Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Exploding and weeping ceramics

Abstract

The systematic tuning of crystal lattice parameters to achieve improved kinematic compatibility between different phases is a broadly effective strategy for improving the reversibility, and lowering the hysteresis, of solid–solid phase transformations1,2,3,4,5,6,7,8,9,10,11. (Kinematic compatibility refers to the fitting together of the phases.) Here we present an apparently paradoxical example in which tuning to near perfect kinematic compatibility results in an unusually high degree of irreversibility. Specifically, when cooling the kinematically compatible ceramic (Zr/Hf)O2(YNb)O4 through its tetragonal-to-monoclinic phase transformation, the polycrystal slowly and steadily falls apart at its grain boundaries (a process we term weeping) or even explosively disintegrates. If instead we tune the lattice parameters to satisfy a stronger ‘equidistance’ condition (which additionally takes into account sample shape), the resulting material exhibits reversible behaviour with low hysteresis. These results show that a diversity of behaviours—from reversible at one extreme to explosive at the other—is possible in a chemically homogeneous ceramic system by manipulating conditions of compatibility in unexpected ways. These concepts could prove critical in the current search for a shape-memory oxide ceramic9,10,11,12.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Lattice correspondences and transformation stretch matrices.
Fig. 2: Frame sequences of martensitic transformation.
Fig. 3: Thermal characterization and kinematic compatibility.

Similar content being viewed by others

Data availability

The raw data that support the findings of this study are available at https://archive.materialscloud.org with the identifierhttps://doi.org/10.24435/materialscloud:6c-hk.

References

  1. Chen, X., Srivastava, V., Dabade, V. & James, R. D. Study of the cofactor conditions: conditions of supercompatibility between phases. J. Mech. Phys. Solids 61, 2566–2587 (2013).

    Article  MathSciNet  CAS  ADS  Google Scholar 

  2. Chluba, C. et al. Ultralow-fatigue shape memory alloy films. Science 348, 1004–1007 (2015).

    Article  CAS  ADS  Google Scholar 

  3. Delville, R. et al. Transmission electron microscopy study of phase compatibility in low hysteresis shape memory alloys. Philos. Mag. 90, 177–195 (2010).

    Article  CAS  ADS  Google Scholar 

  4. Della Porta, F. On the cofactor conditions and further conditions of supercompatibility between phases. J. Mech. Phys. Solids 122, 27–53 (2019).

    Article  MathSciNet  ADS  Google Scholar 

  5. Niitsu, K., Kimura, Y., Omori, T. & Kainuma, R. Cryogenic superelasticity with large elastocaloric effect. NPG Asia Mater. 10, e457 (2018).

    Article  Google Scholar 

  6. Zarnetta, R. et al. Identification of quaternary shape memory alloys with near-zero thermal hysteresis and unprecedented functional stability. Adv. Funct. Mater. 20, 1917–1923 (2010).

    Article  CAS  Google Scholar 

  7. Gu, H., Bumke, L., Chluba, C., Quandt, E. & James, R. D. Phase engineering and supercompatibility of shape memory alloys. Mater. Today 21, 265–277 (2018).

    Article  CAS  Google Scholar 

  8. Jetter, J. et al. Tuning crystallographic compatibility to enhance shape memory in ceramics. Phys. Rev. Mater. 3, 93603 (2019).

    Article  CAS  Google Scholar 

  9. Wegner, M., Gu, H., James, R. D. & Quandt, E. Correlation between phase compatibility and efficient energy conversion in Zr-doped barium titanate. Sci. Rep. 10, 3496 (2020).

    Article  CAS  ADS  Google Scholar 

  10. Liang, Y. G. et al. Tuning the hysteresis of a metal-insulator transition via lattice compatibility. Nat. Commun. 11, 3539 (2020).

    Article  CAS  ADS  Google Scholar 

  11. Pang, E. L., McCandler, C. A. & Schuh, C. A. Reduced cracking in polycrystalline ZrO2-CeO2 shape-memory ceramics by meeting the cofactor conditions. Acta Mater. 177, 230–239 (2019).

    Article  CAS  ADS  Google Scholar 

  12. Lai, A., Du, Z., Gan, C. L. & Schuh, C. A. Shape memory and superelastic ceramics at small scales. Science 341, 1505–1508 (2013).

    Article  CAS  ADS  Google Scholar 

  13. Kohn, R. V. & Müller, S. Branching of twins near an austenite—twinned-martensite interface. Philos. Mag. A 66, 697–715 (1992).

    Article  CAS  ADS  Google Scholar 

  14. Burkart, M. W. & Read, T. A. Diffusionless phase change in the indium-thallium system. Trans. Am. Inst. Min. Metall. Eng. 197, 1516–1524 (1953).

    Google Scholar 

  15. Ball, J. M. & James, R. D. in Analysis and Continuum Mechanics: a Collection of Papers Dedicated to J. Serrin on His Sixtieth Birthday 647–686 (Springer, 1989).

