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.

  • Letter
  • Published:

Formation of monatomic metallic glasses through ultrafast liquid quenching

Abstract

It has long been conjectured that any metallic liquid can be vitrified into a glassy state provided that the cooling rate is sufficiently high1,2,3,4. Experimentally, however, vitrification of single-element metallic liquids is notoriously difficult5. True laboratory demonstration of the formation of monatomic metallic glass has been lacking. Here we report an experimental approach to the vitrification of monatomic metallic liquids by achieving an unprecedentedly high liquid-quenching rate of 1014 K s−1. Under such a high cooling rate, melts of pure refractory body-centred cubic (bcc) metals, such as liquid tantalum and vanadium, are successfully vitrified to form metallic glasses suitable for property interrogations. Combining in situ transmission electron microscopy observation and atoms-to-continuum modelling, we investigated the formation condition and thermal stability of the monatomic metallic glasses as obtained. The availability of monatomic metallic glasses, being the simplest glass formers, offers unique possibilities for studying the structure and property relationships of glasses. Our technique also shows great control over the reversible vitrification–crystallization processes, suggesting its potential in micro-electromechanical applications. The ultrahigh cooling rate, approaching the highest liquid-quenching rate attainable in the experiment, makes it possible to explore the fast kinetics and structural behaviour of supercooled metallic liquids within the nanosecond to picosecond regimes.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Illustration of an ultrafast liquid-quenching approach.
Figure 2: Structure and thermal stability of Ta MG.
Figure 3: Dynamic vitrification process in liquid Ta revealed by AtC computer simulation.
Figure 4: Reversible crystallization–vitrification phase changes of Ta MG.

Similar content being viewed by others

References

  1. Klement, W., Willens, R. H. & Duwez, P. Non-crystalline structure in solidified gold–silicon alloys. Nature 187, 869–870 (1960)

    Article  CAS  ADS  Google Scholar 

  2. Greer, L. A. Metallic glasses. Science 267, 1947–1953 (1995)

    Article  CAS  PubMed  ADS  Google Scholar 

  3. Cohen, M. H. & Turnbull, D. Molecular transport in liquids and glasses. J. Chem. Phys. 31, 1164–1169 (1959)

    Article  CAS  ADS  Google Scholar 

  4. Turnbull, D. Under what conditions can a glass be formed? Contemp. Phys. 10, 473–488 (1969)

    Article  CAS  ADS  Google Scholar 

  5. Bhat, M. H. et al. Vitrification of a monatomic metallic liquid. Nature 448, 787–790 (2007)

    Article  CAS  ADS  PubMed  Google Scholar 

  6. Johnson, W. L. Fundamental aspects of bulk metallic glass formation in multicomponent alloys. Mater. Sci. Forum 225–227, 35–50 (1996)

    Article  Google Scholar 

  7. Ding, S. Y. et al. Combinatorial development of bulk metallic glasses. Nature Mater. 13, 494–500 (2014)

    Article  CAS  ADS  Google Scholar 

  8. Johnson, W. L. Thermodynamic and kinetic aspects of the crystal to glass transformation in metallic materials. Prog. Mater. Sci. 30, 81–134 (1986)

    Article  CAS  Google Scholar 

  9. Davies, H. A., Aucote, J. & Hull, J. B. Amorphous nickel produced by splat quenching. Nature 246, 13–14 (1973)

    CAS  ADS  Google Scholar 

  10. Lin, C.-J. & Spaepen, F. Fe-B glasses formed by picosecond pulsed laser quenching. Appl. Phys. Lett. 41, 721–723 (1982)

    Article  CAS  ADS  Google Scholar 

  11. McCluskey, P. J. & Vlassak, J. J. Combinatorial nanocalorimetry. J. Mater. Res. 25, 2086–2100 (2010)

    Article  CAS  ADS  Google Scholar 

  12. Aga, R. S., Morris, J. R., Hoyt, J. J. & Mendelev, M. Quantitative parameter-free prediction of simulated crystal-nucleation times. Phys. Rev. Lett. 96, 245701 (2006)

