You are on page 1of 51

International Journal of Plasticity

Achieving exceptional strength-ductility synergy in a complex-concentrated alloy via


architected heterogeneous grains and nano-sized precipitates
--Manuscript Draft--

Manuscript Number: INTPLA-D-22-00277

Article Type: Research Paper

Keywords: Complex-concentrated alloy; Heterogeneously-grained structure; Hetero-deformation


induced hardening; Precipitation strengthening; Mechanical properties

Abstract: The development of alloys with excellent strength-ductility synergy is a long-lasting


research theme for material scientists, which also holds true for the newly emerged
complex-concentrated alloys (CCAs). Here, a heterogeneously-grained microstructure
consisting of recrystallized ultrafine and fine grains containing dense nano-sized L12
precipitates, which provides synergic strengthening effects, was intentionally
introduced into a Co-free CCA through appropriate thermomechanical processing
strategy. This CCA exhibits an ultra-high yield strength of 1.5 GPa, a tensile strength of
1.8 GPa and a remarkable uniform elongation of 18.2% at room temperature. During
deformation, dislocation-slip is prevalent in fine grains while stacking faults and
nanotwins are activated in ultrafine grains. In particular, strain partition takes place and
hetero-deformation-induced (HDI) stress is accumulated via pilling up of massive
dislocations at grain boundaries and domain boundaries between fine grain domains
and ultrafine grain domains during straining, resulting in significant HDI hardening. This
HDI hardening along with the interaction between multiple deformation modes offers
the outstanding strain-hardening ability, delaying the onset of necking and hence
enabling high strength and good ductility of alloy.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Highlights

 A unique heterogeneously-grained structure mixed with L12 nanoprecipitates

hardened fine and ultrafine recrystallized grains was tailored in a Co-free

Ni42Fe30Cr12Mn8Al5Ti3 CCA via a single-step annealing after cold-rolling.

 The newly-developed CCA featured ultrahigh yield strength and ultimate strength

of ~1.5 GPa and ~1.8 GPa, respectively, and the product of strength and ductility

could reach 32.7 GPa%.

 The ultrahigh yield strength was mainly attributed to precipitation strengthening,

fine-grained strengthening, and extra hetero-deformation induced (HDI)

strengthening.

 Large strain partitioning among heterogeneous domains, leads to strong HDI

hardening, which coupled with precipitation hardening and forest dislocation

hardening contribute to the exceptional strength-ductility synergy.


Graphical Abstract Click here to download Graphical Abstract Graphical abstract.tif
Manuscript File

1
1 Achieving exceptional strength-ductility synergy in a
2
3
4 2 complex-concentrated alloy via architected heterogeneous
5
6
7 3 grains and nano-sized precipitates
8
9
10 4 Jiantao Fana,b, Xinbo Jia,b, Liming Fua,b*, Jian Wanga,b, Shuo Maa,b, Yanle Suna,b, Mao
11
12 5 Wena,b, and Aidang Shana,b*
13
14 a
15 6 School of Materials Science and Engineering, Shanghai Jiao Tong University,
16
17 7 Shanghai 200240, China
18
19 b
8 Shanghai Key Laboratory of High Temperature Materials and Precision Forming,
20
21
22 9 Shanghai Jiao Tong University, Shanghai, 200240, China
23
24 10 Abstract
25
26
27 11 The development of alloys with excellent strength-ductility synergy is a long-
28
29 12 lasting research theme for material scientists, which also holds true for the newly
30
31
32 13 emerged complex-concentrated alloys (CCAs). Here, a heterogeneously-grained
33
34 14 microstructure consisting of recrystallized ultrafine and fine grains containing dense
35
36
15 nano-sized L12 precipitates, which provides synergic strengthening effects, was
37
38
39 16 intentionally introduced into a Co-free CCA through appropriate thermomechanical
40
41 17 processing strategy. This CCA exhibits an ultra-high yield strength of 1.5 GPa, a tensile
42
43
44 18 strength of 1.8 GPa and a remarkable uniform elongation of 18.2% at room temperature.
45
46 19 During deformation, dislocation-slip is prevalent in fine grains while stacking faults
47
48
49 20 and nanotwins are activated in ultrafine grains. In particular, strain partition takes place
50
51 21 and hetero-deformation-induced (HDI) stress is accumulated via pilling up of massive
52
53
54 22 dislocations at grain boundaries and domain boundaries between fine grain domains
55
56 23 and ultrafine grain domains during straining, resulting in significant HDI hardening.
57
58
24 This HDI hardening along with the interaction between multiple deformation modes
59
60
61 1
62
63
64
65
1 offers the outstanding strain-hardening ability, delaying the onset of necking and hence
1
2 2 enabling high strength and good ductility of alloy.
3
4
5 3 Keywords: Complex-concentrated alloy; Heterogeneously-grained structure; Hetero-
6
7 4 deformation induced hardening; Precipitation strengthening; Mechanical properties;
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61 2
62
63
64
65
1 1. Introduction
1
2 2 Driven by the highly enhanced demand for metallic materials with high strength
3
4
5 3 and ductility in advanced structural applications, complex-concentrated alloys (CCAs),
6
7 4 also termed as high-/medium-entropy alloys, have drawn considerable research interest
8
9
10 5 in recent years (Li et al., 2021b; Sathiyamoorthi and Kim, 2020; Slone et al., 2019).
11
12 6 This is not just because of the new research frontier stimulated by developing CCAs,
13
14
15 7 the large compositional space enables tremendous opportunities for adjusting their
16
17 8 microstructures and properties (Li et al., 2021b). Until now, face-centered-cubic (FCC)
18
19
9 based CCAs are the most studied one owing to their good ductility and toughness at
20
21
22 10 both ambient and cryogenic temperatures (Slone et al., 2019). Besides, their functional
23
24 11 properties, including resistance against oxidation (Kai et al., 2017) and irradiation
25
26
27 12 (Dasari et al., 2021) as well as magnetic performance (Han et al., 2021), have also been
28
29 13 addressed. Unfortunately, the potential application of these FCC-type CCAs is strongly
30
31
32 14 hindered by the insufficient strength at room temperature as the intrinsic character of
33
34 15 FCC structure (Nutor Raymond et al., 2021).
35
36
16 The development of heterostructure is considered as one promising solution to
37
38
39 17 enhance the strength without a severe sacrifice of ductility through the synergistic effect
40
41 18 of domains with distinct structure parameters and mechanical properties
42
43
44 19 (Sathiyamoorthi and Kim, 2020; Wu et al., 2020; Zhu et al., 2021). Recently, various
45
46 20 heterostructures, such as gradient structures (Fu et al., 2022; Hasan et al., 2019; Pan et
47
48
49 21 al., 2021), heterogeneously-grained (HG) (Gwalani et al., 2018; Slone et al., 2019; Yang
50
51 22 et al., 2018) structures, and heterogeneously-laminated structures (Zhang et al., 2019)
52
53
54 23 have been introduced in CCAs. Fu et al. (Fu et al., 2022) and Hasan et al. (Hasan et al.,
55
56 24 2019) successfully fabricated heterogeneous gradient structures in
57
58
25 CrFeCoNiMn0.75Cu0.25 and CoCrFeNiMn via various penning methods, leading to
59
60
61 3
62
63
64
65
1 significant enhancement in tensile properties. It was reported that the HG structures
1
2 2 obtained by proper thermal-mechanical processing (TMP) can also effectively enhance
3
4
5 3 the strength-ductility combination of CCAs, which was ascribed to extra hetero-
6
7 4 deformation induced (HDI) hardening (Sathiyamoorthi and Kim, 2020; Wu et al., 2019;
8
9
10 5 Zhu et al., 2021). For instance, CoCrNi-based CCAs (Gwalani et al., 2018; Slone et al.,
11
12 6 2019) with partially recrystallized HG structures were found to have better strength-
13
14
15 7 ductility synergy over their fine- and coarse-grained counterparts. The strain gradient
16
17 8 generated by the deformation of heterogeneously-structured materials induced the
18
19
9 accumulation of geometrically necessary dislocations (GNDs) near domain boundaries,
20
21
22 10 leading to HDI hardening (Wang et al., 2019; Wu et al., 2020). A recent theoretical
23
24 11 calculation (Li et al., 2020) revealed that: it is the strain partitioning developed by the
25
26
27 12 deformation of soft and hard phases (heterogenous domains) that results in strain
28
29 13 gradient to accommodate numerous GNDs near domain boundaries.
30
31
32 14 Apart from the design of heterostructures, the introduction of nanoprecipitates
33
34 15 (NPs) can effectively block dislocation motion during deformation (Fan et al., 2020;
35
36
16 Gwalani et al., 2021; He et al., 2019; Li et al., 2021b; Liu et al., 2022a), resulting in the
37
38
39 17 so called precipitation hardening in CCAs. Ordered L12 NPs are usually introduced into
40
41 18 CCAs by adding small amount of Al and Ti or Ta elements, such as
42
43
44 19 Ni32.8Fe21.9Co21.9Cr10.9Al7.5Ti5.0 (Fan et al., 2020), Al0.2Ti0.2Co0.7CrFeNi1.7 (Gwalani et
45
46 20 al., 2021), (NiCoCr)92Al6Ta2 (Zhang et al., 2021b) and Fe34.5Co30.3Ni30.2Ta5.0 (Han et
47
48
49 21 al., 2021) CCAs. Accordingly, the strength of FCC-structured CCAs was enhanced.
50
51 22 Our early work (Fan et al., 2022) on Ni47-xFe30Cr12Mn8AlxTi3 CCAs indicated that a
52
53
54 23 desired L12 hardened structure and good strength-ductility combination can be obtained
55
56 24 in the Ni42Fe30Cr12Mn8Al5Ti3 CCA. Moreover, a recent study on the Ni35(CoFe)55V5Nb5
57
58
25 CCA (Liu et al., 2022a) showed that a high YS of 855 MPa and an excellent uniform
59
60
61 4
62
63
64
65
1 elongation (UE) of 50% were achieved by the precipitation of L12-Ni3Nb in the matrix.
1
2 2 These encouraging results indicate that both architecting heterostructures and
3
4
5 3 introducing NPs have great potential to provide superior strength-ductility combination
6
7 4 for CCAs.
8
9
10 5 Since the strengthening provided solely by heterostructures or NPs is limited and
11
12 6 it hardly enhances YS higher than 1 GPa (Liu et al., 2022a; Slone et al., 2019; Wu et
13
14
15 7 al., 2019), many efforts have been paid on fabricating the heterostructures with NPs
16
17 8 recently (Du et al., 2020; He et al., 2021a; He et al., 2021b; Qin et al., 2022a; Zhang et
18
19
9 al., 2020). For example, a combined gradient structure along the depth of grain size and
20
21
22 10 content of NPs was fabricated in the Al0.5Cr0.9FeNi2.5V0.2 CCA by surface mechanical
23
24 11 attrition treatment followed by two-step heat treatments (Qin et al., 2022a). An
25
26
27 12 improved combination of strength and ductility over the typical gradient structure was
28
29 13 obtained. By cold rolling and one-step annealing of the Ni2CoCrFeTi0.24Al0.2 CCA, He
30
31
32 14 et al. (He et al., 2021a) fabricated a complex heterostructure consisting of HG matrix
33
34 15 with a recrystallization volume fraction of 55% and L12 NPs with particle size in the
35
36
16 range of 10-100 nm (termed as HG-NP). Precipitation hardening and HDI hardening
37
38
39 17 were thus effectively coupled, leading to a high yield strength of ~1.3 GPa and tensile
40
41 18 elongation of ~20%. Through similar HG-NP structure design, the YSs of the
42
43
44 19 Co40Cr20Ni20Al5Ti5 (Li et al., 2021a) and Co34.46Cr32.12Ni27.42Al3Ti3 (Du et al., 2020)
45
46 20 CCAs were significantly enhanced to 1.7 GPa and 2.0 GPa with the UEs of 11.4% and
47
48
49 21 13%, respectively. However, although the improved strengthening by HG-NP, the
50
51 22 reported CCAs contain high content of expensive Co, which inhibits their
52
53
54 23 commercialization. So, it is necessary but still challenging to develop Co-free CCAs
55
56 24 with ultrahigh YS of ≥ 1.5 GPa and reasonable ductility at room temperature for
57
58
25 advanced structural applications.
59
60
61 5
62
63
64
65
1 Many recent studies have been focused on the HG and HG-NP CCAs with the HG
1
2 2 matrix composed of non-recrystallized and fully recrystallized grains (Du et al., 2020;
3
4
5 3 He et al., 2021a; He et al., 2021b; Li et al., 2021a; Qin et al., 2022b), hereafter we note
6
7 4 such structures as partially recrystallized HG and HG-NP. The non-recrystallized grains,
8
9
10 5 i.e., the retained deformed grains (RDGs), usually have grain size from tens to hundreds
11
12 6 of micrometers. However, the size of RDGs might be too large to provide good
13
14
15 7 hardening effect by grain refinement. Furthermore, the boundaries between the RDGs
16
17 8 and the ultrafine grained (UFG) matrix are relatively sharp, which is detrimental to
18
19
9 hardening (Wang et al., 2019). Li et al. (Li et al., 2021a) studied the deformation
20
21
22 10 mechanism of the partially recrystallized HG-NP structures, and found that the GND
23
24 11 density in RDGs is sharply declined at the early plastic deformation stage. This
25
26
27 12 indicated that the dislocation annihilation overwhelms dislocation multiplication in this
28
29 13 stage, leading to a softening effect in RDGs. Wu et al. (Wu et al., 2019) ascribed this
30
31
32 14 steep annihilation of dislocations to the interaction between the newly-emitted
33
34 15 dislocations and pre-existed dislocations. Zhang et al. (Zhang et al., 2021a) studied the
35
36
16 effect of different HG structures, including partially recrystallized and fully
37
38
39 17 recrystallized HG structures, on the deformation mechanisms and mechanical
40
41 18 properties of HG β-Ti alloy. They unveiled that the fully recrystallized HG structures
42
43
44 19 manifest stronger HDI hardening effect and better strength-ductility synergy than the
45
46 20 partially recrystallized one (Zhang et al., 2021a). These results demonstrate that
47
48
49 21 fabricating fully recrystallized HG structures other than partially recrystallized ones is
50
51 22 a better design strategy for developing high strength CCAs.
52
53
54 23 Inspired by the aforementioned analysis, in the present study we aimed to develop
55
56 24 an ultra-strong and ductile Co-free CCA. For such purpose, a fully recrystallized HG-
57
58
25 NP structure was architected in a Ni42Fe30Cr12Mn8Al5Ti3 CCA (in unit of atomic
59
60
61 6
62
63
64
65
1 percentage at.%) through a simple TMP, i.e., cold rolling followed by a one-step
1
2 2 annealing treatment. The microstructure was analyzed to provide valuable insight into
3
4
5 3 heterostructures design of high-performance alloys. Tensile properties at room
6
7 4 temperature and deformation microstructures at various strains were systematically
8
9
10 5 investigated. Specially, the strengthening mechanisms, the correlations between
11
12 6 structural incompatibilities, the strain partitioning behavior and the overall mechanical
13
14
15 7 properties were analyzed and discussed.
16
17 8 2. Experiments
18
19
9 2.1 Material preparation
20
21
22 10 The master Ni42Fe30Cr12Mn8Al5Ti3 CCA was selected basing on our recent work
23
24 11 (Fan et al., 2022), wherein a good strength-ductility combination was achieved by
25
26
27 12 fabricating the alloy with a coarse-grained (CG) and L12-type NPs (CG-NP)
28
29 13 microstructure. Three alloy ingots were obtained by arc-melting in Ti-gettered argon
30
31
32 14 atmosphere. An excessive content of 5 wt.% Mn was added to compensate for its
33
34 15 burning loss. Each ingot was inverted and re-melted five times to ensure compositional
35
36
16 homogeneity. The three ingots were finally melted together to make a slab with
37
38
39 17 dimensions of 8 × 20 × 60 mm3. The as-received slab was first homogenized at 1150 °C
40
41 18 for 2 h and water cooled. The homogenized slab was then multi-pass cold rolled (CR)
42
43
44 19 along lengthwise direction until a total thickness reduction of 86% is reached. The slab
45
46 20 was finally cut into several sheets. The resulting CR sheets were annealed at 650 °C for
47
48
49 21 8 h followed by air cooling to introduce HG structure and NPs simultaneously. For
50
51 22 comparison, we also examined the precipitation-containing sample with uniform fine-
52
53
54 23 scale grain distribution (FG-NP). A two-step heat treatment scheme was applied on the
55
56 24 CR sheets to prepare the FG-NP sample, i.e., annealing at high temperature of 900 °C
57
58
25 for 1 h followed by water cooling, and subsequently annealing at 650 °C for 8 h
59
60
61 7
62
63
64
65
1 followed by air cooling. Annealing treatment on the CR sample at 650 °C for 30 min
1
2 2 was performed to further clarify the formation mechanism of HG-NP structure.
3
4
5 3 2.2 Microstructural characterization
6
7 4 Phase structure was characterized by X-ray diffraction (XRD) using a Shimadzu
8
9
10 5 XRD-6000 instrument equipped with Cu-Kα radiation ( = 1.54056 Å), and scanning
11
12 6 was performed over the 2θ range from 20° to 100°with a speed of 1°/min. A slow scan
13
14
15 7 in the range of 89°~92° with a speed of 0.1°/min was conducted carefully to gain the
16
17 8 lattice information by deconvolution of the (311) diffraction peak. Lattice misfit ε was
18
19
20 9 determined according to the equation of ε = 2(aFCC - aL1 )/(aFCC + aL1 ), in which a stands
2 2