  16. Yang, S. et al. A jumping shape memory alloy under heat. Sci. Rep. 6, 21754 (2016).

    Article  CAS  ADS  Google Scholar 

  17. Yadava, K. et al. Extraordinary anisotropic thermal expansion in photosalient crystals. IUCrJ 7, 83–89 (2020).

    Article  CAS  Google Scholar 

  18. Naumov, P., Chizhik, S., Panda, M. K., Nath, N. K. & Boldyreva, E. Mechanically responsive molecular crystals. Chem. Rev. 115, 12440–12490 (2015).

    Article  CAS  Google Scholar 

  19. Tong, F. et al. Photomechanical molecular crystals and nanowire assemblies based on the [2+2] photodimerization of a phenylbutadiene derivative. J. Mater. Chem. C 8, 5036–5044 (2020).

    Article  CAS  Google Scholar 

  20. Chandrasekar, S. & Chaudhri, M. M. The explosive disintegration of Prince Rupert’s drops. Philos. Mag. B 70, 1195–1218 (1994).

    Article  CAS  ADS  Google Scholar 

  21. Gibbs, J.W. On the equilibrium of heterogeneous substances Vol. 1 (Longmans Green and Co., 1879)

  22. Chen, X. et al. Direct observation of chemical short-range order in a medium-entropy alloy. Nature 592, 712–716 (2021).

    Article  CAS  ADS  Google Scholar 

  23. Hayakawa, M., Kuntani, N. & Oka, M. Structural study on the tetragonal to monoclinic transformation in arc-melted ZrO2-2mol.%Y2O3—I. Experimental observations. Acta Metall. 37, 2223–2228 (1989).

    Article  CAS  Google Scholar 

  24. Hayakawa, M. & Oka, M. Structural study on the tetragonal to monoclinic transformation in arc-melted ZrO2-2mol.%Y2O3—II. Quantitative analysis. Acta Metall. 37, 2229–2235 (1989).

    Article  CAS  Google Scholar 

  25. Chen, X. & Song, Y. Structural Phase Transformation Web Tools (StrucTrans, 2014); http://www.structrans.org

  26. Ball, J. M. & James, R. D. A characterization of plane strain. Proc. Math. Phys. Sci. 432, 93–99 (1991).

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

R.D.J. and H.G. were supported by the NSF (DMREF-1629026), the MURI programme (FA9550-18-1-0095 and FA9550-16-1-0566) and a Vannevar Bush Faculty Fellowship. R.D.J. also acknowledges a Mercator Fellowship for the support of this German–US collaboration. E.Q., J.R. and J.J. acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG) through a Reinhart Koselleck Project (313454214), and the project “Search for compatible zirconium oxide-based shape memory ceramics” (453203767). We thank N. Wolff for preliminary TEM measurements, and A. Mill for her assistance in the preparation of specimens for TEM investigation.

Author information

Authors and Affiliations

Authors

Contributions

H.G. and J.R. contributed equally to the manuscript. J.R. synthesized and characterized the samples. E.Q., J.R. and J.J. designed the compositions and J.J. also aided in characterization. L.K. and A.L. contributed to our understanding of the chemical homogeneity in this system. H.G. and R.D.J. developed the theory of compatibility and identified the significance of the equidistance condition with input from all authors. R.D.J. wrote the manuscript with input from all authors. E.Q. and R.D.J. supervised the collaboration.

Corresponding authors

Correspondence to Hanlin Gu, Jascha Rohmer, Justin Jetter, Andriy Lotnyk, Lorenz Kienle, Eckhard Quandt or Richard D. James.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Jian Luo and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Fig. 1 Microstructure and grain size.

(a) SEM image of the sample y = 0.5 with monoclinic twin laminated microstructure, showing the untreated sample surface directly after sintering, and (b) the fractured surface of the same sample. (c) Typical transformed material from a weeping sample that shows separation at the grain boundaries.

Extended Data Fig. 2 Chemical homogeneity and absence of short-range ordering (crushed).

The micrographs of the microstructural and nanoscale investigations are TEM and HRTEM images as well as SAED and NBED patterns of the (Zr0.9 Hf0.1 O2)0.775 (Y0.5 Nb0.5 O2)0.225 (weeping) sample prepared by crushing. No additional reflections are observed in SAED and NBED patterns beyond those due to dynamical double diffraction. Also no additional reflections are seen in the FFT images.

Extended Data Fig. 3 Chemical homogeneity and absence of short-range ordering (FIB).

The micrographs of the microstructural and nanoscale analysis are HAADF-STEM and HAADF-HRSTEM images as well as SAED and NBED patterns of the (Zr0.9 Hf0.1 O2)0.775 (Y0.5 Nb0.5 O2)0.225 (weeping) sample, prepared by FIB. No additional reflections are observed in SAED, NBED patterns (only from dynamical double diffraction) as well as FFT images. Low-magnification HAADFM and atomic-scale HAADF micrographs are raw images, showing no significant intensity variation.

