Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Materials Science & Engineering A 844 (2022) 143170

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

On the adhesion performance of gradient-structured Ni–P metallic coatings


Yan Lin a, 1, Fenghui Duan b, 1, Jie Pan a, b, **, Cheng Zhang a, Qi Chen c, Junyong Lu d, Lin Liu a, *
a
State Key Laboratory of Material Processing and Die & Mould Technology and School of Materials Science and Engineering, Huazhong University of Science and
Technology, Wuhan, 430074, China
b
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang, 110016, China
c
Wuhan National High Magnetic Field Center, Huazhong University of Science and Technology, Wuhan, 430074, China
d
National Key Laboratory of Science and Technology on Vessel Integrated Power System, Naval University of Engineering, Wuhan, 430033, China

A R T I C L E I N F O A B S T R A C T

Keywords: Hard metallic coatings can provide exceptional protection for industrial settings; however, they usually have a
Gradient structure uniform microstructure and suffer from poor adhesion with substrates because of stress/strain incompatibility
Ni–P coatings induced by the sharp interface. Here, a gradient structure that transitions from coarse-grained Ni to nano­
Electrodeposition
crystalline/amorphous Ni–P alloy has been designed to improve the adhesion performance of the electro­
Interface shear strength
Bending strength
deposited Ni–P coatings. Based on two measures of adhesion performance, a high interface shear strength of
~282.6 MPa and high bending strength of ~868.4 MPa have been determined for the gradient-structured Ni–P
coatings on the CuCrZr alloy substrate, which are increased, respectively, by ~90% and ~60% from those of the
conventional monolithic-structured Ni–P coating. Post-mortem fractographic analysis reveals that introducing
the gradient structure accommodates plastic deformation of the coating and provides exceptional crack arrest
capability, contributing to the enhancement of adhesion performances under both shearing and bending. The
present work not only reveals how the adhesion performance can be significantly improved in gradient-
structured Ni–P coatings, but also provides a promising methodology for manufacturing novel GS metallic
coatings with a combination of excellent surface functions and strong interface adhesion.

1. Introduction adhesion strength of the Ni–P coating deposited on the steel is merely in
the range of 85–144 MPa [6]). Ensuring a robust adhesion of the hard
Metallic coatings are workable and vital as they can provide excep­ metallic coatings on the soft substrate has always been a long-term
tional protection for equipment and industrial settings when serviced in challenge.
harsh environments. The advance in engineering technologies has Taking a leaf out of Nature’s playbook, a promising strategy to
imposed increasing demands for superior adhesion of metallic coatings enhance the properties of materials is creating gradient structure
to ensure their durability and to perform their excellent surface prop­ [7–13]. For example, the fiber content in bamboo and other plants in­
erties (e.g., high hardness and high corrosion resistance) [1,2]. For creases from the inner region outwards the outer region, forming a
example, the damage-resistant metallic coating should be well-adhered gradient structure that generates marvelous flexibility and strength [14,
to the Cu alloy rail to avoid the peeling-off failure under mechanical 15]. Recent attempts have utilized the concept of the gradient in coating
impact during electromagnetic launching. However, traditional structure design to avoid abrupt mechanical or chemical changes at the
damage-resistant metallic coatings (e.g., hard Cr coating [3] and interface between two different materials, and achieve a favorable
nanocrystalline/amorphous Ni–P coating [4,5]) have a uniform struc­ adhesion [16–20]. For instance, gradient Ni–Co/SiC coatings exhibited
ture. The significant difference in structure/properties between coating higher adhesion strength, as compared to those of the uniform nano­
and substrate usually causes a stress/strain incompatibility at the structured coatings with a microstructure similar to its outer surface
interface region during deformation, leading to poor adhesion (e.g., the [20]. Concerning metallic coatings, gradient structures have also been

* Corresponding author.
** Corresponding author. State Key Laboratory of Material Processing and Die & Mould Technology and School of Materials Science and Engineering, Huazhong
University of Science and Technology, Wuhan, 430074, China.
E-mail addresses: jpan@hust.edu.cn (J. Pan), lliu2000@mail.hust.edu.cn (L. Liu).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.msea.2022.143170
Received 18 February 2022; Received in revised form 10 April 2022; Accepted 16 April 2022
Available online 20 April 2022
0921-5093/© 2022 Elsevier B.V. All rights reserved.
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

