You are on page 1of 12

REPORTS

Cite as: Q. Pan et al., Science


10.1126/science.abj8114 (2021).

Gradient-cell–structured high-entropy alloy with


exceptional strength and ductility
Qingsong Pan1†, Liangxue Zhang1,2†, Rui Feng3†, Qiuhong Lu1, Ke An3, Andrew Chihpin Chuang4, Jonathan D.
Poplawsky5, Peter K. Liaw6, Lei Lu1*
1Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang, 110016, P.R. China. 2School of Materials Science
and Engineering, University of Science and Technology of China, Shenyang, 110016, P.R. China. 3Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN
37831, USA. 4Advanced Photon Source, Argonne National Laboratory, Lemont, IL 60439, USA. 5Center for Nanophases Materials Sciences, Oak Ridge National Laboratory,
Oak Ridge, TN 37831, USA. 6Department of Materials Science and Engineering, The University of Tennessee, Knoxville, TN 37996, USA.
†These authors contributed equally to this work.
*Corresponding author. Email: llu@imr.ac.cn

Most multicomponent high-entropy alloys (HEAs) lose ductility with increasing strength, similar to
conventional materials. We controllably introduced novel gradient nano-scaled dislocation-cell structures in
one stable single-phase face-centered-cubic HEA, which result in enhanced strength without apparently
losing ductility. The sample-level structural-gradient induces the progressive formation of a high density of

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


tiny stacking-faults (SFs) and twins upon early straining, nucleating from abundant low-angle dislocation
cells. Furthermore, the SF-induced plasticity and the resultant refined structures, coupled with intensively
accumulated dislocations, contribute to the plasticity, extra strengthening and work hardening. Our
findings offer a promising paradigm for tailoring properties with gradient-dislocation cells at the nanoscale
and advance our fundamental understanding of the intrinsic deformation behavior of HEAs.

High-entropy alloys (HEAs), or multi-principal-element al- slip modes (6, 14, 19–21), as well as the enhanced friction
loys, have a near-infinite multicomponent phase space that resistance to motion/accumulation of dislocations (5, 22, 23),
can lead to interesting mechanical properties (1, 2). Good enabled by increasing local concentration fluctuations
strength and ductility, high work hardening, and exceptional or local SRO at the nanoscale (generally <3 nm) (5, 6, 24),
damage tolerance have been achieved in some single-phase are believed to contribute to the improved mechanical prop-
HEAs that have inherent concentration inhomogeneity cre- erties.
ated by tailoring their chemical complexity (3–6). Moreover, We propose a heterogeneous gradient dislocation-cell
engineering a spatial heterogeneous microstructure consist- structure (GDS) in a stable single face-centered-cubic (fcc)
ing of graded grain sizes, nanoclusters, multi-phases, etc. phase Al0.1CoCrFeNi HEA containing randomly-oriented,
could also allow HEAs to achieve superior properties (7–9), equiaxed fine grains (FG) with an average size of ~46 μm.
similar to that having been achieved in traditional hetero- This alloy is a well-studied model material with a locally-
structured metallic materials (10–12). However, the long- varied SFE of 6 - 21 mJ/m2 (25). We found an unexpected and
lasting strength-ductility paradox for conventional metallic extremely high density of tiny stacking fault (SF), twin nucle-
materials still exists for most HEAs (7, 8, 13). ation and accumulation-dominated plastic deformation in
The tradeoff between strength and ductility for HEAs our GDS HEA upon early tensile straining. This feature re-
exists because the fundamental plastic-deformation features sulted in attractive strength and ductility properties, com-
and mechanisms of HEAs reported so far are similar to pared to other HEAs (10, 26, 27).
those of conventional metals (13, 14). Elementary line defects We processed HEA dog-bone-shaped bar specimens with
that carry plasticity, i.e., full dislocations and related interac- a gauge diameter of 4.5 mm and a gauge length of 12 mm,
tions with different structural defects, such as high-angle using a cyclic-torsion (CT) treatment without any surface
grain boundaries (HAGBs) or twin boundaries (TBs), are tooling to form a sample-level multi-scaled hierarchical dis-
well understood in traditional metals (15–18). Interestingly, location structure under the imposed spatially-gradient plas-
some unusual dislocation behaviors have been identified tic strains from the surface to the core (28) (fig. S1). By
in HEAs with highly concentrated solid solutions due to controllably tuning CT parameters, i.e., torsion-angle ampli-
local inhomogeneity with chemical short-range order (SRO) tudes of 20 and 6 degrees, we prepared two different GDS
and spatially variable stacking fault energy (SFE) at the samples with a constant torsion number of 200 cycles, respec-
atomic scale (14). For example, the changing dislocations tively. We refer to these samples as GDS-H (Fig. 1, 20 degrees)

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 1
and GDS-L (fig. S2, 6 degrees), based on the high and low deformation twins in the bulk GDS sample (Fig. 1 and figs. S2
accumulated torsion plastic strains imposed (i.e., 18.4 and 5.2 and S3), indicating the dislocation-controlled plastic defor-
in the topmost layer, fig. S1E, respectively). mation during the CT process, analogous to that observed in
We focus on the GDS-H sample to illustrate the salient numerous HEAs with low SFEs (13, 25, 33, 34).
microstructural feature of a sample-level hierarchical dislo- Synchrotron X-ray diffraction (SXRD) scanning on the as-
cation structure. Grains from the surface to the core of the prepared GDS-H samples from the topmost surface to the
GDS-H sample are still homogeneously distributed, with core showed a spatially-gradient-distributed dislocation den-
characters of faceted grain morphologies, unchanged grain sity, i.e., up to 8.8 × 1014 m−2 in the topmost ~200 μm-depth
size, and random crystallographic orientations (Fig. 1, A and layer (fig. S4). The GDS-H sample still exhibits a stable single
G), using electron backscatter diffraction (EBSD) with scan- fcc phase, as evidenced by EBSD, TEM, and SXRD/neutron
ning electron microscopy (SEM). The same observations were results (Fig. 1 and figs. S5 and S8). Quantitative compositional
made before CT (fig. S1C). The noticeably unchanged grain analysis at the atomic scale shows that the GDS structure is
features after CT processing is in sharp contrast to traditional compositionally homogeneous without detectable elemental
homogeneous or gradient nanostructures with severely-re- segregation at the cell wall, employing three-dimensional
fined grain sizes and higher densities of HAGBs that are pro- atom probe tomography (3D-APT) (fig. S6).
duced by employing conventional severe plastic deformation The hierarchical dislocation-cell structure results in re-
strategies (29, 30). Most grains in the sample core have the markably improved tensile properties (Fig. 2A). Tensile tests
typical planar single-slip dislocation configuration in a rela- of both GDS samples show the higher yield strength (σy, at

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


tively-low dislocation density (fig. S3C), like that reported 0.2% offset) of 362 ± 2 and 539 ± 26 MPa, about 2-3 times
in the same HEA after tension strain (25). Well-developed stronger than those of the FG and coarse-grained (CG) coun-
single-slip-induced dislocation walls with low-angle bounda- terparts (185 ± 5 MPa and 138 ± 3 MPa, respectively) (25). In
ries (LABs) are occasionally observed in several grains (Fig. addition, we measured a high uniform elongation (δu) up to
1H and fig. S3D). Distinctly, massive LABs with misorienta- 42.6 ± 0.2% for GDS-H, slightly reduced, relative to FG coun-
tions <15° were introduced in the topmost grain interior and terparts without GDS (55.4 ± 3.4%). This behavior differs
spatially distributed such that they become smaller in the vol- from the strength gain at the expense of the ductility in
ume fraction and larger in size with increasing the depth most conventional metals and HEAs (7, 10). Specially, the
from the top surface (Fig. 1B, also schematically illustrated in GDS-H sample shows steady strain hardening with a slightly-
Fig. 1C). decreased work-hardening rate, Θ, i.e., from 1.28 GPa at a 3%
We randomly select one grain separated by HAGBs in the strain to 0.99 GPa before necking, distinct from the notably
topmost layer (~100 μm from the surface) of the GDS-H sam- reduced Θ trend upon straining in its FG counterpart and the
ple and investigate its dislocation structure using transmis- concave Θ shape of CG counterpart (Fig. 2B) dominated by
sion electron microscopy (TEM). We show in Fig. 1D an deformation twinning (35).
example of the numerous equiaxed dislocation cells and cell The GDS structure that we introduced leads to an unex-
walls with a small misorientation ranging from 0.7 to 4.8°, pected deformation-induced continuous hardening behavior
corresponding to massive LABs in Fig. 1B. We observed a mis- that we found by measuring the microhardness (Hv) along
orientation gradient in the grain interior along the radial di- the depth from the surface to the cores of GDS and FG
rection (white arrow, Fig. 1D), as a result of the imposed samples before and after tension (Fig. 2C). Owing to the pres-
macro-level plastic-strain gradient upon the CT treatment ence of the sample-level gradient dislocation architecture in
(Fig. 1F) (31). We measure the cell diameter to be ~200 nm in the whole cross-section, we identified a distinct, gradiently-
the top surface, which gradually increases to ~450 nm for the distributed Hv from 3.7 GPa at the topmost surface to 2.2
dislocation-wall structure in the core (fig. S3). Each cell wall GPa in the central region in GDS-H. These values are much
(~40 nm thickness on average) is decorated with a high den- higher than that of GDS-L (from 2.3 to 1.7 GPa). Compared to
sity of dislocations (~1.7 × 1015 m−2) (28), but relatively-fewer the uniform increase in Hv for the FG sample after tensile
dislocations with curved morphologies in the cell interior straining, we found remarkably continuous hardening in
(Fig. 1E and fig. S3). We achieved a similar dislocation struc- both tensioned GDS samples from the top surface to the
ture in the GDS-L sample at a small cumulative plastic strain, core. This hardening feature is quite different from the
exhibiting undeveloped dislocation cells with a larger average deformation-induced continuous softening in conventional
size of ~450 nm in the topmost surface (figs. S1 and S2). The metals with gradient nanograins (10, 36).
dislocation cells in the GDS HEA sample are primarily caused We attributed the pronounced increase in yield strengths
by intensive multi-slip full dislocation interactions under a to the nanoscaled dislocation-cell unit with LABs (Fig. 1).
complex gradient stress/strain state (32) after a large cumu- Specifically, the contribution of LABs to strength has been
lative torsion strain (fig. S1). We observed no visible SFs or demonstrated to be comparable with that of conventional

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 2
HAGBs, and an ultrahigh hardness was observed in a nano- of both SFs and deformation twins, according to conventional
laminated Ni with LABs (18, 37). For the topmost GDS surface wisdom, are associated with the emission and slip of partial
layer of the GDS-H sample, ultra-high Hv that we measured dislocations from GBs (16), in this case most likely from
indicates that the massive low-angle dislocation cells are ef- dislocations-cell walls instead. Besides, numerous short SFs
fective in resisting dislocation motion. This trend results along the other inclined {111} slip plane are widely detected
from their structural features, including a nano-scaled size in between neighboring SF/TBs, thereby forming a unique
and high density of dislocations. We argue that this feature, 3-dimensional SF-twin structure network (Fig. 4C).
together with the obvious continuous hardening in the top- We conducted in-situ neutron-diffraction tensile experi-
most GDS surface and the unusual work-hardening response ments to confirm that the formation of SFs with detectable
of the whole GDS samples, indicates an extra-strengthening splitting of (111) and (222) planes happens at a very small
and ductilizing mechanism in gradient dislocation structures tensile strain of ~5% for GDS-H samples, while setting in
upon straining. at a large strain of ~30% for the FG counterpart (Fig. 4D).
We further characterized the microstructural evolution at Our analysis of the neutron-diffraction results yields for the
the top surface of GDS-H at different strains of 3% (the onset GDS-H sample a higher value of the SF probability, compared
stage of a steady work hardening) and 40% (the later stage of to that of the FG counterpart upon straining (Fig. 4E), which
plasticity) to unravel the intrinsic deformation mechanism of is the evidence of the easier formation of SFs at an early de-
the gradient dislocation-cell-structured HEA. At a strain of formation stage in GDS samples. Ex-situ SXRD measure-
3%, we found almost no detectable changes to the grain-level ments further reveal that the enhanced probability of SFs and

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


characteristics, including the grain shape, grain size, and ori- twins primarily stems from the contribution of GDS (fig. S8),
entation (Fig. 3, A and H). However, the density of LABs in consistent with our ex-situ SEM and TEM observations (Figs.
the subsurface layer and the core were notably reduced. The 3 and 4). By contrast, both SFs and deformation twins are
presence of these dislocation patterns is because they are in rarely observed in FG and at the core of GDS samples even at
a poorly-developed or unstable state at relatively-small cumu- a 40% strain, as confirmed by results in fig. S9, and in agree-
lative plastic strains (38, 39), while the pattern remains al- ment with that reported in numerous single fcc phase HEAs
most unchanged for the topmost grains (Fig. 3, B and I). deformed at ambient temperature (4, 9, 13, 40–42).
Specifically, a widespread class of long parallel lamellae bun- Based on the above results, we rationalize that the
dles, with an average spacing of ~1.7 μm, were detected in the extremely-high density of SFs and TBs-mediated plastic de-
majority of the topmost grain, which decreases in the number formation are primarily responsible for the superior mechan-
density with increasing the depth (Fig. 3, C and D, and fig. ical properties of the gradient dislocation-cell-structured
S7). The main components of these lamellae bundles are SFs, HEA. For the dominance of such dense SFs and TBs, which
with a few TBs, which we observed in the selected area elec- are not achievable in conventional metallic materials or most
tron diffraction (SAED) (Fig. 3E) and high-resolution TEM single fcc phase HEAs in the early deformation stage at am-
image (Fig. 3F). We detected most parallel SFs and TBs (i.e., bient temperature (8, 13, 25, 42), we mainly attributed it to
the dashed lines) across a dislocation wall (i.e., a white con- the chemical features of the HEA, nanoscaled dislocation
trast region). Close-up atomistic-resolution TEM views show cells with LABs, and their spatial gradient distribution.
that each individual long lamellae interface is essentially Plastic deformation of conventional polycrystalline fcc
composed of numerous nanoscaled SFs or twin segments, metals is known to be mainly carried by the full dislocation
several to tens of nanometers in length (Fig. 3G). Measure- slip and their intensive interactions within individual grains
ments of the thickness between adjacent SFs or TBs show an (16). Besides, as traditionally-alternative promising ductiliz-
extremely-smaller value, i.e., 2.9 nm on average, correspond- ing strategies of high-strength metals and alloys, twinning or
ing to a super high volume density (~4.14 × 108 m2/m3) in the phase transformation usually come into play, mostly through
bundle. Moreover, the dislocation cells remain almost un- either tuning compositions to reduce the SFEs or lowering
changed in shape and size. By contrast, the plastic defor- the deformation temperature (16, 43). Under such circum-
mation of most grains in the core of GDS-H is still dominated stances, deformation twinning or phase transformation be-
by intensive parallel full dislocations along {111} primarily comes prevalent, primarily owing to the ready partial
slip planes (Fig. 3, J and K). dislocation glide on consecutive (or every other) {111} atomic
With the tensile strain increased to 40% before necking, planes in fcc metals (16, 43). By contrast, for most single-
the density of planar SF interfaces increases in volume frac- phase fcc HEAs (3, 13, 44), neither twinning nor phase trans-
tion, prevailing in the grain interior (Fig. 4, A and B). Besides, formation dominates. Instead, the planar full dislocations
more numbers of ultrafine nanotwins are also detected (Fig. usually dominate the plastic deformation, as detected in FG
4C). The average spacing between adjacent SFs or TBs that (fig. S9) and the core of the GDS (Fig. 3K) after tension. Spe-
we measured is small, about 4.4 nm (Fig. 4C). The generation cially, the SF-dissociation distance of full dislocations in the

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 3
deformed HEA is wider than that in conventional metals, plastic deformation from the core to the topmost surface (fig.
which can be as large as 34 nm under tension tests at room S11), effectively suppressing the overall strain localization. All
temperature for the same HEA in the literature (25). We ar- nanoscaled dislocation cells jointly mediate plastic defor-
gue that the larger probability for partial dislocations and SFs mation, providing the spatial playground for activating par-
nucleation, rather than twinning or phase transformation, is tial dislocations and the resultant formation of SFs to
closely related to the intrinsically-spatially variable, low SFE undergo plastic deformation. Secondly, the gradient plastic
associated with the saliently-rugged local atomic environ- deformation in the multiscaled architecture is accompanied
ments (14, 22, 23, 25) and chemical SRO (5, 6, 24), inherent to by a complex stress/strain state with the sample-level stress
HEAs, although we do not have the direct evidence. The SF is partition and the presence of the back stress, especially at the
a planar defect whose energy is relative to the lattice energy early stage of plastic deformation (32). As we expected, by
level where it forms. In particular, the increase in the lattice- conducting tensile load-unload-reload tests, we determined
distortion energy induced by high-density dislocations in the the back stress of the GDS-H sample at a strain of 0.6% (close
GDS can more readily adjust the local atomic positions to fur- to 0.2% offset) to be as high as 260 MPa, approximately 50%
ther decrease the SFE (45), making the SF width between par- of the yield strength, much higher than that of the FG coun-
tial dislocations potentially larger. terpart. Consequently, such obvious extra strengthening orig-
In addition to the atomic-scaled compositional effect, the inates from the back-stress hardening associated with the
activity of the dislocations-dominated plastic deformation gradient plastic deformation of GDS, effectively contributing
usually exhibits a strong (grain or cell) size dependence (46). to a higher yield strength than that of the FG counterpart.

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


With decreasing the characteristic size (d), the full disloca- Microscopically, the nanoscaled cells with a higher disloca-
tion activity is gradually inhibited, whereas the partial dislo- tion density at the walls tend to induce a stronger local stress
cation activity becomes more favorable. In terms of the field (than the measured sample-scaled macro-back stress),
classical Orowan relation (15, 47), the critical stress required coupled with that caused by the local atomic distortions, en-
for initiating the full dislocation (τF) and partial dislocation dowing the internal driving force for partial dislocations slip-
(τP) as a function of the cell size, d, can be expressed as: ping away from the equilibrium separation (13, 38), with
2αµb dense SFs stably left behind, as we show in Figs. 3, E to G,
τF ( d ) =τ0 + (1) and 4.
d
Moreover, the ultrahigh density of dislocations at the cell
γ 2αµbp
τP ( d ) = + (2) boundaries naturally serves as abundant sustainable sources
bp d for nucleating partial dislocations. Both large numbers and
where τ0 and γ are the lattice-friction stress (57 MPa) and SFE orientation-independence feature of numerous topmost
(6 - 21 mJ/m2) (25); α is a parameter that reflects the charac- grains containing SFs (fig. S7) indicate that the nucleation of
teristic of dislocations (0.5 for an edge dislocation and 1.5 for SFs is not solely determined by the axial tensile stress state,
a screw dislocation); μ is the shear modulus (80 GPa); b and but mostly controlled by complex multiaxial stress states in
bp are the Burgers vectors of full and partial dislocations GDS upon straining.
(0.253 and 0.146 nm), respectively (28). Using a maximum Taken together, through engineering hierarchical
SFE of 21 mJ/m2 and α = 1, i.e., roughly assuming the disloca- nanoscaled GDS in a stable HEA, the potentially-new defor-
tions in cell interiors have curved morphologies (Fig. 1D and mation mechanism associated with extensive SFs/Twins
fig. S2) belonging to a mixed type, the estimated critical size occurs even after a small tensile strain, responsible for
for the transition from a full to a partial dislocation activity the steady work hardening and large tensile ductility. Such
is ~140 nm (fig. S10), which tends to become larger with de- a super-high density of extremely-fine SFs/TBs themselves
creasing SFEs, evidently being comparable with the cell size (~4.14 ×108 m2/m3) not only effectively mediates plastic
in the topmost GDS layer (Fig. 1 and fig. S3). In essence, both strains, exhibiting SF-induced plasticity (SFIP), but also
the chemical and nanoscaled cell structural characters render substantially helps delocalize plastic deformation, coupled
the high favorability of partial dislocations over full disloca- with the gradient-induced strain delocalization. First,
tions at ambient temperature (Fig. 3, F and G). the penetrations of more and more density SF/TB bundles
The sample-level gradient structure with length scales across dislocation walls progressively subdivide stable dislo-
spanning 6 orders of magnitude from the millimeter to na- cation structures and built-in a unique three-dimensional
noscaled [i.e., compositional fluctuation inhomogeneity or bundle/cell networks (coupled with inherent, spatial local
SRO, like those reported in the similar Al-Co-Cr-Fe-Ni sys- chemical fluctuations) that act as strong obstacles to disloca-
tems (5, 48)] plays a pivotal role in enabling the activation of tion slip (Fig. 3, C to G). In this case, intensive interactions
partial dislocations in our alloy. First of all, the built-in gra- among SFs/TBs and dislocations are promoted to induce con-
dient distribution of dislocation cells causes the progressive siderable strengthening and work-hardening, thus greatly

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 4
counteracting the hastily-reduced Θ trend of the GDS struc- equiatomic multicomponent alloys. Mater. Sci. Eng. A 375–377, 213–218 (2004).
doi:10.1016/j.msea.2003.10.257
ture associated with the rapid recovery of pre-existing full
3. Y. Zhang, T. T. Zuo, Z. Tang, M. C. Gao, K. A. Dahmen, P. K. Liaw, Z. P. Lu,
dislocations and a relatively-limited capacity of dislocation Microstructures and properties of high-entropy alloys. Prog. Mater. Sci. 61, 1–93
multiplications at a small strain (Fig. 2B). (2014). doi:10.1016/j.pmatsci.2013.10.001
In addition, the progressively refined three-dimensional 4. B. Gludovatz, A. Hohenwarter, D. Catoor, E. H. Chang, E. P. George, R. O. Ritchie, A
fracture-resistant high-entropy alloy for cryogenic applications. Science 345,
SF/cell networks also acts as sustainable sources for high-
1153–1158 (2014). doi:10.1126/science.1254581 Medline
density full dislocation storage during uniaxial tension. The 5. Q. Ding, Y. Zhang, X. Chen, X. Fu, D. Chen, S. Chen, L. Gu, F. Wei, H. Bei, Y. Gao, M.
newly-accumulated density of full dislocations in the GDS-H Wen, J. Li, Z. Zhang, T. Zhu, R. O. Ritchie, Q. Yu, Tuning element distribution,
sample is substantially higher than that in the FG counter- structure and properties by composition in high-entropy alloys. Nature 574, 223–
227 (2019). doi:10.1038/s41586-019-1617-1 Medline
parts at the same tensile strain (fig. S4), still consequently
6. R. Zhang, S. Zhao, J. Ding, Y. Chong, T. Jia, C. Ophus, M. Asta, R. O. Ritchie, A. M.
offering a progressive and steady work-hardening response Minor, Short-range order and its impact on the CrCoNi medium-entropy alloy.
of GDS (Fig. 2B). Benefiting from the roles co-played by such Nature 581, 283–287 (2020). doi:10.1038/s41586-020-2275-z Medline
unexpected high-density linear and planar defects, bulk GDS 7. E. Ma, X. Wu, Tailoring heterogeneities in high-entropy alloys to promote strength-
ductility synergy. Nat. Commun. 10, 5623 (2019). doi:10.1038/s41467-019-13311-
exhibits a better strain compatibility without strain localiza-
1 Medline
tion and visible cracks in the grain interior or along GBs, re- 8. P. Sathiyamoorthi, H. S. Kim, High-entropy alloys with heterogeneous
sultantly a comparable high tensile plasticity with FG microstructure: Processing and mechanical properties. Prog. Mater. Sci.
counterparts, at higher stress levels (Fig. 2A). 10.1016/j.pmatsci.2020.100709 (2020). doi:10.1016/j.pmatsci.2020.100709
9. T. Yang, Y. L. Zhao, Y. Tong, Z. B. Jiao, J. Wei, J. X. Cai, X. D. Han, D. Chen, A. Hu, J.
We compare the yield strengths and the product of

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


J. Kai, K. Lu, Y. Liu, C. T. Liu, Multicomponent intermetallic nanoparticles and
strength and ductility of GDS samples with the same compo- superb mechanical behaviors of complex alloys. Science 362, 933–937 (2018).
sitional HEA (40, 41, 49–51) and other existing high- doi:10.1126/science.aas8815 Medline
performance metals and alloys with gradient structures (10, 10. T. H. Fang, W. L. Li, N. R. Tao, K. Lu, Revealing extraordinary intrinsic tensile
plasticity in gradient nano-grained copper. Science 331, 1587–1590 (2011).
36, 52–55). The strength was normalized by its Young’s mod-
doi:10.1126/science.1200177 Medline
ulus of the corresponding material. When plotted this way, 11. X. Wu, M. Yang, F. Yuan, G. Wu, Y. Wei, X. Huang, Y. Zhu, Heterogeneous lamella
we found that the GDS HEA achieves the best combination structure unites ultrafine-grain strength with coarse-grain ductility. Proc. Natl.
of uniform tensile elongation and strength. It is not surpris- Acad. Sci. U.S.A. 112, 14501–14505 (2015). doi:10.1073/pnas.1517193112 Medline
12. Y. M. Wang, T. Voisin, J. T. McKeown, J. Ye, N. P. Calta, Z. Li, Z. Zeng, Y. Zhang, W.
ing that conventional gradient nanograined metals and alloys
Chen, T. T. Roehling, R. T. Ott, M. K. Santala, P. J. Depond, M. J. Matthews, A. V.
with a high-density of HAGBs inevitably suffer from struc- Hamza, T. Zhu, Additively manufactured hierarchical stainless steels with high
tural coarsening, such as GB migration with concomitant strength and ductility. Nat. Mater. 17, 63–71 (2018). doi:10.1038/nmat5021
coarsening (contributing to plasticity) and resultant soften- Medline
13. W. Li, D. Xie, D. Li, Y. Zhang, Y. Gao, P. K. Liaw, Mechanical behavior of high-entropy
ing, albeit which obviously sacrifice their strengths and sta-
alloys. Prog. Mater. Sci. 118, 100777 (2021). doi:10.1016/j.pmatsci.2021.100777
bility to a certain extent (10). By contrast, the monotonic 14. E. Ma, Unusual dislocation behavior in high-entropy alloys. Scr. Mater. 181, 127–
strengthening with high steady work hardening was achieved 133 (2020). doi:10.1016/j.scriptamat.2020.02.021
in the stable single fcc phase GDS HEA with spatially gradi- 15. J. P. Hirth, J. Lothe, Theory of Dislocations, 2nd. (Wiley, New York, USA, 1982).
16. M. A. Meyers, K. K. Chawla, Mechanical Behavior of Materials (Cambridge
ent LABs, contributed by the above SF and twin behavior,
University Press, Cambridge, UK, ed. 2nd, 2009).
with gradual structural refinement. 17. K. Lu, L. Lu, S. Suresh, Strengthening materials by engineering coherent internal
In summary, our observations show that engineering boundaries at the nanoscale. Science 324, 349–352 (2009).
gradient LABs structures on a single fcc-phase Al0.1CoCrFeNi doi:10.1126/science.1159610 Medline
18. K. Lu, Stabilizing nanostructures in metals using grain and twin boundary
HEA can help readily activate a SFIP-strengthening mecha-
architectures. Nat. Rev. Mater. 1, 16019 (2016). doi:10.1038/natrevmats.2016.19
nism, leading to an exceptional strength and ductility. The 19. C. Lee, J. Brechtl, P. K. Liaw, Research on Bulk-metallic Glasses and High-entropy
discovery of such SF/twin behavior in the GDS HEA is essen- Alloys in Peter K. Liaw’s Group and with His Colleagues. Metall. Mater. Trans. A 52,
tial for acquiring the common deformation features inherent 2033–2093 (2021). doi:10.1007/s11661-021-06197-6
20. F. Wang, G. H. Balbus, S. Xu, Y. Su, J. Shin, P. F. Rottmann, K. E. Knipling, J.-C.
to HEAs and widely applicable to other HEA systems to
Stinville, L. H. Mills, O. N. Senkov, I. J. Beyerlein, T. M. Pollock, D. S. Gianola,
achieve better performance with superior properties, which Multiplicity of dislocation pathways in a refractory multiprincipal element alloy.
are of both great fundamental and applied importance for Science 370, 95–101 (2020). doi:10.1126/science.aba3722 Medline
advanced engineering applications, such as automobiles, 21. C. Lee, G. Kim, Y. Chou, B. L. Musicó, M. C. Gao, K. An, G. Song, Y.-C. Chou, V.
Keppens, W. Chen, P. K. Liaw, Temperature dependence of elastic and plastic
power stations, and aeronautic systems.
deformation behavior of a refractory high-entropy alloy. Sci. Adv. 6, eaaz4748
(2020). doi:10.1126/sciadv.aaz4748 Medline
REFERENCES AND NOTES 22. J. Ding, Q. Yu, M. Asta, R. O. Ritchie, Tunable stacking fault energies by tailoring
1. J. W. Yeh, S.-K. Chen, S.-J. Lin, J.-Y. Gan, T.-S. Chin, T.-T. Shun, C.-H. Tsau, S.-Y. local chemical order in CrCoNi medium-entropy alloys. Proc. Natl. Acad. Sci.
Chang, Nanostructured High-Entropy Alloys with Multiple Principal Elements: U.S.A. 115, 8919–8924 (2018). doi:10.1073/pnas.1808660115 Medline
Novel Alloy Design Concepts and Outcomes. Adv. Eng. Mater. 6, 299–303 (2004). 23. Q. J. Li, H. Sheng, E. Ma, Strengthening in multi-principal element alloys with local-
doi:10.1002/adem.200300567 chemical-order roughened dislocation pathways. Nat. Commun. 10, 3563 (2019).
2. B. Cantor, I. T. H. Chang, P. Knight, A. J. B. Vincent, Microstructural development in doi:10.1038/s41467-019-11464-7 Medline

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 5
24. X. Chen, Q. Wang, Z. Cheng, M. Zhu, H. Zhou, P. Jiang, L. Zhou, Q. Xue, F. Yuan, J. and Applications, M. C. Gao, J. W. Yeh, P. K. Liaw, Y. Zhang, Eds. (Springer
Zhu, X. Wu, E. Ma, Direct observation of chemical short-range order in a medium- International Publishing, Switzerland, 2016), pp. 51–113.
entropy alloy. Nature 592, 712–716 (2021). doi:10.1038/s41586-021-03428-z 46. R. J. Asaro, P. Krysl, B. Kad, Deformation mechanism transitions in nanoscale fcc
Medline metals. Philos. Mag. Lett. 83, 733–743 (2003).
25. X. D. Xu, P. Liu, Z. Tang, A. Hirata, S. X. Song, T. G. Nieh, P. K. Liaw, C. T. Liu, M. W. doi:10.1080/09500830310001614540
Chen, Transmission electron microscopy characterization of dislocation structure 47. E. Orowan, in The Symposium on Internal Stresses in Metals and Alloys (Institute
in a face-centered cubic high-entropy alloy Al0.1CoCrFeNi. Acta Mater. 144, 107– of Metals, London, United Kingdom, 1948), pp. 451.
115 (2018). doi:10.1016/j.actamat.2017.10.050 48. L. J. Santodonato, P. K. Liaw, R. R. Unocic, H. Bei, J. R. Morris, Predictive
26. K. S. Kumar, H. Van Swygenhoven, S. Suresh, Mechanical behavior of multiphase evolution in Al-containing high-entropy alloys. Nat. Commun. 9, 4520
nanocrystalline metals and alloys. Acta Mater. 51, 5743–5774 (2003). (2018). doi:10.1038/s41467-018-06757-2 Medline
doi:10.1016/j.actamat.2003.08.032 49. G. Chen, J. W. Qiao, Z. M. Jiao, D. Zhao, T. W. Zhang, S. G. Ma, Z. H. Wang, Strength-
27. B. B. He, B. Hu, H. W. Yen, G. J. Cheng, Z. K. Wang, H. W. Luo, M. X. Huang, High ductility synergy of Al0.1CoCrFeNi high-entropy alloys with gradient hierarchical
dislocation density-induced large ductility in deformed and partitioned steels. structures. Scr. Mater. 167, 95–100 (2019).
Science 357, 1029–1032 (2017). doi:10.1126/science.aan0177 Medline doi:10.1016/j.scriptamat.2019.04.002
28. Materials and methods are available as supplementary materials. 50. N. Kumar, M. Komarasamy, P. Nelaturu, Z. Tang, P. K. Liaw, R. S. Mishra, Friction
29. M. A. Meyers, A. Mishra, D. J. Benson, Mechanical properties of nanocrystalline Stir Processing of a High Entropy Alloy Al0.1CoCrFeNi. JOM 67, 1007–1013
materials. Prog. Mater. Sci. 51, 427–556 (2006). (2015). doi:10.1007/s11837-015-1385-9
doi:10.1016/j.pmatsci.2005.08.003 51. J. Yang, J. W. Qiao, S. G. Ma, G. Y. Wu, D. Zhao, Z. H. Wang, Revealing the Hall-Petch
30. W. Guo, Z. Pei, X. Sang, J. D. Poplawsky, S. Bruschi, J. Qu, D. Raabe, H. Bei, Shape- relationship of Al0.1CoCrFeNi high-entropy alloy and its deformation
preserving machining produces gradient nanolaminate medium entropy alloys mechanisms. J. Alloys Compd. 795, 269–274 (2019).
with high strain hardening capability. Acta Mater. 170, 176–186 (2019). doi:10.1016/j.jallcom.2019.04.333
doi:10.1016/j.actamat.2019.03.024 52. H. W. Huang, Z. B. Wang, J. Lu, K. Lu, Fatigue behaviors of AISI 316L stainless steel