    Article  PubMed  ADS  CAS  Google Scholar 

  13. Trudu, F., Donadio, D. & Parrinello, M. Freezing of a Lennard–Jones fluid: from nucleation to spinodal regime. Phys. Rev. Lett. 97, 105701 (2006)

    Article  PubMed  ADS  CAS  Google Scholar 

  14. Broughton, J. Q., Gilmer, G. H. & Jackson, K. A. Crystallization rates of a Lennard–Jones liquid. Phys. Rev. Lett. 49, 1496–1500 (1982)

    Article  CAS  ADS  Google Scholar 

  15. Fujime, S. Electron diffraction at low temperature. II. Radial distribution analysis of metastable structure of metal films prepared by low temperature condensation. Jpn. J. Appl. Phys. 5, 778–787 (1966)

    Article  CAS  ADS  Google Scholar 

  16. Suslick, K. S., Choe, S.-B., Cichowlas, A. A. & Grinstaff, M. W. Sonochemical synthesis of amorphous iron. Nature 353, 414–416 (1991)

    Article  CAS  ADS  Google Scholar 

  17. Hilsch, R. in Non-Crystalline Solids (ed. Fréchette, V. D. ) 348–373 (Wiley, 1960)

    Google Scholar 

  18. Sheng, H. W. et al. Polyamorphism in a metallic glass. Nature Mater. 6, 192–197 (2007)

    Article  CAS  ADS  Google Scholar 

  19. Ichikawa, T. The assembly of hard spheres as a structure model of amorphous iron. Phys. Status Solidi A 29, 293–302 (1975)

    Article  CAS  ADS  Google Scholar 

  20. Sachdev, S. & Nelson, D. R. Theory of the structure factor of metallic glasses. Phys. Rev. Lett. 53, 1947–1950 (1984)

    Article  CAS  ADS  Google Scholar 

  21. Yamamoto, R., Matsuoka, H. & Doyama, M. Structural relaxation of the dense random packing model for amorphous iron. Phys. Status Solidi A 45, 305–314 (1978)

    Article  CAS  ADS  Google Scholar 

  22. Dzugutov, M. Glass formation in a simple monatomic liquid with icosahedral inherent local order. Phys. Rev. A 46, R2984–R2987 (1992)

    Article  CAS  PubMed  ADS  Google Scholar 

  23. Leung, P. K. & Wright, J. G. Structural investigations of amorphous transition element films. I. Scanning electron diffraction study of cobalt. Phil. Mag. 30, 185–194 (1974)

    Article  CAS  ADS  Google Scholar 

  24. Ichikawa, T. Electron diffraction study of the local atomic arrangement in amorphous iron and nickel films. Phys. Status Solidi A 19, 707–716 (1973)

    Article  CAS  ADS  Google Scholar 

  25. Jones, R. E., Templeton, J. A., Wagner, G. J., Olmsted, D. & Modine, N. A. Electron transport enhanced molecular dynamics for metals and semi-metals. Int. J. Numer. Methods Eng. 83, 940–967 (2010)

    Article  MATH  Google Scholar 

  26. Siegrist, T. et al. Disorder-induced localization in crystalline phase-change materials. Nature Mater. 10, 202–207 (2011)

    Article  CAS  ADS  Google Scholar 

  27. Wuttig, M. & Yamada, N. Phase-change materials for rewriteable data storage. Nature Mater. 6, 824–832 (2007)

    Article  CAS  ADS  Google Scholar 

  28. Stillinger, F. H. & Weber, T. A. Packing structures and transitions in liquids and solids. Science 225, 983–989 (1984)

    Article  CAS  PubMed  ADS  Google Scholar 

  29. Steinhardt, P. J., Nelson, D. R. & Ronchetti, M. Bond-orientational order in liquids and glasses. Phys. Rev. B 28, 784–805 (1983)

    Article  CAS  ADS  Google Scholar 

  30. Bauder, U. & Fromm, E. Absorption of nitrogen and oxygen by vapour-deposited tantalum films. Surf. Sci. 52, 415–425 (1975)