tailored to improve adhesion and other properties [21–26]. For and current density (30–50 mA/cm2); while the chemical composition
example, the gradient-structured (GS) pure Ni coating with the grain gradient can be attained via the continuous additions of phosphorous
size changing from microscale near the substrate to nanoscale on the acid (0–7.5 g/L) and sodium hypophosphite (0–3 g/L). Sodium
outer surface can be bent over 180◦ without cracking or interface saccharin can also be used to adjust internal stress and inhibit cracking.
delamination [23]. In contrast, the nanocrystalline Ni coating ruptures To investigate the influence of gradient profile, two different sets of GS
and separates from the substrate. Likewise, the gradient composition Ni–P coatings (i.e., GS Ni–P I and II) were prepared.
and structure contribute to a decreased internal stress in Ni–Co coatings Monolithic-structured (MS) Ni and MS Ni–P coatings were also prepared
[24], enabling reliable adhesion and enhanced tribological properties. by electrodeposition for comparison. For simplicity, GS Ni–P I, GS Ni–P
The aforementioned studies suggest that metallic coatings having II, MS Ni, and MS Ni–P coatings deposited on the CuCrZr alloy plate
gradient structures can possess superior adhesion performances, in were termed as GS Ni–P/Cu I, GS Ni–P/Cu II, MS Ni/Cu, and MS
comparison with those having a monotonic microstructure or homoge­ Ni–P/Cu, respectively.
neous chemical composition. However, the systematical investigation
on how grain structure affects the adhesion performance is still lacking. 2.2. Material characterization
Furthermore, only a solo-type gradient structure was investigated in
most previous studies due to the limitation in the synthesis technology, Cross-sectional morphologies of various coatings were characterized
and the specific effect of gradient profile on the adhesion performance of by scanning electronic microscope (SEM Phenom XL G2) in the back-
metallic coatings has rarely been explored. These unsolved problems scattered electron (BSE) imaging mode. The P content profile was
hinder the optimization of adhesion in GS metallic coatings, and thus determined using SEM equipped with an energy dispersive X-ray de­
prevent the realization of their full potential in applications. tector. Microstructures of the as-deposited GS Ni–P coating and the
Recently, we reported the successful fabrication of GS Ni with coating/substrate interface were further investigated by transmission
various gradient profiles via electrodeposition [27–30] and revealed the electron microscope (TEM), high-resolution TEM (HRTEM), and energy-
influence of grain size gradient profile on mechanical properties. In the dispersive X-ray spectroscopy (EDS) (FEI Talos F200). The cross-
present work, we go a step further and synthesize GS Ni–P coatings by sectional TEM foils for microstructure characterization were prepared
adding the solute atom P in pure Ni to enhance the hardness of coatings using focused ion beam (FIB) (FEI Helios NanoLab DualBeam 650). The
while maintaining good interface adhesion. Our GS Ni–P coatings microhardness along the thickness direction of various coatings was
encompass dual gradients, i.e., the chemical composition gradient plus measured by a microhardness (Qness Q10 A+) tester using a Vickers
the grain size gradient, transitioning from coarse-grained (CG) pure Ni indenter with a load of 50 g and a dwell time of 10 s. To characterize and
to nanocrystalline/amorphous Ni–P alloy (P content is approximately 9 verify the hardness gradient profile of the coating, one microhardness
at.%). On these grounds, we demonstrate that GS Ni–P coatings can measurement was made at each point with appropriate spacing between
indeed be highly adhesive at the interface and be strong at the surface any two neighboring indentations (~50 μm), and the through-thickness
layer simultaneously, and reveal the effects of gradient structure and hardness profile was validated by repeated measurement.
gradient profile on the adhesion performance.
2.3. Mechanical characterization
2. Experimental methods
The interface shear strength of coatings was determined using a
2.1. Materials preparation home-made testing set-up shown in Fig. 1a. All the specimens have an
identical dimension of ~5.5 (width) × 10 (length) × 1 (thickness) mm3.
GS Ni–P coatings with a thickness of ~0.5 mm were deposited on the For shearing testing, two specimens were symmetrically clamped be­
soft CuCrZr alloy plate through a direct-current (DC) electrodepositing tween two rectangular fixture blocks, where the coating/substrate
set-up, as DC electrodeposition has high processing controllability and interface and the long edge of the lower block were positioned in one
good flexibility, and almost has no limitations on the coating thickness. plane, and the coating parts sandwiched between two blocks were
The basic plating bath was a modified Watts-type electrolyte. The bath forced with a loading rate of 2 mm/min (Fig. 1a). The nominal shear
composition and processing parameters have been given in Table 1. strength was calculated by formula F = P/S (where P is the maximum
According to the previous works [23,27,31–33], the grain size of coating load, and S is the total interface area). At least three tests were con­
can be refined by adding the grain refiner (e.g., sodium saccharin) and ducted for each coating.
increasing the current density; the P content of coating increases with The adhesion performance of the coating/substrate was also tested
the increasing concentrations of phosphorous acid and sodium hypo­ by a three-point bending test at room temperature with a loading rate of
phosphite in the electrolyte. Thus, the grain size gradient can be ach­ 0.5 mm/min (Zwick-Z020 mechanical system). The bending specimen
ieved with gradual increases in sodium saccharin content (0.5–15.5 g/L) (as illustrated in Fig. 1b) has a total thickness (h) of ~5.8 mm (including
the coating thickness (hc) of ~0.5 mm), a width (b) of 6 mm, a total
Table 1 length of 35 mm, and the span (l) between two supporting points is 30
Composition of plating bath and processing parameters for electroplating GS mm. The bending strength was calculated using the equation as follows
Ni–P coatings. [34]:
( )
Bath composition Nickel sulfate hexahydrate 100 g/L 4h Pf l 1 Ec (hc + ζ)
σ b = 0.8 ⋅ 1 − ⋅ ⋅ ⋅ (1)
Nickel chloride hexahydrate 45 g/L 3πl 4 M0 Es (hs − ζ)
Boric acid 40 g/L
where M0 = b(h3s - 3h2s ζ + 3hsζ2)/[3(ζ – hs)] + (EC/ES)b(h3c - 3h2c ζ +
Sodium dodecyl sulfate 0.2 g/L 3hcζ2)/[3(hs - ζ)], ζ= (ECh2s – ESh2c )/[2(EShs+ EChc)]. EC, ES, hc and hs are
Sodium saccharin 0.5–15.5 g/L Young’s modulus of the coating, Young’s modulus of the substrate,
Phosphorous acid 0–7.5 g/L coating thickness and substrate thickness, respectively. Here, ES = 130
GPa, EC for MS Ni–P and MS Ni coatings are 178 GPa [35] and 200 GPa,
Sodium hypophosphite 0–3 g/L
respectively; while EC for GS Ni–P coatings was calculated by a simple
Processing parameters Current density 30–50 mA/cm2 rule-of-mixture type of relationship [36].
Temperature 343 ± 1 K
Initial pH ~3.5
Substrate (CuCrZr alloy) 40 × 60 × 5 mm3

2
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

Fig. 1. (a) Schematic diagram illustrating a measure of interface adhesion strength for the coating under shearing. (b) Schematic diagram illustrating the three-point
bending test of the specimen with coating and substrate.

2.4. Fractographic characterization coatings clearly show a uniform fine grain structure (Fig. 2a) and a
coarse columnar grain structure (Fig. 2b), respectively. For the GS Ni–P
To discern the failure mode for different coatings under shearing and coatings (i.e., GS Ni–P I and GS Ni–P II), a typical grain size gradient
bending loads, the fracture surfaces and crack patterns were examined with microstructure varying from coarse columnar grains to fine grains
using an optical microscope (OM, Leica DFC 450) and SEM. along the thickness direction is presented (Fig. 2c and d). These two GS
Ni–P coatings display distinct gradient profiles, where the grain size
3. Results change in the GS Ni–P I coating is gentler than that in the GS Ni–P II
coating. This result implies that the gradient profile of Ni–P alloys can
3.1. Microstructure of GS coatings also be manipulated in a fashion similar to that of pure Ni reported
previously [27,28].
Backscattered-electron (BSE) images taken from MS Ni–P and MS Ni To provide the microstructure evolution in detail, the representative

Fig. 2. Typical backscattered-electron (BSE) images taken from the coating revealing the various grain structures: (a) MS Ni–P, (b) MS Ni, (c) GS Ni–P I, and (d) GS
Ni–P II, respectively.