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


31. M. F. Ashby, The deformation of plastically non-homogeneous materials. Philos. with a gradient nanostructured surface layer. Acta Mater. 87, 150–160 (2015).
Mag. 21, 399–424 (1970). doi:10.1080/14786437008238426 doi:10.1016/j.actamat.2014.12.057
32. Y. T. Zhu, K. Ameyama, P. M. Anderson, I. J. Beyerlein, H. Gao, H. S. Kim, E. 53. J. Z. Long, Q. Pan, N. Tao, M. Dao, S. Suresh, L. Lu, Improved fatigue resistance of
Lavernia, S. Mathaudhu, H. Mughrabi, R. O. Ritchie, N. Tsuji, X. Zhang, X. Wu, gradient nanograined Cu. Acta Mater. 166, 56–66 (2019).
Heterostructured materials: Superior properties from hetero-zone interaction. doi:10.1016/j.actamat.2018.12.018
Mater. Res. Lett. 9, 1–31 (2020). doi:10.1080/21663831.2020.1796836 54. H. T. Wang, N. R. Tao, K. Lu, Architectured surface layer with a gradient
33. Z. Li, K. G. Pradeep, Y. Deng, D. Raabe, C. C. Tasan, Metastable high-entropy dual- nanotwinned structure in a Fe–Mn austenitic steel. Scr. Mater. 68, 22–27 (2013).
phase alloys overcome the strength-ductility trade-off. Nature 534, 227–230 doi:10.1016/j.scriptamat.2012.05.041
(2016). doi:10.1038/nature17981 Medline 55. J. J. Wang, N. R. Tao, K. Lu, Revealing the deformation mechanisms of nanograins
34. F. Otto, A. Dlouhý, C. Somsen, H. Bei, G. Eggeler, E. P. George, The influences of in gradient nanostructured Cu and CuAl alloys under tension. Acta Mater. 180,
temperature and microstructure on the tensile properties of a CoCrFeMnNi high- 231–242 (2019). doi:10.1016/j.actamat.2019.09.021
entropy alloy. Acta Mater. 61, 5743–5755 (2013). 56. D. Viladot, M. Véron, M. Gemmi, F. Peiró, J. Portillo, S. Estradé, J. Mendoza, N.
doi:10.1016/j.actamat.2013.06.018 Llorca-Isern, S. Nicolopoulos, Orientation and phase mapping in the transmission
35. S. W. Wu, G. Wang, J. Yi, Y. D. Jia, I. Hussain, Q. J. Zhai, P. K. Liaw, Strong grain- electron microscope using precession-assisted diffraction spot recognition:
size effect on deformation twinning of an Al 0.1 CoCrFeNi high-entropy alloy. Mater. State-of-the-art results. J. Microsc. 252, 23–34 (2013). doi:10.1111/jmi.12065
Res. Lett. 5, 276–283 (2017). doi:10.1080/21663831.2016.1257514 Medline
36. Y. Lin, J. Pan, H. F. Zhou, H. J. Gao, Y. Li, Mechanical properties and optimal grain 57. W. Xu, X. C. Liu, K. Lu, Strain-induced microstructure refinement in pure Al below
size distribution profile of gradient grained nickel. Acta Mater. 153, 279–289 100 nm in size. Acta Mater. 152, 138–147 (2018).
(2018). doi:10.1016/j.actamat.2018.04.065 doi:10.1016/j.actamat.2018.04.014
37. X. C. Liu, H. W. Zhang, K. Lu, Strain-induced ultrahard and ultrastable 58. W. Guo, D. A. Garfinkel, J. D. Tucker, D. Haley, G. A. Young, J. D. Poplawsky, An
nanolaminated structure in nickel. Science 342, 337–340 (2013). atom probe perspective on phase separation and precipitation in duplex stainless
doi:10.1126/science.1242578 Medline steels. Nanotechnology 27, 254004 (2016). doi:10.1088/0957-
38. H. Mughrabi, Dislocation wall and cell structures and long-range internal stresses 4484/27/25/254004 Medline
in deformed metal crystals. Acta Metall. 31, 1367–1379 (1983). doi:10.1016/0001- 59. K. Thompson, D. Lawrence, D. J. Larson, J. D. Olson, T. F. Kelly, B. Gorman, In situ
6160(83)90007-X site-specific specimen preparation for atom probe tomography. Ultramicroscopy
39. P. Peralta, C. Laird, “Fatigue of Metals” in Physical Metallurgy (Fifth Edition), D. E. 107, 131–139 (2007). doi:10.1016/j.ultramic.2006.06.008 Medline
Laughlin, K. Hono, Eds. (Elsevier, Oxford, 2014), pp. 1765–1880. 60. M. Miller, K. Russell, Performance of a local electrode atom probe. Surf. Interface
40. M. Komarasamy, N. Kumar, Z. Tang, R. S. Mishra, P. K. Liaw, Effect of Anal. 39, 262–267 (2007). doi:10.1002/sia.2488
Microstructure on the Deformation Mechanism of Friction Stir-Processed Al 0.1 61. X. L. Wang, T. M. Holden, G. Q. Rennich, A. D. Stoica, P. K. Liaw, H. Choo, C. R.
CoCrFeNi High Entropy Alloy. Mater. Res. Lett. 3, 30–34 (2015). Hubbard, VULCAN—The engineering diffractometer at the SNS. Physica B 385-
doi:10.1080/21663831.2014.958586 386, 673–675 (2006). doi:10.1016/j.physb.2006.06.103
41. D. Choudhuri, M. Komarasamy, V. Ageh, R. S. Mishra, Investigation of plastic 62. K. An, H. D. Skorpenske, A. D. Stoica, D. Ma, X.-L. Wang, E. Cakmak, First In Situ
deformation modes in Al0.1CoCrFeNi high entropy alloy. Mater. Chem. Phys. 217, Lattice Strains Measurements Under Load at VULCAN. Metall. Mater. Trans. A 42,
308–314 (2018). doi:10.1016/j.matchemphys.2018.05.050 95–99 (2011). doi:10.1007/s11661-010-0495-9
42. M. Naeem, H. He, F. Zhang, H. Huang, S. Harjo, T. Kawasaki, B. Wang, S. Lan, Z. 63. K. An, VDRIVE- Data Reduction and Interactive Visualization Software for Event
Wu, F. Wang, Y. Wu, Z. Lu, Z. Zhang, C. T. Liu, X.-L. Wang, Cooperative deformation Mode Neutron Diffraction, ORNL Report, Oak Ridge National Laboratory, ORNL-
in high-entropy alloys at ultralow temperatures. Sci. Adv. 6, eaax4002 (2020). TM-2012-621 (2012).
doi:10.1126/sciadv.aax4002 Medline 64. A. Larson, R. V. Dreele, “General Structure Analysis System (GSAS)” (Report
43. B. C. De Cooman, Y. Estrin, S. K. Kim, Twinning-induced plasticity (TWIP) steels. LAUR 86-748, Los Alamos National Laboratory, 1994).
Acta Mater. 142, 283–362 (2018). doi:10.1016/j.actamat.2017.06.046 65. B. Warren, X-ray studies of deformed metals. Prog. Met. Phys. 8, 147–202 (1959).
44. E. P. George, D. Raabe, R. O. Ritchie, High-entropy alloys. Nat. Rev. Mater. 4, 515– doi:10.1016/0502-8205(59)90015-2
534 (2019). doi:10.1038/s41578-019-0121-4 66. E. Kröner, Berechnung der elastischen Konstanten des Vielkristalls aus den
45. J. W. Yeh, “Chaper 3 Physical Metallurgy” in High-Entropy Alloys Fundamentals Konstanten des Einkristalls. Z. Phys. 151, 504–518 (1958). doi:10.1007/BF01337948

