You are on page 1of 14

Letter doi:10.

1038/nature19313

Extreme creep resistance in a microstructurally


stable nanocrystalline alloy
K. A. Darling1, M. Rajagopalan2, M. Komarasamy3, M. A. Bhatia2, B. C. Hornbuckle1, R. S. Mishra3 & K. N. Solanki2

Nanocrystalline metals, with a mean grain size of less than 100 Contrary to conventional wisdom, we have developed a divergent,
nanometres, have greater room-temperature strength than their bulk nanocrystalline copper–tantalum alloy (10 atomic per cent (at%)
coarse-grained equivalents, in part owing to a large reduction tantalum; hereafter Cu–10 at% Ta) that is able to achieve and retain
in grain size1. However, this high strength generally comes with high strength and creep resistance at a high homologous tempera-
substantial losses in other mechanical properties, such as creep ture of 0.64Tm ≈ 600 °C (where Tm is the melting temperature of the
resistance, which limits their practical utility; for example, creep matrix), owing to its unique microstructural architecture. Initially syn-
rates in nanocrystalline copper are about four orders of magnitude thesized through high-energy ball milling and subsequently consoli-
higher than those in typical coarse-grained copper 2,3. The dated via equal-channel angular extrusion (ECAE), the as-processed
degradation of creep resistance in nanocrystalline materials is in microstructure consists of a copper matrix with an average grain size
part due to an increase in the volume fraction of grain boundaries, of 50 ± 17.5 nm and tantalum-based particles that range in size from
which lack long-range crystalline order and lead to processes such atomic nanoclusters (average diameter of 3.18 ± 0.86 nm) to much
as diffusional creep, sliding and rotation3. Here we show that larger precipitates (average diameter of 32 ± 7.5 nm); uncertainties
nanocrystalline copper–tantalum alloys possess an unprecedented given here and elsewhere are 1 s.d. (For a macroscopic view of the
combination of properties: high strength combined with extremely microstructure and additional processing details, see Extended Data
high-temperature creep resistance, while maintaining mechanical Fig. 2.) It has been shown that such ranges in particle size give rise to
and thermal stability. Precursory work on this family of immiscible an extremely stable microstructure4,5,10,11; for example, as compared
alloys has previously highlighted their thermo-mechanical stability to pure nanocrystalline copper, which exhibits rapid grain growth to
and strength4,5, which has motivated their study under more extreme the micrometre scale at only 100 °C (ref. 12), Cu–10 at% Ta powders
conditions, such as creep. We find a steady-state creep rate of less maintain a mean grain size of 167 nm after annealing at 1,040 °C for 4 h
than 10−6 per second—six to eight orders of magnitude lower than (ref. 4). This highly stabilized microstructure could give rise to unusual
most nanocrystalline metals—at various temperatures between 0.5 combinations of mechanical properties, such as creep resistance under
and 0.64 times the melting temperature of the matrix (1,356 kelvin) extreme conditions (high stress and temperature).
under an applied stress ranging from 0.85 per cent to 1.2 per cent We conducted compression creep tests over a wide range of applied
of the shear modulus. The unusual combination of properties in stress and temperature conditions (see Methods), as shown in Fig. 1.
our nanocrystalline alloy is achieved via a processing route that The compression–creep-strain evolution curves shown in Fig. 1a con-
creates distinct nanoclusters of atoms that pin grain boundaries sist of the primary creep region, where the creep strain rate decreased
within the alloy. This pinning improves the kinetic stability of the with time, and the secondary creep region, where the creep strain rate
grains by increasing the energy barrier for grain-boundary sliding remained at steady state. The steady-state creep rates in nanocrystal-
and rotation and by inhibiting grain coarsening, under extremely line Cu–10 at% Ta were all found to be less than 10−6 s−1 at various
long-term creep conditions. Our processing approach should homologous temperatures between 0.5Tm and 0.64Tm under a stress of
enable the development of microstructurally stable structural 1.2%–0.85% of the shear modulus. Note that the creep rates reported
alloys with high strength and creep resistance for various high- for nanocrystalline Cu–10 at% Ta are minimum creep rates. The upper
temperature applications, including in the aerospace, naval, civilian and lower fractions of the shear modulus equate to stress values of
infrastructure and energy sectors. 576 MPa and 319 MPa, respectively, which represent 90% and 65%
Over the past 50 years, the reduction or elimination of intrinsic of the at-temperature (400 °C and 600 °C) yield strength. The yield
topological defects (grain or cell boundaries) has been central to stress values at various temperatures were quantified using a series
the design of creep-resistant materials. Current designs enhance of quasi-static compression tests with a strain rate of 8 × 10−4 s−1;
high-temperature creep performance through the use of single-crystal see Extended Data Fig. 4. To further demonstrate the improvement
alloys, for example, nickel-based, single-crystal superalloys 6,7. in the creep resistance, a creep test was performed under 100% of
Nano-grained materials with grain sizes 7–8 orders of magnitude the yield stress at 0.5Tm (approximately 400 °C), which resulted in a
smaller than, and grain-boundary volume fractions 5–6 orders of creep rate of 5.3 × 10−8 s−1. By contrast, at a rather low homologous
magnitude higher than, the currently used single-crystal superalloys temperature—0.4Tm or about 275 °C, for example—a creep rate of
have never been considered viable for high-temperature creep appli- 10−1 s−1 was reported for an applied stress of 0.12% of the shear mod-
cations. Moreover, nanocrystalline metals exhibit microstructural ulus (57 MPa) with an average grain size of 25 nm in pure nanocrystal-
instability, that is, grain growth via diffusional processes such as line copper2. Compared to pure nanocrystalline copper, nanocrystalline
diffusional creep, sliding and rotation, at moderately low and even Cu–10 at% Ta at 1.5–2 times higher temperature and an order of mag-
room temperatures, sometimes in combination with deformation2,8,9. nitude higher stress has a creep rate that is 6–8 orders of magnitude
Consequently, previous creep studies on nanocrystalline metals have lower (ref. 3). Such a response is reminiscent of, and more comparable
reported creep-stress exponents of 1–3 resulting from grain-size to, that of the creep performance achieved by advanced single-crystal
effects on diffusional (Coble) creep2. nickel-based superalloys (creep rate of about 10−8 s−1)13.

1
Army Research Laboratory, Aberdeen Proving Ground, Maryland 21005, USA. 2School of Engineering of Matter, Transport, and Energy, Arizona State University, Tempe, Arizona 85281, USA.
3
Department of Materials Science and Engineering, University of North Texas, Denton, Texas 76203, USA.

3 7 8 | N A T U R E | V O L 5 3 7 | 1 5 s e p te m b er 2 0 1 6
© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

a 400 °C (0.5Tm) 500 °C (0.57Tm) 600 °C (0.64Tm) b


YS = 640 MPa YS = 520 MPa YS = 490 MPa Nanocrystalline copper (d = 25 nm)
0.70YS 0.65YS 0.45YS Nanocrystalline nickel (d = 40 nm)
0.80YS 0.70YS 0.50YS Nanocrystalline Cu–10 at% Ta (d = 50 nm)
0.90YS 0.80YS 0.55YS Theoretical (d = 50 nm)
1.00YS 0.60YS
0.65YS
16 100 Theoretical shear strength
14 Plasticity
Power-law breakdown

Normalized strength, V/G


12 10–2 10–7 s–1 10–7 s–1
10–9 s–1 Climb
10–8s–1
Creep strain (%)

10
10–8 s–1
8 10–4 10–10 s–1 10–8 s–1

6 1.7 × 10–8 s–1 Coble creep

4 10–6 10 s –6 –1

2 10–7 s–1

0 10–8
0 1 2 3 4 5 6 7 0 0.2 0.4 0.6 0.8 1
Time (105 s) Homologous temperature, T/Tm

Figure 1 | Compressive creep response of nanocrystalline Cu–10 at% Ta. creep rates for a grain size d = 50 nm (circles), experimental creep rates2
a, Conventional creep strain versus time for various applied temperatures in nanocrystalline copper (squares; d = 25 nm) and nickel (diamonds;
and constant-stress conditions (given as fractions of the yield stress, d = 40 nm), and the creep rate we found for nanocrystalline Cu–10 at% Ta
YS), as indicated in the legend. For example, the creep rate for the red (triangles; d = 50 nm) over-plotted. The different coloured symbols
curve corresponds to a value of 1.7 × 10−8 s−1. b, Map of the theoretical correspond to different creep rates, as indicated in the legend. T,
deformation mechanisms of nanocrystalline copper with an average temperature; Tm, melting temperature of the matrix; σ, applied stress; G,
grain size of 50 nm (shaded regions), with theoretical, constant, Coble shear modulus.

In general, creep in nanocrystalline materials has been reported to creep resistance achieved with our nanocrystalline Cu–10 at% Ta
follow the Coble creep mechanism14, whereby creep occurs through alloy outperforms most nanocrystalline materials. To comprehend
the transport of vacancies along grain boundaries14,15 with a low stress this improvement in creep resistance, a compilation of experimental
exponent (of the order of 1–3)2,3. On the other hand, our nanocrys- and theoretical creep-rate data for various nanocrystalline materials is
talline Cu–10 at% Ta alloy exhibits stress exponents that are substan- presented on an Ashby-type deformation mechanism map15 in Fig. 1b,
tially higher than those associated with the diffusional-creep- and which was derived on the basis of creep constants for nanocrystalline
grain-boundary-related mechanisms (see Methods). Therefore, the copper with a mean grain size of 50 nm (see ref. 15 and Methods).

a b c

20 nm 20 nm 5 nm

d e f

93°

5 nm 5 nm 2 nm

Figure 2 | TEM characterization of tantalum-based nanocluster in as- a and b. d, BF-STEM image highlighting the core–shell structure of the
received nanocrystalline Cu–10 at% Ta. a, BF-STEM image highlighting nanocluster. e, Inverse fast Fourier transform image of the red-boxed
the high density of nanoclusters of various sizes. The coloured arrows region in d highlighting the threading dislocations (yellow lines) and
indicate the sizes (radii) of the different coherent or semi-coherent half planes (red arrows) between the matrix semi-coherent clusters.
nanoclusters (red, approximately 1 nm; yellow, approximately, 2.5 nm; f, 3-nm-diameter particle residing at a high-angle (93°) grain boundary
green, 4 nm or greater). b, HAADF-STEM image accentuating tantalum- (red dashed lines). The green arrows correspond to the direction of lattice
rich clusters on the basis of atomic-number contrast; arrows as in planes of the copper matrix.
a. c, High-magnification HAADF image of the green-boxed area in

1 5 s e p te m b er 2 0 1 6 | V O L 5 3 7 | N A T U R E | 3 7 9
© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

a b c
0.33
Before creep testing After creep testing
0.29
Tantalum Tantalum
0.25 Copper Copper

Number fraction 0.21

0.17

0.13

0.08

0.04

0 20 nm
1 2 3 4 5 6 20 50 80 110 1 2 3 4 5 6 20 50 80 110
Diameter (nm) Diameter (nm)
Figure 3 | TEM images showing the microstructures and grain particles before and after creep testing are similar, indicating little if any
distributions indicating stability in nanocrystalline Cu–10 at% Ta after microstructural evolution. c, High-resolution BF-STEM image showing
creep testing. a, b, Number distributions determined from 300 grains the bowing of the grain boundary (green dashed lines) as it interacts with
of the copper matrix (red) and tantalum particles (green) before (a) and tantalum clusters. The coloured arrows highlight the sizes (diameters) of
after (b) creep testing. The distributions of copper grains and tantalum the clusters (red, <1 nm; yellow, <1–2 nm; green, >4 nm).