21
22 10 for the lattice parameter of each phase. Microstructures were characterized by scanning
23
24
11 electron microscopy (SEM) using a JSM-7800F microscope and by electron-
25
26
27 12 backscattered diffraction (EBSD) using a TENSCAN MIRA3 and GAIA3 instruments
28
29 13 with a step size of 0.08 μm. EBSD analysis was performed using the TSL OIM software.
30
31
32 14 Transmission electron microscopy (TEM), selected area electron diffractometer
33
34 15 (SAED), scanning TEM (STEM), and energy dispersive X-ray spectrometer (EDX)
35
36
37 16 were conducted on a Talos F200X and JEM-2100F instruments operating at 200 kV.
38
39 17 TEM slices were cut from the initial (undeformed) and deformed samples,
40
41
42
18 mechanically grounded to ~30 μm, and then twin-jet polished in a solution of 5 vol.%
43
44 19 perchloric acid and 95 vol.% ethanol at around -20 °C. Transmission Kikuchi
45
46 20 diffraction (TKD) analysis was also performed by TENSCAN MIRA3 instrument using
47
48
49 21 the TEM specimens with a step size of 15 nm.
50
51 22 2.3 Mechanical tests
52
53
54 23 Vickers hardness tests were performed by Buehler Micro-Hardness Tester with a
55
56 24 loading force of 500 g and a dwell time of 15 s. Flat dog-bone-shaped tensile samples,
57
58
59 25 with a geometry of 15 mm gauge length, 2 mm width and 1 mm thickness, were
60
61 8
62
63
64
65
1 machined along the longitudinal direction by electrical discharge machining. Quasi-
1
2 2 static uniaxial tensile tests were performed at room temperature and at a normal strain
3
4
5 3 rate of 5 × 10-4 s-1 using a Zwick Roell-Z100 testing machine. Interrupted tensile tests
6
7 4 after several different strains were also conducted to characterize microstructure
8
9
10 5 evolution during deformation. Engineering strain during tensile tests was monitored
11
12 6 using a 10-mm-gauge extensometer. The conventional 0.2% offset plastic strain method
13
14
15 7 was used to determine the yield strength. Loading-unloading-reloading (LUR) tensile
16
17 8 tests with the same conditions as that of monotonic tests were conducted to evaluate
18
19
9 HDI stress. Three samples were tested for each material state to assure the
20
21
22 10 reproducibility of results. Sampling locations for each characterization were
23
24 11 schematically displayed Fig. S1a.
25
26
27 12 3. Results
28
29 13 3.1 Initial microstructures
30
31
32 14 Microstructure of the HG-NP sample is shown in Figs. 1-4. XRD result, as shown
33
34 15 in Fig. 1a, indicates the co-existence of FCC, L12 and BCC phases. The L12 phase is
35
36
16 evidenced by the typical superlattice (100) and (110) peaks shown in the inset of Fig.
37
38
39 17 1a. The lattice constants of aFCC and aL1 , derived from the devolution of the (311)
2

40
41 18 diffraction peak shown in Fig. 1b, are 0.3600 nm and 0.3595 nm, respectively. The BCC
42
43
44 19 phase was identified to be ordered L21 crystal structure by TEM, as shown in Fig. 2 and
45
46 20 Table S1. The SEM image of the microstructure of the HG-NP sample shown in Fig.
47
48
49 21 1c reveals an alternate distribution of light and dark regions. The EBSD IPF map and
50
51 22 phase map of the interest region highlighted in Fig. 1c are shown in Figs. 1d and e.
52
53
54 23 Noticeably, the dark regions comprise FGs with sizes in the range of 1.1-7.0 μm, while
55
56 24 the light regions are filled with UFGs with sizes in the range of 167-942 nm. The
57
58
25 volume fractions of the FG and UFG domains are 52.8% and 47.2%, respectively. This
59
60
61 9
62
63
64
65
1 HG microstructure is fully recrystallized, as evidenced by the kernel average
1
2 2 misorientation (KAM) analysis shown in Fig. S2a. A small amount of BCC particles
3
4
5 3 are observed, as shown in Fig. 1e, in agreement with the XRD result. The detailed
6
7 4 nature of UFG domains were further exploited using TKD, as shown in Figs. 1f and g.
8
9
10 5 We notice that the blocky L21 particles with size ranging from tens to hundreds of
11
12 6 nanometers are mostly formed and located at FCC grain boundaries (GBs) or triple-
13
14
15 7 junctions in UFG domains (Fig. 1f). Typical Kurdjumov–Sachs (K-S) orientation
16
17 8 relationship between the L21 particles and FCC matrix is followed, as shown in Fig. 1g.
18
19
9 Additionally, a large number of annealing twins (ATs) were also formed, as
20
21
22 10 demonstrated by the white lines shown in Figs. 1e and f. In contrast, the FG-NP sample
23
24 11 exhibits a relatively uniform distribution of FGs with average size of 3.6 μm, as shown
25
26
27 12 in Figs. S3a and b. Similarly, the blocky L21 particles with average size of 206 nm at
28
29 13 GBs or triple-junctions are also found in the FG-NP sample (Figs. S3c and d).
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61 10
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 1
35
36
37 2 Fig. 1. Microstructure of the HG-NP sample. (a) XRD pattern. (b) Deconvolution of the (311)
38
39 3 diffraction peak showing the lattice parameter of the FCC and L12 phases as well as the lattice
40
41 4 misfit between them. (c) SEM image. (d) EBSD inverse pole figure (IPF) and (e) corresponding
42
43 5 phase map of the squared area in (c). (f) TKD phase map of UFG domain and (g) corresponding
44
45 6 image quality (IQ) map superimposed by phase color. Red lines in (g) indicate that the FCC/L21
46
47 7 interfaces exhibit a K-S orientation relationship.
48
49
8 TEM analysis was carried out to characterize the detailed nature inside FGs and
50
51
52 9 UFGs, as shown in Figs. 2-4. The alternate distribution of FG and UFG domains can
53
54 10 be observed in the TEM bright-field (BF) image, as displayed in Fig. 2a. TEM dark-
55
56
57 11 field (DF) images show that nanoprecipitates are formed in both FG (Fig. 2b) and UFG
58
59 12 (Fig. 2c) domains. The SAED patterns in the insets of Figs. 2b and c reveal that the
60
61 11
62
63
64
65
1 precipitates are L12 ordered structure. These precipitates are in two different shapes:
1
2 2 nanorods that mostly distribute along GBs and spheroids that uniformly distribute in
3
4
5 3 the grain interior. This is associated with the discontinuous- and continuous-
6
7 4 precipitation behaviors, denoted as DP and CP, respectively (Gwalani et al., 2021; Zhao
8
9
10 5 et al., 2017), commonly observed in Ni3(Al, Ti) strengthened CCAs (Bała et al., 2020;
11
12 6 Gwalani et al., 2021; He et al., 2016; Yang et al., 2020). The average size of nano-rodic
13
14
15 7 DPs and spheroidal CPs are 25.5 nm and 22.0 nm in FG domains, and 19.4 nm and 13.9
16
17 8 nm in UFG domains, respectively, as shown in Figs. 2d and e. The two types of
18
19
9 precipitates are all coherent with FCC matrix, as depicted by the high-resolution TEM
20
21
22 10 (HRTEM) images shown in Figs. 3a and b. The precipitate/matrix lattice misfit,
23
24 11 following the XRD analysis of Fig. 1a), is 0.147%. Qualitative STEM-EDX analysis
25
26
27 12 shown in Fig. 4 indicates that both nano-rodic and spheroidal L12 nanoparticles are
28
29 13 enriched with Ni, Al, and Ti elements. In addition, there are some blocky particles at
30
31
32 14 GBs or triple-junctions in UFG domains, which are 224 nm in size, as shown in Fig.
33
34 15 S4. This is coinciding with the TKD observation shown in Fig. 1f. Characteristic
35
36
37 16 superlattice diffraction spots shown in the inset in Fig. S4a demonstrate that the blocky
38
39 17 particles are in ordered BCC structure, which are enriched with Ni, Al and Mn, as
40
41 18 depicted by the corresponding STEM-EDX maps in Fig. S4a. The quantitative STEM-
42
43
44 19 EDX results listed in Table S1 clarify that these ordered BCC particles are Heusler-type
45
46 20 L21 phase. Some L21 particles with a size of tens of nanometers are also observed inside
47
48
49 21 FGs, as shown in Fig. S4b. Similar precipitation behavior of the coherent L12
50
51 22 nanoparticles inside grains are also observed in the FG-NP sample, as shown by the
52
53
54 23 TEM images in Figs. S5 and S6. Comparing to the CG-NP sample shown in our
55
56 24 previous work (Fan et al., 2022), two distinctly different features can be seen in the HG-
57
58
25 NP and FG-NP samples, irrespective of their grain sizes distribution: firstly, nanorods
59
60
61 12
62
63
64
65
1 with relatively high density are formed, indicating the domination of DPs; secondly,
1
2 2 numerous coarse L21 particles are prevailing at GBs and triple-junctions, demonstrating
3
4
5 3 that CR greatly affects the phase transformation behavior of the alloy upon annealing
6
7 4 process. The detailed microstructure characteristics and precipitation parameters for the
8
9
10 5 HG-NP CCA were summarized in Table 1.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 6
42
43 7 Fig. 2. TEM characterization of the HG-NP sample. (a) Low-magnified bright-field TEM (BF-
44
45 8 TEM) image. Boundaries between UFG and FG domains are highlighted by yellow dot lines, and
46
47 9 annealing twin boundaries are marked by red dash lines. (b) and (c) Magnified dark-field TEM
48
49
50
10 (DF-TEM) images of the FCC grains in FG and UFG domains, respectively. Corresponding
51
52 11 SAED patterns are shown in the insets of (b) and (c). (d) and (e) Particle size distribution of
53
54 12 nanorods and spheroidal nanoparticles in FGs and UFGs, respectively. Note that the particle size
55
56 13 of nanorods was evaluated from interlamellar spacings.
57
58
59
60
61 13
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 1
29
30 2 Fig. 3. HRTEM images and corresponding FFT images of a L12 spheroidal nanoparticle in (a) and
31
32 3 a nanorod in (b) in HG-NP sample. Boundaries between the precipitate and the FCC matrix are
33
34 4 highlighted by yellow dot lines.
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 5
56
57 6 Fig. 4. BF-STEM images of representative DP (a) and CP (b) zones in HG-NP sample, with the
58
59 7 corresponding compositional maps of L12 nanorods and spheroidal nanoparticles, respectively.
60
61 14
62
63
64
65
1 Table 1
1
2 2 Mean grain size (including ATs), particle size and volume fraction of FCC matrix, L12 and L21
3
4 3 particles in various domains. The mean diameter of L12 nanorods was evaluated from interlamellar
5
6 4 spacings. All data are estimated based on TEM and EBSD results, and the detailed evaluation of the
7
8 5 volume fraction of L12 particles was described in Supplementary text.
9
10
11 Sample Domain Phase Grain/particle size Volume fractions (vol.%)
12
13
14 HG-NP FG FCC 2.3 μm 69.6
15
16
17 L21 199 nm 3.1
18
19
20 L12 (nanorod/spheroid) 22.0 nm/25.5 nm 18.4/5.9
21
22
23 UFG FCC 600 nm 63.6
24
25
26 L21 224 nm 9.9
27
28
29 L12 (nanorod/spheroid) 13.9 nm/19.4 nm 14.5/8.6
30
31
32 FG-NP FG FCC 3.7μm 70.6
33
34
35 L21 355 nm 4.2
36
37
L12 (nanorod/spheroid) 28.2 nm/22.8 nm 15.6/5.9
38
39
40
41 6 3.2 Mechanical behaviors
42
43
44 7 Values of hardness corresponding to random 50 indents for HG-NP and FG-NP
45
46 8 samples are presented in Fig. S1b. The hardness of HG-NP exhibits prominent
47
48
9 fluctuation from 405 to 480 HV, demonstrating the mechanical incompatibility between
49
50
51 10 FG and UFG domains. The engineering stress-strain curves of the present studied
52
53 11 specimens are shown in Fig. 5a, and the detailed mechanical properties are summarized
54
55
56 12 in Fig. 5b and Table S3. Compared with CG-NP sample, which has a low YS of 661
57
58 13 MPa and high UE (as a measure of ductility) of 25.3%, FG-NP sample exhibits an
59
60
61 15
62
63
64
65
1 increased YS of 1035 MPa at UE of 20.4%. This is typical for most reported NP CCAs
1
2 2 with uniform grain-size distribution (He et al., 2019; He et al., 2016; Zhao et al., 2018).
3
4
5 3 Surprisingly, the HG-NP exhibits a YS as high as 1547 MPa and a UTS of 1797 MPa.
6
7 4 The YS is about 50% higher than that of the FG-NP sample, and is 1.3 times higher
8
9
10 5 than that of the CG-NP sample. In contrast to the continuous yielding of the CG-NP
11
12 6 sample, discontinuous and obvious yielding occurs in the FG-NP and HG-NP samples.
13
14
15 7 This yielding behavior was found in FG/UFG alloys (Liu et al., 2022b; Sun et al., 2017).
16
17 8 Apart from the significantly enhanced strength, the HG-NP sample maintains a good
18
19
9 ductility of up to 18.2% and exhibits the highest UTS×UE value of 32.7 GPa% among
20
21
22 10 the three samples, as shown in Fig. 5b. The ductile nature of the HG-NP sample is
23
24 11 further confirmed by its dimple fracture characteristics shown in Fig. S7.
25
26
27 12 True stress-strain curves of HG-NP, FG-NP and CG-NP samples, along with their
28
29 13 strain-hardening rates (SHRs), are shown in Fig. 5c. Three-stage hardening occurs in
30
31
32 14 the three samples. The strain-hardening behavior for the CG-NP specimen is similar to
33
34 15 that of L12-precipitate strengthened CG alloys (Ming et al., 2018; Zhao et al., 2017;
35
36
16 Zhao et al., 2018), as detailed in our recent work (Fan et al., 2022). Regarding the SHRs
37
38
39 17 of the HG-NP and FG-NP samples, they drop sharply in stage I and then upturn steeply
40
41 18 in stage II. In stage II, the strain-hardening curves of the HG-NP and FG-NP samples
42
43
44 19 differ markedly from that of the CG-NP sample, which is almost flat, indicating that
45
46 20 there is a marked difference in strain-hardening between them. When the true strain
47
48
49 21 exceeds 6%, the true stress monotonically drops until necking starts in stage III. It is
50
51 22 intriguing that the SHRs of both HG-NP and FG-NP samples are even higher than CG-
52
53
54 23 NP counterpart in the true strain ranges of 2.9-11.5% and 3.8-12.3%, respectively, as
55
56 24 shown in Fig. 5c. This substantial strain-hardening capability allows good UE in the
57
58
25 HG-NP and FG-NP samples at high stress levels. Compared with recently reported
59
60
61 16
62
63
64
65
1 partially recrystallized HG-NP CCAs (Du et al., 2020; He et al., 2021a; He et al., 2021b;
1
2 2 Li et al., 2021a; Qin et al., 2022b), the current CCA with HG-NP structure has best
3
4
5 3 strength-ductility synergy, as shown in Fig. 5d. This indicates that the fully
6
7 4 recrystallized HG-NP structure is more effective than the partially recrystallized one in
8
9
10 5 providing ultrahigh strength and ductility combination for CCAs.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40 6
41
42 7 Fig. 5. Mechanical properties of the HG- NP, FG-NP and CG-NP samples at room temperature. (a)
43
44
45 8 Representative tensile stress-strain curves. (b) YS, UTS, UE and UTS×UE of various
46
47
48 9 microstructures. (c) True stress-strain curves and strain-hardening rate (dt /dt, where the t and
49
50 10 t are the true stress and true strain, respectively) curves. (d) The map of the UTS×UE versus UE,
51
52
53 11 including reported partially recrystallized HG-NP CCAs and current fully recrystallized HG-NP
54
55 12 CCA.
56
57
13 3.3 LUR stress-strain behavior
58
59
60 14 It is generally accepted that HDI stress contributes significantly to the mechanical
61 17
62
63
64
65
1 properties of HG-structured materials (He et al., 2021a; Sathiyamoorthi and Kim, 2020).
1
2 2 Here, LUR tests were conducted to explore the evolution of HDI stress during straining,
3
4
5 3 as shown in Fig. 6a. As can be seen from a close-up view of the 3rd LUR curve shown
6
7 4 in Fig. 6b, both the FG-NP and HG-NP samples exhibit obvious hysteresis loops,
8
9
10 5 implying a strong Bauschinger effect (Zhu and Wu, 2019). More specially, strong strain
11
12 6 gradient develops in the HG-NP sample due to the distinct nature of different domains,
13
14
15 7 which induces the significant accumulation of GNDs near domain boundaries and
16
17 8 results in high HDI stress (Ma and Zhu, 2017; Sathiyamoorthi and Kim, 2020). The
18
19
9 dynamic evolution of HDI stress with true strain, whose values were calculated using
20
21
22 10 the method of Yang et. al (Yang et al., 2016), is given in Fig. 6c. It is clear that HDI
23
24 11 stress in the CG-NP sample is monotonously increased, while the evolution of HDI
25
26
27 12 stress can be divided into two stages in both HG-NP and FG-NP samples. In the first
28
29 13 stage with strains lower than 6%, HDI stress increases rapidly, indicating that HDI
30
31
32 14 hardening is most effective in the early deformation stage (Li et al., 2017). In the second
33
34 15 stage with strain larger than 6%, HDI stress remains almost constant. As a whole, the
35
36
16 HDI stress of the HG-NP sample is the highest among all the three samples at a given
37
38
39 17 strain, followed by the FG-NP and CG-NP samples: for instance, at true strain of 6%,
40
41 18 the HG-NP sample exhibits a HDI stress of ~942 MPa, which is significantly higher
42
43
44 19 than the values exhibited by FG-NP (~534 MPa) and CG-NP (~128 MPa) samples.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61 18
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 1
28
29 2 Fig. 6. (a) The loading-unloading-reloading (LUR) true stress-strain curves for HG-NP, FG-
30
31 3 NP and CG-NP samples. (b) Magnified view of typical hysteresis loop taken from the shadow area
32
33 4 in (a). (c) Dependence of HDI stress on plastic true strain.
34
35 5 3.4 Deformed microstructures
36
37 6 The strain-hardening behavior of the HG-NP sample should be correlated with the
38
39
40 7 microstructural evolution during plastic deformation. To clarify this, the deformed