3
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

microstructure of the GS Ni–P II coating at various locations along the bounds of hardness for the two GS Ni–P coatings are comparable to that
thickness direction was examined, as shown in Fig. 3. In the vicinity of of MS Ni–P and MS Ni coatings, respectively. Although the two GS
the interface (i.e., normalized distance is ~0), columnar grains with a coatings have the same maximum and minimum hardness values, the
width of ~5 μm are characterized (Fig. 3a). As the normalized distance hardness gradient profiles are quite different. The hardness gradient of
shifts to ~0.91, elongated ultrafine grains, ~300 nm in short axis, are GS Ni–P II is much steeper, reminiscent of the depictions of micro­
seen in the bright-field TEM image (Fig. 3b). With a further increase in structure evolution (Fig. 2b and c) and P content distribution (Fig. 3f).
normalized distance (~0.94), nanograins with a size of ~30 nm emerge Fig. 4b shows the grain size profiles as a function of the normalized
(Fig. 3c). The elongated diffraction in the selected area electron distance in various coatings. The width of coarse columnar grain was
diffraction patterns (SAED) indicates the formation of textured struc­ regarded as grain size and measured using the BSE image, while the size
ture, even in the ultrafine-grained regions (insets in Fig. 3b and c). For of fine grain was estimated based on the Hall-Petch relation [38]. It is
the topmost layer (i.e., normalized distance is approximately 1), tiny clear that both MS Ni–P and MS Ni coatings exhibit a homogeneous
grains with a size of ~10 nm are embedded in an amorphous matrix, grain size distribution with the average grain size of ~8 nm and ~7 μm,
showing an amorphous/nanocrystalline hybrid structure (Fig. 3d and e). respectively. For the two GS Ni–P coatings, the grain size continuously
In addition to the grain size gradient, there is also a chemical gradient decreases from ~6 μm to ~8 nm, covering three orders of magnitude in
along the deposition direction, as shown in Fig. 3f. As the normalized size. The upper and lower bounds of grain size in GS Ni–P coatings are
distance shifts from 0 to 1, P content gradually increases from 0 to ~9 at. comparable to the average grain sizes of MS Ni and MS Ni–P coatings,
% for the two GS Ni–P coatings. The P content in the topmost layer of the respectively. Moreover, the grain size change in the fine grain region is
coating falls into the composition range (i.e., 9–15 at.%) for forming an much gentler in GS Ni–P I than that in GS Ni–P II, leading to a higher
amorphous phase [37], which is consistent with the TEM observation in fraction of nano-grains (i.e., grain size is less than 100 nm). Specifically,
Fig. 3e. For the MS Ni–P coating, the P content varies slightly from ~8 to the volume fraction of nano-grains is up to ~70% in the GS Ni–P I, while
~4 at.% in thickness direction due to the excessive consumption of it is less than 10% in the GS Ni–P II.
phosphorous acid and sodium hypophosphite during the long-time
deposition. 3.3. Interface characterization of GS coatings

3.2. Hardness and grain size profiles of GS coatings One of the keys to the adhesion performance of the coatings is the
interface structure. No flaws, such as nano-voids and cracks, are
Fig. 4a presents the microhardness profiles of the two GS Ni–P observed in the TEM image taken from the interface of the GS Ni–P II
coatings. For comparison, the data obtained from MS Ni–P and MS Ni and Cu alloy substrate, as displayed in Fig. 5a. Fig. 5b–d shows a typical
coatings are also included. The microhardness of the MS Ni–P and MS Ni high angle annular dark-field (HAADF) image and the corresponding
coatings are 6.91 ± 0.25 GPa and 1.62 ± 0.13 GPa, respectively. Note EDX mapping around the interface. Apparently, the interface is tortuous
there is a slight fluctuation of hardness values along the thickness di­ and sharp at the nanometer level, without a chemical intermixing layer.
rection in MS Ni–P coating, mainly due to the moderate variation of P A typical HRTEM image in Fig. 5e reveals that the interface is
content (Fig. 3f). atomically bonded. The lattice planes run continuously across the in­
For both GS Ni–P I and II coatings, the microhardness continuously terfaces, indicating the epitaxial growth of the GS Ni–P coating on the Cu
increases from a minimum value of ~1.8 GPa at the interface (normal­ alloy substrate. The GS Ni–P coatings have a gradient structure tran­
ized distance of ~0) to a maximum value of ~7.2 GPa at the outmost sitioning from pure Ni to Ni–P alloy. The initial pure Ni deposits can
surface (normalized distance of ~1). Apparently, the upper and lower epitaxially grow on the (111) lattice plane of the Cu substrate because

Fig. 3. (a) Representative SEM image of coarse columnar grains on the CG side. TEM images at various normalized distances in the GS Ni–P II coating: (b) ~0.91, (c)
~0.94, and (d, e) ~1.0 (i.e., the topmost layer). Insets in (b) and (c) are the corresponding SAED patterns. (f) The P content profiles of MS Ni–P, GS Ni–P I, and GS
Ni–P II coatings.

4
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

Fig. 4. (a) Microhardness profiles of various coatings as a function of the normalized distance ranging from 0 to 1, i.e., from the substrate/coating interface to the
surface of the coating. (b) Grain size profiles of MS Ni–P, GS Ni–P I and GS Ni–P II coatings. Here the width of coarse columnar grain in the pure Ni regime was
determined using the BSE image, while the size of fine grain in the Ni–P alloy regime was calculated based on the Hall-Petch relation reported in Ref. [38].

Fig. 5. (a) TEM observation of the as-deposited interface between the GS Ni–P and Cu alloy substrate. (b) HAADF imaging of an enlarged area around substrate/
coating interface, and EDS mapping of (c) Cu and (d) Ni. (e) Cross-sectional HRTEM image showing the atomistic characteristics of the interface.

