You are on page 1of 11

Acta Materialia 153 (2018) 279e289

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Full length article

Mechanical properties and optimal grain size distribution profile of


gradient grained nickel
Y. Lin a, b, J. Pan a, H.F. Zhou c, H.J. Gao c, **, Y. Li a, *
a
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang, 110016, China
b
University of Chinese Academy of Sciences, 19 Yuquan Road, Beijing, 100039, China
c
School of Engineering, Brown University, Providence, RI, 02912, USA

a r t i c l e i n f o a b s t r a c t

Article history: Gradient structured (GS) materials are ubiquitous in biological systems and now increasingly adopted in
Received 27 February 2018 engineering systems to achieve desirable combinations of mechanical properties. However, how to
Received in revised form control and characterize the gradient structure still remains challenging. In the present work, pure Ni
27 April 2018
samples possessing a gradient structure with a change in the grain size up to three orders of magnitude
Accepted 30 April 2018
Available online 3 May 2018
from 29 nm to 4 mm are prepared by electrodeposition, where the degree of grain size gradient is
accurately controlled. The GS Ni samples exhibit a favorable combination of high strength and high
ductility. An optimal grain size distribution profile is discovered which gives rise to a yield strength of
Keywords:
Gradient structured metals
460 MPa and a uniform elongation of 8.9%, the latter even better than that of the coarse-grained Ni.
Grain size distribution Experimental observations and molecular dynamics (MD) simulations reveal that the surface roughening
Degree of gradient of coarse grains and strain localization of nano-grains can be effectively suppressed by the mutual
Strength constraint between nano-grains and coarse grains, leading to the observed superior uniform elongation.
Ductility This work not only reports a promising methodology of producing materials possessing both high
strength and high ductility, but also provides a model for investigating the deformation mechanisms in
GS materials.
© 2018 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction induced plasticity (TWIP) steel via pre-torsion [10], leading to a


simultaneous improvement of strength and ductility. In addition,
Gradient structured (GS) materials are ubiquitous in biological other mechanical properties such as fatigue life [12e14] and wear
systems [1e4] and now increasingly adopted in engineering sys- resistance [15e19], could also be greatly enhanced in GS materials.
tems to achieve desirable combinations of mechanical properties For a deeper understanding of the deformation mechanisms in
[5e11]. Recently, gradient nano-grained (GNG) Cu prepared GS materials, the relationship between their microstructure and
through surface mechanical grinding treatment (SMGT) was re- mechanical properties has been investigated by computer
ported by Fang et al. [7]. This GNG Cu exhibited a doubling yield modeling and numerical simulation [20e23]. Using a crystal plas-
strength and a tensile plasticity comparable to that of the coarse ticity finite element model, Zeng et al. found that there are large
grained (CG) Cu. Using surface mechanical attrition treatment stress/strain gradients in GNG Cu under uniaxial loading [22]. A
(SMAT), similar gradient structure could also be attained in inter- gradient variation of GNDs was built up to accommodate the
stitial free (IF)-steel, which rendered a unique combination of extra deformation incompatibility, leading to a sustained strain hard-
strain hardening and high ductility [9]. The extra strain hardening ening and enhanced ductility. On the other hand, the extra strain
was attributed to the formation of geometrically necessary dislo- hardening of GS IF steels was successfully described by a dislocation
cations (GNDs) as well as the multiaxial stress state. A gradient density-based continuum plasticity model [21]. The high strain
hierarchical nanotwinned structure can be formed in twinning- hardening capability was attributed to the generation of abundant
GNDs in the gradient structure during plastic deformation.
The aforementioned studies suggest that materials having a
* Corresponding author.
grain size gradient structure could possess superior mechanical
** Corresponding author. properties, in comparison with those having a monotonic grain size
E-mail addresses: huajian_gao@brown.edu (H.J. Gao), liyi@imr.ac.cn (Y. Li). structure. However, most of the previous studies were focused on

https://doi.org/10.1016/j.actamat.2018.04.065
1359-6454/© 2018 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
280 Y. Lin et al. / Acta Materialia 153 (2018) 279e289

only one type of gradient structure due to limitations in the syn- characteristic parameter of gradient materials. Generally, with an
thesis technology. How to control the grain size distribution (i.e. increase in the n value, the volume fraction of coarse grains in-
degree of grain size gradient) and what is the effect of the grain size creases, while the volume fraction of fine grains continuously de-
gradient profile on deformation behaviors were rarely studied. Is creases (Fig. 1a). The mechanical properties of GS Ni samples are
there an optimum gradient profile? What is the key parameter of expected to have a dependency on the n value.
the gradient structure that controls the mechanical properties?
These remaining issues may hinder the development of a thorough
2.2. Sample preparation
understanding of deformation mechanisms in GS materials and
thus realization of their full potential in applications.
The key to electroplating GS Ni plates having various degrees of
This work outlins a systematic investigation of the effect of
gradient is to establish a relationship between the grain size and
degree of gradient on the mechanical properties of GS Ni samples.
deposition conditions. Based on the previous studies [24,25], a
Direct current (DC) electrodeposition was adopted as a bottom-up
quantitative relationship between grain size and current density/
method for preparing GS Ni samples possessing various degrees
additive (sodium saccharin) content had been established, as
of gradient. By regulating processing conditions such as the current
shown in Fig. 2a. Different from the previous works [24,25] where
density and additive concentration, Ni plates up to 2 mm in thick-
the range of variation of grain size was limited to one order of
ness were prepared with the grain size changing continuously from
magnitude, a much wider range of change in grain size from several
29 nm to 4 mm. An optimum grain size distribution was found in GS
microns to a few of tens nanometers has been attained by simul-
Ni samples where ductility is peaked. The underlying mechanism
taneously and continuously adjusting the current density
for such an optimization is also discussed in this work.
(10e100 mA/cm2) and additive content (1e5 g/L) during deposi-
tion. In addition, their changing rates were modified to control the
2. Experimental degree of gradient. Fig. 2b shows a schematic diagram where the
output current and additive content continuously change with the
2.1. The degree of gradient deposition time, contributing to a continuous refinement of the
grain size along the deposition direction.
Fig. 1a shows a schematic diagram of a number of grain size Ni deposits having six different profiles of gradient structure
variations as a function of position along the deposition direction. were designed and synthesized with a thickness up to 2 mm
The maximum and minimum grain sizes, dmax and dmin, are kept the through a conventional rotating disc electrode set-up. The plating
same among different cases, while the distributions of grain size are bath was a modified Watts-type electrolyte containing nickel sul-
quite different. For example, the transition in the grain size of fate (300 g/L), nickel chloride (45 g/L), boric acid (40 g/L) and
gradient A is gentler (Fig. 1b), resulting in a larger volume fraction dodecyl sodium sulfate (0.05 g/L), maintained at pH value of
of fine grains (i.e. hard phases). On the other hand, gradient C has a 4.0 ± 0.2 and temperature of 328 ± 1 K. Monolithic CG and nano-
steeper change in grain size and a smaller volume fraction of fine grained (NG) Ni samples were also prepared by electrodeposition
grains (Fig. 1b). Here, we employ a power law equation to charac- for comparison.
terize the degree of gradient in GS metals with different types of
grain size profile. The relationship between grain size d and
normalized position x can be expressed as follows: 2.3. Elemental analysis and microstructure characterization

d ¼ dmax  ðdmax  dmin Þð1  xÞn (1) The contents of nitrogen, hydrogen, oxygen, carbon and sulfur in
the GS Ni deposits were assessed by gas analysis (TCH-600
where n is the power index for the grain size distribution, which Hydrogen/Nitrogen/Oxygen Analyzer) and IR absorption method
can be considered as the degree of the gradient and the (CS844 Carbon & Sulfur Determinator). The content of light

Fig. 1. (a) Schematic illustration of various grain size gradient profiles as a function of position in gradient materials. (b) Three typical gradient structures with distinct grain size
distributions.
Y. Lin et al. / Acta Materialia 153 (2018) 279e289 281

carried out on an INSTRON-5848 micro-tester system at a strain


rate of 3  104 s1 at room temperature. All GS Ni plates were
annealed at 373 K for 1 h to release the residual stress before cut-
ting and mechanical polishing into dog-bone shaped samples with
a total length of 25 mm, a gauge cross-section of about
0.5  1.2 mm2 and a gauge length of 6 mm. Moreover, these dog-
bone shaped samples were processed with a round of electro-
polishing, to acquire a smooth and residual-stress-free surface. A
laser extensometer was used to measure strain within the sample
gauge upon loading. The surface morphologies around the necking
areas of various GS Ni tensile specimens were examined by SEM
and confocal laser scanning microscope (CLSM) before and after
tension, respectively.

2.5. MD simulations

MD simulations were performed to understand the origin of the


superior ductility of gradient structures. The simulated gradient
sample consists of four regions, each with identical dimensions of
200 (X)  200 (Y)  2 (Z) nm3. To create the gradient grain distri-
bution along the X direction, the average grain size of the four re-
gions were chosen as d ¼ 10, 25, 50 and 100 nm, respectively. The
entire system contains about 26, 000, 000 atoms and 500 randomly
oriented grains in total. For comparison, uniform nanostructured
samples with d ¼ 10 and 100 nm were considered. Periodic
boundary conditions were imposed in Y and Z directions, while free
surfaces were created on the boundaries of the sample in X direc-
tion (i.e., gradient direction). The embedded atom method potential
for Ni [26] was adopted. The integration time-step was fixed at 1 fs.
All simulations were performed at 600 K using a Nose-Hoover
thermostat [27]. The sample was relaxed for 500 ps before strain-
ing. During uniaxial tension, a 20% strain was applied to the sample
at a strain rate of 5  108 s1. To identify plastic deformation, colors
were assigned to the atoms according to their atomic shear strain
[28].
Fig. 2. (a) Grain size as a function of additive content and current density during
deposition. (b) Schematic showing precise control of grain size gradient during
3. Results
electrodeposition.

3.1. Microstructure of GS Ni samples


element impurities was less than 200 ppm (wt.) in all deposits.
Based on Archimedes' principle, the density was measured to be
Fig. 3aed shows the cross-sectional SEM images of four GS Ni
larger than 99% in all deposits, implying that the deposits are dense.
samples, i.e. samples II, III, IV and V, respectively. A microstructural
Cross-sectional morphologies of various GS Ni were character-
evolution from coarse grains to fine grains along the deposition
ized by scanning electronic microscope (SEM FEI Nova NanoSEM
direction without a sharp interface is observed in these specimens,
430). Before cutting and polishing, the GS specimens were coated
indicating a typical gradient structure. In all GS specimens, the
with a protective Ni layer. Microstructure of the NG side was
grain size starts from ~4 mm at the end of columnar coarse grains
investigated by transmission electron microscope (TEM FEI Tecnai
(Fig. 3e), and gradually decreases to 29 nm near the end of fine
G2 F20) operated at 200 kV.
grains (Fig. 3f and g). The range of change in grain size almost spans
To measure the profile of grain size distribution, the GS Ni sheets
three orders of magnitude, comparable to those reported in the
were mechanically polished layer by layer along the deposition
literature [7,9]. Moreover, the change in grain size of the finer grains
direction, then directly tested by X-ray diffraction (XRD, Bruker D2
is much gentler in sample II than that in sample IV, leading to a
phaser) with Cu Ka radiation. Before the XRD measurement, all
higher fraction of nano-grains. This indicates a clear difference in
samples were annealed at 373 K for 1 h to release possible residual
the degree of grain size gradient in samples II and IV.
stresses generated during electrodeposition. Grain size was calcu-
The XRD patterns of sample II display a gradual change in the
lated based on the corresponding full width at half maximum
intensity and FWHM of diffraction peaks at various positions along
(FWHM) using Scherrer formula after correcting the instrumental
the deposition direction, as shown in Fig. 4a. From the substrate to
line broadening. The grain sizes of columnar coarse grains and
the surface, the diffraction intensity of (200) decreases and (111)
nano-grains were measured by SEM and TEM, respectively.
becomes the predominant crystal orientation. The obvious broad-
ening in the diffraction peaks denotes a decrease in the grain size.
2.4. Mechanical testing In addition, the diffraction peaks slightly shift to a higher angle
(~0.2 ) from surface to substrate, which could be attributed to a
The hardness profiles of GS Ni samples along the deposition compressive residual stress induced by the process of mechanical
direction were determined using Qness Q10 A þ microhardness polishing just before the XRD measurement. Detailed grain size
tester with Vickers indenter. The load for measurement was 5 g, profiles of various GS samples as a function of the normalized
with a dwell time of 10 s. Quasi-static uniaxial tensile tests were distance are plotted in Fig. 4b. Based on Eq. (1), the degree of
282 Y. Lin et al. / Acta Materialia 153 (2018) 279e289

Fig. 3. Gradient structures of as-deposited Ni plates. (aed) Cross-sectional SEM images of samples II, III, IV and V, respectively; (e) Representative SEM image of coarse grains on the
CG side. (f) Dark-field TEM image and (g) the corresponding grain size distribution of the topmost surface layer on the NG side.

gradient, estimated by the value of n, ranges from 0.01 in sample I consistent with those obtained from monolithic NG and CG Ni
to 5 in sample VI (Table 1). samples. Similar to the grain size gradient, hardness also exhibits
Grain sizes below 1 mm along the deposition direction in sample distinct gradient profiles for various GS samples. The change in
II were further confirmed by TEM. With an increase in the hardness is much slower in samples II and III than in samples V and
normalized distance from 0.05 to 0.76, the average grain size VI.
gradually increases from about 35 to 224 nm (Fig. 5a). Equiaxed Fig. 7a displays engineering stress-strain curves of the GS Ni
nano-grains were found at normalized distances of 0.05 and 0.58 samples under uniaxial tension at room temperature. For com-
(Fig. 5b and c), while elongated columnar grains appeared at parison, the data obtained from CG (4 mm) and NG (29 nm) Ni
normalized distance of 0.76 (Fig. 5d). The average grain sizes ob- samples are also included. The yield strength (0.2% offset, sy),
tained by TEM are consistent with those calculated by XRD. uniform elongation (εue) and ultimate tensile strength (sUTS) of each
specimen are listed in Table 1. As expected, both strength and
3.2. Mechanical properties of GS Ni samples ductility are strongly dependent on the degree of gradient. From
samples I to VI, sy decreases monotonously from 880 to 409 MPa
Fig. 6 presents the microhardness profiles of various GS Ni with n increasing from 0.01 to 5. More interestingly, εue first in-
samples. The microhardness continuously decreases from a creases from 4.5% in sample I to 8.9% in sample V, and then de-
maximum value of ~5.4 GPa to the minimum value of ~2.5 GPa as creases to 8.0% in sample VI. A maximum uniform elongation is
the normalized distance shifts from 0 to 1. These values are observed in sample V, which is 1.2 times higher than that of the CG
Y. Lin et al. / Acta Materialia 153 (2018) 279e289 283

expense of reduced ductility, known as the strength and ductility


trade-off in conventional metals. It is thus surprising to observe the
simultaneous enhancement in strength and ductility in our GS Ni
samples (Fig. 8a). The true strain at which necking occurs reaches a
peak value of 8.5% at n ¼ 3, which is even larger than that of CG Ni
sample. Our results show that superior uniform elongation is
achievable by optimizing the gradient profile.
To further recognize the influence of degree of gradient on the
mechanical properties, various gradients from previous studies
[7,29,30] are summarized. The gradient structure synthesized by
surface plastic deformation (e.g. SMGT and SMAT) contains a GS
layer with a thickness up to 200 mm, regardless of the sample size.
The degree of gradient (the volume fraction of GS layer) is
controlled by the diameter or thickness of samples. In contrast, the
electrodeposition method developed in the present work for the
synthesis of GS samples imposes no restriction on the sample size
and degree of gradient, and allows for a similar range of variation in
the grain size compared to that by surface plastic deformation. To
eliminate any size effect and offer an intuitive comparison among
various GS materials, both grain size and distance of various GS
materials are normalized in Fig. 8b. The degree of gradient of
sample V is almost the same as that of GNG Cu, i.e. n ¼ 4 [7] (Fig. 8b
and Table 2), despite their distinct profiles in grain size. GS IF-steel
[29] showing favorable strength-ductility synergy has a similar
grain size profile to those of GNG Cu and sample V (Fig. 8b and
Table 2). In addition, there is an optimum range of volume fractions
of GS layer that produces the largest enhancement in strengthening
and extra strain hardening in pure Cu treated by SMAT [30].
Based on the above summary, an optimal grain size gradient
profile with a n value around 3 (such as sample V and GNG Cu [7])
can be reasonably deduced to exist where mechanical properties
Fig. 4. (a) XRD patterns of sample II at various positions along deposition direction; (b)
Average grain size as a function of the normalized distance in samples II, IV and V, can be optimized. The optimization of mechanical properties are
respectively. possibly ascribed to the effective constraint between coarse grains
and nano-grains in the gradient structure.

Table 1 4.2. Countering strain localization via reduced surface roughening


Lists of yield stresse sy, uniform elongation εue, ultimate tensile strength sUTS of CG,
NG, and GS Ni under uniaxial tension at a strain rate of 3  104 s1.
For the purpose of acquiring a large uniform ductility, the main
NG I II III IV V VI CG challenge is to delay strain localization and instability during
n 0.01 0.016 0.4 0.75 3 5 plastic deformation. Surface roughening is an important factor
sy (MPa) 1130 880 790 573 488 460 409 380 associated with strain localization in polycrystalline metals. If
εue (%) 2.96 4.56 5.2 6.72 7.04 8.9 8 7.5 plastic strain can no longer be fully accommodated by slip, the
sUTS (MPa) 1437 1200 1093 851 766 698 627 592 incompatibility between neighboring grains causes grains to move
normal to the surface, leading to roughening on the surface [31].
The unstable growth in surface roughening is considered as the
Ni sample, implying that ductility can be optimized by tuning the onset of critical strain localization, which is a direct precursor to
grain size distribution profile in the GS Ni samples. failure such as formations of cracks or necking [32e34]. Previous
Fig. 7b shows the relationship between strain hardening rate studies have shown that a reduction in strain hardening ability
(Q ¼ ds=dε) and true strain (εT ) for CG, NG, and two typical GS Ni caused by strain localization is accompanied by increased surface
samples. The strain hardening rate of NG Ni sample drops quickly roughness [31]. On the other hand, the inhibition of strain locali-
with straining, resulting in early plastic instability and premature zation is a key to improving the tensile ductility of metals [35].
failure. The strain hardening rates of GS Ni samples depend on the Consequently, it is important to establish a solid causality between
degree of gradient, i.e., the value of n in Eq. (1). Similar to NG Ni ductility and surface roughness in our GS Ni samples.
sample, Q in sample II drops quickly due to the small value of Surface roughness decreases with a decrease in grain size
n ¼ 0.016, leading to a reduced uniform elongation. Sample V has an [36e38]. In GS Ni samples, fine grains are expected to ease surface
appropriate value of n ¼ 3, and thus exhibits an enhanced strain roughening and inhibit strain localization. For gradient structures
hardening rate at the initial stage of straining as well as a slower with relatively high n values (such as sample VI), the gradient is too
decay in Q compared with CG Ni sample. steep and the nano-grains are less effective in reducing surface
roughening. The nano-grains fluctuate with inhomogeneous
deformation of the coarse grains after the onset of plastic insta-
4. Discussion bility. In other words, this kind of gradient structure cannot most
efficiently ease surface roughening and acquire optimal tensile
4.1. Optimal gradient property either, because of the lack of adequate constraint on
plastic instability in the coarse grains.
In general, increase in strength is usually achieved at the For gradient structures with relatively low n values (such as
284 Y. Lin et al. / Acta Materialia 153 (2018) 279e289

Fig. 5. (a) Grain size measured by XRD and TEM as a function of the normalized distance in sample II. Bright-field TEM images at normalized distances of (b) 0.05, (c) 0.58 and (d)
0.76 in sample II, where the corresponding average grain sizes are 35, 85, 224 nm, respectively.

suppressed by the coarse grains, leading to a rapid fracture after


yielding.
In the gradient structure with an optimum n value (i.e. appro-
priate grain size gradient profile), such as sample V with n ¼ 3,
surface roughening is mitigated by the mutual constraint of coarse
grains and nano-grains. The strain localization of nano-grains and
coarse grains can be efficiently suppressed simultaneously,
contributing to the smoothest surface and best accommodated
plastic deformation in the gradient structure.
In order to validate the above hypothesis, the surface roughness
of all samples was investigated. Fig. 9 presents the surface mor-
phologies and height variations of the GS Ni samples after tension.
Fig. 9b shows the morphology of surface valleys [31] observed in
the necking region on the CG side. Strain localization of the coarse
grains is verified by the largest height variation in Fig. 9j.
In contrast, a smoother surface morphology without obvious
cracks is observed on the NG side in sample VI (Fig. 9f and i).
However, the relatively large height variation (Fig. 9j) indicates that
the coarse grains are not effectively constrained. In addition,
intense strain localization in samples I and II leads to a rough sur-
Fig. 6. Hardness distributions of GS Ni having different degrees of grain size gradient.
face in the necking area (Fig. 9c, d and j) and micro-cracks (Fig. 9g)
on the NG side, indicating a rapid rupture of the whole structure.
Sample V has the smoothest surface on the NG side (Fig. 9e and
samples I and II), plastic deformation of GS Ni is dominated by the
j) and no obvious cracks (Fig. 9h) were observed, indicating that the
mechanical behavior of nano-grains, where limited dislocation
coarse grains are effectively constrained and strain localization of
activity [39e41] results in poor strain hardening capacity in the GS
nano-grains is also inhibited. This suppresses tensile instability and
specimens, thus a rougher surface is facilitated by early necking and
results in a superior elongation in sample V (Fig. 7a).
cracking. The strain localization of nano-grains cannot be
Y. Lin et al. / Acta Materialia 153 (2018) 279e289 285

Fig. 7. (a) Engineering stress-strain curves of CG, NG and GS Ni with different degrees
Fig. 8. (a) Yield strength (experimental) versus necking strain for electrodeposited CG,
of gradient obtained at room temperature and a strain rate 3  104s1. (b) Strain
NG, and various GS Ni. (b) Summary of the grain size distribution in grain size gradient
hardening rate (Q ¼ ds/dε) vs. true strain (εT) curves of CG Ni, NG Ni, sample II and V,
materials, including GNG Cu [7], IF steel [35] and the present GS Ni. Normalized grain
respectively.
size is defined as: 1-(dmax-d)/(dmax-dmin).

The dependence of surface roughness and necking strain on n Table 2


values in GS Ni samples is further demonstrated in Fig. 10. The Summary of mechanical properties (yield stress, uniform elongation and ultimate
variation of surface roughness exhibits an opposite trend to that of strength) in gradient materials with various degrees of gradient.
the necking strain. Sample V with n ¼ 3 has the minimum surface n sy (MPa) sUTS (Mpa) εue (%) Authors
roughness and the maximum uniform elongation among the six
4 129 250 31 Fang et al.
samples (Fig. 10).
In spite of numerous previous studies on enhanced strength and 0.01 880 1200 4.5 This work
0.016 790 1093 5.2
ductility in NG and gradient metals [7,9,10,42e44], few reported
0.4 573 851 6.7
superior elongation larger than that of their CG counterpart 0.75 488 766 7.0
(Table 3). Our present work indicates that once the surface 3 460 698 8.9
roughening in a CG metal is effectively reduced by marrying to an 5 409 627 8.0
NG partner, the overall gradient structure could acquire higher 290 310 17 Wu et al.
hardening capacity and larger elongation than those in the case of a 240 280 19
free standing CG metal [42]. On the other hand, in case that the 210 275 26
170 265 30
plastic instability (strain localization) in a NG metal is suppressed
through pairing with a CG partner, the brittle NG metals can also
undergo large tensile deformation [45e48]. This synergy between
nano-grains and coarse grains stabilizes plastic instability by investigations in the future.
introducing gradient structure.
As noted, there is a strong (200) texture existing in the coarse 4.3. MD simulations of shear localization for GS Ni
grains inside the GS Ni specimen. However, the volume fraction of
the coarse grains monotonously changes with the degree of MD simulations were performed to further understand the ef-
gradient, thus the texture should not be the reason for the optimal fect of gradient grain distribution on the underlying plastic defor-
mechanical properties in our GS Ni. This issue deserves further mation mechanism in Ni samples. Deformation patterns of two
uniform nanostructured Ni samples with different grain sizes are
286 Y. Lin et al. / Acta Materialia 153 (2018) 279e289

Fig. 9. 3D surface images of GS Ni, including (a) polished sample before tension, (b) CG side of sample V, and (cef) NG sides of sample I, II, V and VI after tension. (gei) SEM images
of lateral surface at NG side of sample II, V and VI, respectively. (j) Height variation profiles of NG side in various GS Ni after tension, as well as the CG side surface and the polished
surface before tension. All profiles are taken from the regions near the fracture surface.

shown in Fig. 11a and b. In the sample with d ¼ 10 nm, plastic shown in Fig. 11b, where plastic deformation is governed by dis-
deformation is governed by grain boundary associated activities, locations emitted from triple junctions and free surfaces. Large
including grain boundary sliding, grain rotation and grain coales- values of atomic shear strain occur mainly in a few inclined slip
cence. Fig. 11a shows that plastic deformation is localized along a bands, indicating highly localized plastic deformation in the
few planes forming inclination angles of roughly 45 with respect sample.
to the loading axis, which can be attributed to the fact that the In contrast, the deformation patterns of the gradient nano-
driving force for grain boundary sliding, which is the resolved shear structured Ni sample (Fig. 11c) displays several distinct features
stress on the grain boundaries, reaches its maximum value on these compared with those shown in Fig. 11a and b: (1) Shear localization
planes. The result for the sample with d ¼ 100 nm at strain 20% is on both sides of the gradient sample has been effectively
Y. Lin et al. / Acta Materialia 153 (2018) 279e289 287

that the values of atomic shear strain, both at grain boundaries and
in the grain interior, are smaller than those observed in uniform
samples; (3) The size of the surface steps is also reduced due to
suppressed shear localization.
According to the previous works [49e53], NG metals show a
high strain hardening rate but low Q-sustained rate, resulting in
early strain localization. The inhibition of strain localization could
be accomplished by introducing an extra strain hardening mecha-
nism [9,42,44,54] in materials possessing an inhomogeneous
microstructure. The MD simulation results provide atomic scale
evidence in support of the hypothesis that introducing gradient in
grain size could effectively suppress shear localization and help to
achieve enhanced ductility compared with uniform samples.

4.4. Accommodation of mechanical incompatibility in gradient


structure

The mechanical incompatibility is prominent during the plastic


deformation of a heterogeneous structure, whose contributions to
the enhancement of mechanical properties would vary with
microstructural heterogeneity. As a typical heterogeneous struc-
ture, the influence and accommodation of mechanical in-
compatibility are determined by the type of materials and degree of
gradient.
For GS IF steel [9] or GS Cu [30], in order to accommodate a large
degree of non-uniform plastic deformation caused by the me-
chanical incompatibility, an increasing density of GNDs has to be
stored. The GNDs are the origin of back stress [21,55], which pro-
duces remarkable extra strain hardening and extra strengthening
with greater degree of mechanical incompatibility in GS IF steel [9]
and GS Cu [30]. One consequence of this extra strengthening is the
fact that the yield strength of GS IF steel and GS Cu is above the
Fig. 10. (a) Surface roughness (Ra) and (b) necking strain as a function of n. The region
value calculated using rule of mixture (ROM) [29], while the extra
near the fracture surface on the NG side is selected for measurement.
strain hardening should have contributed to the high ductility [9].
In this sense, the GNDs generated for the accommodation of me-
Table 3 chanical incompatibility induced back stress, which can play an
Summary of uniform elongation for heterogeneous structure and CG counterpart. essential role in the improvement of the strength and ductility for
Materials εue (%) Authors GS IF steel and GS Cu.
On the other hand, the extraordinary plasticity in GNG Cu [7,56]
CG Heterogeneous structure
was accommodated by grain coarsening. In contrast to the
Cu 32 31 Fang et al. strengthening/strain hardening caused by back stress in the case of
Ni 7.5 8.9 This work GS IF steel [9] and GS Cu [30], grain coarsening leads to strain
8.0 softening. Previous studies have shown that grain coarsening oc-
IF steel 38 17 Wu et al. curs in nano-grained Ni under plastic deformation [57,58]. We
19 speculate that the superior plasticity observed in our GS Ni samples
26 is induced by grain coarsening, rather than back stress induced
30
strain hardening. The strength of GS Ni samples follows the ROM
TWIP steel 49 70 Wei et al. and no strain hardening up-turn is observed (Fig. 7b), suggesting
58
that the contribution of back stress might not be significant. This
55
57
issue remains to be fully clarified in the future.
Although several deformation mechanisms on enhancing
Ti 8.0 8.2 Wu et al.
ductility in GS metals have been proposed, such as dissociation of
8.5
9.0 dislocations interacting with twin boundaries [10], GNDs intro-
10.0 duced extra work hardening [9,21], grain coarsening [7,56] and
10.5 back stress hardening [42,55], our current observation provides an
Cu 49 30 Wang et al. additional explanation. Furthermore, it should be pointed out that
Cu 40 3 Yang et al.
the present work is only a starting point. The present gradient
14 structure is asymmetric which may be detrimental to the plastic
25 deformation. If the gradient structure or grain size distribution can
29 be further optimized, a better combination of strength and ductility
can be expected.

suppressed at strain of 20%. Though plastic deformation still tends 5. Conclusions


to localize near the free surfaces, it is uniformly dispersed in the
interior of the sample; (2) From the color of the contour, we can tell In conclusion, we have developed a method to synthesize bulk
288 Y. Lin et al. / Acta Materialia 153 (2018) 279e289

Fig. 11. Deformation patterns of (a) 10 nm Ni, (b) 100 nm Ni, and (c) gradient Ni with grain size ranging from 10 to 100 nm at a strain of 20%. Colors are assigned based on local shear
strain. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

and dense GS Ni plates possessing various degrees of grain size material in nacre and bone e perspectives on de novo biomimetic materials,
Prog. Mater. Sci. 54 (2009) 1059e1100.
gradient (i.e., grain size distribution profiles). Microstructural
[3] S. Amada, Hierarchical functionally gradient structures of bamboo barley, and
characterization shows a continuous change in grain sizes spanning corn, MRS Bull. 20 (1995) 35e36.
up to three orders of magnitude from 4 mm to 29 nm and the degree [4] K. Ghavami, Bamboo as reinforcement in structural concrete elements,
of gradient can be accurately controlled. The effect of the degree of Cement Concr. Compos. 27 (2005) 637e649.
[5] K. Lu, Making strong nanomaterials ductile with gradients, Science 345 (2014)
gradient on mechanical properties of GS Ni specimens has been 1455e1456.
systematically investigated. The yield stress monotonically de- [6] F. Lefevre-Schilick, O. Bouaziz, Y. Brechet, J.D. Embury, Compositionally graded
creases with an increase in the n value, while the uniform elon- steels: the effect of partial decarburization on the mechanical properties of
spherodite and pearlite, Mater. Sci. Eng. A 491 (2008) 80e87.
gation reaches a maximum value of 8.9% at n ¼ 3, signifying an [7] T.H. Fang, W.L. Li, N.R. Tao, K. Lu, Revealing extraordinary intrinsic tensile
optimal grain size gradient for enhanced ductility which is even plasticity in gradient nano-grained copper, Science 331 (2011) 1587e1590.
better than that of the CG counterpart. As revealed by our experi- [8] H.N. Kou, J. Lu, Y. Li, High-strength and high-ductility nanostructured and
amorphous metallic materials, Adv. Mater. 26 (2014) 5518e5524.
mental observations and MD simulations, the surface roughening [9] X.L. Wu, P. Jiang, L. Chen, F.P. Yuan, Y.T. Zhu, Extraordinary strain hardening by
of coarse grains and strain localization of nano-grains can be gradient structure, Proc. Natl. Acad. Sci. U.S.A. 111 (2014) 7197e7201.
effectively suppressed due to the mutual constraint between nano- [10] Y.J. Wei, Y.Q. Li, L.C. Zhu, Y. Liu, X.Q. Lei, G. Wang, Y.X. Wu, Z.L. Mi, J.B. Liu,
H.T. Wang, H.J. Gao, Evading the strength-ductility trade-off dilemma in steel
grains and coarse grains. As a result, the GS Ni sample possessing through gradient hierarchical nanotwins, Nat. Commun. 5 (2014) 3580.
the optimal grain size gradient profile undergoes larger plastic [11] R. Thevamaran, O. Lawal, S. Yazdi, S.-J. Jeon, J.-H. Lee, E.L. Thomas, Dynamic
deformation than the homogeneous CG structure, resulting in su- creation and evolution of gradient nanostructure in single-crystal metallic
microcubes, Science 354 (2016) 312e316.
perior plasticity. This work provides a novel strategy for synthe-
[12] T. Roland, D. Retraint, K. Lu, J. Lu, Fatigue life improvement through surface
sizing GS materials possessing various degrees of gradient and an nanostructuring of stainless steel by means of surface mechanical attrition
additional insight into the excellent ductility in GS materials. treatment, Scripta Mater. 54 (2006) 1949e1954.
[13] L. Yang, N.R. Tao, K. Lu, L. Lu, Enhanced fatigue resistance of Cu with a gradient
nanograined surface layer, Scripta Mater. 68 (2013) 801e804.
Acknowledgements [14] H.W. Huang, Z.B. Wang, J. Lu, K. Lu, Fatigue behaviors of AISI 316L stainless
steel with a gradient nanostructured surface layer, Acta Mater. 87 (2015)
150e160.
This work was financially supported by National Natural Science [15] Z.B. Wang, N.R. Tao, S. Li, W. Wang, G. Liu, J. Lu, K. Lu, Effect of surface
Foundation of China under Grant Nos. 51471165 and 51401220. H.Z. nanocrystallization on friction and wear properties in low carbon steel, Mater.
and H.G. acknowledges support by the US National Science Foun- Sci. Eng. A 352 (2003) 144e149.
[16] L. Zhou, G. Liu, Z. Han, K. Lu, Grain size effect on wear resistance of a nano-
dation through grant DMR-1709318. The simulations reported were structured AISI52100 steel, Scripta Mater. 58 (2008) 445e448.
performed on resources provided by the Extreme Science and En- [17] Y. Sun, Sliding wear behaviour of surface mechanical attrition treated AISI 304
gineering Discovery Environment (XSEDE) through grant stainless steel, Tribol. Int. 57 (2013) 67e75.
[18] Y.S. Zhang, Z. Han, Fretting wear behavior of nanocrystalline surface layer of
MS090046.
pure copper under oil lubrication, Tribol. Lett. 27 (2007) 53e59.
[19] X. Chen, Z. Han, X. Li, K. Lu, Lowering coefficient of friction in Cu alloys with
References stable gradient nanostructures, Sci. Adv. 2 (2016), e1601942.
[20] J.J. Li, A.K. Soh, Modeling of the plastic deformation of nanostructured mate-
rials with grain size gradient, Int. J. Plast. 39 (2012) 88e102.
[1] M.A. Meyers, P.Y. Chen, A.Y.M. Lin, Y. Seki, Biological materials: structure and
[21] J.J. Li, G.J. Weng, S.H. Chen, X.L. Wu, On strain hardening mechanism in
mechanical properties, Prog. Mater. Sci. 53 (2008) 1e206.
gradient nanostructures, Int. J. Plast. 88 (2017) 89e107.
[2] H.D. Espinosa, J.E. Rim, F. Barthelat, M.J. Buehler, Merger of structure and
Y. Lin et al. / Acta Materialia 153 (2018) 279e289 289

[22] Z. Zeng, X.Y. Li, D.S. Xu, L. Lu, H. Gao, T. Zhu, Gradient plasticity in gradient Plastic deformation with reversible peak broadening in nanocrystalline nickel,
nano-grained metals, Extreme Mech. Lett. 8 (2016) 213e219. Science 304 (2004) 273e276.
[23] J.J. Li, S.H. Chen, X.L. Wu, A.K. Soh, J. Lu, The main factor influencing the tensile [41] A. Hasnaoui, H. Van Swygenhoven, P.M. Derlet, Dimples on nanocrystalline
properties of surface nano-crystallized graded materials, Mater. Sci. Eng. A fracture surfaces as evidence for shear plane formation, Science 300 (2003)
527 (2010) 7040e7044. 1550e1552.
[24] A.M. Rashidi, A. Amadeh, The effect of current density on the grain size of [42] X.L. Wu, M.X. Yang, F.P. Yuan, G.L. Wu, Y.J. Wei, X.X. Huang, Y.T. Zhu, Het-
electrodeposited nanocrystalline nickel coatings, Surf. Coating. Technol. 202 erogeneous lamella structure unites ultrafine-grain strength with coarse-
(2008) 3772e3776. grain ductility, Proc. Natl. Acad. Sci. U.S.A. 112 (2015) 14501e14505.
[25] A.M. Rashidi, A. Amadeh, The effect of saccharin addition and bath tempera- [43] X.L. Wu, F.P. Yuan, M.X. Yang, P. Jiang, C.X. Zhang, L. Chen, Y.G. Wei, E. Ma,
ture on the grain size of nanocrystalline nickel coatings, Surf. Coating. Tech- Nanodomained nickel unite nanocrystal strength with coarse-grain ductility,
nol. 204 (2009) 353e358. Sci. Rep. 5 (2015).
[26] G.J. Ackland, G. Tichy, V. Vitek, M.W. Finnis, Simple n-body potentials for the [44] Y.M. Wang, M.W. Chen, F.H. Zhou, E. Ma, High tensile ductility in a nano-
noble-metals and nickel, Philos. Mag. A 56 (1987) 735e756. structured metal, Nature 419 (2002) 912e915.
[27] S. Nose, A unified formulation of the constant temperature molecular- [45] Y. Xiang, T. Li, Z.G. Suo, J.J. Vlassak, High ductility of a metal film adherent on a
dynamics methods, J. Chem. Phys. 81 (1984) 511e519. polymer substrate, Appl. Phys. Lett. 87 (2005) 161910.
[28] F. Shimizu, S. Ogata, J. Li, Theory of shear banding in metallic glasses and [46] N.S. Lu, X. Wang, Z. Suo, J. Vlassak, Metal films on polymer substrates
molecular dynamics calculations, Mater. Trans. 48 (2007) 2923e2927. stretched beyond 50%, Appl. Phys. Lett. 91 (2007) 221909.
[29] X.L. Wu, P. Jiang, L. Chen, J.F. Zhang, F.P. Yuan, Y.T. Zhu, Synergetic [47] X.L. Lu, Q.H. Lu, Y. Li, L. Lu, Gradient confinement induced uniform tensile
Strengthening by gradient structure, Mater. Res. Lett. 2 (2014) 185e191. ductility in metallic glass, Sci. Rep. 3 (2013).
[30] X.C. Yang, X.L. Ma, J. Moering, H. Zhou, W. Wang, Y.L. Gong, J.M. Tao, Y.T. Zhu, [48] Z.T. Wang, J. Pan, Y. Li, C.A. Schuh, Densification and strain hardening of a
X.K. Zhu, Influence of gradient structure volume fraction on the mechanical metallic glass under tension at room temperature, Phys. Rev. Lett. 111 (2013)
properties of pure copper, Mater. Sci. Eng. A 645 (2015) 280e285. 135504.
[31] R. Becker, Effects of strain localization on surface roughening during sheet [49] F. Ebrahimi, Q. Zhai, D. Kong, Deformation and fracture of electrodeposited
forming, Acta Mater. 46 (1998) 1385e1401. copper, Scripta Mater. 39 (1998) 315e321.
[32] P.D. Wu, D.J. Lloyd, M. Jain, K.W. Neale, Y. Huang, Effects of spatial grain [50] F. Ebrahimi, G.R. Bourne, M.S. Kelly, T.E. Matthews, Mechanical properties of
orientation distribution and initial surface topography on sheet metal nanocrystalline nickel produced by electrodeposition, Nanostruct. Mater. 11
necking, Int. J. Plast. 23 (2007) 1084e1104. (1999) 343e350.
[33] M.R. Stoudt, J.B. Hubbard, M.A. Iadicola, S.W. Banovic, A study of the funda- [51] M. Legros, B.R. Elliott, M.N. Rittner, J.R. Weertman, K.J. Hemker, Microsample
mental relationships between deformation-induced surface roughness and tensile testing of nanocrystalline metals, Philos. Mag. A 80 (2000) 1017e1026.
strain localization in AA5754, Metall. Mater. Trans. (2009) 1611e1622. [52] P.G. Sanders, J.A. Eastman, J.R. Weertman, Elastic and tensile behavior of
[34] M.R. Stoudt, J.B. Hubbard, Analysis of deformation-induced surface mor- nanocrystalline copper and palladium, Acta Mater. 45 (1997) 4019e4025.
phologies in steel sheet, Acta Mater. 53 (2005) 4293e4304. [53] L. Lu, S.X. Li, K. Lu, An abnormal strain rate effect on tensile behavior in
[35] E. Ma, T. Zhu, Towards strength-ductility synergy through the design of het- nanocrystalline copper, Scripta Mater. 45 (2001) 1163e1169.
erogeneous nanostructures in metals, Mater. Today 20 (2017) 323e331. [54] H. Gao, Y. Huang, W.D. Nix, J.W. Hutchinson, Mechanism-based strain
[36] K. Osakada, M. Oyane, On the roughening of free surface in deformation gradient plasticity - I. Theory, J. Mech. Phys. Solids 47 (1999) 1239e1263.
processes, Bull. JSME 14 (1971) 171e177. [55] M.X. Yang, Y. Pan, F.P. Yuan, Y.T. Zhu, X.L. Wu, Back stress strengthening and
[37] M.R. Stoudt, R.E. Ricker, The relationship between grain size and the surface strain hardening in gradient structure, Mater. Res. Lett. 4 (2016) 145e151.
roughening behavior of Al-Mg alloys, Metall. Mater. Trans. 33 (2002) [56] W. Chen, Z.S. You, N.R. Tao, Z.H. Jin, L. Lu, Mechanically-induced grain
2883e2889. coarsening in gradient nano-grained copper, Acta Mater. 125 (2017) 255e264.
[38] R. Mahmudi, M. Mehdizadeh, Surface roughening during uniaxial and equi- [57] X.Z. Liao, A.R. Kilmametov, R.Z. Valiev, H.S. Gao, X.D. Li, A.K. Mukherjee,
biaxial stretching of 70-30 brass sheets, J. Mater. Process. Technol. 80e1 J.F. Bingert, Y.T. Zhu, High-pressure torsion-induced grain growth in electro-
(1998) 707e712. deposited nanocrystalline Ni, Appl. Phys. Lett. 88 (2006) 021909.
[39] E. Ma, Instabilities and ductility of nanocrystalline and ultrafine-grained [58] Y.B. Wang, B.Q. Li, M.L. Sui, S.X. Mao, Deformation-induced grain rotation and
metals, Scripta Mater. 49 (2003) 663e668. growth in nanocrystalline Ni, Appl. Phys. Lett. 92 (2008) 011903.
[40] Z. Budrovic, H. Van Swygenhoven, P.M. Derlet, S. Van Petegem, B. Schmitt,

You might also like