41
42 8 microstructures of the HG-NP sample at several true strains were investigated by ex-
43
44
9 situ EBSD (Figs. 7 and 8) and TEM (Figs. 9 and 10).
45
46
47 10 3.4.1 Ex-situ EBSD characterization
48
49 11 EBSD KAM maps for the deformed HG-NP samples with true strains of 4%, 10%
50
51
52 12 and 17% (fractured) are shown in Fig. 7. The evolutions of average KAM value in
53
54 13 different domains are listed in Table 2. Two results can be obviously obtained: (i) The
55
56
57 14 average KAM values in FG domains are lower than that in UFG domains among the
58
59 15 whole straining process, that is, the strain in UFG domains is higher than that in FG
60
61 19
62
63
64
65
1 domains; (ii) with increasing tensile strain, the average KAM values in both domains
1
2 2 increase, while the difference of the average KAM values between FG and UFG
3
4
5 3 domains decreases. To elucidate this, detailed characterizations are conducted on the 4%
6
7 4 and 17% strained samples, as shown in the insets in Fig. 7a and c, respectively. Much
8
9
10 5 higher KAM values are found in the areas adjacent to GBs and annealing twin
11
12 6 boundaries (ATBs), which is more evident in FG domains. It is interesting to note that,
13
14
15 7 at true strain of 4%, the KAM value in the FG region adjacent to the domain boundary
16
17 8 is higher than that far away from it, as indicated by the micro-KAM profile in Fig. 7d.
18
19
9 These suggest that the mechanical incompatibility between the FG and UFG domains
20
21
22 10 is more effective than GBs in generating strain gradient. When strain reaches 17%, as
23
24 11 shown in Fig. 7e, the KAM value in the FG region near the domain boundary is found
25
26
27 12 to be almost identical with that at GBs. It appears that the mechanical difference
28
29 13 between FG and UFG domains is reduced with progressively straining.
30
31
32 14 As a comparison, the microstructure of the fractured FG-NP sample is shown in
33
34 15 Figs. 8a. The distributions of KAM values of the HG-NP and FG-NP samples in
35
36
16 undeformed and fractured states are shown in Fig. 8b. The inset in Fig. 8b compares
37
38
39 17 the distribution of KAM values in FG domains in the HG-NP sample and in the FG-NP
40
41 18 sample. Obviously, the average KAM value in both the FG domains and the whole body
42
43
44 19 of the fractured HG-NP sample is higher than that in the fractured FG-NP sample. This
45
46 20 indicates that the HG-NP structure effectively induces strong strain gradient and
47
48
49 21 produce high density of GNDs during deformation, and thus leading to stronger HDI
50
51 22 hardening for better mechanical properties.
52
53
54
55
56
57
58
59
60
61 20
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 1
39
40 2 Fig. 7. Deformed microstructure of HG-NP sample. (a), (b) and (c) EBSD KAM maps at true
41
42 3 strain of 4%, 10% and after failure (equivalent true strain is ~17%), respectively, corresponding to
43
44 4 FG (1) and UFG (2) domains. Insets of (a) and (c) are two enlarged images of squared regions in
45
46 5 each map. (d) and (e) point-to-point misorientation profile (red line) and corresponding KAM
47
48 6 values (green line) along the white lines labelled in the insets of (a) and (c), respectively. White
49
50 7 arrows in (e) indicate the walls of dislocation cells.
51
52
8 Table 2
53
54
55 9 The evolutions of average KAM value and GND (ρGND) density in different domains of the HG-NP
56
57 10 sample during tension at room temperature.
58
59
60
61 21
62
63
64
65
1 True strain (%) Domain KAM (°) ρGND (m-2)
2
3
4
4 FG 0.296 5.08E+14
5
6 UFG 0.504 8.64E+14
7
8
9 10 FG 0.408 6.99E+14
10
11
12 UFG 0.559 9.58E+14
13
14
15 17 FG 0.598 1.02E+15
16
17
18 UFG 0.702 1.20E+15
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 1
34
35 2 Fig. 8. (a) EBSD KAM maps for the fractured FG-NP sample. (b) KAM values of HG-NP and
36
37 3 FG-NP samples at undeformed (solid lines) and fractured (dash lines) states. The inset in (b)
38
39
40
4 compares KAM values of FG-NP sample and of FG domains in HG-NP sample.
41
42 5 3.4.2 Ex-situ TEM characterization
43
44 6 In order to further explore the deformed microstructural evolution, the deformed
45
46
47 7 HG-NP samples at various true strains were also examined by ex-situ TEM, as shown
48
49 8 in Figs. 9-10. It is observed that dislocation gliding and cutting through L12 nanorods
50
51
52
9 occur in the FG domain at the strain of 2%, as seen in Figs. 9a-c. With further increasing
53
54 10 strain up to 4%, as shown in Fig. 9d, dislocation tangles and local clusters are formed.
55
56 11 In addition, dislocation pile-ups at GBs and ATBs are visible, demonstrating that GBs
57
58
59 12 and ATBs effectively inhibit dislocation motion. At the strain of 10%, extended SFs on
60
61 22
62
63
64
65
1 the primary {111} planes are occasionally observed in some FGs that contain ATs, as
1
2 2 shown in Figs. 9e and f. Furthermore, as seen by white dotted lines in Fig. 9e, a
3
4
5 3 dislocation cell in its infancy is also observed. This is in line with the dislocation cell
6
7 4 structures observed in EBSD analysis shown in Figs. 7c and e, wherein the walls of
8
9
10 5 cells exhibit high KAM values. As for the UFG domain, dislocation pile-ups at GBs are
11
12 6 observed at the strain of 2%, as shown in Fig. 10a. At the strain of 4%, the deformed
13
14
15 7 microstructure consists of dislocations, a few deformation-induced nano twins (NTs)
16
17 8 shown in Fig. 10b and numerous of SFs shown in Fig. 10d. HRTEM characterization
18
19
9 combined with the corresponding FFT pattern shown in Fig. 10c further verifies the
20
21
22 10 appearance of NTs. As the strain is further increased to 10%, increasing planar NTs and
23
24 11 SFs are formed while their densities seem to be stable, as demonstrated in Figs. 10e and
25
26
27 12 f.
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 13
59
60
61 23
62
63
64
65
1 Fig. 9. TEM micrographs showing the microstructural evolution of the FG domain in HG-NP
1
2 2 sample. (a-c) ε = 2%. (d) ε = 4%. (e-f) ε = 10%. The insets in (a) and (f) are two-beam SAED
3
4 3 patterns, while in (d) and (e) they are SAED patterns corresponding to the circled regions in each
5
6 4 image. (b-c) Magnified DF-TEM images of the squared region in (a). (f) Magnified DF-TEM
7
8 5 image of the squared region in (e), where green arrows point to SFs.
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 6
35
36 7 Fig. 10. TEM micrographs showing the microstructural evolution of the UFG domain in HG-NP
37
38 8 sample. (a) ε = 2%. (b-d) ε = 4%. (e-f) ε = 10%. The insets in (b) and (f) are SAED patterns
39
40 9 corresponding to the circled regions in each image. (c) HRTEM image of the circled region in (b).
41
42
10 (f) Magnified image of the squared region in (e).
43
44
45 11 4. Discussions
46
47 12 We successfully introduced a fully recrystallized HG microstructure with coherent
48
49
50 13 L12 nanoprecipitates in the Ni42Fe30Cr12Mn8Al5Ti3 CCA. This unique HG-NP structure
51
52 14 makes the alloy with ultrahigh strength and good ductility at ambient temperature.
53
54
15 Therefore, it is of great importance to clarify the formation of the HG-NP
55
56
57 16 microstructure and the underlying mechanisms for the exceptional strength-ductility
58
59 17 synergy, as discussed below.
60
61 24
62
63
64
65
1 4.1 Formation of the HG-NP microstructure
1
2 2 In order to clarify how the HG-NP microstructure is developed from the CR
3
4
5 3 sample, the CR sample and the CR sample annealed at 650 °C for 30 min were studied
6
7 4 by EBSD characterization, as shown in Figs. 11a and b. Severely deformed and
8
9
10 5 elongated grains along rolling direction and numerous shear bands are observed in the
11
12 6 CR sample, as shown in Fig. 11a. These shear bands are the result of nonuniform
13
14
15 7 deformation during CR process, which induces highly localized strain and dense defects
16
17 8 inside the bands (Gwalani et al., 2018; Lins et al., 2007; Su et al., 2019). Benefiting
18
19
9 from the high strain energy stored in shear bands, recrystallization tendency there is
20
21
22 10 stronger than the regions outside shear bands. Upon annealing, recrystallization occurs
23
24 11 preferentially inside shear bands (Gwalani et al., 2018; Su et al., 2019). This is verified
25
26
27 12 by the EBSD analysis on the CR sample annealed at 650 °C for 30 min, as shown in
28
29 13 Fig. 11b. Obviously, full recrystallization occurs in shear bands, leading to the
30
31
32 14 formation of dense UFGs. Outside shear bands, the grains are only partially
33
34 15 recrystallized. With increasing annealing time, the dense recrystallized grains inside
35
36
16 shear bands grow by consuming neighboring strained grains until they impinge on one
37
38
39 17 another. Meanwhile, outside shear bands recrystallization occurs at relatively low rate.
40
41 18 In this way, fully recrystallized HG structure, characterized by UFG bands and FGs, is
42
43
44 19 developed in the HG-NP sample, as revealed by the partitioned EBSD IPF maps in Figs.
45
46 20 11c and d.
47
48
49 21 In addition to recrystallized FCC grains, high-density L12 nanoprecipitates are
50
51 22 formed within grains and some L21 particles are formed at GBs and triple-junctions in
52
53
54 23 the HG-NP sample. As described in our earlier work (Fan et al., 2022), addition of 5
55
56 24 at.% Al and 3 at.% Ti in the Ni42Fe30Cr12Mn8Al5Ti3 alloy allows for the formation of
57
58
25 L12 precipitates. The L12 phase in the HG-NP sample is formed with two states, i.e.,
59
60
61 25
62
63
64
65
1 nanorods and spheroids formed via DP and CP mechanisms (Zhao et al., 2017),
1
2 2 respectively. There exists a competitive relationship between DP and CP, and DP is
3
4
5 3 essentially controlled by interfacial migration (i.e., GB diffusion) (Bała et al., 2020;
6
7 4 Gwalani et al., 2021; Yang et al., 2020). Furthermore, diffusion along GBs at 600-
8
9
10 5 800 °C is substantially faster than that inside grains, facilitating the discontinuous
11
12 6 growth of L12 nanorods (Li et al., 2021a; Yang et al., 2020). Owing to the relatively
13
14
15 7 higher density of GBs in the HG-NP and FG-NP samples, DP is the dominant
16
17 8 mechanism for the formation of L12 precipitates in both samples. The content of L12
18
19
9 nanoprecipitates in the HG-NP sample is close to that in the FG-NP sample, as listed in
20
21
22 10 Table 1, which is in full agreement with the thermodynamic calculation shown in Fig.
23
24 11 S8. The actual reason is that the volume fraction of precipitates is mostly determined
25
26
27 12 by the content of phase-forming elements, e.g., Ni, Al and Ti for L12 phase in the present
28
29 13 case (Yan et al., 2021).
30
31
32 14 As for the L21 particles, which keep K-S orientation relationship with FCC matrix,
33
34 15 are formed primarily at GBs and triple-junctions in UFG domains, as depicted in Figs.
35
36
37
16 1f and g. The FCC/L12 interfacial energy is commonly as low as 10~40 mJ·m-2 (He et
38
39 17 al., 2016). In contrast, the interfacial energy between FCC and BCC phases is generally
40
41 18 in the range of 180~660 mJ·m-2 due to their distinct crystal structures (Gwalani et al.,
42
43
44 19 2018). Thus, the barrier for BCC phase nucleation in FCC matrix should be much
45
46 20 higher than that for the L12 phase. As a result, L21 particles prefer to nucleate at GBs
47
48
49 21 and triple-junctions. Consequently, with the aid of strain energy, the formation of L21
50
51 22 particles at GBs and triple-junctions prevails in the HG-NP and FG-NP samples. It has
52
53
54 23 to be noted that UFG domains contain majority of L21 particles due to the pre-existing
55
56 24 higher density of dislocations and other defects inside shear bands. Additionally, the
57
58
25 formation of both L12 and L21 particles can effectively inhibit the growth of
59
60
61 26
62
63
64
65
1 recrystallized grains due to Zenner pinning effect (Li et al., 2021a; Zhang et al., 2019).
1
2 2 In the end, UFGs are formed inside shear bands and FGs are formed outsides.
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 3
37
38 4 Fig. 11. (a) EBSD IPF map of CR sample. (b) IPF map of the sample annealed at 650 °C for 30
39
40
5 min. (c-d) Partitioned IPF map (corresponding to Fig. 2d) of the HG-NP sample.
41
42
43 6 A schematic diagram for the PH-HG structure, which is a composite of
44
45 7 recrystallized and L12-nanoprecipitate-hardened FGs and UFGs with a small number
46
47
48 8 of ultrafine L21 particles at GBs and triple-junctions, is shown in Fig. 12a. FG and UFG
49
50 9 lamellar domains are alternately stacked, as shown in Figs. 1c-e. Thus, it provides more
51
52
10 effective mutual constraint and more domain boundaries between the soft and hard
53
54
55 11 domains than the randomly distributed one (Wu et al., 2015), which is more effective
56
57 12 in inducing HDI hardening.
58
59
60
61 27
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 1
19
20 2 Fig. 12. Schematic diagram of the initial microstructure (a), and evolution of defects upon tensile
21
22
23 3 deformation at room temperature (b) of the HG-NP sample.
24
25 4 4.2 Multiple strengthening contributions to the yield strength
26
27 5 The strengths of the FG and UFG domains in the HG-NP sample are separately
28
29
30 6 evaluated and then substituted into the general mix rule to quantitatively calculate the
31
32 7 YS (σy), expressed as:
33
34
35 8 𝜎𝑦 = 𝜎𝑦−𝐹𝐺 𝑓𝐹𝐺 + 𝜎𝑦−𝑈𝐹𝐺 𝑓𝑈𝐹𝐺 + 𝜎𝑠𝑦𝑛 (1)
36
37 9 where σy-FG and σy-UFG are the individual YS of FG and UFG domains. σsyn is the extra
38
39
40 10 synergistic strengthening, i.e., extra HDI strengthening, which is usually taken into
41
42 11 account for estimating strength of HG materials (Liu et al., 2020; Ma and Zhu, 2017;
43
44
45 12 Wang et al., 2019). σy-FG and σy-UFG can be evaluated on the basis of the contributions
46
47 13 from individual strengthening mechanisms, expressed as:
48
49
50 14 𝜎𝑦−𝐹𝐺,𝑈𝐹𝐺 = 𝜎0 + 𝜎𝑔 + 𝜎𝑃−𝐿12 + 𝜎𝑃−𝐿12 (2)
51
52 15 where σ0 is the basic strength relating to lattice friction stress and solid-solution
53
54
55 16 strengthening, and the reported value of 132.8 MPa is adopted (Fan et al., 2022). σg, σp-
56
57 17 L12 and σp-L21 are the GB strengthening, and precipitates strengthening resulting from
58
59
18 L12 and L21 particles, respectively.
60
61 28
62
63
64
65
1 The GB strengthening is commonly predicted by the Hall-Petch equation (Nutor
1
2 2 Raymond et al., 2021), as expressed by Eq. (3). Herein we calculate the GB contribution
3
4
5 3 by excluding L21 particles, which is given by Eq. (4).
6
⁄2
7
8
4 σg = ky d-1 (3)
9

10 5 σg = ky d-1 2 (1 − 𝑓𝐿21 ) (4)
11
12
13 6 where ky = 514.4 MPa∙μm1/2 is the Hall-Petch slop (Fan et al., 2022), d stands for the
14
15 7 average grain size of FCC phase (see Table 1). Since there are a lot of ATs in the
16
17
18 8 presently studied samples, the average grain size is evaluated by taking into account of
19
20 9 ATBs (Jang et al., 2021).
21
22
23 10 As for the L12 nanoprecipitates, the particle shearing mechanism is confirmed
24
25 11 directly by TEM observation, as shown in Fig. 9b, consisting with most of previously
26
27
28 12 studied L12-hardened CCAs (Liang et al., 2018; Ming et al., 2018). The strength
29
30 13 enhancement caused by shearing ordered coherent nanoparticles is calculated by (Ming
31
32 14 et al., 2018):
33
34
35 𝛾𝑎𝑝𝑏 3𝜋𝑓 1⁄2
15 𝜎𝑜𝑠 = 0.81𝑀 ( )( ) (5)
36 2𝑏 8
37
38 16 where M = 3.06 is the Taylor factor for a FCC matrix, f is the volume fraction of
39
40
41 17 precipitates (see Table S2), b = (√2/2)⋅a is the magnitude of the Burgers vector of the
42
43 18 FCC matrix and a is the lattice constant (0.3600 nm, obtained by XRD analysis) of the
44
45
46 19 matrix. The anti-phase boundary energy of Ni3(Al, Ti) particles, apb, depends strongly
47
48 20 on the Ti/(Ti+Al) ratio in the particles (Gwalani et al., 2021). It was estimated to be
49
50
51 21 ~130 mJ/m2 based on the composition of L12 shown in Table S1. Kelly (Kelly, 1972)
52
53 22 showed that the strengthening by rod-like particles is about 1.75 times the strengthening
54
55
56 23 by spheroidal particles. Taking into account the different shapes of L12 precipitates,
57
58 24 therefore, a simple rule of mixture is applied to calculate the strengthening contribution
59
60
61 29
62
63
64
65
1 from L12 precipitates, as described by:
1
2
2 𝜎𝑃−𝐿12 = 𝜎𝑜𝑠−𝑠𝑝 + 1.75𝜎𝑜𝑠−𝑟𝑝 (6)
3
4
5 3 where os-sp is the strengthening from spherical precipitates and os-rp stands for the
6
7
8 4 contribution calculated by assuming the rode-like precipitates as spherical ones.
9
10 5 As for the L21 particles with mean radius over 100 nm, Orowan looping is
11
12
6 expected to dominate and can be expressed by the Ashby-Orowan relationship (Qi et
13
14
15 7 al., 2020):
16
17 𝐺𝑏 2𝜆
18 8 𝜎𝑝𝐿21 = 0.4𝑀 ( ) ln ( 𝑏 ) (7)
𝜋𝐿√(1−𝑣)
19
20
21 9 where G = 87.5 GPa is the shear modulus adopted from the Ni50Cr25Fe25 CCA (Lu et
22
23 10 al., 2022). v = 0.31 is the Poisson ration derived from the (FeCoNiCr)94Ti2Al4 CCA (He
24
25
26 11 et al., 2016). λ is the radius of the sphere particles on slip planes and L is the inter-
27
28
29
12 precipitate distance, respectively, which are calculated by (Qi et al., 2020):
30
31
32 13 𝜆 = 𝑟 × √2/3 (8)
33
34
3𝜋
35 14 𝐿 = 𝑟 × (√4𝑓 − 1.64) (9)
36
37
38
15 where r and f are the mean precipitate radius and volume fraction of the L21 particles
39
40
41 16 (see Table 1). Using Eqs. (3-9), the strengthening contributions corresponding to
42
43 17 individual domains are calculated and listed in Table S4. The values of σsyn are obtained
44
45
46 18 by subtracting the strengthening contributions in FG and UFG domains from Eq. (1),
47
48 19 which are approx. 210 MPa and -29 MPa for the HG-NP and FG-NP samples,
49
50
51 20 respectively. Fig. 13a gives the integrated distribution maps of individual strengthening
52
53 21 mechanisms of the HG-NP and FG-NP samples. In comparison to the FG-NP sample,
54
55
56 22 the major enhanced component in the YS of the HG-NP sample is GB strengthening
57
58 23 and the extra synergistic strengthening, which account for approx. 40% and 41%,
59
60
61 30
62
63
64
65
1 respectively.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 2
20
21 3 Fig. 13. (a) Calculated and experimental YS (σy, experimental) of the HG-NP and FG-NP samples. (b)
22
23 4 Calculated and experimental HDI stresses of the HG-NP sample.
24
25 5 4.3 Strain partitioning and HDI hardening
26
27
28 6 In this work, the strain distributions between different domains were qualitatively
29
30 7 evaluated by the strain contouring (SC) map, which is usually used to estimate the
31
32
33 8 deformation or strain in individual grains (Sinha et al., 2015). In the HG-NP sample,
34
35 9 plastic strain incompatibility (He et al., 2021a; Ma and Zhu, 2017), i.e., strain
36
37
38 10 partitioning occurs during deformation due to the distinct nature of different domains,
39
40 11 inducing a strong strain gradient, as shown in Fig. 14a. In contrast, a relatively
41
42
12 homogeneously-grained structure is shown in the FG-NP sample. Thus, a more
43
44
45 13 homogeneous distribution of strain can be observed in Fig. 14b. The statistics shown in
46
47 14 Figs. 14c and d reveal that the HG-NP sample reveals more heterogeneous strain
48
49
50 15 distributions between the FG and UFG domains and a larger average value of maximum
51
52 16 orientation than that of the FG-NP sample. This further indicates that the stronger strain
53
54
55 17 partitioning is established in the HG-NP sample. Accordingly, numerous GNDs are
56
57 18 generated near domain boundaries, resulting in deformation-induced long-range
58
59
19 internal stress, i.e., HDI stress to accommodate structural compatibility (Zhu and Wu,
60
61 31
62
63
64
65
1 2019).
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 2
35
36 3 Fig. 14. The strain contouring maps of the fractured HG-NP (a) and FG-NP (b) samples,
37
38
4 and their corresponding statistic distributions (c). The corresponding statistic distributions in the
39
40
41 5 FG and UFG domains of the HG-NP sample (d).
42
43 6 It is generally accepted that the density of GNDs correlates with the HDI stress as
44
45 7 well as the degree of strain partitioning (Zhong et al., 2022). HDI stress is caused by
46
47
48 8 both the intragranular and intergranular plastic heterogeneities (Keller et al., 2022;
49
50 9 Sauzay, 2008). Thus, there are components contributing to HDI stress: the intergranular
51
52
53 10 GND accumulation and the intragranular GND pile-up against GBs, TBs, precipitates,
54
55 11 SFs and dislocation cells (Li et al., 2021a; Li et al., 2017; Muhammad et al., 2017; Tsuji
56
57
12 et al., 2020). In this work, a simplified HDI stress equation (Zhong et al., 2022) is used
58
59
60 13 to evaluate the individual contribution to HDI stress σHDI:
61 32
62
63
64
65
1 𝜎𝐻𝐷𝐼 = 𝜎𝑔 + ∑𝑖 𝑓𝑚 𝑓𝑖 𝛼𝑀𝐺𝑏√𝜌𝑖−𝐺𝑁𝐷 (10)
1
2
3 2 where fi and ρi-GND are volume fractions and GND densities of heterogeneous domains
4
5 3 (FG and UFG domains). The hardening component of each domain fm = 0.65 is adopted
6
7
8 4 from the normal distribution of strain (Fig. 14d). Here, the first component at the right-
9
10 5 hand site of Eq. (10) stands for the intergranular GNDs strengthening, which remains
11
12
6 constant during straining. The second component represents the intragranular GNDs
13
14
15 7 strengthening from intragranular GNDs in FG and UFG domains.
16
17 8 The local misorientation induced by crystal lattice distortion, which is correlated
18
19
20 9 with GNDs, can be illustrated by the KAM analysis (Kadkhodapour et al., 2011). The
21
22 10 relationship between the ρi-GND and the KAM value is simply related by the strain
23
24
25 11 gradient model (Li et al., 2021a) as:
26
27 2𝜃𝑖
12 𝜌𝑖−𝐺𝑁𝐷 = (11)
28 𝑙𝑏
29
30 13 where θi is the average KAM value in each domain, l = 80 nm is the unit length. The
31
32
33 14 calculated densities of GNDs in each domain at various true strains are listed in Table
34
35 15 2. The individual strengthening contribution to the overall HDI stresses of the HG-NP
36
37
16 sample is displayed in Fig. 13b. Obviously, the strong HDI stress hardening in the HG-
38
39
40 17 NP sample can be ascribed to the synergistic effects of GNDs resulted from neighboring
41
42 18 grains and heterogeneous domains. Generally, the density of GNDs stored in UFGs is
43
44
45 19 much higher than that in FGs during deformation, since the average slip distance of
46
47 20 GNDs is much smaller than that of statistically stored dislocations (Kubin and
48
49
50 21 Mortensen, 2003; Liu et al., 2020). However, the contribution to HDI stress from FG
51
52 22 domains is comparable to that from UFG domains. This can be mainly attributed to the
53
54
55 23 constrained effect from hard UFGs, which strengthens FGs by inducing large amount
56
57 24 of GNDs in the FGs adjacent to the domain boundaries, as seen in Fig. 7d. Overall,
58
59
25 sustainable high HDI stress and strong HDI hardening are developed over the whole
60
61 33
62
63
64
65
1 deformation stage of the HG-NP sample.
1
2 2 4.4 The heterostructure-controlled deformation mechanisms to the strength-
3
4
5 3 ductility synergy
6
7 4 Another exceptional feature of the developed HG-NP sample is that a promising
8
9
10 5 UE of 18.2% at room temperature is achieved even at the ultrahigh YS level. This
11
12 6 behavior originates from the substantial strain-hardening capacity associated with the
13
14
15 7 heterogeneous microstructure. As mentioned above, the exceptional strain-hardening
16
17 8 capability of the HG-NP sample is closely related to the HDI hardening caused by the
18
19
9 heterogeneous-deformation induced strain gradient. To elucidate this, the strain-
20
21
22 10 hardening behavior corelated with the underlying deformation mechanism is discussed
23
24 11 based on the synergistic effect of the deformed substructures and stress/strain
25
26
27 12 partitioning owing to the plastic strain incompatibility.
28
29 13 The evolution of the deformed substructures in different domains are
30
31
32 14 schematically summarized in Fig. 12b. At the early plastic deformation stage (true strain
33
34 15 <= 6%), the deformed FGs are characterized by plenty of dislocations. In UFGs, in
35
36
16 addition to the more densely accumulated dislocations, numerous SFs and NTs are
37
38
39 17 formed. Combining the EBSD analysis shown in Fig. 7, it can be clarified that the
40
41 18 following three factors result in the rapid increase in HDI stress (Fig. 6c) of the HG-NP
42
43
44 19 sample at the early deformation stage: (i) GNDs steadily accumulate near hot spots like
45
46 20 GBs, triple junctions and ATBs; (ii) deformation-induced SFs and NTs within UFGs
47
48
49 21 hinder the subsequent dislocation slip by shortening the mean free path of dislocations
50
51 22 (Li et al., 2021a); and (iii) high strain gradient around domain boundaries helps the
52
53
54 23 densification of GNDs. The sharp increasing stage of HDI stress just corresponds to the
55
56 24 stage where SHRs increase, as seen in Fig. 5c. When the true strain exceeds 6%,
57
58
25 dislocation pile-up near GBs and ATBs gradually saturates as strain increases. On the
59
60
61 34
62
63
64
65
1 other hand, the mechanical incompatibility between FG and UFG domains is reduced
1
2 2 with continuous deformation, as evidenced by the variation of KAM values shown in
3
4
5 3 Figs. 7d and e. Moreover, dislocation multiplication is slowed down and dislocation
6
7 4 mobility is decreased upon straining as the suppression effect from HDI stress.
8
9
10 5 Therefore, the deformation of the HG-NP sample gradually turns into less
11
12 6 heterogeneous, leading to the gradually saturated HDI stress, as seen in Fig. 6c. These
13
14
15 7 overall effects result in the declined SHRs in stage III shown in Fig. 5c.
16
17 8 The stress partitioning between the FG and UFG domains during deformation
18
19
9 could influence the activation of deformation modes in different domains. As shown in
20
21
22 10 Fig. 14d, the average value of the maximum orientation in UFG domains is higher than
23
24 11 that in FG domains, demonstrating that higher partitioned stress is induced in UFG
25
26
27 12 domains. This localized partitioned stress favors the nucleation of SFs and NTs in UFGs
28
29 13 (Fig. 10) once it reaches a critical level (He et al., 2017). However, owing to the multi-
30
31
32 14 axial stress state induced by the surrounding hard UFGs, dislocation clusters and tangles
33
34 15 prevail in FGs (Fig. 9). The activation of multiple deformation modes in UFGs,
35
36
16 including dislocations, SFs and NTs, as well as interactions between them, help increase
37
38
39 17 the strain-hardening capability. This is consistent with the plaston concept proposed by
40
41 18 Tsuji et al., that different deformation modes could be activated for maintaining the
42
43
44 19 strain-hardening capability of the fully recrystallized UFG alloys (Tsuji et al., 2020).
45
46 20 This along with the precipitation hardening by L12 nanoparticles compensate the
47
48
49 21 diminished strain hardening caused by grain refinement in UFG domains. Therefore,
50
51 22 the HG-NP sample exhibits similar strain-hardening behavior with the FG-NP sample
52
53
54 23 (Fig. 5c), yet with significant enhanced strength-ductility synergy. Moreover, the
55
56 24 coherency interfaces between L12 precipitates and FCC matrix are also beneficial to the
57
58
25 good ductility of the HG-NP sample. This is due to the fact that dislocations can pass
59
60
61 35
62
63
64
65
1 through the FCC/L12 interfaces, thus reduce dislocation aggregation and stress
1
2 2 concentration near the interfaces (Gwalani et al., 2021; Jiang et al., 2017; Liang et al.,
3
4
5 3 2018). Premature crack initiation is therefore effectively delayed.
6
7 4 A direct comparison of YS versus UE of the current CCA with other reported FCC
8
9
10 5 CCAs is shown in Fig. 15a. Apparently, the current HG-NP CCA appears to have an
11
12 6 outstanding strength-ductility synergy and to be comparable to that of recently
13
14
15 7 developed ultrastrong and ductile CCAs, even though it maintains low raw material
16
17 8 cost, as shown in Fig. 15b. In particular, our HG-NP CCA far surpasses previously
18
19
9 reported FCC-type Co-free CCAs in its exceptional combination between YS and UE,
20
21
22 10 as indicated in red symbols in Fig. 15a. This suggests the advantage of the HG-NP
23
24 11 microstructural design strategy for substantially enhancing the synergic mechanical
25
26
27 12 properties of Co-free CCAs.
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 13
60
61 36
62
63
64
65
1 Fig. 15. (a) YS-UE profiles of the present HG-NP CCA compared with those of previously
1
2 2 reported FCC-type CCAs that with various microstructures, including single FCC-phased solid
3
4 3 solutions (SS), HG, CG/FG/UFG-NP (NP) and HG-NP microstructures. (b) Comparison of raw-
5
6 4 material cost between the present CCA and other CCAs underlined with red dot lines in (a) with
7
8 5 high YS (≥1 GPa). The tensile properties and raw-material cost of reported CCAs were acquired
9
10
6 from Table S5.
11
12
13 7 5. Conclusions
14
15 8 In summary, we developed a Co-free CCA with exceptional strength-ductility
16
17
18 9 combination. The microstructural features, deformation mechanisms and tensile
19
20 10 properties at room temperature were extensively investigated. The following
21
22
11 conclusions are drawn:
23
24
25 12 (1) By employing a facile TMP method, a unique fully recrystallized HG-NP structure
26
27 13 consisting of alternately stacked FG and UFG lamellar domains, with coherent L12
28
29
30 14 nanoparticles precipitated in FGs and UFGs and L21 particles precipitated at GBs,
31
32 15 is introduced into the Co-free Ni42Fe30Cr12Mn8Al5Ti3 CCA. Deformation-induced
33
34
35 16 shear bands in the cold-rolled state mainly contribute to the formation of the HG
36
37 17 structure. The processing strategy described here is relatively simple and desirable
38
39
40 18 for large scale process in modern industry at an economic cost.
41
42 19 (2) The HG-NP structure leads to a great combination of ultrahigh YS of 1.5 GPa and
43
44 20 UTS of 1.8 GPa while retaining a decent UE of 18.2%. Specially, the product of
45
46
47 21 UTS×UE for the HG-NP CCA is 32.7 GPa%, superior to that of the homogeneous-
48
49 22 grained FG-NP, CG-NP counterparts, and recently reported partially recrystallized
50
51
52 23 HG-NP CCAs. This indicates the strong effectiveness of the fully recrystallized
53
54 24 HG-NP structure in providing ultrahigh strength and ductility combination for
55
56
57 25 CCAs. The ultrahigh YS of the HG-NP structure is mainly attributed to
58
59 26 precipitation strengthening, GB strengthening and extra HDI strengthening, which
60
61 37
62
63
64
65
1 are ~43%, ~30% and ~14% of the YS, respectively.
1
2 2 (3) The plastic deformation of the HG-NP structure is dominated by dislocation slip
3
4
5 3 within FGs, while dislocations, SFs and NTs dominate the plastic deformation of
6
7 4 UFGs. During deformation, the accumulation of dislocations at GBs and
8
9
10 5 deformation-induced SFs, NTs and dislocation cells, leads to a large amount of
11
12 6 GNDs. A strong stress/strain partitioning between FG and UFG domains results in
13
14
15 7 the generation of SFs and NTs in UFGs, and substantial pile-ups of GNDs at domain
16
17 8 boundaries. As a consequence, a sustainable strong HDI hardening is obtained.
18
19
9 Along with forest dislocation hardening and precipitation hardening, HDI
20
21
22 10 hardening contributes to the good strain hardening capacity and thus leads to the
23
24 11 exceptional strength-ductility synergy in the current HG-NP CCA.
25
26
27 12 CRediT authorship contribution statement
28
29 13 Aidang Shan and Liming Fu designed this study; Jiantao Fan, Xinbo Ji, Jian Wang
30
31
32 14 and Shuo Ma performed the research; Jiantao Fan and Yanle Sun analyzed the data, and
33
34 15 Jiantao Fan wrote the paper; Mao Wen, Liming Fu and Aidang Shan reviewed and
35
36
16 edited the paper. All authors discussed the results and commented on the manuscript.
37
38
39 17 Acknowledgements
40
41 18 This work was supported by the National Natural Science Foundation of China
42
43
44 19 (No. 52071212 & 52001201). The Young Researchers Startup Fund for Youngman
45
46 20 Research at Shanghai Jiao Tong University (SFYR at SJTU: No. 18X100040023) also
47
48
49 21 financially supported this work.
50
51 22 Appendix.
52
53
54 23 The abbreviations of academic words used in this paper.
55
56 Abbreviation Full name
57
58 AT Annealing twin
59
60
61 38
62
63
64
65
CG Coarse-grained
1
2 CR Cold rolling
3
4 EBSD Electron backscattering diffracting
5
6 EDX Energy dispersive X-ray spectrometer
7
8
FG Fine-grained
9
10
11 GB Grain boundary
12
13 GND Geometrically necessary dislocation
14
15 KAM Kernel average misorientation
16
17 HG Heterogeneously-grained
18
19 NP Nanoprecipitate
20
21 NT Nano twin
22
23 SAED Selected area electron diffractometer
24
25
SEM Scanning electron microscopy
26
27
28 SHR Strain-hardening rate
29
30 STEM Scanning transmission electron microscopy
31
32 UE Uniform elongation
33
34 UFG Ultrafine-grained
35
36 UTS Ultimate tensile strength
37
38 XRD X-ray diffraction
39
40
SC Strain contouring
41
42
43 SEM Scanning electron microscopy
44
45 SF Stacking fault
46
47 TEM Transmission electron microscopy
48
49 TE Total elongation
50
51 TKD Transmission Kikuchi diffraction
52
53
54
YS Yield strength
55
56
57
58
59
60
61 39
62
63
64
65
1 References
1
2 2 Bała, P., Górecki, K., Bednarczyk, W., Wątroba, M., Lech, S., Kawałko, J., 2020. Effect
3
4
5 3 of high-temperature exposure on the microstructure and mechanical properties of the
6 4 Al5Ti5Co35Ni35Fe20 high-entropy alloy. J. Mater. Res. Technol. 9, 551-559.
7
8 5 https://doi.org/10.1016/j.jmrt.2019.10.084
9
10 6 Dasari, S., Sharma, A., Byers, T.A., Glass, G.A., Srivilliputhur, S., Rout, B., Banerjee,
11
12 7 R., 2021. Proton irradiation induced chemical ordering in an Al0.3CoFeNi high entropy
13
14 8 alloy. Appl. Phys. Lett. 119, 161907. https://doi.org/10.1063/5.0065875
15
16
9 Du, X.H., Li, W.P., Chang, H.T., Yang, T., Duan, G.S., Wu, B.L., Huang, J.C., Chen,
17 10 F.R., Liu, C.T., Chuang, W.S., Lu, Y., Sui, M.L., Huang, E.W., 2020. Dual
18
19 11 heterogeneous structures lead to ultrahigh strength and uniform ductility in a Co-Cr-Ni
20
21 12 medium-entropy alloy. Nat. Commun. 11, 2390. https://doi.org/10.1038/s41467-020-
22
23 13 16085-z
24
25 14 Fan, J., Fu, L., Sun, Y., Xu, F., Ding, Y., Wen, M., Shan, A., 2022. Unveiling the
26
15 precipitation behavior and mechanical properties of Co-free Ni47-xFe30Cr12Mn8AlxTi3
27
28 16 high-entropy alloys. J. Mater. Sci. Technol. 118, 25-34.
29
30 17 https://doi.org/10.1016/j.jmst.2021.11.058
31
32 18 Fan, L., Yang, T., Zhao, Y., Luan, J., Zhou, G., Wang, H., Jiao, Z., Liu, C.-T., 2020.
33
34 19 Ultrahigh strength and ductility in newly developed materials with coherent
35
36 20 nanolamellar architectures. Nat. Commun. 11, 6240. https://doi.org/10.1038/s41467-
37
21 020-20109-z
38
39 22 Fu, W., Huang, Y., Sun, J., Ngan, A.H.W., 2022. Strengthening
40
41 23 CrFeCoNiMn0.75Cu0.25 high entropy alloy via laser shock peening. Int. J. Plast. 154,
42
43 24 103296. https://doi.org/10.1016/j.ijplas.2022.103296
44
45 25 Gwalani, B., Dasari, S., Sharma, A., Soni, V., Shukla, S., Jagetia, A., Agrawal, P.,
46
47 26 Mishra, R.S., Banerjee, R., 2021. High density of strong yet deformable intermetallic
48 27 nanorods leads to an excellent room temperature strength-ductility combination in a
49
50 28 high entropy alloy. Acta Mater. 219, 117234.
51
52 29 https://doi.org/10.1016/j.actamat.2021.117234
53
54 30 Gwalani, B., Gorsse, S., Choudhuri, D., Styles, M., Zheng, Y., Mishra, R.S., Banerjee,
55
56 31 R., 2018. Modifying transformation pathways in high entropy alloys or complex
57
58 32 concentrated alloys via thermo-mechanical processing. Acta Mater. 153, 169-185.
59 33 https://doi.org/10.1016/j.actamat.2018.05.009
60
61 40
62
63
64
65
1 Han, L., Rao, Z., Souza Filho, I.R., Maccari, F., Wei, Y., Wu, G., Ahmadian, A., Zhou,
1
2 2 X., Gutfleisch, O., Ponge, D., Raabe, D., Li, Z., 2021. Ultrastrong and Ductile Soft
3
4 3 Magnetic High-Entropy Alloys via Coherent Ordered Nanoprecipitates. Adv. Mater.,
5
4 2102139. https://doi.org/10.1002/adma.202102139
6
7 5 Hasan, M.N., Liu, Y.F., An, X.H., Gu, J., Song, M., Cao, Y., Li, Y.S., Zhu, Y.T., Liao,
8
9 6 X.Z., 2019. Simultaneously enhancing strength and ductility of a high-entropy alloy via
10
11 7 gradient hierarchical microstructures. Int. J. Plast.
12
13 8 https://doi.org/10.1016/j.ijplas.2019.07.017
14
15 9 He, B.B., Hu, B., Yen, H.W., Cheng, G.J., Wang, Z.K., Luo, H.W., Huang, M.X., 2017.
16 10 High dislocation density–induced large ductility in deformed and partitioned steels.
17
18 11 Science 357, 1029-1032. https://doi.org/10.1126/science.aan0177
19
20 12 He, F., Chen, D., Han, B., Wu, Q., Wang, Z., Wei, S., Wei, D., Wang, J., Liu, C.T., Kai,
21
22 13 J.-j., 2019. Design of D022 superlattice with superior strengthening effect in high
23
24 14 entropy alloys. Acta Mater. 167, 275-286.
25
26 15 https://doi.org/10.1016/j.actamat.2019.01.048
27 16 He, F., Yang, Z., Liu, S., Chen, D., Lin, W., Yang, T., Wei, D., Wang, Z., Wang, J., Kai,
28
29 17 J.-j., 2021a. Strain partitioning enables excellent tensile ductility in precipitated
30
31 18 heterogeneous high-entropy alloys with gigapascal yield strength. Int. J. Plast. 144,
32
33 19 103022. https://doi.org/10.1016/j.ijplas.2021.103022
34
35 20 He, J.Y., Wang, H., Huang, H.L., Xu, X.D., Chen, M.W., Wu, Y., Liu, X.J., Nieh, T.G.,
36
37
21 An, K., Lu, Z.P., 2016. A precipitation-hardened high-entropy alloy with outstanding
38 22 tensile properties. Acta Mater. 102, 187-196.
39
40 23 https://doi.org/10.1016/j.actamat.2015.08.076
41
42 24 He, Z., Jia, N., Yan, H., Shen, Y., Zhu, M., Guan, X., Zhao, X., Jin, S., Sha, G., Zhu, Y.,
43
44 25 Liu, C.T., 2021b. Multi-heterostructure and mechanical properties of N-doped
45
46 26 FeMnCoCr high entropy alloy. Int. J. Plast. 139, 102965.
47
27 https://doi.org/10.1016/j.ijplas.2021.102965
48
49 28 Jang, T.J., Choi, W.S., Kim, D.W., Choi, G., Jun, H., Ferrari, A., Körmann, F., Choi, P.-
50
51 29 P., Sohn, S.S., 2021. Shear band-driven precipitate dispersion for ultrastrong ductile
52
53 30 medium-entropy alloys. Nat. Commun. 12, 4703. https://doi.org/10.1038/s41467-021-
54
55 31 25031-6
56
57 32 Jiang, S., Wang, H., Wu, Y., Liu, X., Chen, H., Yao, M., Gault, B., Ponge, D., Raabe,
58
33 D., Hirata, A., Chen, M., Wang, Y., Lu, Z., 2017. Ultrastrong steel via minimal lattice
59
60
61 41
62
63
64
65
1 misfit and high-density nanoprecipitation. Nature 544, 460-464.
1
2 2 https://doi.org/10.1038/nature22032
3
4 3 Kadkhodapour, J., Schmauder, S., Raabe, D., Ziaei-Rad, S., Weber, U., Calcagnotto, M.,
5
4 2011. Experimental and numerical study on geometrically necessary dislocations and
6
7 5 non-homogeneous mechanical properties of the ferrite phase in dual phase steels. Acta
8
9 6 Mater. 59, 4387-4394. https://doi.org/10.1016/j.actamat.2011.03.062
10
11 7 Kai, W., Li, C.C., Cheng, F.P., Chu, K.P., Huang, R.T., Tsay, L.W., Kai, J.J., 2017. Air-
12
13 8 oxidation of FeCoNiCr-based quinary high-entropy alloys at 700–900°C. Corros. Sci.
14
15 9 121, 116-125. http://dx.doi.org/10.1016/j.corsci.2017.02.008
16 10 Keller, C., Calvat, M., Flipon, B., Barbe, F., 2022. Experimental and numerical
17
18 11 investigations of plastic strain mechanisms of AISI 316L alloys with bimodal grain size
19
20 12 distribution. Int. J. Plast. 153, 103246. https://doi.org/10.1016/j.ijplas.2022.103246
21
22 13 Kelly, P., 1972. The effect of particle shape on dispersion hardening. Scr. Metall. 6, 647-
23
24 14 656. https://doi.org/10.1016/0036-9748(72)90120-2
25
26 15 Kubin, L.P., Mortensen, A., 2003. Geometrically necessary dislocations and strain-
27 16 gradient plasticity: a few critical issues. Scr. Mater. 48, 119-125.
28
29 17 https://doi.org/10.1016/S1359-6462(02)00335-4
30
31 18 Li, J., Lu, W., Chen, S., Liu, C., 2020. Revealing extra strengthening and strain
32
33 19 hardening in heterogeneous two-phase nanostructures. Int. J. Plast. 126, 102626.
34
35 20 https://doi.org/10.1016/j.ijplas.2019.11.005
36
37
21 Li, W., Chou, T.-H., Yang, T., Chuang, W.-S., Huang, J.C., Luan, J., Zhang, X., Huo,
38 22 X., Kong, H., He, Q., Du, X., Liu, C.-T., Chen, F.-R., 2021a. Design of ultrastrong but
39
40 23 ductile medium-entropy alloy with controlled precipitations and heterogeneous grain
41
42 24 structures. Appl. Mater. Today 23, 101037. https://doi.org/10.1016/j.apmt.2021.101037
43
44 25 Li, W., Xie, D., Li, D., Zhang, Y., Gao, Y., Liaw, P.K., 2021b. Mechanical behavior of
45
46 26 high-entropy alloys. Prog. Mater. Sci. 118, 100777.
47
27 https://doi.org/10.1016/j.pmatsci.2021.100777
48
49 28 Li, Y., Li, W., Min, N., Liu, W., Jin, X., 2017. Effects of hot/cold deformation on the
50
51 29 microstructures and mechanical properties of ultra-low carbon medium manganese
52
53 30 quenching-partitioning-tempering steels. Acta Mater. 139, 96-108.
54
55 31 https://doi.org/10.1016/j.actamat.2017.08.003
56
57 32 Liang, Y.-J., Wang, L., Wen, Y., Cheng, B., Wu, Q., Cao, T., Xiao, Q., Xue, Y., Sha, G.,
58
33 Wang, Y., Ren, Y., Li, X., Wang, L., Wang, F., Cai, H., 2018. High-content ductile
59
60
61 42
62
63
64
65
1 coherent nanoprecipitates achieve ultrastrong high-entropy alloys. Nat. Commun. 9,
1
2 2 4063. https://doi.org/10.1038/s41467-018-06600-8
3
4 3 Lins, J.F.C., Sandim, H.R.Z., Kestenbach, H.J., Raabe, D., Vecchio, K.S., 2007. A
5
4 microstructural investigation of adiabatic shear bands in an interstitial free steel. Mater.
6
7 5 Sci. Eng. A 457, 205-218. https://doi.org/10.1016/j.msea.2006.12.019
8
9 6 Liu, L., Zhang, Y., Li, J., Fan, M., Wang, X., Wu, G., Yang, Z., Luan, J., Jiao, Z., Liu,
10
11 7 C.T., Liaw, P.K., Zhang, Z., 2022a. Enhanced strength-ductility synergy via novel
12
13 8 bifunctional nano-precipitates in a high-entropy alloy. Int. J. Plast. 153, 103235.
14
15 9 https://doi.org/10.1016/j.ijplas.2022.103235
16 10 Liu, M., Gong, W., Zheng, R., Li, J., Zhang, Z., Gao, S., Ma, C., Tsuji, N., 2022b.
17
18 11 Achieving excellent mechanical properties in type 316 stainless steel by tailoring grain
19
20 12 size in homogeneously recovered or recrystallized nanostructures. Acta Mater. 226,
21
22 13 117629. https://doi.org/10.1016/j.actamat.2022.117629
23
24 14 Liu, Y., Cao, Y., Mao, Q., Zhou, H., Zhao, Y., Jiang, W., Liu, Y., Wang, J.T., You, Z.,
25
26 15 Zhu, Y., 2020. Critical microstructures and defects in heterostructured materials and
27 16 their effects on mechanical properties. Acta Mater. 189, 129-144.
28
29 17 https://doi.org/10.1016/j.actamat.2020.03.001
30
31 18 Lu, W., Yan, K., Luo, X., Wang, Y., Hou, L., Li, P., Huang, B., Yang, Y., 2022. Superb
32
33 19 strength and ductility balance of a Co-free medium-entropy alloy with dual
34
35 20 heterogeneous structures. J. Mater. Sci. Technol. 98, 197-204.
36
37
21 https://doi.org/10.1016/j.jmst.2021.05.023