Extended Data Fig. 4 Chemical homogeneity, crushed sample.

Nanoscale chemical study of a (Zr0.9 Hf0.1 O2)0.775 (Y0.5 Nb0.5 O2)0.225 (weeping) sample, prepared by crushing. The images are HAADF-HRSTEM micrograph and high resolution EDX elemental maps, suggesting a uniform distribution of elements.

Extended Data Fig. 5 Chemical homogeneity, FIB sample.

Nanoscale chemical analysis of a (Zr0.9 Hf0.1 O2)0.775 (Y0.5 Nb0.5 O2)0.225 (weeping) sample prepared by FIB. The images are HAADF-HRSTEM micrograph and high resolution EDX elemental maps, suggesting a uniform distribution of elements.

Extended Data Fig. 6 Chemical homogeneity and structure at grain boundaries (FIB).

Nanoscale study and local chemical analysis of a grain boundary of a (Zr0.9 Hf0.1 O2)0.775 (Y0.5 Nb0.5 O2)0.225 (weeping) sample prepared by FIB. The micrographs are HAADF-STEM and HAADF-HRSTEM images as well as high resolution EDX elemental maps of the sample. Atomic-resolution HAADF micrographs are raw images, showing no significant intensity variation along the grain boundary (GB). The high resolution EDX maps suggest no significant element segregation at the grain boundary.

Extended Data Fig. 7 Structure by X-Ray Diffraction with Rietveld Refinement.

XRD diffraction pattern and calculated fit after Rietveld refinement (Top) with Topas software. The diagram at the bottom (Residual) shows the difference in intensity between the measured and calculated diffraction pattern. The sample in plot a) has a phase transformation above room temperature (RT) and is in the monoclinic phase, whereas the phase transformation of the sample in plot b) is below RT, so the pattern shows the tetragonal crystal structure. The low Rwp values, representing the goodness of the fit, indicate the quality of the Rietveld refinement. Temperature dependent measurements were conducted with a graphite domed heating stage. In c) the sample y = 0.9 is in its monoclinic phase, while in d) the measurement was taken at 415 °C following the monoclinic to tetragonal phase transition. The stage appears in the measured XRD pattern with additional peaks that were identified and excluded from the refinement done on the parameters of the physical phases. The resulting higher Rwp values compared to a) and b) can be explained by the reduced intensities due to limited transmissibility of the graphite dome used in these cases. Each measurement is refined up to 1000 times with different varied starting parameters with only the best fit being used for further calculations.

Extended Data Fig. 8 Temperature dependent XRD measurements.

Temperature dependent XRD measurement of the sample y = 0.8 with domed heating stage showing the phase transformation on heating and cooling. Upon heating, the characteristic tetragonal peak starts to grow at the austenitic start temperature and the monoclinic peaks are vanishing. At temperatures far above Af we force transformation of the residual phase. During cooling to 30 °C, we observe the reverse transformation (t-to-m) of the sample. These measurements are the basis to obtain the lattice parameter of the monoclinic and tetragonal phases by Rietveld refinement, to determine the temperature dependent change of these lattice parameters and to calculate the middle eigenvalues λ2 of the transformation stretch matrix for the lattice Correspondences 1a, 1b and 2.

Extended Data Fig. 9 Frame sequence of explosive behaviour.

In a sequence of frames, the Fig. shows the path of a jumping ceramic confined to a cylinder. This jump is also shown in Supplementary Video 2.

Supplementary information

Supplementary Information

This file contains Supplementary Sections 1–3 including equations for the analysis of compatibility and a heat transfer analysis and Tables 1–4.

Supplementary Video 1

A sample at a high Zr/Hf ratio exhibiting explosive behaviour on cooling to the transformation temperature. See the main text for details.

Supplementary Video 2

The path of a sample at a high Zr content exhibiting jumping, confined to a graduated cylinder. See also the main text and Extended Data Fig. 9.

Supplementary Video 3

A sample at a high Zr/Hf ratio that exhibits weeping (that is, steady falling apart at the grain boundaries). See also the main text and Fig. 2.

Supplementary Video 4

Reversible motion of the interface in the context of a general cooling protocol. The sample closely satisfies the equidistance condition. See the main text and the caption of Fig. 2.

Supplementary Video 5

This video shows reversible motion of an interface close up, so the interface position is clear. The sample closely satisfies the equidistance condition. See also the main text and the caption of Fig. 2.

Supplementary Video 6

A sample that was held at a temperature a little above the transformation temperature for 94 days, and then cooled to the point of transformation. This supports the arguments given in the Methods (“Possible rate effects”) on the absence of a significant rate effect.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gu, H., Rohmer, J., Jetter, J. et al. Exploding and weeping ceramics. Nature 599, 416–420 (2021). https://doi.org/10.1038/s41586-021-03975-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03975-5

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing