Next Article in Journal
Ionic Mobility and Charge Carriers Recombination Analyzed in Triple Cation Perovskite Solar Cells
Previous Article in Journal
Electrochemical Properties of Carbon Nanobeads and Mesophase-Pitch-Based Graphite Fibers as Anodes for Rechargeable Lithium-Ion Batteries
Previous Article in Special Issue
The Self-Catalyzed Growth of GaAsSb Nanowires on a Si (111) Substrate Using Molecular-Beam Epitaxy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Bias Voltage and Substrate Temperature on the Mechanical Properties and Oxidation Behavior of CrWSiN Films

1
Department of Materials Engineering, Ming Chi University of Technology, New Taipei 243303, Taiwan
2
Center for Plasma and Thin Film Technologies, Ming Chi University of Technology, New Taipei 243303, Taiwan
3
Department of Optoelectronics and Materials Technology, National Taiwan Ocean University, Keelung 202301, Taiwan
4
Center of Excellence for Ocean Engineering, National Taiwan Ocean University, Keelung 202301, Taiwan
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(10), 1672; https://doi.org/10.3390/coatings13101672
Submission received: 31 July 2023 / Revised: 19 September 2023 / Accepted: 22 September 2023 / Published: 24 September 2023

Abstract

:
CrWSiN films were prepared through the co-sputtering technique, and the process variables were substrate bias voltage and temperature. The mechanical properties of hardness and elastic modulus of the CrWSiN films were dominantly affected by their average crystallite size and by substrate bias voltage and temperature. Moreover, the effect of substrate temperature was more evident than that of substrate bias. The highest hardness and elastic modulus of 42.6 and 459 GPa, respectively, were obtained for the Cr20W28Si9N43 film fabricated at a substrate temperature of 400 °C, which exhibits an evident advantage over the 25.0 and 323 GPa values for the Cr21W28Si9N42 film fabricated at room temperature. In contrast, an increase in negative bias voltage to −100 V on the substrate decreased the mechanical properties compared to one prepared using a similar process without applying the negative bias voltage. The oxidation resistance of the Cr-enriched Cr37W4Si10N49 and Cr37W5Si10N48 films was superior to that of the Cr20W28Si9N43 films with near-equal Cr and W contents annealed at 900 °C in air. The formation of a surficial Cr2O3 layer plays a vital role in restricting subsequent oxidation for CrWSiN films.

