at surfaces and interfaces. In: Seitz F, Turnbull D (eds.) Solid state physics, Academic Press, New York 47:287–365

  61. Lainé SJ, Knowles KM (2015) {1124} deformation twinning in commercial purity titanium at room temperature. Phil Mag 95:2153–2166

    Article  Google Scholar 

  62. Newman WR (1991) The summa perfectionis of Pseudo-Geber. A critical edition, translation and study. E.J. Brill, Leyden

  63. Mellor JW (1927) Comprehensive treaty on inorganic and theoretical chemistry, vol 7. Longmans Green and Co., New York, pp 279–296

    Google Scholar 

  64. Smith CS, Gnudi MT (1942) The Pirotechnia of Vannoccio Biringuccio: the classic sixteenth-century treatise on metals and metallurgy. Am Inst Min Metall Eng, republished in 1943 and 1959, and in 1990 by Dover, New York

  65. Kalisher S (1881) Ueber den Einfluss der Wärme auf die Molekularstruktur des Zinks. Ber dtsch chem Ges 14:2747–2753

    Article  Google Scholar 

  66. Mathewson CH, Phillips AJ (1927) Plastic deformation of coarse-grained zinc. Proc. IMD, Am Inst Min Metall Eng 143–194

  67. Mathewson CH, Edmunds GH (1928) The Neumann Bands in Ferrite. Trans Am Inst Min Metall Eng 80:311–333

    Google Scholar 

  68. Wadley HNG, Scruby CB, Speake JH (1980) Acoustic emission for physical examination of metals. Int Met Rev 2:41–64

    Google Scholar 

  69. Hahn Th, Klapper H (2014) Twinning of Crystals. In: Authier A (ed) International tables of crystallography, vol. D, 2006, 2014, pp. 413–487

Download references

Acknowledgements

I thank the Library Department of the École Polytechnique from where I could obtain some old articles not available with the World Wide Web medium. I also thank my two Reviewers for their careful reading of my typewritten manuscript and their suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to O. B. M. Hardouin Duparc.

Ethics declarations

Conflict of interest

The author declares that he has no conflict of interest.

Appendices

Appendix 1

Pseudo-Geber and Vannocio Biringuccio. Creak or cry tests on tin: From written knowledge, the first scientists who intentionally produced mechanical twinning, even if they didn’t know it was twinning, are the pseudo-Geber and Vannoccio Biringuccio with their creak or cry tests on tin,Footnote 22 between his teeth with Biringuccio for whom it was a test of purity.Footnote 23 Twins or twin crystals, however, were not yet conceived in those times,Footnote 24 let alone twinning in a metal, so that these mentions are only of anecdotical value. These observations by Pseudo-Geber and by Biringuccio did not have any influence on the advancement of mechanical twinning.

Pseudo-Geber is an anonymous European alchemist born in the thirteenth century once confused with Jabir ibn Hayyan, an eigth century Islamic alchemist. In his summa perfectionis (Geberi regis Arabum Summa Perfectionis ministerii), probably written in the thirteenth century, one finds: “I intimate to the sons of learning that tin is a metallic substance which is white, but not pure white; it has a little ring, and emits a creaking sound [sonans parum stridorem];” as translated by Joseph William Mellor [63] and see [62] for both a latin version and a commented translation; Pseudo-Geber also tried to draw conclusions from experiments since he could not be aware that the state of knowledge was too crude in his time to lead to good conclusions. He noted that “After its double calcination, it [tin] does not creak [post vero illius duplicem calcinationem non stridet]”, but believed that the substance causing the creak in tin was mercury [62].

Vannoccio Biringuccio (1480–c.1539) is better known than Pseudo-Geber and he can be considered as a father of the foundry industry, thanks to his famous book De la pirotechnia published a little after his death in 1540 and where he wrote, speaking about tin: “That metal is known to be purer… if when some thin part of it is bent or squeezed by the teeth it gives its natural cracking noise, like that which water makes when it is frozen by cold [Cogno∫ce∫∫i que∫to tanto e∫∫er piu puro,… piegandolo, in qualche parte ∫ottile, o col dente ∫tringendolo, ∫i ∫ente un natural ∫uo ∫tridore, come fa l’acqua dal freddo gelata Footnote 25]” as translated by Martha Teach Gnudi and Cyril Stanley Smith [64]. Oddly, C.S. Smith, albeit a great metallurgist and historian of science, makes no footnote about the cry of tin.

If the creaks emitted by tin or zinc under bending are not a sign of purity, it is a sign of cristallinity as it appeared to the German metallurgist Kalisher in 1881: “The assumption that the “cry” of tin [dasSchreiendes Zinns] is a consequence of the crystalline structure finds a new confirmation with zinc” [65]. Mellor’s interpretation of the phenomenon in 1927 was not yet exactly correct, as a kind of internal friction between crystals in polycrystalline tin: “The crystalline structure of tin is further evidenced by the fact noted by Geber in (…); this is supposed to be produced by the grinding of the crystals against one another during the bending of the metal” [63]. Also in 1927, Champion Herbert Mathewson and Albert J. Phillips noticed that “Although satisfactory commercial fine-grained soft strip does not emit a perceptible zinc cry when bent, somewhat coarser-grained strip does, and apparently the intensity of the cry depends on the size of the crystals and might even be recognized in very fine-grained material by using a suitable audiometric method” [66].

Mathewson and G.H. Edmunds probably were the first to attribute Biringuccio’s experience to twinning in 1928, at least implicitly, in their paper about the production of Neumann bands in silicon ferrite, which they clearly recognised as twinning lamellae: “Neumann bands were produced in this material by bending, stretching, rolling and hammering. Their appearance was accompanied by crackling sound similar to the tin or zinc “cry” [67]. (It is amusing to see Mathewson and Edmunds speaking of Mügge’s Einfache Schiebung, not knowing it was a German translation of the English ‘simple shear’ introduced by Thomson and Tait, see the second section about Thomson–Tait, Liebisch and Mügge, and Appendix 3).

The acoustic emission from solids under deformation is nowadays a well-recognised experimental technique, see for instance [68].

Appendix 2

Vocabulary Problems. Twin as a group of two or as a part of a group of two: ‘Twin’ came to be an ambiguous noun: a twin, or twin crystal, or mascle, as originally defined by Romé de L’Isle in 1783 designed a whole, a special group of two crystals: “When, in a given crystal, there is one or several entering angles, one must conclude that this is not an elementary crystal but a group of two or several crystals, or even two inverted halves of a similar crystal. This crystal is then named a MACLE”, see [1]. That definition of a twin as a group of two crystals was repeated by William Hallowes Miller in his 1839 Treatise on Crystallography: “A TWIN crystal is composed of two crystals joined together (in such a manner that…Footnote 26)”. Yet, for veins which were observed in calcite and determined to be in twinned orientation with respect to the matrix (see Brewster and Reusch for instance), the noun ‘twin crystal’ or ‘twin’ was given to these veins themselves, as a synecdoche. The same synecdoche is used for the more or less extended lenticular veins observed in deformed metals for instance.

One must thus take care that the word ‘twin’ may designate a group of two crystals or just one part of such a group, specially when that part clearly looks embedded in the other part which can be considered as the matrix or parent part. Charles Frank, inspired by the Greek mythology proposed to name Castor and Pollux the two parts which do look equal and tightly related at the atomic level (see [43]). In most articles dealing with disconnections, one finds one crystal labelled with the Greek letter λ and the other with the Greek letter μ. The λ part is normally drawn with white atoms, since λ is the first letter of leucos which means white in Greek, with the μ part normally drawn with black atoms since μ is the first letter of melanos meaning black in Greek [Bob Pond, personal communication].

Our familiar symmetric tilt grain boundaries (GBs) have been first described by René-Just Haüy and by Auguste Bravais as 180° twist GBs: half-turned crystals with respect to an axis perpendicular to the boundary surface: twins by hemitropy.

Appendix 3

Otto Mügge’s original texts in 1889: Otto Mügge [2], p. 281: “Es sind zweierlei Deformationen beobachtet. Beides sind „einfache Schiebungen”; und zwar führt die eine, im folgenden als α bezeichnet, die Krystalle in Zwillingsstellung nach {010} über, die zweite, im folgenden mit β bezeichnet, bewirkt Zwillingsbildung nach [010]”. Which I translate as: “There are two types of observed deformations [with the triclinic salt Mügge studied]. Both are “simple shears”, to wit: the first one, referred as α in the following, brings the crystal in twinning position along the crystal in twinning position along {010}, the second, referred as β in the following, causes twinning by [010]”.

Otto Mügge [2], p. 286: “Die im Vorstehenden beschriebenen Deformationen sind offenbar homogene, und zwar einfache Schiebungen. Bei der ersten (α), deren allgemeine Verhältnisse wir zunächst untersuchen wollen, ist {010} die rationale Gleitfläche: wir bezeichnen dieselbe allgemein mit k1 = {k11, k12, k13}; sie ist zugleich die erste Kreisschnittsebene und Zwillingsfläche. Für einfache Schiebungen ist weiter charakteristisch eine andere, zweite Kreisschnittsebene genannte Ebene; ihr allgemeines Zeichen sei k2 = {k21, k22, k23}”. I again freely translate as: “The previously mentioned deformations are obviously homogeneous, namely simple shears. For the first one, (α), the general relations of which we want to examine first, the rational gliding surface is {101}: we designate it in general as k1 = {k11,k12,k13}. It is at the same time the first circle cutting plane (circular cross-sectional plane) and twin surface. For simple shears there is still another, characteristic circle cutting plane, called second cutting plane. Let its general description be k2 = {k21, k22, k23}”.

Appendix 4

Mügge’s (K 1,η 1,K 2,η 2) description of mechanical twins and Friedel’s classification of twins: The aim of Georges Friedel’s classification of twins was to classify all observed twins whatever their origins, during crystal growth or after crystal growth. Otto Mügge’s (K 1, η 1, K 2, η 2) only deals with mechanical twins, twins obtained by mechanical deformation of crystals. It is clear, as already noticed by Robert Cahn [14], that Friedel’s ‘twins by merohedry’, for which the lattice has the same orientation for the two crystals (only the atomic motif is changed from one crystal to the other, and the boundary does not need to be planar), are not mechanical twins. Only ‘twins by reticular (i.e. lattice) merohedry’ can be formally described by a (K 1, η 1, K 2, η 2) set, even if they have not been obtained by deformation, such as annealing twins in metals (i.e. twins observed after annealing a polycrystalline metal). Georges Friedel was concerned about finding a common lattice between the two crystals of the bicrystals, the crystal lattice itself for ‘twins by merohedry’ or a small multiple lattice of the two disoriented crystal lattices, with as small a multiplicity (the twin index) as possible in mineral twins (see [1]). Otto Mügge was concerned with a proper description of the mechanical twins, in terms of simple shear thanks to Thomson, Tait, and Liebisch. We now know that, except for cubic crystals, many mechanical twins actually do not have any meaningful coincidence lattice. For instance, for mechanical twins observed in hexagonal metals, a ‘twin index’ could only be defined for approached values of the physical ratio c/a which differs from one hexagonal metal to another when the (K 1, η 1, K 2, η 2) description remains the same. It is probably pointless to discuss from the point of view of mechanical twinning the twins which correspond to approximate (pseudo, see end of footnote 23) merohedry or to approximate (pseudo) reticular merohedry within the angular limits of tolerance defined by Georges Friedel, viz. around five or six degrees, nor all the other classes of twins which have been defined since (see [69]).

Appendix 5

Tentative illustration of the meaning of the K 1, K 2, η 1, and η 2 symbols: This appendix is intended to illustrate for a modern non-specialist reader the concepts and the K 1, K 2, η 1, and η 2 symbols introduced in the present paper. It is illustrated with a realist, simplified, special case, trying to be both not too complicated and not too simple, clearly an impossible task and I apologise on both sides. It should also be kept in mind that although Otto Mügge knew his crystals were made of atoms, he only knew of lattice and crystalline symmetries and did not know the positions of the atoms within the motifs of his complicated minerals (with chemical units such as BaCdCl4·4H2O or K2(Cd,Mn)(SO4)2·H2O).