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 6
67. H. Diao, D. Ma, R. Feng, T. Liu, C. Pu, C. Zhang, W. Guo, J. D. Poplawsky, Y. Gao, P. (DOE) Office of Science User Facility operated by the Oak Ridge National
K. Liaw, Novel NiAl-strengthened high entropy alloys with balanced tensile Laboratory (ORNL). Synchrotron diffraction was conducted at the Advanced
strength and ductility. Mater. Sci. Eng. A 742, 636–647 (2019). Photon Source (APS), a U.S. DOE Office of Science User Facility operated for the
doi:10.1016/j.msea.2018.11.055 DOE Office of Science by the Argonne National Laboratory under the Contract
68. ASM Handbook Volume 08 Mechanical Testing & Evaluation (ASM International, No. of DE-AC02-06CH11357. APT was conducted at ORNL’s Center for
2000). Nanophase Materials Sciences (CNMS), which is a U.S. DOE Office of Science
69. Y. L. Lee, J. Pan, R. Hathaway, M. Barkey, Fatigue Testing and Analysis: Theory and User Facility. R. Feng thanks for the support from Material Engineering Initiative
Practice (Elsevier, Burlington, MA, USA, 2005). (MEI) at SNS, ORNL. Author contributions: L.L and P.K.L. initiated the project.
70. S. Bagheri, M. Guagliano, Review of shot peening processes to obtain L.L. supervised the project. Q.P. and L.Z. prepared the sample and performed
nanocrystalline surfaces in metal alloys. Surf. Eng. 25, 3–14 (2009). the experimental tests. Q.L. conducted TEM observations. R.F., K.A., and A.C.
doi:10.1179/026708408X334087 performed the neutron/synchrotron diffraction experiments. R.F. carried out the
71. N. A. Fleck, G. M. Muller, M. F. Ashby, J. W. Hutchinson, Strain gradient plasticity: diffraction line profile analysis. J.D.P. and R.F. performed the APT experiments
Theory and experiment. Acta Metall. Mater. 42, 475–487 (1994). and analysis. Q.P. and L.L. drafted the manuscript and all the authors
doi:10.1016/0956-7151(94)90502-9 contributed to the discussions and revised the manuscript. Competing
72. D. A. Hughes, N. Hansen, D. J. Bammann, Geometrically necessary boundaries, interests: A Chinese patent (Application number. 201911044516.5) on the
incidental dislocation boundaries and geometrically necessary dislocations. Scr. cyclic-torsion processing is being under review. Data and materials availability:
Mater. 48, 147–153 (2003). doi:10.1016/S1359-6462(02)00358-5 All data generated or analyzed during this study are included in this published
73. H. Mughrabi, Dual role of deformation-induced geometrically necessary article and its supplementary materials.
dislocations with respect to lattice plane misorientations and/or long-range
internal stresses. Acta Mater. 54, 3417–3427 (2006). SUPPLEMENTARY MATERIALS
doi:10.1016/j.actamat.2006.03.047 science.org/doi/10.1126/science.abj8114
74. H. J. Gao, Y. G. Huang, W. D. Nix, J. W. Hutchinson, Mechanism based strain Materials and Methods

Downloaded from https://www.science.org at Northeastern University on September 24, 2021


gradient plasticity – I. Theory. J. Mech. Phys. Solids 47, 1239–1263 (1999). Supplementary Text
doi:10.1016/S0022-5096(98)00103-3 Figs. S1 to S11
75. L. P. Kubin, A. Mortensen, Geometrically necessary dislocations and strain- References (56–83)
gradient plasticity: A few critical issues. Scr. Mater. 48, 119–125 (2003).
doi:10.1016/S1359-6462(02)00335-4 3 June 2021; accepted 1 September 2021
76. L. Balogh, G. Ribárik, T. Ungár, Stacking faults and twin boundaries in fcc crystals Published online 23 September 2021
determined by x-ray diffraction profile analysis. J. Appl. Phys. 100, 023512 10.1126/science.abj8114
(2006). doi:10.1063/1.2216195
77. G. Ribárik, B. Jóni, T. Ungár, Global optimum of microstructure parameters in the
CMWP line-profile-analysis method by combining Marquardt-Levenberg and
Monte-Carlo procedures. J. Mater. Sci. Technol. 35, 1508–1514 (2019).
doi:10.1016/j.jmst.2019.01.014
78. G. Ribárik, B. Jóni, T. Ungár, The Convolutional Multiple Whole Profile (CMWP)
Fitting Method, a Global Optimization Procedure for Microstructure
Determination. Crystals 10, 623 (2020). doi:10.3390/cryst10070623
79. W. C. Hinds, Aerosol Technology: Properties, Behavior, and Measurement of
Airborne Particles (John Wiley & Sons, 1999).
80. M. Wilkens, The determination of density and distribution of dislocations in
deformed single crystals from broadened X-ray diffraction profiles. Phys. Stat.
Sol. A 2, 359–370 (1970). doi:10.1002/pssa.19700020224
81. T. Ungár, I. Dragomir, Á. Révész, A. Borbély, The contrast factors of dislocations in
cubic crystals: The dislocation model of strain anisotropy in practice. J. Appl.
Cryst. 32, 992–1002 (1999). doi:10.1107/S0021889899009334
82. M. Treacy, J. Newsam, M. Deem, A general recursion method for calculating
diffracted intensities from crystals containing planar faults. Proc. R. Soc. London
A 433, 499–520 (1991). doi:10.1098/rspa.1991.0062
83. J. Chen, L. Lu, K. Lu, Hardness and strain rate sensitivity of nanocrystalline Cu.
Scr. Mater. 54, 1913–1918 (2006). doi:10.1016/j.scriptamat.2006.02.022
ACKNOWLEDGMENTS
We thank Dr. S.Y. He for performing HAADF-STEM experiments, Dr. M.X. Yang for
measuring the back stress, Dr. J. Xu for analyzing EBSD results, Dr. C.J. Zhang,
Dr. L.F. Cheng for performing raw sample forging treatment and Dr. J. Burns for
performing APT experiments. Funding: Q. Pan and L. Lu acknowledge the
financial support by National Science Foundation of China (NSFC, Grant
Numbers. 51931010, 92163202, 52122104 and 52071321), the Key Research
Program of Frontier Science and International partnership program (GJHZ2029)
and Youth Innovation Promotion Association (2019196), Chinese Academy of
Sciences (CAS), and LiaoNing Revitalization Talents Program (XLYC1802026).
P. K. Liaw very much appreciates the support from the National Science
Foundation (DMR-1611180 and 1809640) and the US Army Research Office
(W911NF-13-1-0438 and W911NF-19-2-0049). The present research used
resources at the Spallation Neutron Source (SNS), a U.S. Department of Energy

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 7
Downloaded from https://www.science.org at Northeastern University on September 24, 2021
Fig. 1. Typical microstructure and structural gradient of gradient
dislocation structure. Cross-sectional EBSD (A, B) images of the GDS-
H Al0.1CoCrFeNi HEA processed by cyclic-torsion processing at a torsion
angle amplitude of 20° showing the distributions of a grain-scaled
morphology, orientation (A), and three types of boundaries (i.e., HAGB,
LAB and TB) with different misorientation angles (B) within an
approximately 1.2 mm depth from the surface, compared with that in the
core (G, H). (C) Schematic of GDS with a gradiently-distributed low-
angle dislocation structure. The corresponding bright-field TEM images
of the dislocation structures at the topmost surface (D) of treated
samples (indicated in A, C). The misorientation angle of each cell wall in
Fig. 1D is measured and indicated in (D), using an electron procession
diffraction technique (28) in TEM. The up-left inset in (D) is the
corresponding SAED patterns. (E) A closer view of a typical dislocation-
cell structure. (F) Plots of the misorientation-angle variation, measured
with respect to the origin, across multiple cells at the topmost surface of
the GDS-H HEA, along the solid white line arrow in (D).

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 8
Downloaded from https://www.science.org at Northeastern University on September 24, 2021
Fig. 2. Mechanical properties of the GDS Al 0.1CoCrFeNi HEA.
(A) Tensile engineering stress-strain relations of GDS, CG, and FG
samples. (B) Work-hardening rate and true strain relations of GDS
samples compared with their homogeneous components. (C) The
variations of measured microhardness along the distance from the top
surface to interior of GDS samples after cyclic-torsion processing and
those after a tensile strain of 40%. (D) The product of strength
and ductility versus yield strength normalized by Young’s modulus of
the GDS Al0.1CoCrFeNi HEA, compared with the counterparts with
homogenous and gradient-grained structures and other metals and
alloys with gradiently-distributed nanograins and nanotwins reported
in the literature (10, 36, 52–55). The error bars in (C, D) represent
standard deviations from 10 independent hardness measurements and
more than 3 independent tensile tests. GNG and GNT denote the
gradient nanograin and nanotwin, respectively; TWIP denotes twinning-
induced plasticity.

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 9
Downloaded from https://www.science.org at Northeastern University on September 24, 2021
Fig. 3. Deformation microstructure of the GDS-H Al0.1CoCrFeNi HEA
at a tensile strain of 3%. Cross-sectional EBSD (A, B) images showing
the distributions of the grain-scaled morphology, orientation (A), and
three types of boundaries with different misorientation angles (B) within
~1.2 mm depths from the surface, compared with that in the core
(H, I) after tension. The corresponding SEM (C, D), bright-field TEM
(E) images presenting the widespread occurrence of dense tens of
micron-length SF bundles, indicated by the white arrows, cutting
through multiple dislocation-cell structures. The inset in (E) is
the corresponding SAED patterns containing parallel streaks (along
[111] direction, noted by the white arrow) from SFs. (F) An aberration-
corrected high-angle dark-field scanning transmission electron
microscopy (HAADF-STEM) image taken from bundles in the vicinity
of dislocation-cell walls with a relative low dislocation density, revealing
an ultrahigh density of SFs and TBs. (G) Close-up HAADF-STEM image
exhibiting numerous nanoscaled SFs or twin segments. The solid
and dashed lines in (F-G) denote SF/TBs and the (111) plane,
respectively. (J-K) The corresponding SEM (J) and TEM (K) images at
the sample core presenting planar-slip-induced parallel dislocation
morphologies. The white line with double arrows in (C) denotes loading
axis (LA).

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 10
Downloaded from https://www.science.org at Northeastern University on September 24, 2021
Fig. 4. Deformation feature at 40% tensile strain and in-situ neutron-
diffraction measurements during uniaxial tension of the GDS-H
Al0.1CoCrFeNi HEA. (A-B) SEM and TEM images of a high density of
SF bundles (indicated by white arrows) in the whole grain interior.
(C) Typical atomic-resolution HAADF-STEM image showing an ultrahigh
density of SFs and TBs with inclined short SFs. The solid and dashed
lines in (C) denote SF/TBs and the (111) plane, respectively. (D) The
evolution of the lattice strains for (111)//LA and (222)//LA grains in both
GDS-H and FG samples versus engineering strain measured by in-situ
neutron-diffraction experiments. The error bars are obtained from the
uncertainties of the single-peak fitting on hkl diffraction peaks. The black
arrows in (D) indicate the occurrence of (111) and (222) splitting event.
(E) The variation of the calculated SF probability (SFP) as a function of
engineering strain in both GDS-H and FG samples.

First release: 23 September 2021 science.org (Page numbers not final at time of first release) 11
Gradient-cell–structured high-entropy alloy with exceptional strength and ductility
Qingsong Pan, Liangxue Zhang, Rui Feng, Qiuhong Lu, Ke An, Andrew Chihpin Chuang, Jonathan D. Poplawsky, Peter K.
Liaw, and Lei Lu

Science, Ahead of Print • DOI: 10.1126/science.abj8114

View the article online


https://www.science.org/doi/10.1126/science.abj8114
Permissions
https://www.science.org/help/reprints-and-permissions

Downloaded from https://www.science.org at Northeastern University on September 24, 2021

Use of think article is subject to the Terms of service

Science (ISSN 1095-9203) is published by the American Association for the Advancement of Science. 1200 New York Avenue NW,
Washington, DC 20005. The title Science is a registered trademark of AAAS.
Copyright © 2021, American Association for the Advancement of Science

You might also like