1. Introduction

The mechanical and antioxidative properties of transition metal nitride thin films are enhanced by the addition of Si, according to the literature on CrSiN [1,2], WSiN [3,4], and CrAlSiN [5,6] films. Si atoms doped into transition metal nitride lattices play the roles of substitutional or interstitial atoms, resulting in the solid solution strengthening of mechanical properties [1,3]. Moreover, the formation of amorphous SiNx establishes nanocomposite structures, reduces crystallite sizes, and hinders the movement of dislocations, which enhances the mechanical properties related to specific Si contents [7]. In the work of Lee and Chang [8], the CrSiN film with 10.1 at.% Si formed a nanocomposite structure and had a hardness of 24.6 GPa. Moreover, Ju et al. [3] reported that nanocomposite WSiN films with a Si content of 23.5 at.% presented a high hardness value of 40 GPa. In a previous study [9], CrWSiN thin films were explored to combine the mechanical properties of WSiN films and the antioxidative properties of CrSiN films. The formation enthalpy at 298 K is −22 kJ/mol [10] for W2N, which is not as large as the values of −744.8 and −117.2 kJ/mol for Si3N4 and CrN [11], respectively. These CrWSiN films are under-stoichiometric due to the low affinity between W and N. The structures of CrWSiN thin films are categorized into crystalline CrSiN-, amorphous WSiN-, and amorphous SiNx-dominated structures depending on their chemical compositions, and the films mentioned above exhibited distinct mechanical and antioxidative properties. The Cr21W28Si9N42 films with comparable Cr and W contents exhibit a CrSiN-dominated structure, accompanied by a hardness of 25 GPa and an elastic modulus of 323 GPa, which are the highest mechanical properties among those of the investigated CrWSiN films [9]. In contrast, the CrSiN-dominated Cr39W5Si9N47 films with Cr-enriched content exhibit a hardness of 8.6 GPa and an elastic modulus of 208 GPa, demonstrating the lowest mechanical properties [9]. Moreover, an improved oxidation resistance at 800 °C was achieved for the Cr21W28Si9N42 films due to the formation of surficial Cr2WO6 scale [9,12]. Mechanical and tribological properties are crucial for films utilized as hard coatings [13,14], whereas antioxidative properties are vital for films applied at high temperatures [5,6]. Applying substrate bias voltage (VB) and elevated substrate temperature (TS) are essential process parameters in sputtering technology that improve the characteristics of nitride thin films [15,16]. In a previous study [16], the phase, texture, and bonding characteristics of CrSiN films with a Si level of 14 at.% were not influenced by VB and TS, whereas the lattice constants and crystallite sizes were affected by VB and TS. The CrSiN films fabricated at a VB of −100 V and a TS of 400 °C demonstrated a hardness of 22.2 GPa and an elastic modulus of 299 GPa, which are superior to the values of 13.4 and 223 GPa for films fabricated at the grounded state and room temperature. The competition between grain growth and nucleation rates of the CrSiN films affects their crystallite sizes and mechanical properties [16,17]. Generally, raising the substrate temperature causes higher adatom mobility, resulting in a higher growth rate and crystallite size [18,19]. However, at much higher temperatures, high kinetic energy enhances atomic diffusion and collision, resulting in higher nucleation rates and reduced crystallite size [18,20]. Therefore, exploring the influences of VB and TS on the mechanical and antioxidative properties of CrWSiN films is vital. The aforementioned Cr21W28Si9N42 and Cr39W5Si9N47 films can be recognized as W and Si co-incorporated CrN films, but they exhibit distinct mechanical properties. Modifying the sputtering processes for fabricating Cr21W28Si9N42 and Cr39W5Si9N47 films by applying VB and TS has become a key research topic. In this study, CrWSiN films were fabricated through sputtering with VB values of 0 to −150 V and TS values of room temperature to 400 °C. The fabricated films’ mechanical properties and wear resistance were examined, and their oxidation behavior was inspected in ambient air at 800 and 900 °C.

2. Materials and Methods

The preparation of CrWSiN films with a Cr interlayer through co-sputtering was described in a previous study [9]. Direct current powers were applied to Cr, W, and Si targets. The targets mentioned above were 50.8 mm in diameter. Two batches (Batch 50 and Batch 90) of CrWSiN films were fabricated. The Batch 50 films, including samples A50, A50B(50), A50B(100), A50B(150), A50T(300), A50T(400), and A50BT, were prepared by applying DC powers of 50 W on each target (PCr = PW = PSi = 50 W). In contrast, the Batch 90 films, including A90, A90B(100), A90T(400), and A90BT, were fabricated with PCr = 90 W, PW = 10 W, and PSi = 50 W. Samples A50 and A90 were prepared using electrically grounded substrates without additional heating. Samples A50B(100) and A90B(100) were manufactured with a substrate bias (VB) of −100 V and without heating. Samples A50T(400) and A90T(400) were fabricated in a grounded state and a substrate temperature (TS) of 400 °C, whereas samples A50BT and A90BT were produced at VB = −100 V and TS = 400 °C. Three more Batch 50 samples, A50B(50), A50B(150), and A50T(300), prepared at a VB of −50 V, VB of −150 V, and a TS of 300 °C, respectively, were included in the discussion of chemical compositions, deposition rates, and mechanical properties. The antioxidative properties of CrWSiN films were evaluated after annealing in air at 800 and 900 °C for 2 h.
The chemical compositions of the CrWSiN films were determined using a field-emission electron probe microanalyzer (JXA-iHP200F, JEOL, Tokyo, Japan). The phases and texture in the films were identified using grazing incident and Bragg–Brentano (θ–2θ) X-ray diffraction (X’Pert PRO MPD, PANalytical, Almelo, the Netherlands), respectively. The texture coefficients Tc(111) and Tc(200) were calculated using the Bragg–Brentano X-ray diffraction (BBXRD) patterns as follows:
T c h k l = I m h k l I 0 h k l 1 n n = 1 n I m h k l I 0 h k l 1 ,
where Im and I0 are the measured and standard relative intensities of the (hkl) reflection [21], respectively, and n = 2 as (111) and (200) reflections were considered. The average crystallite sizes of the CrWSiN films were calculated following Scherrer’s formula [22] using the (111) reflections in the BBXRD patterns. The samples with a protective Pt layer for transmission electron microscopy (TEM) observation were prepared using a focused ion beam system (NX2000, Hitachi, Tokyo, Japan). The mechanical properties, hardness and reduced elastic modulus (Er), of the films were measured using a nanoindentation tester (TI-900 Triboindenter, Hysitron, Eden Prairie, MN, USA) and determined using the Oliver–Pharr method [23]. The elastic modulus (E) of a film was derived from
1 E r = 1 ν 2 E + 1 ν i 2 E i ,
where Ei (1141 GPa) is the elastic modulus and νi (0.07) is Poisson’s ratio of the indenter [24]. The Poisson ratio of the CrWSiN films (v) was assumed to be 0.25 [9]. The Poisson ratios of CrN, W2N, and Si3N4 were 0.24 [25], 0.29 [26], and 0.26 [27], respectively. The residual stress in the films was estimated using the curvature method [28]. The wear resistance of the films was evaluated through the pin-on-disk test [9].

3. Results and Discussion

3.1. Chemical Compositions and Phase Structures

Table 1 lists the chemical compositions of the CrWSiN films. The O contents were 0–0.8 and 1.7–3.4 at.% in the Batch 50 and 90 samples, respectively, which implies that the O pollution was contributed by the high chemical affinity between Cr and residual O in the sputter chamber since the Batch 90 samples were Cr-enriched. The O content was ignored in this study. The Batch 50 samples exhibited a chemical composition of 20–26 at.% Cr—20–28 at.% W—7–10 at.% Si—42–47 at.% N, denoted as Cr20–26W20–28Si7–10N42–47, whereas the Batch 90 samples showed Cr37–41W4–5Si8–10N47–49. The process parameters, VB and TS, exhibited limited effects on these samples’ constitutions. The process with a negative bias on the substrate first increased and then decreased the deposition rate by increasing the bias voltage related to that prepared with a grounded substrate. For example, the deposition rate of the A50B(50) sample was 6.4 nm/min, which is higher than the 5.9 nm/min of the A50 sample. Moreover, the deposition rate decreased from 6.4 to 6.1 and 5.2 nm/min with an increase in the bias voltage from −50 to −100 and −150 V due to resputtering [16,29]. The A50T(300) films exhibited a deposition rate of 6.0 nm/min, similar to that of the A50 films, whereas the A50T(400) sample showed a lower deposition rate of 5.4 nm/min. The higher substrate temperature enhances the adatom mobility and raises the evaporation rate, which decreases the deposition rate [20].
Figure 1a,b displays the grazing incident X-ray diffraction (GIXRD) patterns of the Batch 50 and 90 samples, respectively. All the CrWSiN films exhibited a face-centered cubic (fcc) solid solution consisting of CrN and W2N components with evident (111), (200), (220), and (311) reflections. Moreover, (222) reflections were observed for the A50T(400) and A50BT samples. The lattice constants were 0.4188 and 0.4146 nm for the A50 and A90 samples. The lattice constants decreased to lower levels after applying the substrate bias or substrate temperature, except for in the A50B(150) sample (Table 2). The lattice constants decreased from 0.4188 nm for the A50 sample to 0.4186, 0.4184, 0.4182, 0.4186, and 0.4182 nm for the A50B(50), A50B(100), A50T(300), A50T(400), and A50BT samples, respectively, whereas the lattice constants decreased from 0.4146 nm for the A90 sample to 0.4143, 0.4139, and 0.4139 nm for the A90B(100), A90T(400), and A90BT samples, respectively. The reflection shifts in XRD patterns were related to the variation in lattice constants of the films. These changes were not evident in the samples within batches due to small deviations in their chemical compositions. However, the lattice constants of the Batch 90 samples (0.4139–0.4146 nm) were lower than those of the Batch 50 samples (0.4182–0.4193 nm), which was attributed to a lower W content for the Batch 90 samples. Moreover, Benkahoul et al. [7] reported that CrSiN films with a Cr content of 6.7–11.6 at.% exhibited a nanocomposite structure in which amorphous SiNx phase surrounded the crystallites. Figure 2a,b displays the BBXRD patterns of the Batch 50 and 90 samples, respectively. All the CrWSiN films exhibited an fcc phase, and the Cr (110) reflections contributed from the interlayers were observed. The Cr (200) reflection was observed at a two-theta angle of 64°, not shown in Figure 2. The texture coefficients indicate that all the films exhibited a (111) orientation (Table 2). Table 2 shows the average crystallite sizes of CrWSiN films determined using the full width at half maximum (FWHM) of the (111) reflections. The crystallite size of the A50 sample was 44 nm, whereas the A50B(50) and A50B(100) exhibited a larger size of 51 nm, and the A50B(150) sample exhibited a crystallite size of 43 nm. The A50T(300) and A50T(400) showed crystallite sizes of 47 and 37 nm, respectively, which reveals distinct effects contributed by grain growth and nucleation [16]. The CrSiN films prepared at room temperature, 300, and 400 °C exhibited average crystallite sizes of 16.2, 19.2, and 14.1 nm [16], respectively. The crystallite sizes of the Batch 90 samples revealed a similar variation trend. The A90 sample exhibited an average crystallite size of 62 nm, whereas the A90B(100) prepared through a biased process exhibited a higher value of 77 nm, and the A90T(400) produced at an elevated temperature of 400 °C revealed a lower value of 37 nm. The crystallite size affected the mechanical properties of the films, which is further discussed in Section 3.2.
Figure 3a exhibits a cross-sectional TEM (XTEM) image of the A50T(400) sample, which reveals a tight columnar structure. The selected area electron diffraction (SAED) pattern of the A50T(400) sample (Figure 3b) shows an fcc phase with diffraction spots (111), (200), and (220), which represent d-spacing values of 0.242, 0.211, and 0.150 nm, respectively. Figure 3c exhibits a high-resolution TEM (HRTEM) image of the A50T(400) sample, which displays lattice fringes corresponding to d-spacing values of 0.210–0.215 nm for the fcc (200) planes. Figure 4a depicts an XTEM image of the A90B(100) sample, revealing a columnar structure with stripes between the grains. The SAED pattern of the A90B(100) sample (Figure 4b) reveals an fcc phase similar to that of the A50T(400) sample (Figure 3b). Moreover, the stripes at the interfaces between crystalline grains were identified to be amorphous in the HRTEM image (Figure 4c). These amorphous stripes could affect the residual stress and mechanical properties of the films. The amorphous SiNx phase with low density and disordered structure decreased the compressive residual stress and hardness [3].

3.2. Mechanical Properties

Table 3 shows the mechanical properties of the CrWSiN films. The Batch 50 samples with a similar chemical composition of Cr20–26W20–28Si7–10N42–47 exhibited distinct mechanical properties. The A50 sample possessed a hardness (H) of 25.0 GPa and an elastic modulus (E) of 323 GPa, whereas the A50B(100) sample prepared under a VB of −100 V revealed lower H and E values of 20.4 and 255 GPa, respectively. In contrast, the A50T(400) sample fabricated at an elevated TS of 400 °C exhibited higher H and E values of 42.6 and 459 GPa, respectively. Moreover, the A50BT sample prepared at a VB of −100 V and a TS of 400 °C showed H and E values of 40.6 and 439 GPa, respectively, which are higher than those of A50 and A50B(100) and slightly lower than those of the A50T(400) sample. Moreover, the Batch 90 films demonstrated a variation tendency in mechanical properties similar to that of the Batch 50 films, i.e., the mechanical properties revealed that A90T(400) > A90BT > A90 > A90B(100).
Crystalline structure, crystalline texture, residual stress, and crystallite size affected the mechanical properties of the nitride films. The Batch 50 and 90 samples exhibited chemical compositions of Cr20–26W20–28Si7–10N42–47 and Cr37–41W4–5Si8–10N47–49, respectively, and such CrWSiN films belonged to CrSiN-dominated structures [9]. CrSiN [30] and WSiN [4] films crystallized into a rock salt structure, and these films with a (200) texture were reported to be harder than the films with a (111) texture. Figure 5a displays the hardness values of CrWSiN films related to their texture coefficients of reflection (200) [Tc(200)]. The hardness value increased from 20.4 GPa for the A50B(100) film with a Tc(200) value of 0.01 to 20.8, 25.0, 25.3, 26.5, and 42.6 GPa for the A50B(50), A50, A50B(150), A50T(300), and A50T(400) films with Tc(200) values of 0.015, 0.04, 0.04, 0.16, and 0.29, then slightly decreased to 40.6 GPa for the A50BT film with a Tc(200) value of 0.56. In a previous study [16], CrSiN prepared at a VB of 0 to −75 V exhibited a hardness level of 13–14 GPa, and a higher H value of 19.5 GPa was obtained when the CrSiN film was prepared at −100 V. The hardness and Tc(200) values of the A50B(50) films were similar to those of the A50B(100) film, although A50B(50) and A50B(100) were prepared at VB values of −50 V and −100 V, respectively. The increase in H to 25.3 GPa by changing VB was observed for the A50B(150) prepared at a VB of −150 V. However, the increased hardness value as a result of applying a substrate bias was far lower than the effect caused by an elevated substrate temperature. Moreover, the hardness level increased from 6.6 GPa for the A90B(100) film with a Tc(200) value of 0 to 8.6, 24.4, and 26.2 GPa for the A90, A90BT, and A90T(400) films with Tc(200) values of 0.06, 0.09, and 0.13, respectively. In a previous study [16], CrSiN films with Tc(200) values higher than 1.5 exhibited hardness values of 13–22 GPa. The texture of CrWSiN films was not the primary factor that affected their mechanical properties. Figure 5b depicts the hardness values of CrWSiN films related to their residual stress values. The Batch 50 samples exhibited compressive residual stresses, whereas the Batch 90 samples exhibited tensile stresses. However, no specific relationship was obeyed between the hardness and stress values of these CrWSiN films. For example, the A50 sample possessed an H value of 25.0 GPa accompanied by a compressive stress of –0.84 GPa; in contrast, the A90T(400) sample showed a higher H value of 26.2 GPa and tensile stress of 0.58 GPa, which disagrees with the conventional judgment, that is, that compressive stress increases film hardness [31]. Figure 6a depicts the relationship between the hardness and crystallite size values of the CrWSiN films. The films with smaller crystallite sizes had higher hardness values. CrWSiN films prepared at 400 °C (A50T(400), A50BT, A90T(400), and A90BT) exhibited smaller crystallite sizes compared to the films prepared without heating the substrate holder, and this phenomenon was consistent with that observed in the study of CrSiN films due to the increased nucleation rate at 400 °C [16]. In contrast, the film A50T(300) fabricated at 300 °C exhibited a larger crystallite size than those prepared at 400 °C. The main factor affecting the crystallite size could be attributed to the formation of the amorphous SiNx tissue phase in the nanocomposite nitride coatings. Benkahoul et al. [7] proposed that Si atoms segregate on the CrN grain surface and form amorphous SiNx, which restricts the crystallite sizes. High-temperature processes provide higher mobility for the adatoms to move on the free surface. When amorphous SiNx was distributed homogeneously, the crystallized CrN/W2N phase exhibited a smaller nucleation size and a larger nucleation number related to the films deposited at a low substrate temperature. Figure 6b displays the relationship between elastic modulus and crystallite size, demonstrating a tendency similar to that between hardness and crystallite size. In summary, a VB of −100 V negatively affected the mechanical properties, whereas a TS of 400 °C increased the mechanical properties of the CrWSiN films. The average crystallite size dominantly affected the mechanical properties (hardness and elastic modulus) of CrWSiN films.
The H/E ratio [32] was used as an indicator for estimating the toughness of the films. Moreover, the H3/E2 ratio [33] was used as a gauge for evaluating the wear-resistant property of films. However, the wear resistance of the CrSiN films was not equal to that predicted by H/E and H3/E2 values [16]. The usefulness of H/E and H3/E2 in predicting film toughness was restricted for films with low plasticity [34]. The characteristics of interfaces, defects, and residual stresses on nanoscale films could affect their fracture toughness. Figure 7 depicts the relationship between H and E for CrSiN [2,16], WSiN [4,35], and CrWSiN [9] films fabricated using the same sputter apparatus. The data for Batches 50 and 90 samples are identified in this figure. The A50T(400) and A50BT samples possessed high H/E ratios of 0.092–0.093 (Table 3), whereas the A50, A50B(100), A90T(400), and A90BT samples showed median H/E ratios of 0.069–0.080, and the A90 and A90B(100) samples revealed low H/E levels of 0.037–0.041. Table 3 shows that the Batch 50 and 90 samples can be categorized into high, median, and low ranges of H3/E2 and elastic recovery (We), the same as those classified by the H/E ratio. Figure 8 shows the wear scars of the selected Batch 50 and 90 samples prepared on SUS420 substrates. The A90B(100) sample exhibited the lowest H, E, H/E, and H3/E2 values among the surveyed CrWSiN films in this study. The wear track of the A90B(100) sample revealed plastically deformed morphology, similar to that observed for the A90 sample [9]. For the films with median H/E, H3/E2, and We values, except for the A90T(400) sample, cracks occurred on the wear scars. Table 4 lists the wear test results. The wear rate of the A90BT sample was underestimated due to crack formation, and these crack fragments still adhere to the worn part. Crack formation and propagation could be induced by the residual tensile stress of the films [36]. The A90BT sample exhibited a tensile stress of 0.49 GPa, and the A50B(100) sample possessed a near-zero stress of −0.05 GPa, which could be the reason for the formation of cracks during wear tests. However, the wear track of the A90T(400) sample was similar to those of the A50T(400) and A50BT samples, which were smooth in most portions, although the A90T(400) and A90BT samples possessed similar mechanical properties and residual tensile stresses. The deviation between the thickness values of the A90T(400) (701 nm) and A90BT (937 nm) samples should be considered and requires further investigation. The wear rates of the A90T(400), A50T(400), and A50BT samples were 1.2 × 10−7, 8.5 × 10−6, and 6.3 × 10−6 mm3/(N·m), respectively. In summary, the CrWSiN films with low, median, and high mechanical properties and their H/E and H3/E2 ratios exhibited plastically deformed, crack-formed, and flattened worn morphologies, respectively. Figure 9 exhibits the coefficients of friction (COFs) recorded during the wear tests. Table 4 shows the average COFs at a 50–100 m wear distance. The COF of A90BT exhibited a high value of 0.63, indicating the effect of crack formation during testing. In contrast, the COFs of A90 and A90B(100) were 0.39 and 0.45, respectively, which corresponds to the occurrence of plastic deformation. The COFs of A90T(400), A50T(400), and A50BT were in the range of 0.46–0.50.