Cu and Ni have the same face-centered cubic structure and almost reaches about ~2890 N and then drops rapidly as interface fracture
identical atomic size (atomic radii of Ni and Cu are 0.124 nm and 0.128 ensues. This result indicates that the interface region of the MS Ni–P
nm, respectively) [39,40]. This epitaxial growth enables an excellent coating cannot accommodate any plastic strain. Conversely, MS Ni
metallurgical bonding for the GS Ni–P coatings in this study. coating exhibits a large plastic displacement beyond yielding. The
maximum load is ~5820 N, ~100% higher than that of the MS Ni–P
3.4. Shear-adhesion properties of various coatings coating. For both GS Ni–P coatings, the load-displacement responses are
analogous to the MS Ni coating, i.e., the load slowly ascends with a
Fig. 6a presents the load-displacement responses of the various remarkable plastic displacement beyond yielding. In particular, the
coatings upon shear testing. For MS Ni–P coating, a maximum load maximum loads for GS Ni–P I and II coatings are ~5960 and ~6490 N,

5
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

Fig. 6. (a) Typical load-displacement curves and (b) the corresponding interface shear strengths of the MS Ni–P/Cu, GS Ni–P/Cu I, GS Ni–P/Cu II, and MS Ni/Cu
obtained from shearing tests at room temperature.

respectively, which is close to that of the MS Ni coating (the slight dif­


ference is due to the specimen geometry). Based on the maximum load,
the interface shear strength of various coatings is calculated and sum­
marized in Fig. 6b and Table 2. The MS Ni–P coating has a low interface
shear strength of 154.7 ± 44.3 MPa, approximately half of that of the MS
Ni coating (289.3 ± 7.6 MPa). In contrast, GS Ni–P I and II coatings
exhibit the interface shear strengths of 280.1 ± 24.3 MPa and 282.6 ±
36.3 MPa, respectively, much higher than the MS Ni–P coating and
comparable to that of the MS Ni coating. These interface shear strengths
are even approximately equal to the shear strength of Cu alloy (~261.5
± 6.7 MPa). For the two GS Ni–P coatings, the gradient profile has no
influence on the interface shear strength, as both coatings have an
identical initial pure Ni deposit contiguous to the interface.
Fig. 7 presents fracture morphologies of the MS Ni–P, GS Ni–P II, and
MS Ni coatings after shearing testing. Due to the significant differences
in microstructure/property between the coating and the substrate and
the brittleness of the MS Ni–P alloy, neither the interface region nor the
coating could accommodate shear strain. As a result, local interface
delamination and cracking within the coating occurred in the MS Ni–P/
Cu after shear testing (Fig. 7b). Unlike the rupture of the MS Ni–P
coating, the replication of gradient structure can keep the integrity of
coating after shear testing (Fig. 7c), similar to that of MS Ni coating
(Fig. 7d). Furthermore, a thin film of Cu alloy substrate adheres to the
sheared-off coating (Fig. 7c), suggesting failure occurs within the Cu
alloy instead of the interface of the GS Ni–P and Cu alloy substrate. Fig. 7. (a) OM image showing the interface region between the coating and
substrate before shearing. Fracture morphologies for the (b) MS Ni–P, (c) GS
3.5. Bending behaviors of various coatings Ni–P II, and (d) MS Ni coatings after shearing testing. Insets in (b) show the
coating cracking and interface delamination.
To evaluate the adhesion performance of the coatings suffering from
more complicated and harsh load-bearing conditions, the three-point coating. The subsequent load increment is mainly due to strain-
bending test was also performed at room temperature. Fig. 8a shows hardening of the Cu alloy substrate upon plastic deformation. There­
the corresponding load-displacement curves of various coating/sub­ fore, the determined failure load for the MS Ni–P/Cu is ~2500 N. In
strate system (i.e., coating + substrate). For the MS Ni–P/Cu (i.e., the MS contrast, no load drop is detected in the load-displacement curve of the
Ni–P coating + the Cu alloy substrate, similarly hereinafter), a signifi­ MS Ni/Cu (Fig. 8a), which manifests that no coating failure occurred
cant load drop is observed at a displacement of 0.35 mm beside the small even after large plastic deformation. The load at the maximum
serrations at the early stage, indicating the cracking/failure of the displacement (~3800 N) is thus determined as the failure load.
The GS Ni–P/Cu I and II have distinct load-displacement curves
Table 2 under bending (Fig. 8a). The load-displacement curve of the GS Ni–P/Cu
Coating adhesion properties measured by shearing and three-point bending I is slightly serrated, and the magnitude of load serration is much smaller
tests. than that in MS Ni–P/Cu. The load increases firstly and then starts to
Specimens Shear strength Bending strength Surface hardness decline slowly, thus the peak load (~3610 N) is regarded as the failure
(MPa) (MPa) (GPa) load. Conversely, the load-displacement curve of the GS Ni–P/Cu II is
MS Ni–P 154.7 ± 44.3 534.5 6.72 ± 0.11
smooth without noticeable serrations. The load increases monotonously
MS Ni 289.3 ± 7.6 812.8 1.62 ± 0.13 during loading, thus the load at the maximum displacement (~4010 N)
GS Ni–P I 280.1 ± 24.3 736.7 7.04 ± 0.24 is the failure load. In addition, the load to sustain bending of the GS
GS Ni–P II 282.6 ± 36.3 868.4 6.91 ± 0.25 Ni–P/Cu II exceeds that of the MS Ni/Cu, implying a higher bending
Note that the surface hardness was measured at the normalized distance of 1.0. strength.

6
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

Fig. 8. (a) Typical load-displacement curves and (b) bending strengths of the MS Ni–P/Cu, GS Ni–P/Cu I, GS Ni–P/Cu II, and MS Ni/Cu obtained from three-point
bending tests at room temperature.

Based on Eq. (1), the bending strength of the coating/substrate is in the CG region of the GS coatings. More importantly, the interface of
calculated and presented in Fig. 8b and Table 2. Clearly, the MS Ni–P/Cu GS coating/substrate keeps intact, showing the good adhesion perfor­
exhibits a low bending strength of 534.5 MPa, ~50% lower than that of mance under bending.
the MS Ni/Cu (812.8 MPa). The GS Ni–P/Cu exhibits high bending
strength, but depending on the gradient profile. In particular, the 4. Discussion
bending strength of the GS Ni–P/Cu II is 868.4 MPa, which not only far
surpasses that of the MS Ni–P/Cu but exceeds that of the MS Ni/Cu. The present work reveals that conventional electrodeposition can
Fig. 9 shows fracture morphologies of the bending-tested coating/ yield a practical advance in creating gradient architecture with a broad
substrate specimens. For the MS Ni–P/Cu, there is only one vertical grain-size span and accurate control over gradient profile in Ni–P
crack crossing through the coating (Fig. 9a), without any plastic defor­ coatings. Introducing the gradient structure in the coating could
mation (Fig. 9b). With a further increase in bending strain, the stress significantly improve interfacial adhesion while maintaining high sur­
concentration around the vertical crack tip also results in interfacial face hardness, and tailoring the gradient profile could further enhance
crack initiation and propagation, responsible for the coating delamina­ the adhesion performances. In the following, how gradient structure and
tion (Fig. 9a). Conversely, the MS Ni/Cu could bear a high bending gradient profile affect the adhesion performances under shearing and
strain without cracking at the interface or within the coating (Fig. 9c), bending will be discussed. Moreover, given that our GS metallic coatings
showing excellent interface adhesion. What is visible are multiple slip combine the excellent interface adhesion and the superior coating sur­
bands at coating sides (Fig. 9c) and surface (Fig. 9d), indicating that the face properties, we offer a brief prospect for the engineering application
ductile MS Ni coating experienced a large plastic deformation. of the GS Ni–P coatings.
For the two GS Ni–P/Cu, cracks with a V-shaped profile are also
observed in the coatings (Fig. 9e and g), akin to the MS Ni–P coating. 4.1. Effect of gradient structure on the interface shear strength
However, the crack depths in the GS Ni–P I and GS Ni–P II coatings are
~430 and ~50 μm, respectively, indicating the evading through- According to the experimental results, the following reasons could
thickness cracking. The crack tips locate at the CG region with a grain contribute to the improved adhesion performance in the GS coatings. (1)
size of ~3 μm and ~1 μm for the GS Ni–P I and II coatings, respectively There is excellent metallurgical bonding between the GS Ni–P coating
(see Figs. 2 and 4b), implying that the propagation of cracks is arrested and the Cu alloy substrate. Epitaxial growth was allowed for our GS Ni–P

Fig. 9. Side-view and plan-view fracture morphologies of the (a, b) MS Ni–P/Cu, (c, d) MS Ni/Cu, (e, f) GS Ni–P/Cu I, and (g, h) GS Ni–P/Cu II under bending. Insets
in (a, b, e, f, g, h) showing the morphology neighboring crack path. Inset in (c) is an enlarged SEM image showing slip bands left by plastic deformation of MS
Ni coating.

7
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

coatings electrodeposited on the Cu alloy substrate due to the unique can have extraordinary adhesion performance, and is much better than
gradient structure design (i.e., transitioning from pure Ni to Ni–P alloy) the MS Ni–P coating and even superior to the MS Ni coating.
and the low lattice mismatch between the initial pure Ni deposits and
the Cu substrate. (2) Gradient structure has an advantage in accom­
modating stress/strain incompatibility between the coating and the 4.3. Hardness-adhesion synergy of GS Ni–P coatings: implication for
substrate during deformation. For the MS Ni–P coating, a larger hard­ engineering applications
ness difference (the hardness of substrate and coating are ~1.5 GPa and
~7.2 GPa, respectively, at two sides of the interface) causes the stress/ Hardness is of great importance for most protective metal coatings;
strain incompatibility at the interface during shearing and bending, however, the traditional hard coating with a homogeneous structure is
resulting in the crack initiation at the interface or interface delamina­ usually inclined to encounter poor adhesion. In other words, there is a
tion. Conversely, for the GS Ni–P coating, minor hardness discontinuity trade-off between hardness and adhesion performance in metallic
across the interface (the hardness of substrate and coating are ~1.5 GPa coatings. In the present work, the surface hardness of the MS Ni–P
and ~1.8 GPa, respectively, at two sides of the interface) is beneficial for coating is ~6.7 GPa, 400% higher than that of MS Ni coating. However,
accommodating the deformation of the interface region during shearing the interface shear and bending strengths are 154.7 MPa and 534.5 MPa,
and bending. (3) The deformation ability of Ni–P alloy is improved due 50% and 20%, respectively, lower than those of the MS Ni coating,
to the gradient structure, contributing to keeping the intactness of the following the above trade-off (Fig. 10). In contrast, our GS Ni–P coatings
coating during deformation. This item will be further discussed below. simultaneously possess a high surface hardness of ~7.2 GPa, an inter­
As revealed in the thin film/substrate system [41,42], the deforma­ face shear strength of ~280 MPa, and a bending strength over 730 MPa.
tion capability of the interface region is vital to interface adhesion and These results show that the adhesion performances of the Ni–P coating