    Article  CAS  ADS  Google Scholar 

  31. Haas, T. W., Jackson, A. G. & Hooker, M. P. Adsorption on niobium (110), tantalum (110), and vanadium (110) surfaces. J. Chem. Phys. 46, 3025–3033 (1967)

    Article  CAS  ADS  Google Scholar 

  32. Pollard, J. H. & Danforth, W. E. LEED observations of epitaxially grown thorium on (100) tantalum. J. Appl. Phys. 39, 4019–4020 (1968)

    Article  CAS  ADS  Google Scholar 

  33. Hopkins, B. J., Leggett, M. & Watts, G. D. A. RHEED study of the adsorption of oxygen, hydrogen, nitrogen and water vapour on the (110) face of tantalum. Surf. Sci. 28, 581–597 (1971)

    Article  CAS  ADS  Google Scholar 

  34. Daw, M. S. & Baskes, M. I. Embedded-atom method: derivation and application to impurities, surfaces, and other defects in metals. Phys. Rev. B 29, 6443–6453 (1984)

    Article  CAS  ADS  Google Scholar 

  35. Sheng, H. W., Kramer, M. J., Cadien, A., Fujita, T. & Chen, M. W. Highly optimized embedded-atom-method potentials for fourteen fcc metals. Phys. Rev. B 83, 134118 (2011)

    Article  ADS  CAS  Google Scholar 

  36. Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999)

    Article  CAS  ADS  Google Scholar 

  37. Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994)

    Article  ADS  Google Scholar 

  38. Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 47, 558–561 (1993)

    Article  CAS  ADS  Google Scholar 

  39. Brommer, P. & Gähler, F. Potfit: effective potentials from ab initio data. Model. Simul. Mater. Sci. Eng. 15, 295–304 (2007)

    Article  CAS  ADS  Google Scholar 

  40. Cheng, Y. Q., Ma, E. & Sheng, H. W. Atomic level structure in multicomponent bulk metallic glass. Phys. Rev. Lett. 102, 245501 (2009)

    Article  CAS  PubMed  ADS  Google Scholar 

  41. Li, D. Y., Wu, Y. Y., Fan, R., Yang, P. D. & Majumdar, A. Thermal conductivity of Si/SiGe superlattice nanowires. Appl. Phys. Lett. 83, 3186–3188 (2003)

    Article  CAS  ADS  Google Scholar 

  42. Schelling, P. K., Phillpot, S. R. & Keblinski, P. Comparison of atomic-level simulation methods for computing thermal conductivity. Phys. Rev. B 65, 144306 (2002)

    Article  ADS  CAS  Google Scholar 

  43. Ho, C. Y., Powell, R. W. & Liley, P. E. Thermal conductivity of the elements. J. Phys. Chem. Ref. Data 1, 279–421 (1972)

    Article  CAS  ADS  Google Scholar 

  44. Tari, A. The Specific Heat of Matter at Low Temperatures 71 (Imperial College Press, 2003)

    Book  Google Scholar 

  45. Allen, P. B. Theory of thermal relaxation of electrons in metals. Phys. Rev. Lett. 59, 1460–1463 (1987)

    Article  CAS  PubMed  ADS  Google Scholar 

  46. Lin, Z. B., Zhigilei, L. V. & Celli, V. Electron–phonon coupling and electron heat capacity of metals under conditions of strong electron–phonon nonequilibrium. Phys. Rev. B 77, 075133 (2008)

    Article  ADS  CAS  Google Scholar 

  47. Tang, C. G. & Harrowell, P. Anomalously slow crystal growth of the glass-forming alloy CuZr. Nature Mater. 12, 507–511 (2013)

    Article  CAS  ADS  Google Scholar 

  48. Kim, Y.-W., Lin, H.-M. & Kelly, T. F. Amorphous solidification of pure metals in submicron spheres. Acta Metall. 37, 247–255 (1989)