Experimental creep-rate data from nanocrystalline metals such as cop- transmission electron microscopy (HAADF-STEM) image (Fig. 2b)
per (25-nm grain size) and nickel (40-nm grain size)2 along with the points to the shell portion of the particle being tantalum-rich, with
theoretical, constant, Coble creep-rate lines for copper with an average the core generating less contrast, possibly owing to an element with a
grain size of 50 nm (green and blue circles in Fig. 1b) are also presented. lower atomic number or to structural defects that would not generate
The reported creep properties of nanocrystalline copper and nickel contrast (such as vacancies). The high-resolution transmission electron
fall within the Coble region. This is mainly due to grain-boundary microscopy (HR-TEM) image of the nanocluster (Fig. 2d) provides
diffusional processes; that is, vacancy diffusion and self-diffusion in further evidence that the loss in contrast could be partly due to the
copper and nickel, occurring through both the grain boundaries and presence of vacancies within the core region. The inverse fast Fourier
the lattice are faster at elevated temperatures and, hence, the diffusional transform image of one such nanocluster is shown in Fig. 2e. Note the
creep controls the creep behaviour3. Therefore, in these conventional distortion of the lattice as it approaches and enters the nanocluster, with
nanocrystalline copper and nickel metals, the grain coarsening cre- the yellow arrows indicating the insertion of extra half planes of atoms
ates powerful kinetics that constantly evolves the microstructure. By into the lattice to minimize distortion. Finally, Fig. 2f shows a HR-TEM
contrast, the creep rates of our nanocrystalline Cu–10 at% Ta show characterization of semi-coherent bonding between the nanocluster
a marked departure from convention, with the measured creep rates and the copper matrix at a high-angle (93°) grain boundary with an
primarily in the dislocation–climb region (as shown by the triangles average misfit strain of 5.8%, indicating strong interfacial bonding that
in Fig. 1b). In other words, the diffusional creep processes have been can lead to enhanced mechanical properties. Quasi-static and dynamic
suppressed (or were absent) in our nanocrystalline Cu–10 at% Ta alloy. strengths of greater than 1.2 GPa have previously been measured for the
To understand the observed enhancement of the creep property, we nanocrystalline Cu–10 at% Ta alloy (see Methods); these strengths are
turn our attention to the large number (density of 6.5 × 1023 m−3) of greater than double that predicted by Hall–Petch hardening for nano-
coherent or semi-coherent (diameters of 1–4 nm) nanoclusters (see crystalline copper4 and presented with an apparent linear temperature
Fig. 2a, b). These small nanoclusters have a core–shell-type structure dependence of flow stress5. Core–shell-type nanoclusters have recently
that can be seen in Fig. 2b, c, which shows that the contrast within been reported in oxide-dispersion strengthened (ODS) ferritic alloys16
and across the individual nanoclusters varies, indicating a composi- and molybdenum alloys17, and are responsible for the excellent strength
tional gradient. In addition, the high-angle annular dark-field scanning and ductility therein.

a b

Initial
Initial

Final Final

10 nm 10 nm

Figure 4 | Modelling data indicating stability in nanocrystalline the extent of coarsening associated with plastic deformation under
Cu–10 at% Ta after creep testing. a, b, Two-dimensional slices through constant load and temperature conditions. In b, tantalum atoms (blue)
three-dimensional atomistic creep simulations of pure nanocrystalline formulate a random distribution of grain-boundary clusters and localized
copper (a) and nanocrystalline Cu–10 at% Ta (b) at 600 °C and 295 MPa growth is observed (circled in black) owing to insufficient Zener pinning
of applied stress. White atoms represent the initial grain-boundary in some grains.
configurations (average grain size of 8 nm); red and green atoms represent

3 8 0 | N A T U R E | V O L 5 3 7 | 1 5 s e p te m b er 2 0 1 6
© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

To understand the underlying mechanisms of creep resistance and 4. Darling, K. A. et al. Microstructure and mechanical properties of bulk
nanostructured Cu–Ta alloys consolidated by equal channel angular extrusion.
to determine the enhancement of the creep property induced by the Acta Mater. 76, 168–185 (2014).
nanoclusters, atomistic simulations and post-HR-TEM characteri- 5. Darling, K. A., Huskins, E. L., Schuster, B. E., Wei, Q. & Kecskes, L. J. Mechanical
zations were performed (see Figs 3 and 4). First, the HR-TEM char- properties of a high strength Cu–Ta composite at elevated temperature.
acterizations of post-deformed creep samples (at 600 °C and 50% Mater. Sci. Eng. A 638, 322–328 (2015).
6. Giamei, A. F. Development of single crystal superalloys: a brief history.
yield stress) were performed, as shown in Fig. 3. The stability of the Adv. Mater. Process. 171, 26–30 (2013).
nanoclusters, which is crucial for enhanced properties and can be 7. Reed, R. C. The Superalloys: Fundamentals and Applications (Cambridge Univ.
seen in Fig. 3a, b, is indicated by the coarsening rate of nanoclusters Press, 2008).
8. Meyers, M. A., Mishra, A. & Benson, D. J. Mechanical properties of
during creep at elevated temperatures being negligible, mainly owing nanocrystalline materials. Prog. Mater. Sci. 51, 427–556 (2006).
to the coherency of such dispersions. Further, owing to highly stabi- 9. Choi, I.-C. et al. Nanoscale room temperature creep of nanocrystalline nickel
lized nanoclusters, bowing of the grain boundary as it interacts with pillars at low stresses. Int. J. Plast. 41, 53–64 (2013).
10. Hornbuckle, B. C. et al. Effect of Ta solute concentration on the microstructural
numerous nanoclusters can clearly be identified in a high-resolution evolution in immiscible Cu-Ta Alloys. JOM 67, 2802–2809 (2015).
bright-field STEM (BF-STEM) image of the post-creep sample, as 11. Bhatia, M., Rajagopalan, M., Darling, K. A., Tschopp, M. A. & Solanki, K. N. The
shown in Fig. 3c. The bowing indicates that clusters located at grain role of Ta on twinnability in nanocrystalline Cu-Ta alloys. Mater. Res. Lett.
http://dx.doi.org/10.1080/21663831.2016.1201160 (2016).
boundaries are likely to increase the energy barrier for grain-boundary 12. Huang, Y., Menovsky, A. & De Boer, F. Calorimetric analysis of the grain
sliding and rotation, both of which are crucial creep mechanisms in growth in nanocrystalline copper samples. Nanostruct. Mater. 2, 587–595
nanocrystalline metals. In addition, such nanoclusters pin the grain (1993).
13. Pollock, T. M. & Tin, S. Nickel-based superalloys for advanced turbine engines:
boundaries (Zener pinning18), thereby preventing substantial grain chemistry, microstructure and properties. J. Propuls. Power 22, 361–374
coarsening, consistent with the atomistic calculations (Fig. 4a, b). (2006).
Here, atomistic simulations were performed using the molecular- 14. Coble, R. A model for boundary diffusion controlled creep in polycrystalline
dynamics code LAMMPS19 and an embedded atom potential20 (see materials. J. Appl. Phys. 34, 1679–1682 (1963).
15. Ashby, M. F. A first report on deformation-mechanism maps. Acta Metall. 20,
Methods). Therefore, highly stabilized nanoclusters with strong struc- 887–897 (1972).
tural affinity within the matrix and along the grain boundary are the 16. Hirata, A. et al. Atomic structure of nanoclusters in oxide-dispersion-
microstructural features that governing the unusual combination of strengthened steels. Nat. Mater. 10, 922–926 (2011).
17. Liu, G. et al. Nanostructured high-strength molybdenum alloys with
materials properties—high strength, extreme thermal stability and unprecedented tensile ductility. Nat. Mater. 12, 344–350 (2013).
creep resistance. 18. Manohar, P., Ferry, M. & Chandra, T. Five decades of the Zener equation.
Our results will lead to advances in designing nanocrystalline alloys ISIJ Int. 38, 913–924 (1998).
19. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics.
with many simultaneously enhanced high-temperature properties, J. Comput. Phys. 117, 1–19 (1995).
similar to those exhibited by creep-resistant single crystals, but with 20. Purja Pun, G. P., Darling, K. A., Kecskes, L. J. & Mishin, Y. Angular-dependent
the additional benefit of much higher strength. For example, we show interatomic potential for the Cu–Ta system and its application to structural
stability of nano-crystalline alloys. Acta Mater. 100, 377–391 (2015).
that a steady-state creep rate of less than 10−6 s−1 is attained at even
0.64Tm under a high applied stress (1.2% of the shear modulus). The
Supplementary Information is available in the online version of the paper.
creep rates in nanocrystalline Cu–10 at% Ta reported here are 6–8
orders of magnitude lower than most of the previously reported creep Acknowledgements M.R., and K.N.S. acknowledge the use of facilities within
rates in nanocrystalline metals. We expect that the divergent creep the LeRoy Eyring Center for Solid State Science at Arizona State University.
This work was supported by US Army Research Laboratory under contract
behaviour reported here will change the theoretical understand- W911NF-15-2-0038. K.A.D. acknowledges A. J. Roberts and T. Luckenbaugh for
ing and expectations of the ways in which nanocrystalline metals synthesis of the Cu–Ta powder.
deform at high temperatures and will result in new applications and
Author Contributions K.A.D., K.N.S. and R.S.M. equally contributed to the idea.
capabilities. K.A.D. and B.C.H. processed the nanocrystalline materials. M.K. performed the
creep experiments. M.R. characterized the microstructure data and analysed
Online Content Methods, along with any additional Extended Data display items and
deformation-mechanisms maps. M.A.B. performed the modelling work. K.N.S.,
Source Data, are available in the online version of the paper; references unique to
K.A.D., M.R. and R.S.M. analysed the data. K.N.S., K.A.D., M.R. and R.S.M. wrote
these sections appear only in the online paper.
the paper. K.A.D. and M.R. edited the figures. K.A.D. supervised B.C.H., R.S.M.
supervised M.K., and K.N.S. supervised M.R. and M.A.B.
received 4 May; accepted 8 July 2016.
Author Information Reprints and permissions information is available at
1. Gleiter, H. Nanostructured materials: basic concepts and microstructure. www.nature.com/reprints. The authors declare no competing financial
Acta Mater. 48, 1–29 (2000). interests. Readers are welcome to comment on the online version of the
2. Mohamed, F. A. & Li, Y. Creep and superplasticity in nanocrystalline materials: paper. Correspondence and requests for materials should be addressed to
current understanding and future prospects. Mater. Sci. Eng. A 298, 1–15 K.N.S. (kiran.solanki@asu.edu).
(2001).
3. Chokshi, A. H. Unusual stress and grain size dependence for creep in Reviewer Information Nature thanks J. Cormier, S. Forest and T. Perez Prado for
nanocrystalline materials. Scr. Mater. 61, 96–99 (2009). their contribution to the peer review of this work.

1 5 s e p te m b er 2 0 1 6 | V O L 5 3 7 | N A T U R E | 3 8 1
© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

Methods to quantify the grain sizes and microstructure, transmission electron microscopy
Powder processing and consolidation via ECAE. For the preparation of nano- (TEM) was used. TEM characterizations were carried out in the as-received and
crystalline (NC) Cu–10 at % Ta powder, the powder was generated through high- post-deformed conditions using aberration-corrected ARM-200F and JEOL-2010F
energy cryogenic mechanical alloying. The desired composition was obtained by at 200 kV. Several images were captured in bright-field and high-resolution TEM
loading elemental Cu and Ta powders (−325 mesh and 99.9% purity) into a hard- as well as STEM mode to analyse the microstructure and quantify the statistics
ened steel vial along with the milling media (440C stainless steel balls) inside a such as the grain-size distribution. The TEM samples were prepared through con-
glove box with an Ar atmosphere (oxygen and H2O are <1 p.p.m.). The vials were ventional thinning procedures whereby a 3-mm disk from the bulk specimen was
loaded with 10 g of the Cu–Ta powder as well as the appropriate amount of media thinned to about 70 μm and then dimpled to a thickness of about 5 μm. Ion milling
to ensure a ball-to-powder ratio of 5:1 by weight. A SPEX 8000 M shaker mill was performed under liquid nitrogen temperatures to obtain electron-transparent
was used to perform the milling at cryogenic temperature (verified to be about regions in the specimens. The samples were also plasma cleaned in Ar before TEM
−196 °C) for 4 h (14.4 ks) using liquid nitrogen. To ensure the vial remained at observations to minimize contamination.
cryogenic temperature, a thick polymer sleeve was retrofitted to fit around the As-received microstructure. The primary microstructural characterization
vial in the SPEX mill with an inlet and outlet vent to flow the liquid nitrogen. (Extended Data Fig. 2) of the as-received NC Cu–10 at% Ta ECAE processed
Before starting the milling process, the vial was placed in the polymer sleeve with at 700 °C revealed the presence of binary phases of Cu and Ta consistent with
the liquid nitrogen flowing for approximately 20 min (1.2 ks) to ensure the vial the X-ray diffraction measurements. The TEM characterization and precession
approached −196 °C. Once the milling was completed, the vials were placed back diffraction data are illustrated in Extended Data Fig. 2, and demonstrate a high
into the glove box, opened and stored. This milling procedure was performed degree of randomness in the orientation relationship between the grains of NC
until 100 g of NC Cu–10 at% Ta powder was generated. The resulting powder Cu matrix with an average equiaxed grain diameter of 50 ± 17.5 nm. Orientation
after cryogenic mechanical milling was an unagglomerated mass of powder with details were extracted from a region in the sample using the TOPSPIN software
particulates ranging in size from about 20 μm to 100 μm. (resolution of 2 nm) on the TEM where a precession diffraction technique was
For consolidating the NC Cu–10 at% Ta powder to bulk, equal-channel angu- used. In this technique, the incident electron beam is tilted and precessed along a
lar extrusion (ECAE) was selected as the consolidation process. Billets of Ni 201 conical surface, having a common axis with the TEM optical axis24. Even though
with dimensions of 25.4 mm × 25.4 mm × 90 mm had cylindrical chambers with our NC material was consolidated to bulk, through severe plastic deformation,
diameters of 10 mm and lengths of 50 mm made within them for housing the pow- at 700 °C with a total accumulated strain of 4.6 (460%), the as-received averaged
der. The powder was loaded into the chamber followed by press-fitting a Ni 201 grain sizes were still in a NC regime (Extended Data Fig. 2). Further, Extended Data
plug into the open end to seal the chamber. Both of these steps were performed Fig. 2 shows a histogram, taken from multiple images similar to Extended Data
within the glove box. Before starting the ECAE process, the die assembly used for Fig. 2a, b, indicating that the Ta particle size distribution has an average diameter
processing the billets was preheated to 350 °C to minimize thermal loss during of 32 ± 7.5 nm. A lower-magnification bright-field TEM image (Extended Data
the ECAE processing. Additionally, the billets containing the powder were held Fig. 2c) provides more insight into the much smaller Ta-based particles (diameters
at 700 °C in a box furnace purged with Ar for 40 min (2.4 ks) to ensure that they of <32 nm) and the presence of nano-twins within the NC Cu grains. Twinning is
reach the desired extrusion temperature. The heated billets were dropped into the another important deformation mechanism in NC Cu that can be suppressed by
ECAE tooling as quickly as possible from the furnace and extruded at an extrusion the presence of fine nanoclusters11. Further, the processing route produces a wide
rate of 25.5 mm s−1. This step was repeated four times following route ‘BC’21–23 to range of Ta particle sizes, ranging from atomic nanoclusters (average diameter of
prevent imparting a texture to the consolidated powder. As a result of the extru- 3.18 ± 0.86 nm) to much larger precipitates (see Fig. 3 and Extended Data Fig. 2d).
sion channel having an angle of 90°, a total strain of 460% was imparted onto The energy of the interface25 between the nanoclusters and the Cu matrix can be
the powder-containing billet as a result of processing. The creep specimens were used to quantify the type of coherency and the cluster diameters over which the
then machined from these billets, within the region containing the consolidated degree of coherency persists. Characterizing the coherency has indicated that this
powder, via a wire electric discharge machine. Finally, SEM imaging confirmed the material has coherent, semi-coherent and incoherent nanoclusters (diameters of
creep specimens to be fully consolidated after the ECAE process with no porosity <3.898 nm, 3.898–15.592 nm and >15.592 nm, respectively). The nanoclusters
or as-milled particle boundaries being present. Note that a change in processing also have misfit lattice dislocations at the interface that are indicative of the misfit
conditions or steps, such as in ECAE process temperatures, will result in different strain present, which was identified using inverse fast Fourier transform analysis
microstructural statistics such as grain-size distributions; however, as shown pre- (Extended Data Fig. 3). On average, the misfit strain in the as-received sample is
viously, the nanocluster density depends mainly on the Ta concentrations, which about 5.8%, but can be as high as 11%.
are the primary features resulting in an enhanced creep behaviour10. Mechanical characterization at quasi-static conditions. Quasi-static compres-
Impurity levels. Impurities are a concern for all material processing techniques, sion and tension tests of specimens over a temperature range from ambient up
including mechanical alloying via ball milling. During ball milling, the powder to 1,000 °C were performed using an Instron load frame equipped with a 10 kN
can pick-up impurities as a result of being exposed to the atmosphere and from and a 50 kN load cell, respectively, and an Applied Test Systems (ATS) clam-shell
the milling media themselves. To minimize O contamination, all powders were heating furnace capable of reaching maximum temperature of 1,500 °C. The speci-
stored (before and after processing) and loaded into vials and billets under an Ar mens for compression were cylinders of 3 mm in diameter and length (aspect ratio
atmosphere (O and H2O are <1 p.p.m.) inside a glove box. Additionally, to reduce of 1.0), whereas rectangular dogbones with lengths, widths and thicknesses of
the level of Fe contamination, the milling vial and bearings were coated in Cu by 3 mm, 1 mm and 1 mm, respectively, were used for tension. Tests were conducted
pre-milling the vial and the required bearings with high-purity Cu powder for 2 h at 24 °C, 200 °C, 400 °C, 600 °C, 800 °C, 900 °C and 1,000 °C with a strain rate of
at cryogenic temperatures. This process was repeated several times and produced 8 × 10−4 s−1 for compression5 and 1 × 10−3 s−1 for tension. The system was held
a smooth, even coating of Cu over all milling surfaces (that is, the interior vial and at the testing temperature for 15 min before loading to provide uniform temper-
exterior bearing surfaces). Despite these steps, energy-dispersive X-ray spectros- ature within the specimen. The push rods of the load frame were constructed out
copy (EDS) analysis detected approximately 0.75 at% O in the bulk of the alloy10. of precision-machined ZrO2 rods to minimize heat losses. Boron nitride lubri-
To verify this O level, atom probe tomography (APT) was performed on as-milled cated, polished tungsten carbide (WC) disks were used as platens for compression
powder and as-milled powder that was annealed for 1 h at 450 °C under a reducing testing. Specimens were loaded under displacement control with an across head
atmosphere and NC Cu–10 at% Ta ECAE processed at 700 °C. APT results found all displacement of 0.15 mm min−1. The force–displacement data were compliance
conditions to contain less than 1.25 at% O (ref. 10). Consequently, the O contam- corrected for all tests. The stress–strain responses are provided in Extended Data
ination in the alloy was minimized by following the procedural steps highlighted Fig. 4. The compressive curves in Extended Data Fig. 4a exhibit behaviour from
earlier. Finally, Fe contamination from the milling media was also detected via elastic to nearly perfectly plastic over the entire temperature with no substantial
EDS, but could not be accurately measured; therefore, atom probe tomography strain hardening. Furthermore, the flow stress presented with an apparent linear
was used again. From the APT analysis, the Fe contamination was found to vary temperature dependence4,5 as compared to the expected sigmoidal manifestation
between atom probe tips; however, the highest Fe content found was 1 at% and expected for pure coarse-grained Cu. Moreover, the Cu grain size after testing at
the lowest was 0.05 at% (ref. 10). This range indicates that the contamination from 800 °C was estimated to be about 90 nm, indicating that grain coarsening is very
Fe is also minimal. limited and the reduction in observed yield and flow stress is a result of increased
Microstructural characterization. X-ray diffraction was performed on as-received thermal softening only5. Therefore, NC Cu–10 at% Ta exhibits an extremely sta-
samples using an X’Pert PRO PANalytial MPD X-ray diffractometer with a Cu Kα ble microstructure and unusual mechanical properties. In general, face-centred
(λ = 0.1542 nm) radiation source. Owing to the resolution limit, the grain-size cubic materials such as Cu should not show any tension–compression asymmetry,
estimates from Scherrer’s equation for the Cu matrix and Ta phase were inaccu- which is evident from Extended Data Fig. 4b. The response in tension is perfectly
rate. Extended Data Fig. 1 indicates the X-ray reflections from Cu and Ta for the elastic–plastic in nature with negligible strain hardening, identical to the compres-
as-received condition, for which a random texture can be identified. Therefore, sion tests. This response has implications for the tensile creep behaviour, where

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

this material will be expected to behave in a similar way for tensile-type creep tests experimental data covers a range of strain rates. The stress exponent value of five
as for compression. for threshold correction was deemed appropriate for this study, and corresponds
Mechanical characterization at creep conditions. Compressive cylindrical creep to a dislocation-climb-based deformation mechanism whereby the threshold stress
experiments were performed using a 2320 series lever arm creep tester (Applied arises owing to the influence of dislocations in the creep process. Using the approx-
test systems) with a 5:1 lever arm ratio. Both the diameter and the height of the imation, a threshold stress of 165 MPa was deduced and subtracted from the
cylindrical creep specimens were about 3 mm. The specimens were kept at the applied stress to illustrate the relation between normalized stress and creep rate
centre of a 3210 series split tube furnace to maintain a constant temperature across (Extended Data Fig. 5c). The stress exponent values were computed to be between
the sample height. A heating rate of 200 °C h−1 and a soak time of 0.5 h were four and eight at various temperatures after threshold correction, indicating the
used for the creep tests. For the best temperature measurement and control, a absence of diffusional creep processes during creep. The data obtained through
thermocouple was always wrapped around the creep specimens to maintain good threshold correction for NC Cu–10 at% Ta and literature data2 for NC Cu and NC
contact. An ST 1278 incremental length gauge with ± 1 μm accuracy was used Ni are plotted in Extended Data Fig. 5d. It is evident that NC Cu–10 at% Ta exhib-
to measure the conventional creep strain. The compression creep experiments its extreme creep resistance, with an increase in stress by an order of magnitude
were conducted in air at 873 K and with fractions of 0.45, 0.50, 0.55, 0.60 and 0.65 resulting in a 6–8-fold decrease in ε .
of the yield stress, at 773 K and with fractions of 0.70, 0.75 and 0.80 of the yield Atomistic modelling. The qualitative atomistic simulations were performed using
stress, and at 673 K and with fractions of 0.70, 0.80, 0.90 and 1.00 of the yield stress. a large-scale atomic/molecular massively parallel simulator (LAMMPS)19 along
The specimens were first coated with a thin layer of boron nitride for lubrication with a semi-empirical embedded atom potential (EAM) reported in ref. 20. This
and then placed between the compression platens. Creep test temperatures were EAM potential was parameterized using an extensive database of energies and
attained at a constant heating rate followed by soaking at the set temperature (for configurations from density functional theory (DFT) calculations of energy dif-
0.5 h) to avoid the temperature fluctuation during the test. After the soaking stage, ferences between various crystal structures of pure Cu and pure Ta, the formation
the loading begins automatically, followed by the start of the creep test. These tests energies of coherent Cu–Ta interfaces, and the binding energy of several ordered
were typical constant-force tests. All the creep data were recorded from the test compounds, such as L12–Cu3Ta, L10–CuTa, L11–CuTa, B2–CuTa and L12–Ta3Cu
start to finish. Further, specimens did not reach failure because tests were stopped (ref. 20). See ref. 20 for more details on the validation of the EAM potential at
before the strain rate exponentially increases with stress (before the tertiary creep different temperatures. The Voronoi tessellation method29 was used to construct
domain), and our primary objective was to characterize the secondary creep rates. 3D NC Cu with an average grain size of 8 nm. Further, in the same NC Cu sample,
For most of the creep tests, the total strain values did not exceed about 6%. All the spherical Ta particles with random distribution and size (average sphere radius
crept samples were quenched in water immediately after unloading to preserve the of 0.7 nm) were doped to obtain the 10% Ta concentrations. The total number
crept microstructure. The physical dimensions of the crept samples were measured of atoms in a simulation cell was 1.4 million (with approximate box sizes of
after the test and compared with the extensometer measurements. The stress was 35 nm × 38 nm × 15 nm). The samples were first relaxed at the desired temper-
determined after the test by taking into account the amount of strain. Further, ature using an NVT (conserving the number of atoms, volume and temperature)
during loading the initial strain of the creep test specimen can include both elastic ensemble for 5 ns, followed by an independent relaxation in three directions using
and plastic strains. The minimum creep rate was calculated from the slope of the an NPT (conserving the number of atoms, pressure and temperature to mimic
curve of conventional creep strain versus time. bulk behaviour) ensemble for another 5 ns with zero pressure in all the directions.
Theoretical deformation map. Theoretical deformation maps identify the defor- These relaxations were performed to uniformly distribute the excess free energy
mation modes by which a polycrystalline material can deform15. In the case of through the whole system. Atomistic simulations were carried out using a molec-
creep deformation maps, the dominant mechanism is defined by considering the ular dynamics time step of 1 fs. Periodic boundary conditions were adopted in all
stress and temperature values for a particular value of the steady-state creep rate. directions. Then, the samples were loaded under tension along the y axis with a
The upper bound, that is, the theoretical shear strength, depicts the limit beyond strain rate of 108 s−1 and at 600 °C while maintaining periodic and pressure-free
which flow is possible even in defect-free crystals. This value of stress is of the order boundary conditions along the x and z directions, respectively. The tensile simu-
of the shear modulus and is independent of temperature. In the case of materials lations (up to 3% strain) were performed to increase the defect (dislocation twin,
with defects, the motion of dislocations contributes toward plastic deformation; stacking fault) densities before the creep simulation to mimic the as-received exper-
that is, the dislocation mechanisms are glide, climb and temperature-dependent imental sample microstructure, as seen in Extended Data Fig. 6.
dislocation creep. In the case of dislocation glide creep, impurities, solutes, precip- Finally, the NC and NC Cu–10 at% Ta simulation models were crept at 600 °C
itates and so on that are present in the material provide obstacles to plastic flow. and 295 MPa applied stress along the y direction, whereas deformation in the
At high temperatures, the dislocation creep mechanism is predominant where the other two directions was carried out by maintaining a zero pressure. The desired
deformation is controlled by diffusion and the strain rate is a nonlinear function stresses were applied in the incremental form (a 5-MPa step) until 295 MPa, and
of stress. Further, the motion of point defects leads to plastic deformation through then simulations were run at a constant applied stress for 5 ns or until failure. To
either the grains (Nabarro–Herring26,27) or the grain boundaries (Coble)14. These overcome the relatively short time interval of the molecular dynamics simulation,
diffusional processes are independent of each other and depend on only the tem- we performed simulations at elevated temperatures, at which the distinct effects
perature15. The mechanisms that relate the steady-state creep rate ε to the applied of the resultant liquid-like, fast grain-boundary diffusion (such as grain growth
stress can be depicted using the equation15 and microstructural instability), if present, were clearly identifiable (see refs 30,
31). Extended Data Fig. 7 shows minimum topological changes with the addition
p
 Q  Gb  b   σ n of Ta. Grain growth typically observed in NC materials due to both the stress-
ε = AD0exp −    
 RT  kT  d   G  induced grain-boundary diffusive fluxes and grain-boundary sliding is hindered
in Cu–Ta alloys3. The creep simulation of NC Cu and NC Cu–10 at% Ta in shown
where A is a dimensionless constant, D0 is a frequency factor, Q is the activation in Supplementary Video 1.
energy, R is the gas constant, T is the temperature, G is the shear modulus at the Microstructural stability and creep mechanisms. Reports of room-temperature
particular temperature, b is the burgers vector, k is the Boltzmann constant, d is grain growth—a common feature unique to highly pure NC metals—has been
the grain size, p is the grain-size exponent, n is the stress exponent and σ is the reported numerous times and is in stark contrast to the growth that takes place at
applied stress. The value of the constant A and exponents p and n depend on the much higher temperatures in coarse-grained metals (such as in the experiments on
mechanism considered. The values for the constants can be found in ref. 15. After NC Cu reported in ref. 32). Considerable research has been undertaken to address
incorporating the threshold stress, the rate-controlling creep deformation mech- this specific limitation, culminating in two main methods, one based on thermo-
anism in the high stress and temperature regime was identified from the deforma- dynamics and the other on kinetics33,34. The thermodynamic approach deals with
tion map (Extended Data Fig. 5d) to be dislocation climb, where the apparent stress reducing the excess free energy of the grain boundaries through solute segregation,
exponents of 10–18 were reduced to 4–8 (true stress exponents). whereas the kinetic approach deals with reducing grain-boundary mobility. The
Creep rate data as a function of applied load with and without threshold thermodynamic approach is considered to be more promising because it attenuates
correction. NC Cu–10 at% Ta processed at 700 °C and subjected to creep exhibits the driving force for grain growth, whereas kinetic approaches based on solute
a high stress exponent n, as evident from the plot in Extended Data Fig. 5. To drag, chemical ordering and Zener pinning continually fight against the system
rationalize the high n values, appropriate threshold stress values were determined reaching equilibrium with an Arrhenius temperature dependence.
using a standard linear extrapolation method28 whereby creep rate curves were In light of these two competing mechanisms there has been much discussion
plotted as a function of applied instantaneous stress (ε1 / n versus σ) at various on which method may provide a more successful path in bringing about the reali-
temperatures (Extended Data Fig. 5b). The data points fit onto a straight line that, zation of commercially available bulk NC metals10. In many cases, such debate has
on extrapolation to zero strain rates, yields a threshold value. The inadequacy with been fostered by the fact that it is not always possible to fully separate or delineate
this method is that multiple straight lines exist for different n values provided the the contributions of these two competing stabilization mechanisms in preventing

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

grain growth in NC metals. For instance, thermodynamic stabilization of NC grain 21. Segal, V. M. Materials processing by simple shear. Mater. Sci. Eng. A 197,
size involves examining the energetic penalty associated with the high volume 157–164 (1995).
fraction of the grain boundaries, and the possibility of solute segregation driving 22. Furukawa, M., Horita, Z., Nemoto, M. & Langdon, T. G. Review: processing of
metals by equal-channel angular pressing. J. Mater. Sci. 36, 2835–2843
this associated excess free energy to zero35. However, intertwined in this scenario
(2001).
are the kinetic aspects of solute drag, and its role in reducing grain growth in this 23. Zhu, Y. T. & Lowe, T. C. Observations and issues on mechanisms of grain
thermodynamic stabilization construct, which has been an area of active research34. refinement during ECAP process. Mater. Sci. Eng. A 291, 46–53 (2000).
Additionally, the precipitation of secondary, solute-rich phases have been experi- 24. Vincent, R. & Midgley, P. A. Double conical beam-rocking system for
mentally observed to disrupt the stabilization set in place by the thermodynamic measurement of integrated electron diffraction intensities. Ultramicroscopy 53,
mechanism10. However, recent research has shown that the occurrence of phase 271–282 (1994).
25. Eshelby, J. D. The determination of the elastic field of an ellipsoidal inclusion,
separation or precipitation does not necessarily mean that a stabilized NC system
and related problems. Proc. R. Soc. Lond. A 241, 376–396 (1957).
does not exist10. Recent theoretical work36 predicts the existence of stable duplex 26. Nabarro, F. R. N. Dislocations in a simple cubic lattice. Proc. Phys. Soc. 59,
systems, wherein both grain-boundary segregation and phase separation occurs, 256–272 (1947).
resulting in a stable NC grain size (that is, grain-boundary energy of zero) and a 27. Herring, C. Diffusional viscosity of a polycrystalline solid. J. Appl. Phys. 21,
precipitate structure coexisting with one another. These types of microstructures 437–445 (1950).
are currently under investigation. In reference to these particular immiscible NC 28. Lagneborg, R. & Bergman, B. The stress/creep rate behaviour of precipitation-
Cu–10 at% Ta alloys, the nature of their thermal decomposition and formation of hardened alloys. Met. Sci. 10, 20–28 (1976).
29. Du, Q., Faber, V. & Gunzburger, M. Centroidal Voronoi tessellations: applications
an extremely high density of clusters, occurring primarily along grain boundaries, and algorithms. SIAM Rev. 41, 637–676 (1999).
gives rise to an unusually stable microstructure. Additionally, we have reason to 30. Bhatia, M. A., Mathaudhu, S. N. & Solanki, K. N. Atomic-scale investigation of
believe the exact mechanisms of Zener pinning in this system may be more com- creep behavior in nanocrystalline Mg and Mg–Y alloys. Acta Mater. 99,
plicated than conventional theory18. Nevertheless, the NC Cu–10 at% Ta alloys pri- 382–391 (2015).
marily stabilized kinetically by small-scale coherent clusters shown here provide a 31. Yamakov, V., Wolf, D., Phillpot, S. & Gleiter, H. Grain-boundary diffusion creep in
design route to the development of advanced structural materials for various appli- nanocrystalline palladium by molecular-dynamics simulation. Acta Mater. 50,
61–73 (2002).
cations including high-strength, high-temperature applications. Many of the pro- 32. Chokshi, A. H., Rosen, A., Karch, J. & Gleiter, H. On the validity of the Hall-Petch
cessing and consolidation challenges that have haunted NC metals are now more relationship in nanocrystalline materials. Scr. Metall. 23, 1679–1683 (1989).
fully understood, opening the door for bulk NC metals and parts to be produced. 33. Schiøtz, J., Di Tolla, F. D. & Jacobsen, K. W. Softening of nanocrystalline metals
This has been made possible by the advancement of thermodynamic, kinetic and at very small grain sizes. Nature 391, 561–563 (1998).
thermo-kinetic methods of stabilizing their microstructures. The Cu–Ta family 34. Koch, C. C. Structural nanocrystalline materials: an overview. J. Mater. Sci. 42,
of alloys are currently one of the few systems that have been shown to retain NC 1403–1414 (2007).
35. Chookajorn, T., Murdoch, H. A. & Schuh, C. A. Design of stable nanocrystalline
grain sizes in a fully dense part, allowing the study of these microstructures under
alloys. Science 337, 951–954 (2012).
extreme environments. To the best of our knowledge, our work represents the first 36. Murdoch, H. A. & Schuh, C. A. Estimation of grain boundary segregation
time a stable NC metal has been studied under conditions of high-temperature, enthalpy and its role in stable nanocrystalline alloy design. J. Mater. Res. 28,
high-stress creep. 2154–2163 (2013).

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

Extended Data Figure 1 | X-ray diffraction analysis. X-ray diffraction


plot showing Cu and Ta reflections from the as-received NC Cu–10 at% Ta
sample processed at 700 °C.

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

Extended Data Figure 2 | As-received microstructure characterization. distinct phases: Cu (red) and Ta (green). c, Bright-field TEM micrograph
TEM characterization of as-received NC Cu–10 at% Ta. a, Precession showing the microstructure. d, Size distributions of Cu and Ta grains
diffraction TEM micrograph revealing orientation detail of nanometre-sized determined from 300 grains.
grains. b, Phase map of the region in the TEM micrograph in a showing two

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

Extended Data Figure 3 | TEM characterization of a Ta-rich nanocluster in a highlighting the misfit dislocations (indicated by the
nanocluster. a, HR-TEM micrograph of a Ta-rich nanocluster with yellow boxes) across the interface indicative of the presence of misfit
misfit lattice dislocations. b, Inverse fast Fourier transform image of the strain.

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

Extended Data Figure 4 | Mechanical behaviour of NC Cu–10 at% Ta for samples tested at a strain rate of 1 × 10−3 s−1. The blue line and red
at quasi-static strain rates. Stress–strain response of NC Cu–10 at% Ta diamonds corresponds to tensile and compressive data, respectively. The
samples. a, Compressive stress–strain curve tested at a strain rate of curves indicate behaviour ranging from elastic to nearly plastic with no
8 × 10−4 s−1 at various temperatures5. b, Tension–compression curve strain hardening; tension–compression asymmetry is absent.

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

Extended Data Figure 5 | Creep response of NC Cu–10 at% Ta with and respectively. d, Theoretical deformation map with threshold-corrected
without threshold correction. a, Creep rate versus applied stress without data (n = 5) of the NC Cu–10 at% Ta alloy, along with experimental creep
threshold correction for various temperatures. b, Creep rate versus rates of NC Cu and Ni (ref. 2). Theoretical Coble creep rates for a grain
applied stress with a stress exponent of five for various temperatures, size d = 50 nm (circles), experimental creep rates2 in nanocrystalline
used to extract the threshold stress. c, Creep rate versus normalized copper (squares; d = 25 nm) and nickel (diamonds; d = 40 nm), and the
stress, obtained by subtracting the threshold stress (σth) from the applied creep rate we found for NC Cu–10 at% Ta (triangles; d = 50 nm) are
stress (σ) for various temperatures. The red, green and blue circles in plotted.
a–c correspond to NC Cu–10 at% Ta tested at 400 °C, 500 °C and 600 °C,

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH Letter

Extended Data Figure 6 | Atomistic models used for creep study. (face-centred cubic) atoms, red atoms are stacking faults (hexagonal close-
a, Initial microstructure of pure NC Cu. b, Microstructure of pure NC Cu packed) atoms, white atoms are grain-boundary (other) atoms and blue
after 5 ns. c, Initial microstructure of NC Cu–10 at% Ta. d, Microstructure atoms are Ta (body-centred cubic) particles.
of NC Cu–10 at% Ta after 5 ns. In all panels, green atoms are grain-interior

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letter RESEARCH

Extended Data Figure 7 | Isolated single grain illustrating the effect represent the initial grain-boundary configurations and blue atoms
of Ta nanoclusters on grain-boundary motion. a, b, 2D atomistic slices represent Ta; red and green atoms represent the extent of coarsening
of a single grain obtained from pure NC Cu (a) and NC Cu–10 at% Ta associated with the plastic deformation under constant load and
(b) highlighting the effect of nanoclusters in coarsening. White atoms temperature conditions.

© 2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

You might also like