Before the bicrystal state schematically shown in white and black in Fig. 5, what originally existed, as in the top of Fig. 2, was a monocrystal (white circles, corresponding to atoms or simply to lattice nodes). What exists at a subsequent time is a bicrystal with unchanged white atoms below a K 1 plane and moved atoms above it. These moved atoms are shown in black. They are in symmetrical positions with respect to the lower white atoms with respect to K 1. The drawing specifically corresponds to one (\( 1\bar{1}0 \)) plane of a {111} twin in an fcc crystal (such a twin is known to have a very small interfacial energy in copper for instance. It could occur as a growth twin or as an annealing twin. We do not assume such an origin here). One can obviously not believe that the positions of the black atoms are the result of a physical half-turn (hemitropic) rotation of the upper white atoms with respect to a well-chosen axis perpendicular to K 1 (viz. the prolongation in the upper part of the [111] vector drawn in Fig. 5) even if it is true from a pure geometrical point of view. One can try to imagine a more plausible physical way, in terms of a homogeneous (simple) shear motion along x, with s x (y) = s·y, where y is the distance to K 1, to be applied to every white i atom having y i  > 0. Several possibilities still exist. The one which would transform the set of K 2b (half-)planes ((110) planes) into K 2b (half-)planes would imply much larger sb·y i atomic motions than the one which transforms the set of K 2a (half-)planes [(\( 11\bar{1} \)) planes] into K 2 (half-)planes (with the fcc {111} twin choice, one has s a = 1/√2 versus 2√2 for s b). One thus chooses K 2K 2a, and the horizontal arrows show the corresponding shear atomic motions. The plane of shear is the (x, y) plane of the sheet [here a (\( 1\bar{1}0 \)) plane]. One then notes η 1, the shear direction vector. It belongs to K 1 and the plane of shear (here η 1∝[\( 11\bar{2} \)]). One also defines a vector η 2 which belongs to K 2 and the plane of shear and which may, by conventional choice, reflect or not the periodicity of the lattice (here η 2∝[112]). We thus have the (K 1, η 1, K 2, η 2) set that defines a twinning mode. Yet, even such a simple shear motion is unlikely: depending on how the deformation stress is applied, the real atomic motion may occur only progressively from the top, such as shown in Fig. 2 and certainly locally and progressively from left to right, pushing rightwards the upper part of the crystal, with a twinning dislocation such as schematically shown in Fig. 4. Some atomic shuffles also certainly occur at the atomic level (see [56] and [18] with illustrations from atomic simulations).

Figure 5
figure 5

Tentative illustration of a possibly mechanical twin and its twinning elements. The white (open) circles below the trace of the K 1 plane, together with the black (filled) circles may correspond to the atoms of a (\( 1\bar{1}0 \)) plane of a {111} twin in an elemental face-centred cubic crystal such as copper. See text in Appendix 5 for more comments

The phrase “invariant” planes corresponds to the fact that the set of all planes parallel to K 1 are not distorted by the simple s x (y) = s·y shear. They are just translated. The set of all (upper half-)planes parallel to K 2 are not distorted either by the simple shear, in the sense that they are just rotated into their corresponding K 2 planes (Fig. 5 shows two such K 2 planes). Any other plane will be distorted.

Figure 6, slightly modified from the 1954 book by Eric Ogilvie Hall (better known for sharing his name with Norman Petch for a famous relationship in plasticity) [10] shows the relations of a sphere and its half upper part homogeneously sheared into a half-ellipsoid. The ‘planes of the circular sections’, or Kreisschnittebenen, are clearly visible. The bicrystal with its nodes and atoms are not visible.

Figure 6
figure 6

“Invariant” plane relations between a sphere and its half upper part homogeneously sheared into a half-ellipsoid (from Hall 1954 [10])

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hardouin Duparc, O.B.M. A review of some elements for the history of mechanical twinning centred on its German origins until Otto Mügge’s K 1 and K 2 invariant plane notation. J Mater Sci 52, 4182–4196 (2017). https://doi.org/10.1007/s10853-016-0513-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10853-016-0513-4

Keywords

Navigation