interface fracture resistance. The unique characteristic of the GS Ni–P have significantly been improved by tailoring a gradient structure
coatings is that their gradient structure transitions from soft-yet-ductile without compromising its high hardness. Namely, the superior hardness-
Ni to hard-yet-brittle Ni–P alloy. The initial soft Ni layer can be adhesion synergy has been achieved in our GS Ni–P coatings.
co-deformed well with the soft substrate, giving rise to the notable strain Regarding the coating structure design for practical application, the
hardening during deformation and ductile shearing in the Cu alloy side findings in this study show that the gradient architecture has advantages
(Fig. 7c). In contrast, no plastic deformation occurs in the interface re­ over the conventional MS nanocrystalline/amorphous and CG struc­
gion of MS Ni–P coating during shearing (Fig. 7b) because of the brit­ tures. For instance, the conventional hard chromium deposits have
tleness of nanocrystalline/amorphous Ni–P alloy, leading to the failure relatively good adhesion (i.e., the interface shear is over 200 MPa [45]),
of MS Ni–P coating and interface delamination. In a word, the failure but are prone to suffer from interfacial cracking and through-thickness
mode transits from cracking or interface delamination to ductile tearing, cracking subjected to bending [46,47], significantly limiting their ap­
enabling the GS Ni–P coatings to evade poor adhesion of the MS Ni–P plications in the fields evolving bending impact. In contrast, our GS Ni–P
coating and achieve the same strong adhesion as the MS Ni coating II coating, with a high surface hardness close to that of the hard chro­
during shearing. mium deposit (~7–9 GPa [48,49]), can possess a superior adhesion
For the coatings under bending, the brittleness of MS Ni–P coating performance under shearing and bending, thus widening the application
leads to easy initiation of cracks at the surface, followed by the through- fields.
thickness cracking and interface delamination (Fig. 9a). For the GS Ni–P It should be pointed out that the present work is only a starting point,
coating, although bending-induced cracking occurs, the interface keeps and the gradient structure of the GS Ni–P coatings can be further opti­
intact as all the cracks are not able to pass through the entire GS Ni–P mized for withstanding more complicated service conditions. Taking the
coatings (Fig. 9e and g). The crack arrest is attributed to the crack GS Ni–P I coating as an example, it is still susceptible to the bending load
blunting induced by plastic deformation of the CG region. This positive to some extent due to a higher volume fraction of brittle nanocrystalline
role of gradient structure in arresting crack contributes to the excep­ structure. This problem can be readily resolved by tuning the gradient
tional damage tolerance for GS Ni–P coatings under bending [28,30,43, profile to that shown in the GS Ni–P II coating (i.e., a smaller volume
44]. fraction of brittle nanocrystalline structure). The accurate control over
the gradient profile is one of the most attractive findings of this work.
4.2. Effect of gradient profile on adhesion performances Furthermore, the material class of the coating (e.g., Ni–W alloy [50,51],
Ni–W–P alloy [52], Ni–Co alloy [53] and Ni–Mo alloy [54, 55]) can be
The formation of gradient structure is the only first step while readily enriched to acquire a more favorable surface function. As such,
patterning gradient profile plays a more significant role in manipulating we believe that our work provides a promising strategy for producing GS
properties or functions. Regarding the shear loading, the interface metallic coatings with a combination of excellent surface properties (e.
strength is mainly governed by the interface region. Recall that both GS g., high hardness) and strong interface adhesion. These GS metallic
Ni–P I and II coatings have a ~100 μm thick layer of CG Ni that is coatings have the potential to replace hard chrome coatings and can be
contiguous to the substrate, the shear strengths of two GS coatings are used in various engineering fields (e.g., protective coatings of electro­
identical and analogous to that of the MS Ni coating (Fig. 6). Therefore, magnetic launching rail) that traditional coatings were incompetent.
the gradient profile has no impact on the interface shear strength of GS
Ni–P coatings. 5. Conclusions
Conversely, gradient profile makes a significant difference in the
bending performance of GS Ni–P coatings. The interface region of two In this work, the gradient-structured (GS) Ni–P coatings, with a wide
GS coatings remains intact; however, the deformation behavior of the gradient spanning from the inner pure Ni to the outer Ni–P alloy, have
coating is completely different. In contrast to the few deep cracks and been successfully fabricated via electrodeposition. Based on the exper­
smooth surface presented in the bending-tested GS Ni–P I coating imental results of the adhesion performance for the coating/substrate (i.
(Fig. 9e and f), numerous shallow cracks and multiple shear bands near e., the GS Ni–P coating on the CuCrZr alloy substrate), the main con­
the crack are observed in the GS Ni–P II coating (Fig. 9g and f), implying clusions can be obtained:
that the surface layer has undergone plastic deformation before crack
initiation. Combining the excellent interface adhesion and the superior 1. A dual-gradient architecture was achieved in the Ni–P coatings, in
plastic deformability, GS Ni–P/Cu II exhibits a bending strength larger which the P content changes from 0 to ~9 at.%, and the corre­
than that of the GS Ni–P/Cu I, and even surpasses that of the MS Ni/Cu. sponding microstructure transitions from coarse columnar grains of
In short, the GS Ni–P coating with proper regulation of gradient profile ~6 μm to nanocrystalline/amorphous hybrids.

8
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

Fig. 10. Plots showing (a) the interface shear strength and (b) the bending strength as a function of the hardness of the coating surface.

2. Two measures of the adhesion performance showed that the GS Ni–P References
II coating possessed an interface shear strength of ~280 MPa and a
bending strength of ~870 MPa, which are ~80% and ~60% higher, [1] A. Karimzadeh, M. Aliofkhazraei, F.C. Walsh, A review of electrodeposited Ni-Co
alloy and composite coatings: microstructure, properties and applications, Surf.
respectively, than those of the MS Ni–P coating. Coating. Technol. 372 (2019) 463–498.
3. The generation of gradient structure can reduce stress/strain in­ [2] S. Wang, C. Ma, F. Walsh, Alternative tribological coatings to electrodeposited hard
compatibility and accommodate plastic deformation of the interface chromium: a critical review, Trans. IMF 98 (4) (2020) 173–185.
[3] D. Wu, J. Zhang, J. Huang, H. Bei, T.G. Nieh, Grain-boundary strengthening in
region under shear loading, and can impart the exceptional crack nanocrystalline chromium and the Hall–Petch coefficient of body-centered cubic
arrest capability to the coating, leading to enhanced adhesion per­ metals, Scripta Mater. 68 (2) (2013) 118–121.
formances under both shearing and bending. [4] R. Cao, J. Pan, Y. Lin, Y. Li, Mechanical properties and optimum layer thickness in
an amorphous Ni–P/coarse-grained Ni bi-layered structure, Mater. Sci. Eng., A 760
4. Gradient profile has no influence on the interface shear strength but (2019) 458–468.
dramatically impacts the bending behavior of GS Ni–P coatings. [5] A. Yonezu, M. Niwa, J. Ye, X. Chen, Contact fracture mechanism of electroplated
5. This work provides a promising methodology for producing GS Ni–P coating upon stainless steel substrate, Mater. Sci. Eng., A 563 (2013)
184–192.
metallic coatings with the combinations of excellent surface prop­
[6] K. Kanamori, Y. Kimoto, S. Toriumi, A. Yonezu, Evaluation of adhesion durability
erties (e.g., the high surface hardness) and strong interface adhesion, of Ni–P coating using repeated Laser Shock Adhesion Test, Surf. Coating. Technol.
and hence extends their engineering applications. 396 (2020) 125953.
[7] Z. Liu, M.A. Meyers, Z. Zhang, R.O. Ritchie, Functional gradients and
heterogeneities in biological materials: design principles, functions, and
Data availability bioinspired applications, Prog. Mater. Sci. 88 (2017) 467–498.
[8] S.E. Naleway, M.M. Porter, J. McKittrick, M.A. Meyers, Structural design elements
in biological materials: application to bioinspiration, Adv. Mater. 27 (37) (2015)
Experimental data from this study are available from Prof. Lin Liu
5455–5476.
and Prof. Jie Pan from Huazhong University of Science and Technology [9] U.G. Wegst, H. Bai, E. Saiz, A.P. Tomsia, R.O. Ritchie, Bioinspired structural
upon reasonable request. materials, Nat. Mater. 14 (1) (2015) 23–36.
[10] D. Kokkinis, F. Bouville, A.R. Studart, 3D printing of materials with tunable failure
via bioinspired mechanical gradients, Adv. Mater. 30 (19) (2018) 1705808.
CRediT authorship contribution statement [11] S. Zhang, J. Xie, Q. Jiang, X. Zhang, C. Sun, Y. Hong, Fatigue crack growth
behavior in gradient microstructure of hardened surface layer for an axle steel,
Mater. Sci. Eng., A 700 (2017) 66–74.
Yan Lin: Investigation, Methodology, experiments, Writing – orig­
[12] T.H. Fang, W.L. Li, N.R. Tao, K. Lu, Revealing extraordinary intrinsic tensile
inal draft, Funding acquisition. Fenghui Duan: experiments, Writing – plasticity in gradient nano-grained copper, Science 331 (6024) (2011) 1587–1590.
original draft. Jie Pan: Formal analysis, Visualization, Writing – review [13] K. Lu, Making strong nanomaterials ductile with gradients, Science 345 (6203)
& editing, Funding acquisition. Cheng Zhang: Formal analysis, Visu­ (2014) 1455–1456.
[14] S. Amada, Hierarchical functionally gradient structures of bamboo, barley, and
alization. Qi Chen: Formal analysis, Visualization. Junyong Lu: Formal corn, MRS Bull. 20 (1) (1995) 35–36.
analysis. Lin Liu: Resources, Supervision, Writing – review & editing, [15] S. Amada, Y. Ichikawa, T. Munekata, Y. Nagase, H. Shimizu, Fiber texture and
Funding acquisition. mechanical graded structure of bamboo, Composites, Part B 28 (1–2) (1997)
13–20.
[16] Y.J. Won, H. Ki, Effect of film gradient profile on adhesion strength, residual stress
and effective hardness of functionally graded diamond-like carbon films, Appl.
Declaration of competing interest Surf. Sci. 311 (2014) 775–779.
[17] S. Nißen, J. Heeg, M. Warkentin, D. Behrend, M. Wienecke, The effect of deposition
The authors declare that they have no known competing financial parameters on structure, mechanical and adhesion properties of aC: H on Ti6Al4V
with gradient Ti-aC: H: Ti interlayer, Surf. Coating. Technol. 316 (2017) 180–189.
interests or personal relationships that could have appeared to influence
[18] Z.Y. Chen, J. Zhao, X.H. Meng, J.M. Li, Evaluation of fatigue resistance of a
the work reported in this paper. gradient CrNx coating applied to turbine blades, Mater. Sci. Eng., A 527 (6) (2010)
1436–1443.
[19] M. Sathish, N. Radhika, B. Saleh, A critical review on functionally graded coatings:
Acknowledgement methods, properties, and challenges, Composites, Part B 225 (2021) 109278.
[20] S.L. Baghal, M.H. Sohi, A. Amadeh, A functionally gradient nano-Ni–Co/SiC
This work is supported by the National Natural Science Foundation composite coating on aluminum and its tribological properties, Surf. Coating.
Technol. 206 (19–20) (2012) 4032–4039.
of China (Nos. 92066202, 52001075, and 52022100), and funded by
[21] L.P. Wang, Y. Gao, Q.J. Xue, H.W. Liu, T. Xu, A novel electrodeposited Ni-P
China Postdoctoral Science Foundation (No. 2021M701290). Y. L. is also gradient deposit for replacement of conventional hard chromium, Surf. Coating.
grateful for support from GDAS’ Projects of the Construction of the Technol. 200 (12–13) (2006) 3719–3726.
[22] W. Sun, P. Zhang, F. Zhang, W. Hou, K. Zhao, Microstructure and corrosion
Domestic First-class Research Organization (2019GDASYL-0103074). J.
resistance of Ni–P gradient coatings, Trans. IMF 93 (4) (2015) 180–185.
P. is also grateful for support from the Youth Innovation Promotion
Association of the Chinese Academy of Sciences (No. 2020194).

