16
materials Article Effects of Bi Addition on the Microstructure and Mechanical Properties of Nanocrystalline Ag Coatings Yuxin Wang 1,2 ID , Guang Cheng 3, * ID , See Leng Tay 2 , Yunxia Guo 1 , Xin Sun 4 and Wei Gao 2 1 School of Materials Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, Jiangsu, China; [email protected] (Y.W.); [email protected] (Y.G.) 2 Department of Chemical & Materials Engineering, the University of Auckland, PB 92019, Auckland 1142, New Zealand; [email protected] (S.L.T.); [email protected] (W.G.) 3 Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, USA 4 Energy and Transportation Science Division, Oak Ridge National Laboratory, 1 Bethel Valley Road, Oak Ridge, TN 37830, USA; [email protected] * Correspondence: [email protected]; Tel.: +1-509-375-4436 Received: 15 July 2017; Accepted: 7 August 2017; Published: 10 August 2017 Abstract: In this study we investigated the effects of Bi addition on the microstructure and mechanical properties of an electrodeposited nanocrystalline Ag coating. Microstructural features were investigated with transmission electron microscopy (TEM). The results indicate that the addition of Bi introduced nanometer-scale Ag-Bi solid solution particles and more internal defects to the initial Ag microstructures. The anisotropic elastic-plastic properties of the Ag nanocrystalline coating with and without Bi addition were examined with nanoindentation experiments in conjunction with the recently-developed inverse method. The results indicate that the as-deposited nanocrystalline Ag coating contained high mechanical anisotropy. With the addition of 1 atomic percent (at%) Bi, the anisotropy within Ag-Bi coating was very small, and yield strength of the nanocrystalline Ag-Bi alloy in both longitudinal and transverse directions were improved by over 100% compared to that of Ag. On the other hand, the strain-hardening exponent of Ag-Bi was reduced to 0.055 from the original 0.16 of the Ag coating. Furthermore, the addition of Bi only slightly increased the electrical resistivity of the Ag-Bi coating in comparison to Ag. Results of our study indicate that Bi addition is a promising method for improving the mechanical and physical performances of Ag coating for electrical contacts. Keywords: electrodeposited nanocrystalline Ag; Bi addition; microstructure; nanoindentation; mechanical properties; electrical conductivity 1. Introduction Silver (Ag) has been widely applied in the semiconductor industry because of its excellent thermal and electrical conductivity [13]. However, the wear resistance of Ag is not as desirable as those of nickel (Ni) or copper (Cu) [35] due to the fact that bulk Ag or Ag coatings are soft, with low hardness. Various methods were attempted to improve the mechanical performance of Ag coatings, such as introducing hard nanoparticles to form Ag-based composites [68], creating new Ag-based alloys with different crystal structures [1,5,913], and reducing the grain size to the hundred nanometer scale [9,1418]. In addition, different coating/film deposition techniques, i.e., electrodeposition, physical vapor deposition (PVD), chemical vapor deposition (CVD), and magnetron sputtering, have been employed to prepare Ag coatings with improved properties [19]. Among those techniques, Materials 2017, 10, 932; doi:10.3390/ma10080932 www.mdpi.com/journal/materials

1,2 ID 3, ID 2

  • Upload
    others

  • View
    17

  • Download
    0

Embed Size (px)

Citation preview

Page 1: 1,2 ID 3, ID 2

materials

Article

Effects of Bi Addition on the Microstructure andMechanical Properties of NanocrystallineAg Coatings

Yuxin Wang 1,2 ID , Guang Cheng 3,* ID , See Leng Tay 2, Yunxia Guo 1, Xin Sun 4 and Wei Gao 2

1 School of Materials Science and Engineering, Jiangsu University of Science and Technology,Zhenjiang 212003, Jiangsu, China; [email protected] (Y.W.); [email protected] (Y.G.)

2 Department of Chemical & Materials Engineering, the University of Auckland, PB 92019, Auckland 1142,New Zealand; [email protected] (S.L.T.); [email protected] (W.G.)

3 Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, P.O. Box 999,Richland, WA 99352, USA

4 Energy and Transportation Science Division, Oak Ridge National Laboratory, 1 Bethel Valley Road,Oak Ridge, TN 37830, USA; [email protected]

* Correspondence: [email protected]; Tel.: +1-509-375-4436

Received: 15 July 2017; Accepted: 7 August 2017; Published: 10 August 2017

Abstract: In this study we investigated the effects of Bi addition on the microstructure andmechanical properties of an electrodeposited nanocrystalline Ag coating. Microstructural featureswere investigated with transmission electron microscopy (TEM). The results indicate that the additionof Bi introduced nanometer-scale Ag-Bi solid solution particles and more internal defects to the initialAg microstructures. The anisotropic elastic-plastic properties of the Ag nanocrystalline coating withand without Bi addition were examined with nanoindentation experiments in conjunction with therecently-developed inverse method. The results indicate that the as-deposited nanocrystalline Agcoating contained high mechanical anisotropy. With the addition of 1 atomic percent (at%) Bi, theanisotropy within Ag-Bi coating was very small, and yield strength of the nanocrystalline Ag-Bialloy in both longitudinal and transverse directions were improved by over 100% compared to thatof Ag. On the other hand, the strain-hardening exponent of Ag-Bi was reduced to 0.055 from theoriginal 0.16 of the Ag coating. Furthermore, the addition of Bi only slightly increased the electricalresistivity of the Ag-Bi coating in comparison to Ag. Results of our study indicate that Bi additionis a promising method for improving the mechanical and physical performances of Ag coating forelectrical contacts.

Keywords: electrodeposited nanocrystalline Ag; Bi addition; microstructure; nanoindentation;mechanical properties; electrical conductivity

1. Introduction

Silver (Ag) has been widely applied in the semiconductor industry because of its excellent thermaland electrical conductivity [1–3]. However, the wear resistance of Ag is not as desirable as those ofnickel (Ni) or copper (Cu) [3–5] due to the fact that bulk Ag or Ag coatings are soft, with low hardness.Various methods were attempted to improve the mechanical performance of Ag coatings, such asintroducing hard nanoparticles to form Ag-based composites [6–8], creating new Ag-based alloyswith different crystal structures [1,5,9–13], and reducing the grain size to the hundred nanometerscale [9,14–18]. In addition, different coating/film deposition techniques, i.e., electrodeposition,physical vapor deposition (PVD), chemical vapor deposition (CVD), and magnetron sputtering, havebeen employed to prepare Ag coatings with improved properties [19]. Among those techniques,

Materials 2017, 10, 932; doi:10.3390/ma10080932 www.mdpi.com/journal/materials

Page 2: 1,2 ID 3, ID 2

Materials 2017, 10, 932 2 of 16

electrodeposition has attracted a great deal of research interest since this technique could preparemetallic alloys with unique compositions and novel micro/nanostructures, and, therefore, improveperformance significantly [7,14,20].

An ideal electrical contact material should possess high electrical conductivity and high wearresistance/hardness. However, among the various electrodeposited Ag alloy systems, the electricalconductivity and strength of alloys have always been two competing factors. In order to obtain strongerAg alloys, higher alloying compositions are typically employed which increase electrical resistivityand, therefore, compromise conductivity [21]. Naturally, an Ag alloy system with low solubilitycould become a good candidate for realizing desired performance. Among different Ag-based alloysystems, the solubility of bismuth in solid-state silver is very low (~1.5 at%), according to the phasediagram [22], as well as the silver-bismuth (Ag-Bi) systems that have been attempted in previousstudies with mechanical alloying [11] and electrodeposition [3,23]. The small amount (<1 at%) of Biaddition has the potential for improving the mechanical performances with only a slight decrease ofelectrical conductivity when the solid solutions are formed. However, so far no study reported anyimproved mechanical properties of the Ag-Bi alloy systems in comparison with those of pure Ag.

In addition to the alloy system design, the electrodeposition process typically generates columnaror lamellar coating structures, depending on the processing parameters, e.g., pulsed wave forms,current density, temperature, bath composition and constituents, additive type and concentration, andmetal ion concentration. Previous studies have shown that coatings prepared by different processingmethods have directionally-dependent elastic-plastic properties due to their microstructural-levelanisotropy [24–27]. Nevertheless, considering the thin nature of the coatings (with thickness from~1 to 100 µm), only small-scale testing and characterization methods, such as free-standing (FS) thinfilm tensile [28–30], micro-cantilever deflection [31,32], nanoindentation [31,33–37], and micro-pillarcompression [38–40], can be applied to characterize the coating mechanical properties. Some studieshave addressed the anisotropic behaviors of coatings [24–27]; however, very few studies quantifiedthe elastic-plastic anisotropy of thin coatings, i.e., thickness below 10 µm, due to the difficulty ofsmall-scale sample preparation. Meanwhile, an additional question should be clarified: how low of anat% element addition will influence the anisotropy of the alloy coatings, from the perspectives of bothmicrostructural features and mechanical properties? Thus, it is necessary to conduct microstructuraland mechanical characterizations in order to quantify the properties of coatings, as well as to illustratethe effects of new alloy element addition on the coating anisotropy.

In this study, we first characterized the microstructural features of the coatings using scanningelectron microscopy (SEM) and X-ray diffraction (XRD). Next, we investigated the microstructures ofthe Ag and Ag-Bi nanocrystalline using TEM and high-resolution transmission electron microscopy(HRTEM) to characterize the effects of Bi addition on the microstructural features at the nano-scale.Then, we applied the recently-reported nanoindentation-based technique [27] to interrogate theanisotropic elastic-plastic properties of the Ag-Bi coating. The technique was developed by theauthors to extract the anisotropic elastic-plastic properties of electrodeposited Ag coating [27] withnanoindentation in conjunction with three-dimensional (3-D) finite element analyses (FEA). Previousstudies have revealed that the inverse flow stress from the nanoindentation is consistent with thetensile flow curves measured by free-standing (FS) Ag film [29]. In the current study, we applied thenano-indentation-based technique [27] to the electrodeposited Ag-Bi coating in order to calculate theflow stress of Ag-Bi to compare with previously studied Ag coating, hence, the effects of Bi addition onthe mechanical properties of Ag coating were quantified. Finally, we presented a systematic discussionof the mechanical properties and electrical conductivity.

Page 3: 1,2 ID 3, ID 2

Materials 2017, 10, 932 3 of 16

2. Materials and Methods

2.1. Sample Preparation and Characterization

Nanocrystalline Ag and Ag-Bi coatings were electrodeposited onto a Ni-coated brass substrate(20 × 20 × 0.6 mm) as illustrated in Figure 1a. The solution composition and operating conditionsare listed in Table 1. The detailed description of the deposition process was reported by Tay [14].The surface morphology and composition of coatings were analyzed using a field emission scanningelectron microscope (FESEM) equipped with an energy dispersive spectroscope (EDS) in Figure 1b,d.The surface morphology of the Ag coating is ambiguous compared with that of the Ag-Bi coating dueto the low contrast and relatively plain surface. The at% of Bi within the Ag-Bi coating was measuredas ~1%. The thickness of Ag-Bi and Ni was ~10 µm and ~7 µm, respectively, as shown in Figure 1c,e.

Materials 2017, 10, 932 3 of 16

listed in Table 1. The detailed description of the deposition process was reported by Tay [14]. The surface morphology and composition of coatings were analyzed using a field emission scanning electron microscope (FESEM) equipped with an energy dispersive spectroscope (EDS) in Figure 1b,d. The surface morphology of the Ag coating is ambiguous compared with that of the Ag-Bi coating due to the low contrast and relatively plain surface. The at% of Bi within the Ag-Bi coating was measured as ~1%. The thickness of Ag-Bi and Ni was ~10 μm and ~7 μm, respectively, as shown in Figure 1c,e.

Figure 1. (a) The sketch of Ag/Ag-Bi coatings; (b) scanning electron microscope (SEM) image of Ag coating on as-deposited surfaces; (c) SEM image of Ag coating on the polished surfaces perpendicular to the T direction; (d) SEM image of Ag-Bi coating on the as-deposited surfaces; and (e) SEM image of Ag-Bi coating on the polished surfaces perpendicular to the T direction.

Table 1. Solution composition and operating parameters of electrodeposited Ag and Ag-Bi coatings.

Bath Composition and Plating Parameters QuantityAg Deposition

Silver metal 30 g/L KCN free 120 g/L

KOH 10 g/L Temperature 30 °C

Current density 10 mA/cm2 Agitation speed 200 rpm

Plating time 30 min Ag-Bi Deposition

Silver metal 30 g/L KCN free 120 g/L

KOH 10.35 g/L Bi(NO3)3·5H2O 0.2425 g/L Tartaric acid 0.075 g/L Temperature 30 °C

Current density 10 mA/cm2 Agitation speed 200 rpm

Plating time 30 min

The crystal structure was characterized by X-ray diffraction (XRD) with Cu Kα radiation (D2 Phaser Bruker, Karlsruhe, Germany, V = 30 KV, I = 10 mA). Diffraction patterns were recorded in the 2 theta range from 35 to 85° with a scanning step of 0.01°, as shown in Figure 2.

The electrical resistivity was measured by the four-point probe method after the Ag and Ag-Bi coatings were peeled off from the substrate and were layered on a silicon substrate. An electric current was passed through the two outer probes, and the potential was measured between two inner

Figure 1. (a) The sketch of Ag/Ag-Bi coatings; (b) scanning electron microscope (SEM) image of Agcoating on as-deposited surfaces; (c) SEM image of Ag coating on the polished surfaces perpendicularto the T direction; (d) SEM image of Ag-Bi coating on the as-deposited surfaces; and (e) SEM image ofAg-Bi coating on the polished surfaces perpendicular to the T direction.

Table 1. Solution composition and operating parameters of electrodeposited Ag and Ag-Bi coatings.

Bath Composition and Plating Parameters Quantity

Ag DepositionSilver metal 30 g/L

KCN free 120 g/LKOH 10 g/L

Temperature 30 CCurrent density 10 mA/cm2

Agitation speed 200 rpmPlating time 30 min

Ag-Bi DepositionSilver metal 30 g/L

KCN free 120 g/LKOH 10.35 g/L

Bi(NO3)3·5H2O 0.2425 g/LTartaric acid 0.075 g/LTemperature 30 C

Current density 10 mA/cm2

Agitation speed 200 rpmPlating time 30 min

The crystal structure was characterized by X-ray diffraction (XRD) with Cu Kα radiation(D2 Phaser Bruker, Karlsruhe, Germany, V = 30 kV, I = 10 mA). Diffraction patterns were recorded inthe 2 theta range from 35 to 85 with a scanning step of 0.01, as shown in Figure 2.

Page 4: 1,2 ID 3, ID 2

Materials 2017, 10, 932 4 of 16

The electrical resistivity was measured by the four-point probe method after the Ag and Ag-Bicoatings were peeled off from the substrate and were layered on a silicon substrate. An electric currentwas passed through the two outer probes, and the potential was measured between two inner probeswith a Keithley 2602 meter (Cleveland, OH, USA). The electrical resistivity was calculated using astandard method [41]. The peeled-off coatings were thinned using an ion beam milling system and theTEM images were taken by a high-resolution TEM (FEI TECNAI G2 F20, Hillsboro, OR, USA, 200 kV).

Materials 2017, 10, 932 4 of 16

probes with a Keithley 2602 meter (Cleveland, OH, USA). The electrical resistivity was calculated using a standard method [41]. The peeled-off coatings were thinned using an ion beam milling system and the TEM images were taken by a high-resolution TEM (FEI TECNAI G2 F20, Hillsboro, OR, USA, 200 kV).

Figure 2. (a) XRD pattern of Ag and Ag-Bi coatings; (b) the magnified peak of Ag (111); and (c) the magnified peak of Ag (200).

2.2. Nanoindentation

The nanoindentation tests were conducted using a Hysitron TI950 triboindenter (Minneapolis, MN, USA) along the L and T directions with a diamond Berkovich tip (Minneapolis, MN, USA) as shown in Figure 3a,b. The maximum indentation loading was 5 mN so that the maximum indentation depth was around 300 nm in both directions to avoid the substrate effect. The maximum load was held for 10 s to avoid creep behaviors in both indentation directions. Multiple indents were attempted, and at least three P-h curves in each direction were obtained. The surfaces perpendicular to the T directions were polished using a standard method so the influence of surface roughness could be ignored. The typical P-h curves are shown as the solid red curve in Figure 3c,d where the upper bound (the maximum P-h curve) and lower bound (the minimum P-h curve) of each direction are shown. The different indentation responses could be attributed to the microstructural orthotropy of the polycrystalline coating. In our study, the distance of each indent was kept at ~2 μm to reduce the possible residual stress from the surrounding indents [42]. Meanwhile, the indents near the interface were not considered in order to avoid the potential effects of Ni substrate on the indents in the T direction.

The surface perpendicular to the L direction, i.e., the as-deposited surface, could not be polished to minimize surface roughness due to the very thin nature of the coatings. To reduce the influence of surface roughness on the P-h curves, the indents were observed under scanning probe microscopy (SPM). As shown in Figure 3a, the color contour illustrates the height differences within the 30 × 30 μm observation region. Hence, the P-h curves in the L direction would be influenced by the surface roughness if indents were located on uneven surfaces [43]. If significant differences were observed on the loading curvature in the L direction, the P-h curves would be rejected. Additionally, a series of low depth (~120 nm) indents in the L direction were conducted before deep depth indentation (~300 nm). The P-h curves were obtained with relatively small scatter as the dotted green curves in Figure 3c,d, and these curves were utilized to guide the selection of the back curve (~300 nm) in the L direction. It should be noted that the challenges of surface roughness cannot be

Figure 2. (a) XRD pattern of Ag and Ag-Bi coatings; (b) the magnified peak of Ag (111); and (c) themagnified peak of Ag (200).

2.2. Nanoindentation

The nanoindentation tests were conducted using a Hysitron TI950 triboindenter (Minneapolis,MN, USA) along the L and T directions with a diamond Berkovich tip (Minneapolis, MN, USA) asshown in Figure 3a,b. The maximum indentation loading was 5 mN so that the maximum indentationdepth was around 300 nm in both directions to avoid the substrate effect. The maximum load washeld for 10 s to avoid creep behaviors in both indentation directions. Multiple indents were attempted,and at least three P-h curves in each direction were obtained. The surfaces perpendicular to the Tdirections were polished using a standard method so the influence of surface roughness could beignored. The typical P-h curves are shown as the solid red curve in Figure 3c,d where the upperbound (the maximum P-h curve) and lower bound (the minimum P-h curve) of each direction areshown. The different indentation responses could be attributed to the microstructural orthotropy ofthe polycrystalline coating. In our study, the distance of each indent was kept at ~2 µm to reducethe possible residual stress from the surrounding indents [42]. Meanwhile, the indents near theinterface were not considered in order to avoid the potential effects of Ni substrate on the indents inthe T direction.

The surface perpendicular to the L direction, i.e., the as-deposited surface, could not be polishedto minimize surface roughness due to the very thin nature of the coatings. To reduce the influence ofsurface roughness on the P-h curves, the indents were observed under scanning probe microscopy(SPM). As shown in Figure 3a, the color contour illustrates the height differences within the 30 × 30 µmobservation region. Hence, the P-h curves in the L direction would be influenced by the surfaceroughness if indents were located on uneven surfaces [43]. If significant differences were observedon the loading curvature in the L direction, the P-h curves would be rejected. Additionally, a seriesof low depth (~120 nm) indents in the L direction were conducted before deep depth indentation(~300 nm). The P-h curves were obtained with relatively small scatter as the dotted green curves in

Page 5: 1,2 ID 3, ID 2

Materials 2017, 10, 932 5 of 16

Figure 3c,d, and these curves were utilized to guide the selection of the back curve (~300 nm) in the Ldirection. It should be noted that the challenges of surface roughness cannot be completely overcomeat such small indentation depths. However, we have taken special care to choose the P-h curves withminimal surface roughness influence for the inverse calculations of Ag and Ag-Bi coatings, as the redand orange curves shown in Figure 3c,d, respectively. The averaged P-h curves from all the indentsalong a single direction were used for the subsequent inverse analyses.

Materials 2017, 10, 932 5 of 16

completely overcome at such small indentation depths. However, we have taken special care to choose the P-h curves with minimal surface roughness influence for the inverse calculations of Ag and Ag-Bi coatings, as the red and orange curves shown in Figure 3c,d, respectively. The averaged P-h curves from all the indents along a single direction were used for the subsequent inverse analyses.

Figure 3. (a) Indents (highlighted in circles) in the L direction under SPM with contour color bars indicating the surface height (the indents highlighted in green circles denoted an indentation ~100 nm and the indent highlighted in the blue dotted circle denoted an indentation with P_max = 5 mN); (b) indents in the T direction under SPM (the indents highlighted in the white box are used for inverse calculation); (c) The upper bound (solid) and lower bound (dotted) of the P-h curves obtained in the two directions for the Ag coating and the small depth (~150 nm) P-h curves in the L direction; and (d) the upper bound (solid) and lower bound (dotted) of the P-h curves obtained in the two directions for Ag-Bi coating and the small depth (~120 nm) P-h curves in the L direction.

3. Results

3.1. Microstructure of the Ag and Ag-Bi Coatings

The peak intensities of Ag and Ag-Bi coatings are shown in Figure 2; no new phases were generated. However, the Bi element addition would generate the smaller crystallite size [11] and introduce intrinsic lattice distortions to the Ag matrix [44]. The lower-intensity peaks for the Ag-Bi coating were observed. There is a very small shift of peaks to the lower diffraction angle observed in Figure 2b. Then, the microstructures of the Ag and Ag-Bi electrodeposited coatings were characterized using TEM and HRTEM. Equiaxed fine-grains were observed in the plane view TEM images, as shown in Figure 4a. The grain sizes are smaller than 200 nm including nano-sized twins of different thickness. Equiaxed fine-grains of smaller sizes are shown in Figure 4b for the Ag-Bi coating. However, no nano-sized twins were observed in the Ag-Bi coating.

Further detailed microstructures of these two coatings are shown in Figures 5 and 6. The nano-sized twins were observed within the Ag coating as shown in Figure 5a,b. The orientations of these nano-sized twins are random. The selected area diffraction (SAD) in the Figure 5a pattern from the red dotted region demonstrates a twinning pattern, and the width of these nano-sized twins could be as small as 20 nm. In addition, few dislocations were observed with the Ag grains and are

Figure 3. (a) Indents (highlighted in circles) in the L direction under SPM with contour color barsindicating the surface height (the indents highlighted in green circles denoted an indentation ~100 nmand the indent highlighted in the blue dotted circle denoted an indentation with P_max = 5 mN);(b) indents in the T direction under SPM (the indents highlighted in the white box are used for inversecalculation); (c) The upper bound (solid) and lower bound (dotted) of the P-h curves obtained in thetwo directions for the Ag coating and the small depth (~150 nm) P-h curves in the L direction; and(d) the upper bound (solid) and lower bound (dotted) of the P-h curves obtained in the two directionsfor Ag-Bi coating and the small depth (~120 nm) P-h curves in the L direction.

3. Results

3.1. Microstructure of the Ag and Ag-Bi Coatings

The peak intensities of Ag and Ag-Bi coatings are shown in Figure 2; no new phases weregenerated. However, the Bi element addition would generate the smaller crystallite size [11] andintroduce intrinsic lattice distortions to the Ag matrix [44]. The lower-intensity peaks for the Ag-Bicoating were observed. There is a very small shift of peaks to the lower diffraction angle observed inFigure 2b. Then, the microstructures of the Ag and Ag-Bi electrodeposited coatings were characterizedusing TEM and HRTEM. Equiaxed fine-grains were observed in the plane view TEM images, as shownin Figure 4a. The grain sizes are smaller than 200 nm including nano-sized twins of different thickness.Equiaxed fine-grains of smaller sizes are shown in Figure 4b for the Ag-Bi coating. However, nonano-sized twins were observed in the Ag-Bi coating.

Further detailed microstructures of these two coatings are shown in Figures 5 and 6.The nano-sized twins were observed within the Ag coating as shown in Figure 5a,b. The orientations

Page 6: 1,2 ID 3, ID 2

Materials 2017, 10, 932 6 of 16

of these nano-sized twins are random. The selected area diffraction (SAD) in the Figure 5a patternfrom the red dotted region demonstrates a twinning pattern, and the width of these nano-sized twinscould be as small as 20 nm. In addition, few dislocations were observed with the Ag grains and arehighlighted in Figure 5a. A HRTEM image, as shown in Figure 5c, was taken to calculate the latticeparameters in and out of the twinning grains as two highlighted regions in the orange box and blue box,respectively. The purposes of calculating the lattice parameters were to analyze the possible reasonsfor generating nano-sized twins in Ag [45] and to verify the existence of Bi in the Ag-Bi coating [11].The distance between atoms was 2.041 Å from the (200) face and the lattice parameter in the blue boxwas calculated as 4.082 Å, which is quite similar to the 4.09 Å for pure silver [15]. In comparison, thedistance between atoms was 2.325 Å from the (111) face in the orange box and the lattice parameterwas calculated as 4.027 Å, which is smaller than the 4.082 Å obtained outside the nano-sized twins.The smaller lattice parameters in the nano-sized twins confirmed the existence of the shear stress in theelectrodeposition process, and the shear stress value could be as high as 65 MPa according to densityfunctional theory (DFT) calculation for generating the nanocrystalline twinning [46].

Materials 2017, 10, 932 6 of 16

highlighted in Figure 5a. A HRTEM image, as shown in Figure 5c, was taken to calculate the lattice parameters in and out of the twinning grains as two highlighted regions in the orange box and blue box, respectively. The purposes of calculating the lattice parameters were to analyze the possible reasons for generating nano-sized twins in Ag [45] and to verify the existence of Bi in the Ag-Bi coating [11]. The distance between atoms was 2.041 Å from the (200) face and the lattice parameter in the blue box was calculated as 4.082 Å, which is quite similar to the 4.09 Å for pure silver [15]. In comparison, the distance between atoms was 2.325 Å from the (111) face in the orange box and the lattice parameter was calculated as 4.027 Å, which is smaller than the 4.082 Å obtained outside the nano-sized twins. The smaller lattice parameters in the nano-sized twins confirmed the existence of the shear stress in the electrodeposition process, and the shear stress value could be as high as 65 MPa according to density functional theory (DFT) calculation for generating the nanocrystalline twinning [46].

Figure 4. TEM images of the nanocrystalline: (a) Ag coating and (b) Ag-Bi coating from the plane view with individual grains highlighted in the white dotted box.

Figure 5. (a) The bright field TEM image from the plane view for Ag with selected area diffractions (SAD) patterns to illustrate the FCC structures ([110]) within the white area and nanocrystalline twining within the red area; (b) the TEM bright field image of nanocrystalline Ag with dislocations; and (c) the HRTEM images with areas in and out of the nanocrystalline twining.

No nano-sized twins were observed within the Ag-Bi coating, as shown in the high-magnification TEM images in Figure 6a,b. The local SAD pattern images in Figure 6a from the highlighted dark area show the face centered cubic (FCC) structure from the [001] direction, which demonstrated the Ag matrix. However, some random spots were observed, and these spots should

Figure 4. TEM images of the nanocrystalline: (a) Ag coating and (b) Ag-Bi coating from the plane viewwith individual grains highlighted in the white dotted box.

Materials 2017, 10, 932 6 of 16

highlighted in Figure 5a. A HRTEM image, as shown in Figure 5c, was taken to calculate the lattice parameters in and out of the twinning grains as two highlighted regions in the orange box and blue box, respectively. The purposes of calculating the lattice parameters were to analyze the possible reasons for generating nano-sized twins in Ag [45] and to verify the existence of Bi in the Ag-Bi coating [11]. The distance between atoms was 2.041 Å from the (200) face and the lattice parameter in the blue box was calculated as 4.082 Å, which is quite similar to the 4.09 Å for pure silver [15]. In comparison, the distance between atoms was 2.325 Å from the (111) face in the orange box and the lattice parameter was calculated as 4.027 Å, which is smaller than the 4.082 Å obtained outside the nano-sized twins. The smaller lattice parameters in the nano-sized twins confirmed the existence of the shear stress in the electrodeposition process, and the shear stress value could be as high as 65 MPa according to density functional theory (DFT) calculation for generating the nanocrystalline twinning [46].

Figure 4. TEM images of the nanocrystalline: (a) Ag coating and (b) Ag-Bi coating from the plane view with individual grains highlighted in the white dotted box.

Figure 5. (a) The bright field TEM image from the plane view for Ag with selected area diffractions (SAD) patterns to illustrate the FCC structures ([110]) within the white area and nanocrystalline twining within the red area; (b) the TEM bright field image of nanocrystalline Ag with dislocations; and (c) the HRTEM images with areas in and out of the nanocrystalline twining.

No nano-sized twins were observed within the Ag-Bi coating, as shown in the high-magnification TEM images in Figure 6a,b. The local SAD pattern images in Figure 6a from the highlighted dark area show the face centered cubic (FCC) structure from the [001] direction, which demonstrated the Ag matrix. However, some random spots were observed, and these spots should

Figure 5. (a) The bright field TEM image from the plane view for Ag with selected area diffractions(SAD) patterns to illustrate the FCC structures ([110]) within the white area and nanocrystalline twiningwithin the red area; (b) the TEM bright field image of nanocrystalline Ag with dislocations; and (c) theHRTEM images with areas in and out of the nanocrystalline twining.

Page 7: 1,2 ID 3, ID 2

Materials 2017, 10, 932 7 of 16

Materials 2017, 10, 932 7 of 16

be caused by the Bi atoms embedded in the Ag matrix. In the HRTEM image, as shown in Figure 6c, the dark spot without clear atomic structure indicates a possible nanoparticle formed in the electrodeposition process. The atom distance from the (111) face was averaged as 2.397 Å in the orange box, and the lattice parameter was calculated as 4.152 Å. The atom space from the (111) face was measured as 2.336 Å in the highlighted blue box, and the lattice parameter was calculated as 4.045 Å.

Figure 6. (a) The bright field TEM image from the plane view for Ag-Bi with selected area diffractions (SAD) patterns to illustrate a FCC structures ([001]) within highlighted area; (b) the TEM bright field image of Ag-Bi alloy with plenty of defects; and (c) the HRTEM images with areas in and out of the black Ag-Bi solid solutions.

3.2. Nanoindentation and Inverse Calculation with Three-Dimensional FEA

For the purpose of comparison, the thin Ag-Bi coating was assumed to be transversely isotropic, similar to the Ag coating [27], with the isotropic mechanical property in the plane parallel to the surface. Table 2 summarizes the constitutive relationship and parameters used to describe the anisotropic mechanical properties of Ag-Bi coating.

Table 2. Definitions of the terms and the summary of the constitutive equations used for the inverse calculation.

(a) Symbol Descriptions and Hardness in the L and T directions

and Nominal elastic modulus in the L and T directions

and Maximum indentation load in the L and T directions ℎ and ℎ Maximum indentation depth in the L and T directions ℎ∗ Indentation size effect (ISE) characteristic length

The ratio of yield strength

and Maximum indentation load with modification in the L and T directions

and Maximum indentation load from FEA in the L and T directions

and Shear modulus in the L and T directions

Reference strength

Figure 6. (a) The bright field TEM image from the plane view for Ag-Bi with selected area diffractions(SAD) patterns to illustrate a FCC structures ([001]) within highlighted area; (b) the TEM bright fieldimage of Ag-Bi alloy with plenty of defects; and (c) the HRTEM images with areas in and out of theblack Ag-Bi solid solutions.

No nano-sized twins were observed within the Ag-Bi coating, as shown in the high-magnificationTEM images in Figure 6a,b. The local SAD pattern images in Figure 6a from the highlighted dark areashow the face centered cubic (FCC) structure from the [001] direction, which demonstrated the Agmatrix. However, some random spots were observed, and these spots should be caused by the Bi atomsembedded in the Ag matrix. In the HRTEM image, as shown in Figure 6c, the dark spot without clearatomic structure indicates a possible nanoparticle formed in the electrodeposition process. The atomdistance from the (111) face was averaged as 2.397 Å in the orange box, and the lattice parameter wascalculated as 4.152 Å. The atom space from the (111) face was measured as 2.336 Å in the highlightedblue box, and the lattice parameter was calculated as 4.045 Å.

3.2. Nanoindentation and Inverse Calculation with Three-Dimensional FEA

For the purpose of comparison, the thin Ag-Bi coating was assumed to be transversely isotropic,similar to the Ag coating [27], with the isotropic mechanical property in the plane parallel to the surface.Table 2 summarizes the constitutive relationship and parameters used to describe the anisotropicmechanical properties of Ag-Bi coating.

Detailed descriptions of the constitutive relationship can be found in our previous study [27].To estimate the mechanical properties of Ag-Bi, i.e., the elastic modulus and flow stress from theload-depth curves of the T and L directions, a three-step method proposed recently was utilized, asshown in Figure 7a. The elastic modulus values of two directions were assumed as the value obtainedfrom the Oliver-Pharr method [47]. The reference strength was determined with empirical equations.Thus, n and γ were determined using 3-D finite element indentation modeling as sketched in Figure 7b.

(1) With the Oliver-Pharr (OP) method [47], the nominal elastic modulus was calculated asET = 94.29± 3.82 GPa and EL = 87.83 ± 7.52 GPa for the two directions, respectively.The hardness values in the two directions were calculated as HL = 2.66 ± 0.04 GPa andHT = 2.71± 0.07 GPa.

(2) Since the difference between ET and EL is ~5%, which is quite similar to the previous observationof the Ag coating, ET and EL were replaced with ET and EL respectively. Meanwhile, withthe previous yield strength-hardness relationship at 5 mN [48], σTo and σLo were calculated as480 and 410 MPa, respectively. Regarding that the actual yield strength in the two directions

Page 8: 1,2 ID 3, ID 2

Materials 2017, 10, 932 8 of 16

(σTo + σLo) could be approximated by the sum of the nominal yield strength (σTo + σLo) [25,27], σo

was calculated as 445 MPa.(3) Finally, a 3-D finite element indentation model was developed within the commercial finite

element package, ABAQUS, to simulate the indentation process of the Ag-Bi coating in orderto inversely determine n and γ. A parametric FEA-based indentation study was carried out tocalculate the maximum indentation loads in the two directions to compare with the experimentalmeasurements as target values. As mentioned earlier, the indentation depth was ~300 nm, andthe strong indentation size effect (ISE) could be expected considering the thin nature of the Ag-Bicoating. Thus, we first removed the ISE based on the characteristic length (h∗ = 450 nm [49,50])for the electrodeposited Ag-Bi coating. It should be noted that the target load values for thesimulation are PS

Tmax or PSLmax rather than 5 mN [51]. With one set of n and γ values, the Berkovich

indenter penetrates the Ag-Bi coating to the same depth as hLmax, and the corresponding PFEALmax in

the L direction was calculated. Next, the input materials properties were rotated by 90 aroundthe T direction, the indenter penetrated the material to hTmax, and the corresponding PFEA

Tmax wascalculated in the T direction. Thus, different sets of PFEA

Lmax vs. PFEATmax values can be obtained with

corresponding n and γ values. The most plausible combination of γ and n can be determined bychoosing the corresponding PFEA

Lmax and PFEATmax set closest to the target experimentally measured

the data point.

Table 2. Definitions of the terms and the summary of the constitutive equations used for theinverse calculation.

(a) Symbol Descriptions

HL and HT Hardness in the L and T directions

EL and ET Nominal elastic modulus in the L and T directions

PLmax and PTmaxMaximum indentation load in the L and Tdirections

hLmax and hTmaxMaximum indentation depth in the L and Tdirections

h∗ Indentation size effect (ISE) characteristic length

γ The ratio of yield strength

PSLmax and PS

TmaxMaximum indentation load with modification inthe L and T directions

PFEALmax and PFEA

TmaxMaximum indentation load from FEA in the Land T directions

GL and GT Shear modulus in the L and T directions

σo Reference strength

εL and εT Isotropic plastic strain in the L and T directions

n Strain hardening exponent

σLo and σTo Nominal yield strength in the L and T directions

n′Lo and n′ToNominal strain hardening exponent in the L andT directions

(b) Constitutive equations

Elastic Plastic

εxxεyyεzzγyzγzxγxy

=

1/ET − vLTEL

− vTET

− vTLEL

1/EL − vTLET

− vTET

− vLTET

1/ET1/GL

1/GL1/GT

σxxσyyσzzσyzσzxσxy

ELET

= νLTνTL

GT = ET2(1+νT)

GL = ET+EL

4∗(

1+(

νLT+νTL2

))νT = 0.3 νLT + νTL = 2 νT

σL = σLo

(1 + ( ELεL

σLo))n

for σL > σLo

σT = σTo

(1 + ( ET εT

σTo))n

for σT > σToγ = σLo

σTo

Assumptions for inverse calculation

ET = ET and EL = ELσLo + σTo = 2σo = σ′Lo + σ′To

Page 9: 1,2 ID 3, ID 2

Materials 2017, 10, 932 9 of 16Materials 2017, 10, 932 9 of 16

Figure 7. (a) The algorithm of inverse calculation and (b) the sketch of the FEA model with a Berkovich indenter.

The estimated average properties of the Ag-Bi coating are: = 87.83GPa , = 92.46GPa, = 445MPa, = 1.03, and = 0.055. The obtained flow stress of the two coatings in the two directions are listed and plotted in Figure 8.

3.3. Electrical Resistivity of Ag and Ag-Bi Coatings

The standard electrical resistivity and electrical conductivity of Ag coating are 1.78 ± 0.02 × 10 Ω ∙ m and 97.1 ± 1.2 (%IACS, International Annealed Copper Standard), respectively. As expected, the addition of Bi to the Ag caused the increase of electrical resistivity, and the electrical resistivity of Ag-Bi coating was measured as 1.88 ± 0.02 × 10 Ω ∙ m, which indicates a 6% increase of the electrical resistivity.

4. Discussion

4.1. Microstructure of the Ag and Ag-Bi Coatings

The main microstructural features to describe a coating include (1) grain size and grain morphology, (2) morphology of grain boundaries and intergranular defects/phases, (3) density of the intergranular defects, and (4) composition distribution across grains and grain boundaries. These features determine the entire elastic-plastic properties of the coatings [52], and the detailed observations are discussed below in conjunction with the inverse results of Ag and Ag-Bi coatings.

4.1.1. Significant Improvement of Flow Stress

A significant improvement on the flow stress was observed for the Ag-Bi coating in both directions: the yield strength of the Ag-Bi coating was improved by 300 MPa (~150%) and 200 MPa (~100%) compared to those of the Ag coating in the L direction and T direction, respectively. This improvement can be explained by the nanoscale composite microstructures with a large amount of stack faults, i.e., dislocations generated during the electrodeposited process. The TEM images in Figure 5a,b show clean Ag grains with few dislocations for the Ag coating. On the other hand, plenty of internal defects, i.e., dislocations, were observed in the Ag-Bi coating microstructures, as shown in Figure 6a,b. The two images in Figure 6 are quite similar to the TEM images of the mechanically-milled Ag-5.1% Bi alloys as reported by Chithra et al. [11]. The existence of Bi in the Ag matrix cannot be easily distinguished from XRD patterns as compared to other Ag-based alloy systems, i.e., Ag-Al [9], since no new phase was generated. With EDS, the average at% of Bi across the coating thickness would be ~1%.

Figure 7. (a) The algorithm of inverse calculation and (b) the sketch of the FEA model with aBerkovich indenter.

The estimated average properties of the Ag-Bi coating are: EL = 87.83 GPa, ET = 92.46 GPa,σo = 445 MPa, γ = 1.03, and n = 0.055. The obtained flow stress of the two coatings in the twodirections are listed and plotted in Figure 8.

3.3. Electrical Resistivity of Ag and Ag-Bi Coatings

The standard electrical resistivity and electrical conductivity of Ag coating are1.78± 0.02× 10−8 Ω·m and 97.1 ± 1.2 (%IACS, International Annealed Copper Standard),respectively. As expected, the addition of Bi to the Ag caused the increase of electrical resistivity, andthe electrical resistivity of Ag-Bi coating was measured as 1.88± 0.02 × 10−8Ω·m, which indicates a6% increase of the electrical resistivity.

4. Discussion

4.1. Microstructure of the Ag and Ag-Bi Coatings

The main microstructural features to describe a coating include (1) grain size and grainmorphology, (2) morphology of grain boundaries and intergranular defects/phases, (3) densityof the intergranular defects, and (4) composition distribution across grains and grain boundaries.These features determine the entire elastic-plastic properties of the coatings [52], and the detailedobservations are discussed below in conjunction with the inverse results of Ag and Ag-Bi coatings.

4.1.1. Significant Improvement of Flow Stress

A significant improvement on the flow stress was observed for the Ag-Bi coating in both directions:the yield strength of the Ag-Bi coating was improved by 300 MPa (~150%) and 200 MPa (~100%)compared to those of the Ag coating in the L direction and T direction, respectively. This improvementcan be explained by the nanoscale composite microstructures with a large amount of stack faults, i.e.,dislocations generated during the electrodeposited process. The TEM images in Figure 5a,b show cleanAg grains with few dislocations for the Ag coating. On the other hand, plenty of internal defects, i.e.,dislocations, were observed in the Ag-Bi coating microstructures, as shown in Figure 6a,b. The twoimages in Figure 6 are quite similar to the TEM images of the mechanically-milled Ag-5.1% Bi alloysas reported by Chithra et al. [11]. The existence of Bi in the Ag matrix cannot be easily distinguishedfrom XRD patterns as compared to other Ag-based alloy systems, i.e., Ag-Al [9], since no new phasewas generated. With EDS, the average at% of Bi across the coating thickness would be ~1%.

Page 10: 1,2 ID 3, ID 2

Materials 2017, 10, 932 10 of 16

In addition, the local HRTEM images illustrate different lattice parameters. The black region in thelower left corner of the orange box can be discerned as the Ag-Bi nanoparticle, which is more difficultto mill in the sample preparation for TEM characterization. Near the nanoparticle in the highlightedorange box as shown in Figure 6c, the lattice parameter was calculated as 4.152 Å, which is muchhigher than the 4.082 Å [15]. This result indicates that the addition of Bi atoms in the electrodepositedprocess can expand the Ag lattice, which has been observed in the Ag-5.1% Bi alloys prepared by usinghigh-energy mechanical alloying [11]. Since the possible lattice shrink would also occur, as shownin highlighted blue box of Figure 6c, the introduction of Bi would lead to the variations of the Aglattice in the current electrodeposited Ag-Bi coating. Meanwhile, within certain regions of the Ag-Bicoating, the lattice parameter was calculated as 4.082 Å, which indicates that there are no solid-solutionAg-Bi nanoparticles in the adjacent regions. Hence, a composite of Ag and Ag-Bi nanostructures wasproduced with the current electrodeposition process. Those dark spots can act as barriers or obstaclesfor dislocations to overcome under plastic deformation, and the associated strengthening mechanismresulted from dislocation-particle interaction has been well studied using discrete dislocation dynamics(DD) simulations for nanoscale metallic (NMM) composites [53,54]. Hence, the internal defects andthe stiff nanoparticles serves as the main reasons for the improvement of flow stress in Ag-Bi coating.

Additionally, the grain highlighted in Figure 4a for the Ag coating is slightly larger (~181 nm)than the highlighted grain of Ag-Bi coating (~131 nm). This difference indicates that the addition ofBi would also lead to a smaller grain size. The smaller crystalline size would lead to the higher yieldstrength according to the Hall-Petch relationship [55–57]. Meanwhile, it is worth pointing out that thereis no intermetallic compound in the current electrodeposited process and we did not observe brightspots from the high-angle annular dark-field (HAAFD) TEM tomography as previously prepared Ni-Bicoatings [14].

4.1.2. Significant Decrease of the Hardening Exponent (n)

The hardening exponents of two coatings are also determined by the internal microstructure-levelfeatures. As the results listed in Figure 8, the hardening exponent of Ag would be as high as0.16 regarding the larger grains and clean internal structure, compared to those of the Ag-Bi coating.The deformation mechanism of nanocrystalline FCC metals has been well studied and discussedin the previous literature [18,30,45,58]. The twin-containing microstructure in the Ag coating hassufficient space for the storage of dislocations regarding the much fewer nano-sized twins withinthe current Ag coating in contrast to high-density nano-sized twins a Cu coating prepared byYou et al. [30]. The dislocation intersection actions and twin boundaries resulted in the formationof locks in which dislocation trapping and absorption along twin boundaries occurred [18,30,58].Meanwhile, the rearrangement and annihilation of dislocations led to a low dynamic recovery ratein comparison with that of ordinary grain boundaries [30]. Thus, the high hardening exponent wasachieved. The addition of Bi reduced the hardening exponent from 0.16 to 0.055 according to theinverse calculation. Compared to the Ag coating, the Ag-Bi coating has already carried plenty ofdefects. Therefore, limited spaces for newly-generated dislocations led to a much lower hardeningexponent. The high hardening rate, the high uniform elongation, and the high total elongation havebeen obtained for electrodeposited nanocrystalline materials with thicknesses around 500 µm [30,58].However, similar behaviors might not be obtained for the current coating with thicknesses less than10 µm in the FS tensile tests [29] considering inhomogeneous properties [59,60].

Page 11: 1,2 ID 3, ID 2

Materials 2017, 10, 932 11 of 16Materials 2017, 10, 932 11 of 16

Figure 8. The inverse calculated flow stress for Ag and Ag-Bi in L and T directions (Ag_T, Ag_L, Ag-Bi_T, and Ag-Bi_L denote the flow stress of Ag coating in the T direction, Ag coating in the L direction, Ag-Bi coating in the T direction, and Ag-Bi coating in the L direction, respectively).

4.1.3. Slight Improvement of Elastic Modulus

Compared to the Ag coating, an improvement of the elastic modulus was observed for the Ag-Bi coating by 15.3 GPa (~21%) and 23.7 GPa (~34%) in the L and T directions, respectively. Additionally, the obtained values are higher than the calculated value of 83.8 GPa for the <111> orientation of pure Ag in the previous study [29]. Two factors can lead to the improvement of the elastic modulus: the newly-formed Ag-Bi solid solution and the smaller grain size. Generally speaking, the solid solution can increase the elastic modulus, e.g., more carbon within iron matrix will generate martensite with higher elastic modulus than pure iron [61]. The smaller grain size within Ag-Bi can increase the lattice parameter as shown in Figure 6c, leading to a higher elastic modulus [11]. The inversely calculated elastic modulus of Ag-Bi are consistently higher than those of the Ag coating in two directions. Further studies using DFT will be helpful to support our current results.

4.1.4. Significant Reduced Plastic Anisotropy (γ)

Generally speaking, the electrodeposition process will generate a columnar structure, as previously reported, for FCC metals. Our previous study on Ag coatings confirmed this structure using a nanoindentation-based approach and determined = 0.6 for Ag coatings with microstructure information from both the plane view and the cross-section view images [30]. The value of the currently prepared Ag-Bi coating was calculated as 1.03, which indicates a more homogenous microstructure as compared to the Ag coating. A schematic sketch is shown in Figure 9 to illustrate the effect of a small amount of Bi addition on the structure of the electrodeposited Ag coating: without Bi addition, a columnar structure with larger grains, including nano-sized twins, was formed; with Bi addition, the coating was turned to a finer homogeneous microstructure with Ag-Bi nanoparticles in the Ag matrix.

Since the current Ag and Ag-Bi coatings were prepared with grain size below 1 μm, it would be difficult to directly observe the microstructure by SEM. Hence, further studies should be conducted using TEM and HRTEM to obtain the local microstructural details of the Ag and Ag-Bi coatings in the cross-section view to illustrate the effect of small Bi additions to the grain morphologies.

Figure 8. The inverse calculated flow stress for Ag and Ag-Bi in L and T directions (Ag_T, Ag_L,Ag-Bi_T, and Ag-Bi_L denote the flow stress of Ag coating in the T direction, Ag coating in the Ldirection, Ag-Bi coating in the T direction, and Ag-Bi coating in the L direction, respectively).

4.1.3. Slight Improvement of Elastic Modulus

Compared to the Ag coating, an improvement of the elastic modulus was observed for the Ag-Bicoating by 15.3 GPa (~21%) and 23.7 GPa (~34%) in the L and T directions, respectively. Additionally,the obtained values are higher than the calculated value of 83.8 GPa for the <111> orientation of pureAg in the previous study [29]. Two factors can lead to the improvement of the elastic modulus: thenewly-formed Ag-Bi solid solution and the smaller grain size. Generally speaking, the solid solutioncan increase the elastic modulus, e.g., more carbon within iron matrix will generate martensite withhigher elastic modulus than pure iron [61]. The smaller grain size within Ag-Bi can increase the latticeparameter as shown in Figure 6c, leading to a higher elastic modulus [11]. The inversely calculatedelastic modulus of Ag-Bi are consistently higher than those of the Ag coating in two directions.Further studies using DFT will be helpful to support our current results.

4.1.4. Significant Reduced Plastic Anisotropy (γ)

Generally speaking, the electrodeposition process will generate a columnar structure, aspreviously reported, for FCC metals. Our previous study on Ag coatings confirmed this structureusing a nanoindentation-based approach and determined γ = 0.6 for Ag coatings with microstructureinformation from both the plane view and the cross-section view images [30]. The γ value ofthe currently prepared Ag-Bi coating was calculated as 1.03, which indicates a more homogenousmicrostructure as compared to the Ag coating. A schematic sketch is shown in Figure 9 to illustrate theeffect of a small amount of Bi addition on the structure of the electrodeposited Ag coating: without Biaddition, a columnar structure with larger grains, including nano-sized twins, was formed; with Biaddition, the coating was turned to a finer homogeneous microstructure with Ag-Bi nanoparticles inthe Ag matrix.

Since the current Ag and Ag-Bi coatings were prepared with grain size below 1 µm, it would bedifficult to directly observe the microstructure by SEM. Hence, further studies should be conductedusing TEM and HRTEM to obtain the local microstructural details of the Ag and Ag-Bi coatings in thecross-section view to illustrate the effect of small Bi additions to the grain morphologies.

Page 12: 1,2 ID 3, ID 2

Materials 2017, 10, 932 12 of 16Materials 2017, 10, 932 12 of 16

Figure 9. The sketch of Bi addition on the microstructure of the nanocrystalline Ag coating in the the electrodeposition process.

So far, the effects of Bi addition on the mechanical properties of Ag on the flow stress, hardening exponent, elastic modulus, and the plastic anisotropy are quantitatively obtained. The mechanism behind these effects is explained with the micro- and nanostructures obtained under TEM and HRTEM. Multiple factors, i.e., coating thickness, hardness, and the surface roughness, will determine the final wear performances. Hence, the tribology/wear resistance of coatings is generally characterized by the friction coefficient, volume/weight loss, and track morphology. Among them, volume loss is a characteristic value affected by the mechanical responses from both L and T directions. According to Archard’s law, volume loss during sliding wear is inversely proportional to the hardness of the coating regardless of different coating compositions, electrodeposited processing parameters, micro/nanostructures, as well as the different contact mechanisms between an abrasive ball and the alloys [62]. The relative reduction of volume loss for the Ag-Bi coating compared to that of Ag coating can then be deduced based on the hardness improvement: with the addition of Bi, a 60% hardness increase is observed from Ag-Bi to Ag coating (~2.67 GPa for Ag-Bi versus ~1.65 GPa for Ag), so a 40% lower volume loss can be expected. Hence, a small amount of Bi addition can greatly improve the strength and the wear resistance of Ag coatings.

4.2. Effects of Bi Addition on the Electrical Resistivity

As expected, the addition of Bi to Ag matrix increased the electrical resistivity. The electrical resistivity of Ag and Ag-Bi coatings were measured as 1.78 ± 0.02 × 10 Ω ∙ m and 1.88 ± 0.02 ×10 Ω ∙ m, respectively. The ~6% increase in the electrical resistivity is quite small compared to the 60% increase in the hardness. The low electrical resistivity and the high hardness of Ag-Bi coating indicate that the current Ag-Bi nanocomposite is a good candidate for electrical contacting materials. In addition to Ag-Bi alloy, four other Ag alloys were prepared before, and the electrical resistivity of these alloys were measured [1,9,10,13]. The relationship between normalized electrical resistivity at room temperature (the electrical resistivity of Ag alloys, Ω ∙ m(Ag − ), over the electrical resistivity of pure Ag, Ω ∙ m(Ag) and alloy at% is shown in Figure 10.

Figure 9. The sketch of Bi addition on the microstructure of the nanocrystalline Ag coating in the theelectrodeposition process.

So far, the effects of Bi addition on the mechanical properties of Ag on the flow stress, hardeningexponent, elastic modulus, and the plastic anisotropy are quantitatively obtained. The mechanismbehind these effects is explained with the micro- and nanostructures obtained under TEM and HRTEM.Multiple factors, i.e., coating thickness, hardness, and the surface roughness, will determine the finalwear performances. Hence, the tribology/wear resistance of coatings is generally characterized bythe friction coefficient, volume/weight loss, and track morphology. Among them, volume loss is acharacteristic value affected by the mechanical responses from both L and T directions. Accordingto Archard’s law, volume loss during sliding wear is inversely proportional to the hardness ofthe coating regardless of different coating compositions, electrodeposited processing parameters,micro/nanostructures, as well as the different contact mechanisms between an abrasive ball and thealloys [62]. The relative reduction of volume loss for the Ag-Bi coating compared to that of Ag coatingcan then be deduced based on the hardness improvement: with the addition of Bi, a 60% hardnessincrease is observed from Ag-Bi to Ag coating (~2.67 GPa for Ag-Bi versus ~1.65 GPa for Ag), so a40% lower volume loss can be expected. Hence, a small amount of Bi addition can greatly improve thestrength and the wear resistance of Ag coatings.

4.2. Effects of Bi Addition on the Electrical Resistivity

As expected, the addition of Bi to Ag matrix increased the electrical resistivity. The electricalresistivity of Ag and Ag-Bi coatings were measured as 1.78 ± 0.02 × 10−8 Ω·m and1.88± 0.02 × 10−8 Ω·m, respectively. The ~6% increase in the electrical resistivity is quite smallcompared to the 60% increase in the hardness. The low electrical resistivity and the high hardnessof Ag-Bi coating indicate that the current Ag-Bi nanocomposite is a good candidate for electricalcontacting materials. In addition to Ag-Bi alloy, four other Ag alloys were prepared before, and theelectrical resistivity of these alloys were measured [1,9,10,13]. The relationship between normalizedelectrical resistivity at room temperature (the electrical resistivity of Ag alloys, Ω·m (Ag− X), overthe electrical resistivity of pure Ag, Ω·m (Ag) and alloy at% is shown in Figure 10.

In spite of different preparation processing, grain/crystalline size, thermal-mechanical treatment,and geometry shape/size (i.e., bulk, coating, or thin films), a general relationship was found that thehigher at% of alloying elements, the higher the electrical resistivity. This relationship is very consistentin the binary alloy system of Ag when the at% of alloying elements is no more than 20%. Meanwhile,an upper bound from Ag-Al and Ag-Ti systems and a lower bound from Ag-Pd and Ag-Au systemsare denoted as the green and red dashed curves in Figure 10, respectively. The different phenomenacould be attributed to the different atom radii between Ag and the alloy elements. The differencesbetween atom radii in the four alloys are 25.0%, 16.9%, 3.5%, and 5.2% for Ag-Ti (172 pm vs. 215 pm),Ag-Al (172 pm vs. 143 pm), Ag-Au (172 pm vs. 166 pm), and Ag-Pd (172 pm vs. 163 pm), respectively.

Page 13: 1,2 ID 3, ID 2

Materials 2017, 10, 932 13 of 16

The total electrical resistivity of one material at room temperature can be attributed to differentfactors including structural defects (dislocations, vacancies, alloying elements, and impurities) andgeometrical scattering (internal and external interfaces) [21]. In the current Ag binary alloys, thealloying elements (atom radii) and the at% of alloying elements within the Ag matrix are two dominantfactors determining the electrical resistivity. Although the difference between the atom radii betweenBi (230 pm) and Ag (172 pm) is significant, the amount of Bi within the Ag is small (~1%). The increaseof electrical resistivity of Ag-Bi is rather small, and the black square in Figure 10 is quite close to thelower bound.Materials 2017, 10, 932 13 of 16

Figure 10. The relationships of alloy at% in Ag matrix versus normalized (Norm.) electrical resistivity ( ∙ ( )∙ ( ) ) of Ag-Bi, Ag-Al [9], Ag-Pd [10], Ag-Ti [12], and Ag-Au [13] alloys.

In spite of different preparation processing, grain/crystalline size, thermal-mechanical treatment, and geometry shape/size (i.e., bulk, coating, or thin films), a general relationship was found that the higher at% of alloying elements, the higher the electrical resistivity. This relationship is very consistent in the binary alloy system of Ag when the at% of alloying elements is no more than 20%. Meanwhile, an upper bound from Ag-Al and Ag-Ti systems and a lower bound from Ag-Pd and Ag-Au systems are denoted as the green and red dashed curves in Figure 10, respectively. The different phenomena could be attributed to the different atom radii between Ag and the alloy elements. The differences between atom radii in the four alloys are 25.0%, 16.9%, 3.5%, and 5.2% for Ag-Ti (172 pm vs. 215 pm), Ag-Al (172 pm vs. 143 pm), Ag-Au (172 pm vs. 166 pm), and Ag-Pd (172 pm vs. 163 pm), respectively. The total electrical resistivity of one material at room temperature can be attributed to different factors including structural defects (dislocations, vacancies, alloying elements, and impurities) and geometrical scattering (internal and external interfaces) [21]. In the current Ag binary alloys, the alloying elements (atom radii) and the at% of alloying elements within the Ag matrix are two dominant factors determining the electrical resistivity. Although the difference between the atom radii between Bi (230 pm) and Ag (172 pm) is significant, the amount of Bi within the Ag is small (~1%). The increase of electrical resistivity of Ag-Bi is rather small, and the black square in Figure 10 is quite close to the lower bound.

5. Conclusions

In this study, the effects of Bi addition on the microstructures, mechanical performances, and electrical resistivity of an electrodeposited Ag coating were investigated. Using nanoindentation in two directions and the inverse calculation, we found that the Ag-Bi coating was much stronger than Ag with a significantly higher yield strength and flow stress. Meanwhile, the anisotropy of the Ag-Bi was significantly reduced as compared to that of the Ag coating, which indicates a more homogenous microstructure. The TEM images demonstrate that the improvement in the mechanical response was originated from the new Ag-Bi solid solution nanoparticles and more defects were generated during the electrodepositing process as compared to the clean structures of Ag, including nano-sized twins. Additionally, the HRTEM results confirm the difference from the lattice parameters of Ag-Bi and nanocrystalline Ag. Since the current Ag-Bi coating presents a higher hardness and low electrical resistivity, this promising method will provide a good solution for electrical contact applications. The methodology used in this study can guide the selection of processing parameters of electrodeposition in optimizing the mechanical properties of nanocrystalline coatings for intended applications.

Acknowledgments: This study was carried out by the Pacific Northwest National Laboratory (PNNL), an organization operated by Battelle Memorial Institute for the US Department of Energy (DOE) under Contract No. DE-AC06-76RL01830. The University of Auckland performed all the experimental characterization reported

Figure 10. The relationships of alloy at% in Ag matrix versus normalized (Norm.) electrical resistivity( Ω·m (Ag−X)

Ω·m (Ag) ) of Ag-Bi, Ag-Al [9], Ag-Pd [10], Ag-Ti [12], and Ag-Au [13] alloys.

5. Conclusions

In this study, the effects of Bi addition on the microstructures, mechanical performances, andelectrical resistivity of an electrodeposited Ag coating were investigated. Using nanoindentation intwo directions and the inverse calculation, we found that the Ag-Bi coating was much strongerthan Ag with a significantly higher yield strength and flow stress. Meanwhile, the anisotropyof the Ag-Bi was significantly reduced as compared to that of the Ag coating, which indicates amore homogenous microstructure. The TEM images demonstrate that the improvement in themechanical response was originated from the new Ag-Bi solid solution nanoparticles and moredefects were generated during the electrodepositing process as compared to the clean structuresof Ag, including nano-sized twins. Additionally, the HRTEM results confirm the difference fromthe lattice parameters of Ag-Bi and nanocrystalline Ag. Since the current Ag-Bi coating presents ahigher hardness and low electrical resistivity, this promising method will provide a good solutionfor electrical contact applications. The methodology used in this study can guide the selection ofprocessing parameters of electrodeposition in optimizing the mechanical properties of nanocrystallinecoatings for intended applications.

Acknowledgments: This study was carried out by the Pacific Northwest National Laboratory (PNNL),an organization operated by Battelle Memorial Institute for the US Department of Energy (DOE) under ContractNo. DE-AC06-76RL01830. The University of Auckland performed all the experimental characterization reportedin this study. This study was also supported by National Natural Science Foundation of China (51601073) andJiangsu Distinguished Professor Project (1064901601). We would like to extend our appreciation for the assistanceprovided by the technical staff in the Department of Chemical and Materials Engineering and the Research Centerof Surface and Materials Science at the University of Auckland. We also would like to express our gratitude toGlen Slater, Chris Goode, and the technical staff in Rigg Electroplating Ltd, New Zealand.

Author Contributions: Yuxin Wang and Guang Cheng conceived and designed the experiments; See Leng Tayand Yunxia Guo performed the experiments; Xin Sun and Wei Gao analyzed the data; Guang Cheng wrote thefirst manuscript; all the authors revised the manuscript.

Page 14: 1,2 ID 3, ID 2

Materials 2017, 10, 932 14 of 16

Conflicts of Interest: The authors declare no conflict of interest.

References

1. Antler, M. Electrical effects of fretting connector contact materials: A review. Wear 1985, 106, 5–33. [CrossRef]2. Findik, F.; Uzun, H. Microstructure, hardness and electrical properties of silver-based refractory contact

materials. Mater. Des. 2003, 24, 489–492. [CrossRef]3. Krastev, I.; Valkova, T.; Zielonka, A. Structure and properties of electrodeposited silver–bismuth alloys.

J. Appl. Electrochem. 2004, 34, 79–85. [CrossRef]4. Schlesinger, M.; Paunovic, M. Modern Electroplating, 5th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2011.5. Krastev, I.; Valkova, T.; Zielonka, A. Effect of electrolysis conditions on the deposition of silver-bismuth

alloys. J. Appl. Electrochem. 2003, 33, 1199–1204. [CrossRef]6. Yu, B.; Leung, K.M.; Guo, Q.; Lau, W.M.; Yang, J. Synthesis of Ag-TiO2 composite nano thin film for

antimicrobial application. Nanotechnology 2011, 22, 115603. [CrossRef] [PubMed]7. Gay, P.-A.; Bercot, P.; Pagetti, J. Electrodeposition and characterisation of Ag–ZrO2 electroplated coatings.

Surf. Coat. Technol. 2001, 140, 147–154. [CrossRef]8. Pasricha, R.; Gupta, S.; Srivastava, A.K. A facile and novel synthesis of Ag–graphene-based nanocomposites.

Small 2009, 5, 2253–2259. [CrossRef] [PubMed]9. Mao, F.; Taher, M.; Kryshtal, O.; Kruk, A.; Czyrska-Filemonowicz, A.; Ottosson, M.; Andersson, A.M.;

Wiklund, U.; Jansson, U. Combinatorial study of gradient Ag–Al thin films: Microstructure, phase formation,mechanical and electrical properties. ACS Appl. Mater. Interfaces 2016, 8, 30635–30643. [CrossRef] [PubMed]

10. Butler, W.H.; Stocks, G.M. Calculated electrical-conductivity and thermopower of silver-palladium alloys.Phys. Rev. B 1984, 29, 4217–4223. [CrossRef]

11. Chithra, S.; Lele, S.; Chattopadhyay, K. Structure evolution and phase change in Ag–5.1 at. % Bi alloy duringmechanical alloying. Acta Mater. 2011, 59, 2009–2019. [CrossRef]

12. Alford, T.; Adams, D.; Laursen, T.; Manfred Ullrich, B. Encapsulation of Ag films on SiO2 by ti reactionsusing Ag–Ti alloy/bilayer structures and an NH3 ambient. Appl. Phys. Lett. 1996, 68, 3251–3253. [CrossRef]

13. Crisp, R.; Rungis, J. Thermoelectric power and thermal conductivity in the silver-gold alloy system from3–300 k. Philos. Mag. 1970, 22, 217–236. [CrossRef]

14. Tay, S.L. Electrodeposition Coatings by Ionic Co-Discharge Technique. Ph.D. Thesis, University of Auckland,Auckland, New Zealand, 2016.

15. Kumar, M.; Deka, S. Multiply twinned AgNi alloy nanoparticles as highly active catalyst for multiplereduction and degradation reactions. ACS Appl. Mater. Interfaces 2014, 6, 16071–16081. [CrossRef] [PubMed]

16. Adamik, M.; Barna, P.B.; Tomov, I. Correlation between texture and average grain size in polycrystalline Agthin films. Thin Solid Films 2000, 359, 33–38. [CrossRef]

17. Zhong, S.; Koch, T.; Wang, M.; Scherer, T.; Walheim, S.; Hahn, H.; Schimmel, T. Nanoscale twinned coppernanowire formation by direct electrodeposition. Small 2009, 5, 2265–2270. [CrossRef] [PubMed]

18. Kumar, K.; Suresh, S.; Chisholm, M.; Horton, J.; Wang, P. Deformation of electrodeposited nanocrystallinenickel. Acta Mater. 2003, 51, 387–405. [CrossRef]

19. Navinšek, B.; Panjan, P.; Milošev, I. PVD coatings as an environmentally clean alternative to electroplatingand electroless processes. Surf. Coat. Technol. 1999, 116, 476–487. [CrossRef]

20. Wang, Y.; Ju, Y.; Wei, S.; Lu, W.; Yan, B.; Gao, W. Mechanical properties and microstructure of Au–Ni–TiO2

nano-composite coatings. Mater. Charact. 2015, 102, 189–194. [CrossRef]21. Strehle, S.; Bartha, J.; Wetzig, K. Electrical properties of electroplated Cu (Ag) thin films. Thin Solid Films

2009, 517, 3320–3325. [CrossRef]22. Elliott, R.P.; Shunk, F.A. The Ag-Bi (silver-bismuth) system. Bull. Alloy Phase Diagr. 1980, 1, 62–64. [CrossRef]23. Krastev, I.; Valkova, T.; Zielonka, A. Internal stress in multilayer silver–bismuth coatings. J. Appl. Electrochem.

2005, 35, 539–544. [CrossRef]24. Manzano, C.V.; Abad, B.; Rojo, M.M.; Koh, Y.R.; Hodson, S.L.; Martinez, A.M.L.; Xu, X.; Shakouri, A.;

Sands, T.D.; Borca-Tasciuc, T. Anisotropic effects on the thermoelectric properties of highly orientedelectrodeposited Bi2Te3 films. Sci. Rep. 2016, 6, 19129. [CrossRef] [PubMed]

25. Nakamura, T.; Gu, Y. Identification of elastic–plastic anisotropic parameters using instrumented indentationand inverse analysis. Mech. Mater. 2007, 39, 340–356. [CrossRef]

Page 15: 1,2 ID 3, ID 2

Materials 2017, 10, 932 15 of 16

26. Pei, Y.T.; Song, G.M.; Sloof, W.G.; De Hosson, J.T.M. A methodology to determine anisotropy effects innon-cubic coatings. Surf. Coat. Technol. 2007, 201, 6911–6916. [CrossRef]

27. Cheng, G.; Sun, X.; Wang, Y.; Tay, S.L.; Gao, W. Nanoindentation study of electrodeposited Ag thin coating:An inverse calculation of anisotropic elastic-plastic properties. Surf. Coat. Technol. 2017, 310, 43–50.[CrossRef]

28. Xiang, Y.; Tsui, T.Y.; Vlassak, J.J. The mechanical properties of freestanding electroplated Cu thin films.J. Mater. Res. 2006, 21, 1607–1618. [CrossRef]

29. Huang, H.; Spaepen, F. Tensile testing of free-standing Cu, Ag and Al thin films and Ag/Cu multilayers.Acta Mater. 2000, 48, 3261–3269. [CrossRef]

30. You, Z.S.; Lu, L.; Lu, K. Tensile behavior of columnar grained Cu with preferentially oriented nanoscaletwins. Acta Mater. 2011, 59, 6927–6937. [CrossRef]

31. Baker, S.P.; Nix, W.D. Mechanical properties of compositionally modulated Au-Ni thin films:Nanoindentation and microcantilever deflection experiments. J. Mater. Res. 1994, 9, 3131–3145. [CrossRef]

32. Fang, W.; Wickert, J. Determining mean and gradient residual stresses in thin films using micromachinedcantilevers. J. Micromech. Microeng. 1996, 6, 301. [CrossRef]

33. Nili, H.; Cheng, G.; Venkatesh, T.A.; Sriram, S.; Bhaskaran, M. Correlation between nanomechanical andpiezoelectric properties of thin films: An experimental and finite element study. Mater. Lett. 2013, 90, 148–151.[CrossRef]

34. Su, F.-H.; Huang, P. Microstructure and tribological property of nanocrystalline Co–W alloy coating producedby dual-pulse electrodeposition. Mater. Chem. Phys. 2012, 134, 350–359. [CrossRef]

35. Suresh, S.; Nieh, T.G.; Choi, B.W. Nano-indentation of copper thin films on silicon substrates. Scr. Mater.1999, 41, 951–957. [CrossRef]

36. Cheng, G.; Choi, K.S.; Hu, X.; Sun, X. Determining individual phase properties in a multi-phase Q&P steelusing multi-scale indentation tests. Mater. Sci. Eng. A 2016, 652, 384–395. [CrossRef]

37. Cheng, G.; Sriram, S.; Bhaskaran, M.; Venkatesh, T.A. Nanoindentation response of piezoelectric nano-islands.Appl. Phys. Lett. 2014, 105, 122902. [CrossRef]

38. Williams, J.J.; Walters, J.L.; Wang, M.Y.; Chawla, N.; Rohatgi, A. Extracting constitutive stress–strain behaviorof microscopic phases by micropillar compression. JOM 2013, 65, 226–233. [CrossRef]

39. Buzzi, S.; Dietiker, M.; Kunze, K.; Spolenak, R.; Löffler, J.F. Deformation behavior of silversubmicrometer-pillars prepared by nanoimprinting. Philos. Mag. 2009, 89, 869–884. [CrossRef]

40. Stewart, J.L.; Jiang, L.; Williams, J.J.; Chawla, N. Prediction of bulk tensile behavior of dual phase stainlesssteels using constituent behavior from micropillar compression experiments. Mater. Sci. Eng. A 2012, 534,220–227. [CrossRef]

41. Bautista, K. Four-Point Probe Operation; SP2004-TF-005 Erik Jonsson School of Engineering, The University ofTexas: Dallas, TX, USA, 2003; Available online: http://www.utdallas.edu/~gpp052000/Docs/4PointProbe_Manual.pdf (accessed on 30 April 2015).

42. Chen, J.; Bull, S. On the relationship between plastic zone radius and maximum depth duringnanoindentation. Surf. Coat. Technol. 2006, 201, 4289–4293. [CrossRef]

43. Fischer-Cripps, A.C. Introduction to Contact Mechanics, 2nd ed.; Springer: New York, NY, USA, 2007.44. Yeh, J.-W.; Chang, S.-Y.; Hong, Y.-D.; Chen, S.-K.; Lin, S.-J. Anomalous decrease in X-ray diffraction intensities

of Cu–Ni–Al–Co–Cr–Fe–Si alloy systems with multi-principal elements. Mater. Chem. Phys. 2007, 103, 41–46.[CrossRef]

45. Liu, L.; Wang, J.; Gong, S.; Mao, S. High resolution transmission electron microscope observation ofzero-strain deformation twinning mechanisms in Ag. Phys. Rev. Lett 2011, 106, 175504. [CrossRef] [PubMed]

46. Kibey, S.; Liu, J.; Johnson, D.; Sehitoglu, H. Predicting twinning stress in FCC metals: Linking twin-energypathways to twin nucleation. Acta Mater. 2007, 55, 6843–6851. [CrossRef]

47. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using loadand displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [CrossRef]

48. Elmustafa, A.A.; Stone, D.S. Indentation size effect in polycrystalline FCC metals. Acta Mater. 2002, 50,3641–3650. [CrossRef]

49. Ma, Q.; Clarke, D.R. Size dependent hardness of silver single crystals. J. Mater. Res. 1995, 10, 853–863.[CrossRef]

Page 16: 1,2 ID 3, ID 2

Materials 2017, 10, 932 16 of 16

50. Stegall, D.E.; Mamun, M.A.; Crawford, B.; Elmustafa, A. Indentation size effect in FCC metals: Anexamination of experimental techniques and the bilinear behavior. J. Mater. Res. 2012, 27, 1543–1552.[CrossRef]

51. Cheng, G.; Barker, E.I.; Stephens, E.V.; Choi, K.S.; Sun, X. Quantifying grain level stress-strain behavior forAM40 via instrumented microindentation. MRS Adv. 2016, 1, 761–772. [CrossRef]

52. Gurrappa, I.; Binder, L. Electrodeposition of nanostructured coatings and their characterization—A review.Sci. Technol. Adv. Mater. 2008, 9, 043001. [CrossRef] [PubMed]

53. Askari, H.; Zbib, H.M.; Sun, X. Multiscale modeling of inclusions and precipitation hardening in metalmatrix composites: Application to advanced high-strength steels. J. Nanomech. Micromech. 2012, 3, 24–33.[CrossRef]

54. Rhee, M.; Hirth, J.; Zbib, H. A superdislocation model for the strengthening of metal matrix composites andthe initiation and propagation of shear bands. Acta Mater. 1994, 42, 2645–2655. [CrossRef]

55. Liang, X.; Chen, J.; Mora, M.T.; Urdaneta, J.F.; Zeng, Q. Effect of precipitation on the hardness of ternarymetallic glass. Adv. Mater. Phys. Chem. 2017, 7, 255–262. [CrossRef]

56. Hansen, N. Hall–petch relation and boundary strengthening. Scr. Mater. 2004, 51, 801–806. [CrossRef]57. Schuh, C.A.; Nieh, T.G.; Yamasaki, T. Hall-petch breakdown manifested in abrasive wear resistance of

nanocrystalline nickel. Scr. Mater. 2002, 46, 735–740. [CrossRef]58. Ma, E.; Wang, Y.; Lu, Q.; Sui, M.; Lu, L.; Lu, K. Strain hardening and large tensile elongation in

ultrahigh-strength nano-twinned copper. Appl. Phys. Lett. 2004, 85, 4932–4934. [CrossRef]59. Hu, X.; Jain, M.; Wilkinson, D.; Mishra, R. Microstructure-based finite element analysis of strain localization

behavior in AA5754 aluminum sheet. Acta Mater. 2008, 56, 3187–3201. [CrossRef]60. Cheng, G.; Choi, K.S.; Hu, X.; Sun, X. Predicting deformation limits of dual-phase steels under complex

loading paths. JOM 2017, 69, 1046–1051. [CrossRef]61. Cheng, G.; Zhang, F.; Ruimi, A.; Field, D.P.; Sun, X. Quantifying the effects of tempering on individual phase

properties of DP980 steel with nanoindentation. Mater. Sci. Eng. A 2016, 667, 240–249. [CrossRef]62. Archard, J. Contact and rubbing of flat surfaces. J. Appl. Phys. 1953, 24, 981–988. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons Attribution(CC BY) license (http://creativecommons.org/licenses/by/4.0/).