38 22 Ma, E., Zhu, T., 2017. Towards strength–ductility synergy through the design of
39
40 23 heterogeneous nanostructures in metals. Mater. Today 20, 323-331.
41
42 24 https://doi.org/10.1016/j.mattod.2017.02.003
43
44 25 Ming, K.S., Bi, X.F., Wang, J., 2018. Realizing strength-ductility combination of
45
46 26 coarse-grained Al0.2Co1.5CrFeNi1.5Ti0.3 alloy via nano-sized, coherent precipitates. Int.
47
27 J. Plast. 100, 177-191. https://doi.org/10.1016/j.ijplas.2017.10.005
48
49 28 Muhammad, W., Brahme, A.P., Kang, J., Mishra, R.K., Inal, K., 2017. Experimental
50
51 29 and numerical investigation of texture evolution and the effects of intragranular
52
53 30 backstresses in aluminum alloys subjected to large strain cyclic deformation. Int. J.
54
55 31 Plast. 93, 137-163. https://doi.org/10.1016/j.ijplas.2016.11.003
56
57 32 Nutor Raymond, K., Cao, Q., Wei, R., Su, Q., Du, G., Wang, X., Li, F., Zhang, D., Jiang,
58
33 J.-Z., 2021. A dual-phase alloy with ultrahigh strength-ductility synergy over a wide
59
60
61 43
62
63
64
65
1 temperature range. Sci. Adv. 7, eabi4404. https://doi.org/10.1126/sciadv.abi4404
1
2 2 Pan, Q., Zhang, L., Feng, R., Lu, Q., An, K., Chuang Andrew, C., Poplawsky Jonathan,
3
4 3 D., Liaw Peter, K., Lu, L., 2021. Gradient cell–structured high-entropy alloy with
5
4 exceptional strength and ductility. Science 374, 984-989.
6
7 5 https://doi.org/10.1126/science.abj8114
8
9 6 Qi, Y., Wu, Y., Cao, T., He, L., Jiang, F., sun, J., 2020. L21-strengthened face-centered
10
11 7 cubic high-entropy alloy with high strength and ductility. Mater. Sci. Eng. A 797,
12
13 8 140056. https://doi.org/10.1016/j.msea.2020.140056
14
15 9 Qin, S., Yang, M., Jiang, P., Wang, J., Wu, X., Zhou, H., Yuan, F., 2022a. Designing
16 10 structures with combined gradients of grain size and precipitation in high entropy alloys
17
18 11 for simultaneous improvement of strength and ductility. Acta Mater. 230, 117847.
19
20 12 https://doi.org/10.1016/j.actamat.2022.117847
21
22 13 Qin, S., Yang, M., Jiang, P., Yuan, F., Wu, X., 2022b. Excellent tensile properties
23
24 14 induced by heterogeneous grain structure and dual nanoprecipitates in high entropy
25
26 15 alloys. Mater. Charact. 186, 111779. https://doi.org/10.1016/j.matchar.2022.111779
27 16 Sathiyamoorthi, P., Kim, H.S., 2020. High-entropy alloys with heterogeneous
28
29 17 microstructure: Processing and mechanical properties. Prog. Mater. Sci., 100709.
30
31 18 https://doi.org/10.1016/j.pmatsci.2020.100709
32
33 19 Sauzay, M., 2008. Analytical modelling of intragranular backstresses due to
34
35 20 deformation induced dislocation microstructures. Int. J. Plast. 24, 727-745.
36
37
21 https://doi.org/10.1016/j.ijplas.2007.07.004
38 22 Sinha, S., Szpunar, J.A., Kiran Kumar, N.A.P., Gurao, N.P., 2015. Tensile deformation
39
40 23 of 316L austenitic stainless steel using in-situ electron backscatter diffraction and
41
42 24 crystal plasticity simulations. Mater. Sci. Eng. A 637, 48-55.
43
44 25 https://doi.org/10.1016/j.msea.2015.04.005
45
46 26 Slone, C.E., Miao, J., George, E.P., Mills, M.J., 2019. Achieving ultra-high strength and
47
27 ductility in equiatomic CrCoNi with partially recrystallized microstructures. Acta Mater.
48
49 28 165, 496-507. https://doi.org/10.1016/j.actamat.2018.12.015
50
51 29 Su, J., Raabe, D., Li, Z., 2019. Hierarchical microstructure design to tune the
52
53 30 mechanical behavior of an interstitial TRIP-TWIP high-entropy alloy. Acta Mater. 163,
54
55 31 40-54. https://doi.org/10.1016/j.actamat.2018.10.017
56
57 32 Sun, S.J., Tian, Y.Z., Lin, H.R., Dong, X.G., Wang, Y.H., Zhang, Z.J., Zhang, Z.F., 2017.
58
33 Enhanced strength and ductility of bulk CoCrFeMnNi high entropy alloy having fully
59
60
61 44
62
63
64
65
1 recrystallized ultrafine-grained structure. Mater. Des. 133, 122-127.
1
2 2 https://doi.org/10.1016/j.matdes.2017.07.054
3
4 3 Tsuji, N., Ogata, S., Inui, H., Tanaka, I., Kishida, K., Gao, S., Mao, W., Bai, Y., Zheng,
5
4 R., Du, J.-P., 2020. Strategy for managing both high strength and large ductility in
6
7 5 structural materials–sequential nucleation of different deformation modes based on a
8
9 6 concept of plaston. Scr. Mater. 181, 35-42.
10
11 7 https://doi.org/10.1016/j.scriptamat.2020.02.001
12
13 8 Wang, Y.F., Wang, M.S., Fang, X.T., Guo, F.J., Liu, H.Q., Scattergood, R.O., Huang,
14
15 9 C.X., Zhu, Y.T., 2019. Extra strengthening in a coarse/ultrafine grained laminate: Role
16 10 of gradient interfaces. Int. J. Plast. 123, 196-207.
17
18 11 https://doi.org/10.1016/j.ijplas.2019.07.019
19
20 12 Wu, S.W., Wang, G., Wang, Q., Jia, Y.D., Yi, J., Zhai, Q.J., Liu, J.B., Sun, B.A., Chu,
21
22 13 H.J., Shen, J., Liaw, P.K., Liu, C.T., Zhang, T.Y., 2019. Enhancement of strength-
23
24 14 ductility trade-off in a high-entropy alloy through a heterogeneous structure. Acta Mater.
25
26 15 165, 444-458. https://doi.org/10.1016/j.actamat.2018.12.012
27 16 Wu, X., Yang, M., Yuan, F., Wu, G., Wei, Y., Huang, X., Zhu, Y., 2015. Heterogeneous
28
29 17 lamella structure unites ultrafine-grain strength with coarse-grain ductility. Pro. Natl.
30
31 18 Acad. Sci. U.S.A. 112, 14501-14505. https://doi.org/10.1073/pnas.1517193112
32
33 19 Wu, X., Zhu, Y., Lu, K., 2020. Ductility and strain hardening in gradient and lamellar
34
35 20 structured materials. Scr. Mater. 186, 321-325.
36
37
21 https://doi.org/10.1016/j.scriptamat.2020.05.025
38 22 Yan, C., Xin, Y., Chen, X.-B., Xu, D., Chu, P.K., Liu, C., Guan, B., Huang, X., Liu, Q.,
39
40 23 2021. Evading strength-corrosion tradeoff in Mg alloys via dense ultrafine twins. Nat.
41
42 24 Commun. 12, 4616. https://doi.org/10.1038/s41467-021-24939-3
43
44 25 Yang, M., Pan, Y., Yuan, F., Zhu, Y., Wu, X., 2016. Back stress strengthening and strain
45
46 26 hardening in gradient structure. Mater. Res. Lett. 4, 145-151.
47
27 https://doi.org/10.1080/21663831.2016.1153004
48
49 28 Yang, M., Yan, D., Yuan, F., Jiang, P., Ma, E., Wu, X., 2018. Dynamically reinforced
50
51 29 heterogeneous grain structure prolongs ductility in a medium-entropy alloy with
52
53 30 gigapascal yield strength. Pro. Natl. Acad. Sci. U.S.A. 115, 7224-7229.
54
55 31 https://doi.org/10.1073/pnas.1807817115
56
57 32 Yang, T., Zhao, Y.L., Fan, L., Wei, J., Luan, J.H., Liu, W.H., Wang, C., Jiao, Z.B., Kai,
58
33 J.J., Liu, C.T., 2020. Control of nanoscale precipitation and elimination of intermediate-
59
60
61 45
62
63
64
65
1 temperature embrittlement in multicomponent high-entropy alloys. Acta Mater. 189,
1
2 2 47-59. https://doi.org/10.1016/j.actamat.2020.02.059
3
4 3 Zhang, C., Zhu, C., Cao, P., Wang, X., Ye, F., Kaufmann, K., Casalena, L., MacDonald,
5
4 B.E., Pan, X., Vecchio, K., Lavernia, E.J., 2020. Aged metastable high-entropy alloys
6
7 5 with heterogeneous lamella structure for superior strength-ductility synergy. Acta Mater.
8
9 6 199, 602-612. https://doi.org/10.1016/j.actamat.2020.08.043
10
11 7 Zhang, C., Zhu, C., Vecchio, K., 2019. Non-equiatomic FeNiCoAl-based high entropy
12
13 8 alloys with multiscale heterogeneous lamella structure for strength and ductility. Mater.
14
15 9 Sci. Eng. A 743, 361-371. https://doi.org/10.1016/j.msea.2018.11.073
16 10 Zhang, C.L., Bao, X.Y., Zhang, D.D., Chen, W., Zhang, J.Y., Kuang, J., Liu, G., Sun,
17
18 11 J., 2021a. Achieving superior strength-ductility balance in a novel heterostructured
19
20 12 strong metastable β-Ti alloy. Int. J. Plast. 147, 103126.
21
22 13 https://doi.org/10.1016/j.ijplas.2021.103126
23
24 14 Zhang, D.D., Zhang, J.Y., Kuang, J., Liu, G., Sun, J., 2021b. Superior strength-ductility
25
26 15 synergy and strain hardenability of Al/Ta co-doped NiCoCr twinned medium entropy
27 16 alloy for cryogenic applications. Acta Mater. 220, 117288.
28
29 17 https://doi.org/10.1016/j.actamat.2021.117288
30
31 18 Zhao, Y.L., Yang, T., Tong, Y., Wang, J., Luan, J.H., Jiao, Z.B., Chen, D., Yang, Y., Hu,
32
33 19 A., Liu, C.T., Kai, J.J., 2017. Heterogeneous precipitation behavior and stacking-fault-
34
35 20 mediated deformation in a CoCrNi-based medium-entropy alloy. Acta Mater. 138, 72-
36
37
21 82. https://doi.org/10.1016/j.actamat.2017.07.029
38 22 Zhao, Y.L., Yang, T., Zhu, J.H., Chen, D., Yang, Y., Hu, A., Liu, C.T., Kai, J.J., 2018.
39
40 23 Development of high-strength Co-free high-entropy alloys hardened by nanosized
41
42 24 precipitates. Scr. Mater. 148, 51–55. https://doi.org/10.1016/j.scriptamat.2018.01.028
43
44 25 Zhong, S., Xu, C., Li, Y., Li, W., Luo, H., Peng, R., Jia, X., 2022. Hierarchy
45
46 26 modification induced exceptional cryogenic strength, ductility and toughness
47
27 combinations in an asymmetrical-rolled heterogeneous-grained high manganese steel.
48
49 28 Int. J. Plast. 154, 103316. https://doi.org/10.1016/j.ijplas.2022.103316
50
51 29 Zhu, Y., Ameyama, K., Anderson, P.M., Beyerlein, I.J., Gao, H., Kim, H.S., Lavernia,
52
53 30 E., Mathaudhu, S., Mughrabi, H., Ritchie, R.O., Tsuji, N., Zhang, X., Wu, X., 2021.
54
55 31 Heterostructured materials: superior properties from hetero-zone interaction. Mater.
56
57 32 Res. Lett. 9, 1-31. https://doi.org/10.1080/21663831.2020.1796836
58
33 Zhu, Y., Wu, X., 2019. Perspective on hetero-deformation induced (HDI) hardening and
59
60
61 46
62
63
64
65
1 back stress. Mater. Res. Lett. 7, 393-398.
1
2 2 https://doi.org/10.1080/21663831.2019.1616331
3
4 3
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61 47
62
63
64
65
Supplementary Material

Click here to access/download


Supplementary Material
Supplementary material-IJP.docx

You might also like