    Article  CAS  Google Scholar 

  49. Egerton, R. F. Electron Energy-Loss Spectroscopy in the Electron Microscope 3rd edn (Springer, 2011)

    Book  Google Scholar 

Download references

Acknowledgements

We thank A. J. Coley and Y. Liu for their assistance in designing the electrical circuit for our experiment, and Y. He for performing the liquid-quenching experiment on iridium and rhodium. H.W.S. acknowledges helpful discussions with L. V. Zhigilei. S.X.M. acknowledges National Science Foundation (NSF) grant CMMI 0928517 through the support of the University of Pittsburgh. Work at George Mason University was supported by the US NSF under grant no. DMR-0907325. This work was performed in part at the Center for Integrated Nanotechnologies (CINT), an Office of Science User Facility operated for the US Department of Energy (DOE) Office of Science. The computational work was conducted on the SR10000-K1/52 supercomputing facilities of the Institute for Materials Research, Tohoku University.

Author information

Authors and Affiliations

Authors

Contributions

L.Z. and J.W.W. carried out the TEM experiments under the direction of S.X.M. S.X.M. created the experimental protocols. H.W.S. performed the computer simulation. L.Z., H.W.S. and S.X.M. performed the experimental data analysis. All the authors (L.Z., J.W.W., H.W.S., Z.Z. and S.X.M.) contributed to the discussion of the results. L.Z., S.X.M. and H.W.S. wrote the manuscript.

Corresponding authors

Correspondence to Hongwei Sheng or Scott X. Mao.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Ta nano-tips and Ta MG obtained by liquid quenching.

a, Triangular nano-tips derived by pulling a metal substrate apart. b, High-resolution TEM image of a Ta nano-tip viewed along the 〈001〉 zone axis with no oxidation layer identified at the surface. c, Controllable size by tuning the electric pulse. A Ta MG with diameter close to 100 nm is obtainable under an electric pulse with voltage above 2 V. d, Tunable aspect ratio by coupling with tensile or compressive stress. A nanowire with an aspect ratio of 4 was formed by applying tensile loading during liquid quenching. The glass–crystal interfaces (GCIs) are indicated by yellow dotted curves.

Extended Data Figure 2 Formation and room-temperature stability of V MG.

a, TEM image of two crystalline V nano-tips in contact with each other. b, Formation of a V MG 75 nm long and 80 nm thick under a 1.2-V, 3.6-ns electric pulse. The V MG as formed was sandwiched by two crystalline substrates with two GCIs (denoted by yellow dotted curves). c, d, Thermal stability of V MG at room temperature. Amorphous V was found stable after relaxation for 6 h (c) and even 56 h (d). To protect V from oxidation, an amorphous carbon layer 5 nm thick was coated on the surface of the V MG. The reduced length of the V MG in d is due to a slight change in the viewing angle. e, High-resolution image of the V MG relaxed for 56 h, showing typical amorphous characteristics. f, Electron diffraction pattern of V MG, showing diffusive amorphous halos and a diffusive background. The bright diffraction spots originate from bcc V because the aperture was not small enough to exclude the neighbouring crystalline V substrate.

Extended Data Figure 3 Formation of Mo MG by ultrafast quenching.

a, TEM image of two contacting Mo nano-tips with a well-defined bcc structure viewed along the 〈111〉 zone axis (inset in a). b, Formation of Mo MG under the first vitrification pulse. The GCIs are indicated by yellow dotted curves. c–e, Growth of the Mo MG under subsequent vitrification pulses. The GCIs moved one step away from each other (denoted by two yellow arrows in b) after the second (c), third (d) and fourth (e) pulse, respectively, resulting in the growth of the Mo MG. The amorphous structure is corroborated by the diffuse halos in the fast Fourier transformation (inset in e).

Extended Data Figure 4 Formation and spontaneous crystallization of W MG.

a, High-resolution TEM image of the original crystalline W nano-tip viewed along the 〈100〉 zone axis. b, Formation of W MG in the W nano-tip under a vitrification pulse. An atomically rough and diffuse GCI was identified, where the transition zone from amorphous to bcc W was about 1–2 nm thick (the region between the two dotted yellow curves). c–e, Time-lapse images of spontaneous crystallization of W MG. W MG is unstable at room temperature, undergoing spontaneous crystal growth at the GCI to a well-defined bcc structure.

Extended Data Figure 5 Cooling rate and crystal growth rate estimated by AtC computer simulation.

a, Atomic temperature distribution of a Ta nanowire with dimensions 85 × 40.8 × 13.6 nm3 at time zero when Joule heating is stopped instantly. The 32 × 12 × 6 finite-element meshes were used to simulate the electron temperature field in the TTM. The electron temperature on both sides of the nanowire was kept constant at 300 K. b, Evolution of atomic temperature distribution along the x direction of the Ta nanowire during the cooling process. c, Cooling rate as a function of temperature in liquid Ta (the middle section within 5 nm along the x direction). The highest cooling rate at the initial stage of quenching reaches as high as 1014 K s−1. d, Crystal growth rate at the LCIs of the (100) (cyan circles) and (110) (orange squares) crystallographic planes, on the basis of classic MD simulation. The simulation details for crystal growth from the melt are similar to those in ref. 47.

Extended Data Figure 6 Glass transition temperature Tg of Ta from MD simulation.

Both the enthalpy change and the volume change as a function of temperature indicate that Tg is close to 1,650 K, which is in good agreement with other theoretical work48.

Extended Data Figure 7 Vitrification of liquid Ta under a heating flux terminated within 0.4 ns by AtC computer simulation.

a, Atomic configuration of the Ta nanowire after Joule heating. A melting zone 35 nm long was formed before quenching (atoms coloured with red). b, An 18-nm region in the middle of the melting zone is vitrified to a glassy state after being quenched to room temperature. c, Cooling rate as a function of temperature in the middle region of the melting zone during quenching. The cooling rate varied between 3 × 1013 K s−1 at 4,200 K and 1013 K s−1 at Tg.

Extended Data Figure 8 Electron energy-loss spectroscopy (EELS) spectra of oxygen in Ta.

An oxygen K-edge was identified at about 530 eV energy loss from Ta that had been exposed to air for 10 min (red curve). In contrast, no oxygen K-edge was detected in both Ta MG and crystal (cyan curve) when the Ta nano-tips were processed in a helium-protected environment and deoxidized by Joule heating before the liquid-quenching experiment, indicating that the oxygen concentration in the Ta MG was below the detection limit of EELS, which is 1,000 p.p.m. in atomic ratio49 (that is, 100 p.p.m. in weight percentage).

Extended Data Figure 9 Tensile mechanical test on Ta MG.

(See also Supplementary Video 2.) a, A hybrid Ta nanowire: the left half is Ta MG and the right half is bcc Ta. This hybrid nanowire was fabricated by forming a Ta MG nanowire at first, followed by controlled gradual crystallization to move one of the GCIs (outlined by yellow dotted curves) to the middle of the nanowire. b, Deformation of Ta MG along a major shear band, which began shortly after tensile loading. No plasticity was observed in bcc Ta on the right. The major shear proceeded along 52° off the tensile loading direction, as denoted by two dark red arrows.

Extended Data Table 1 Purities of the bcc metals studied here

Supplementary information

Controlled crystallization under a series of crystallization pulses

Gradual crystallization in a 180 nm long, 85 nm thick Ta MG proceeded by migration of the GCIs (indicated by red arrow heads) a step forward towards each other after each pulse. (MOV 11511 kb)

Tensile mechanical test on Ta MG

In a half-MG, half-crystal hybrid Ta nanowire, deformation was localized in the MG by a major shear along ~52° off the loading direction, which initiated shortly after tensile straining at a rate of 10-3 s-1 while no sign of plasticity was observed in bcc Ta. This movie is played at 8× (see Extended Data Figure 9). (MOV 8489 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhong, L., Wang, J., Sheng, H. et al. Formation of monatomic metallic glasses through ultrafast liquid quenching. Nature 512, 177–180 (2014). https://doi.org/10.1038/nature13617

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature13617

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