9
Y. Lin et al. Materials Science & Engineering A 844 (2022) 143170

[23] L. Qin, J. Xu, J. Lian, Z. Jiang, Q. Jiang, A novel electrodeposited nanostructured [40] T. Yan, Y.C. Huang, Y.C. Hou, L. Chang, Epitaxial growth and microstructural
Ni coating with grain size gradient distribution, Surf. Coating. Technol. 203 (1–2) evolution of nickel electrodeposited on a polycrystalline copper substrate,
(2008) 142–147. J. Electrochem. Soc. 165 (14) (2018) D743.
[24] L. Wang, Y. Gao, Q. Xue, H. Liu, T. Xu, Graded composition and structure in [41] M. Lane, R.H. Dauskardt, A. Vainchtein, H. Gao, Plasticity contributions to
nanocrystalline Ni–Co alloys for decreasing internal stress and improving interface adhesion in thin-film interconnect structures, J. Mater. Res. 15 (12)
tribological properties, J. Phys. D Appl. Phys. 38 (8) (2005) 1318. (2000) 2758–2769.
[25] L. Jiang, X. Cui, G. Jin, H. Tian, Z. Tian, X. Zhang, S. Wan, Synthesis and [42] M. Cordill, N. Moody, D. Bahr, The effects of plasticity on adhesion of hard films on
microstructure, properties characterization of Ni-Ti-Cu/Cu-Al functionally graded ductile interlayers, Acta Mater. 53 (9) (2005) 2555–2562.
coating on Mg-Li alloy by laser cladding, Appl. Surf. Sci. 575 (2022) 151645. [43] F. Brenne, T. Niendorf, Damage tolerant design by microstructural
[26] Q. Wang, Q. Li, L. Zhang, D.X. Chen, H. Jin, J.D. Li, J.W. Zhang, C.Y. Ban, gradation–influence of processing parameters and build orientation on crack
Microstructure and properties of Ni-WC gradient composite coating prepared by growth within additively processed 316L, Mater. Sci. Eng., A 764 (2019) 138186.
laser cladding, Ceram. Int. 48 (6) (2022) 7905–7917. [44] R. Cao, Q. Yu, Y. Li, R.O. Ritchie, Dual-gradient structure leads to optimized
[27] Y. Lin, J. Pan, H.F. Zhou, H.J. Gao, Y. Li, Mechanical properties and optimal grain combination of high fracture resistance and strength-ductility synergy with
size distribution profile of gradient grained nickel, Acta Mater. 153 (2018) minimized final catastrophic failure, J. Mater. Res. Technol. 15 (2021) 901–910.
279–289. [45] E. Zmihorski, Shear tests of the adhesion of electrodeposited chromium to steel,
[28] Y. Lin, Q. Yu, J. Pan, F. Duan, R.O. Ritchie, Y. Li, On the impact toughness of Trans. IMF 23 (1) (1947) 203–213.
gradient-structured metals, Acta Mater. 193 (2020) 125–137. [46] J. Jiang, M. Du, Z. Pan, M. Yuan, X. Ma, B. Wang, Effects of oxidation and inter-
[29] Y. Lin, J. Pan, Z. Luo, Y. Lu, K. Lu, Y. Li, A grain-size-dependent structure evolution diffusion on the fracture mechanisms of Cr-coated Zry-4 alloys: an in situ three-
in gradient-structured (GS) Ni under tension, Nano Mater. Sci. 2 (1) (2020) 39–49. point bending study, Mater. Des. (2021) 110168.
[30] R. Cao, Q. Yu, J. Pan, Y. Lin, A. Sweet, Y. Li, R.O. Ritchie, On the exceptional [47] J. Jiang, M. Yuan, M. Du, X. Ma, On the crack propagation and fracture properties
damage-tolerance of gradient metallic materials, Mater. Today 32 (2019) 94–107. of Cr-coated Zr-4 alloys for accident-tolerant fuel cladding: in situ three-point
[31] A.M. Pillai, A. Rajendra, A.K. Sharma, Electrodeposited nickel–phosphorous (Ni–P) bending test and cohesive zone modeling, Surf. Coating. Technol. (2021) 127810.
alloy coating: an in-depth study of its preparation, properties, and structural [48] Y. Liang, Y.S. Li, Q.Y. Yu, Y.X. Zhang, W.J. Zhao, Z.X. Zeng, Structure and wear
transitions, J. Coating Technol. Res. 9 (6) (2012) 785–797. resistance of high hardness Ni-B coatings as alternative for Cr coatings, Surf.
[32] T. Mahalingam, M. Raja, S. Thanikaikarasan, C. Sanjeeviraja, S. Velumani, Coating. Technol. 264 (2015) 80–86.
H. Moon, Y.D. Kim, Electrochemical deposition and characterization of Ni–P alloy [49] Z. Zeng, L. Wang, L. Chen, J. Zhang, The correlation between the hardness and
thin films, Mater. Char. 58 (8–9) (2007) 800–804. tribological behaviour of electroplated chromium coatings sliding against ceramic
[33] Y.E. Sknar, O. Savchuk, I. Sknar, Characteristics of electrodeposition of Ni and Ni-P and steel counterparts, Surf. Coating. Technol. 201 (6) (2006) 2282–2288.
alloys from methanesulfonate electrolytes, Appl. Surf. Sci. 423 (2017) 340–348. [50] T. Yamasaki, P. Schlomacher, K. Ehrlich, Y. Ogino, Formation of amorphous
[34] H. Deng, H. Shi, S. Tsuruoka, Influence of coating thickness and temperature on electrodeposited Ni-W alloys and their nanocrystallization, Nanostruct. Mater. 10
mechanical properties of steel deposited with Co-based alloy hardfacing coating, (3) (1998) 375–388.
Surf. Coating. Technol. 204 (23) (2010) 3927–3934. [51] K. Sriraman, S.G.S. Raman, S. Seshadri, Synthesis and evaluation of hardness and
[35] Y. Zhou, U. Erb, K. Aust, G. Palumbo, The effects of triple junctions and grain sliding wear resistance of electrodeposited nanocrystalline Ni–W alloys, Mater. Sci.
boundaries on hardness and Young’s modulus in nanostructured Ni–P, Scripta Eng., A 418 (1–2) (2006) 303–311.
Mater. 48 (6) (2003) 825–830. [52] K. Sriraman, S.G.S. Raman, S. Seshadri, Corrosion behaviour of electrodeposited
[36] M.A. Meyers, K.K. Chawla, Mechanical Behavior of Materials, Cambridge nanocrystalline Ni–W and Ni–Fe–W alloys, Mater. Sci. Eng., A 460 (2007) 39–45.
University Press2009. [53] C. Gu, J. Lian, Q. Jiang, Z. Jiang, Ductile–brittle–ductile transition in an
[37] X. Yuan, D. Sun, H. Yu, H. Meng, Z. Fan, X. Wang, Preparation of amorphous- electrodeposited 13 nanometer grain sized Ni–8.6 wt.% Co alloy, Mater. Sci. Eng.,
nanocrystalline composite structured Ni–P electrodeposits, Surf. Coating. Technol. A 459 (1–2) (2007) 75–81.
202 (2) (2007) 294–300. [54] E. Beltowska-Lehman, A. Bigos, P. Indyka, M. Kot, Electrodeposition and
[38] L. Chang, P. Kao, C.H. Chen, Strengthening mechanisms in electrodeposited Ni–P characterisation of nanocrystalline Ni–Mo coatings, Surf. Coating. Technol. 211
alloys with nanocrystalline grains, Scripta Mater. 56 (8) (2007) 713–716. (2012) 67–71.
[39] A. Haseeb, J.P. Celis, J. Roos, Dual-Bath electrodeposition of Cu/Ni [55] J. Hu, Y. Shi, X. Sauvage, G. Sha, K. Lu, Grain boundary stability governs hardening
compositionally modulated multilayers, J. Electrochem. Soc. 141 (1) (1994) 230. and softening in extremely fine nanograined metals, Science 355 (6331) (2017)
1292–1296.

10

You might also like