3.3. Oxidation Behavior

Figure 10a,b depicts the GIXRD patterns of the A50T(400), A50BT, A90T(400), and A90BT samples annealed in ambient air for 2 h at 800 and 900 °C, respectively. In a previous study [9], the GIXRD patterns of the A50 and A90 samples after 2 h of annealing at 800 °C exhibited a Cr2WO6 phase. The oxide phases of the A50T(400) and A50BT samples after annealing at 800 °C were dominated by Cr2WO6, and transformed to mixed Cr2WO6 and Cr2O3 when annealed at 900 °C. In contrast, the A90T(400) and A90BT samples exhibited a Cr2O3 phase accompanied by the original fcc nitride phase after annealing at 800 and 900 °C, which implies that only the upper part of the A90T(400) and A90BT films were oxidized. Figure 11 depicts the surficial morphologies of the 800 °C annealed CrWSiN films. Pores were observed in the crystallite corners of the A50T(400) sample due to nitrogen-loss behavior during annealing [4]. Moreover, larger oxide grains and pores are present on the surface of the A50BT sample. In contrast, the oxide scales of the A90T(400) and A90BT samples comprised needle Cr2O3 grains, and the Cr2O3 grains on the A90BT sample were more significant than those that grew on the surface of the A90T(400) sample. Figure 12 shows cross-sectional SEM (XSEM) images of the 800 °C annealed CrWSiN films. A bi-layered oxide structure formed above the columnar A50T(400) sample, similar to that observed on the A50 sample after annealing at 800 °C for 2 h [9]. The outer oxide layer of approximately 180 nm formed above the original surface of the A50T(400) sample, whereas the inner oxide layer of 220 nm indicated the inward diffusion of O. A detailed observation on the oxide layers was performed using the XTEM image, as discussed later. In contrast, locally grown oxides were observed on the surface of the A50BT sample (Figure 11), which implies that the oxide layer thickness observed from the XESM image is uncertain. The surface oxide layers of the A90T(400) and A90BT samples were 150–180 nm, with embedded pore channels within stacked Cr2O3 grains.
Figure 13a displays an XTEM image of the 800 °C annealed A50T(400) sample comprising three regions, A, B, and C, with distinct morphologies. The outer oxide layer (Region A) consisted of larger Cr2WO6 grains and rod-like Cr2O3 grains (Figure 13b). The rod-like grains grew from beneath medium grains. Figure 13c,d displays HRTEM images related to Cr2WO6 and Cr2O3 grains, respectively. The lattice fringes with d-spacing values of 0.247–0.248 nm in the larger grains indicate Cr2WO6 (103) planes. The rod-like grains exhibited lattice fringes with d-spacing values of 0.266–0.267 and 0.363–0.364 nm representing the Cr2O3 (104) and (012) planes. The portion interlaid with larger and rod-like grains had d-spacing values of 0.247–0.248 nm, which could be either the Cr2WO6 (103) or Cr2O3 (110) planes. Region B, located beneath the original surface, comprised smaller Cr2WO6 grains (Figure 13e). Region C exhibited columnar CrN/W2N grains (Figure 13f).
Figure 14a depicts an XTEM image of the 800 °C annealed A90T(400) sample. Three regions, I, II, and III, are distinguished. Regions I and III are also shown in the XSEM image (Figure 12), whereas Region II could not be indicated in the XSEM image. Region I shows that precipitations extrude above the original surface, whereas the columnar structure accompanied by amorphous boundaries is observed in Region III. Figure 14b,c exhibits the SAED patterns of Regions I and III, which exhibit the Cr2O3 and CrN/W2N phases, respectively. The Cr2O3 precipitations are large, so its SAED pattern represents a single crystal. Figure 14d displays an HRTEM image of Region I and related lattice fringes, indicating that these precipitations are Cr2O3. Also observed in Figure 11 is that these extruded Cr2O3 precipitations stack on top of each other, and voids appear that cause difficulty in protecting from inward oxidation. Figure 14e shows an HRTEM image of Region II that exhibits crystalline Cr2O3 domains with sizes of several nanometers. The dense Region II plays the role of advancing the oxidation resistance of CrWSiN films.
Figure 15 and Figure 16 show the XSEM images and surficial morphologies of the 900 °C annealed CrWSiN films. Oxide grains with dimensions of several microns were observed on the surfaces of the A50T(400) and A50BT samples. Circular dark parts on the A50BT sample imply the formation and removal of buckle oxides. In contrast, the oxidation behavior of the A90T(400) and A90BT samples at 900 °C was similar to that of these samples annealed at 800 °C. A Cr2O3 layer restricted the inward diffusion of oxygen. However, the Cr2O3 grains were round (Figure 16) but not rod-like as observed for the 800 °C annealed A90T(400) sample (Figure 11). Figure 17a depicts an XTEM image of the A90T(400) sample after annealing at 900 °C for 2 h. Compared with Figure 14a, Figure 17a displays that the three regions (I, II, and III) are sustained, and the thicknesses of Regions I and II are enlarged. Figure 17b,c displays the SAED pattern and HRTEM image of Region I of the 900 °C annealed A90T(400) sample, respectively, implying that the Cr2O3 phase dominates Region I. Region II exhibited mixed oxides of Cr2O3, Cr2WO6, and WO3 (Figure 17d). The SAED pattern of Region III exhibited Cr2O3, WO3, and fcc (CrN/W2N) phases (Figure 17e). In summary, A90T(400) and A90BT samples exhibited superior oxidation resistance compared to the A50T(400) and A50BT samples at 900 °C.

4. Conclusions

The effects of substrate bias voltage and temperature on the characteristics of CrWSiN films were studied. Two batches of CrWSiN films were fabricated, Cr20–26W20–28Si7–10N42–47 and Cr37–41W4–5Si8–10N47–49, which were near-equal Cr and W and Cr-enriched films, respectively. The effects of VB and TS on macroscopic properties, such as the films’ chemical compositions and macrostructures were insignificant, as observed from their EPMA results and XRD patterns. In contrast, the mechanical properties, hardness and elastic modulus, of the CrWSiN films were dominantly affected by their average crystallite size. The effect of TS was more evident than that of VB. The Cr20W28Si9N43 and Cr21W28Si10N42 samples with near-equal Cr and W contents prepared at a TS of 400 °C exhibited small crystallite sizes of 37 and 25 nm, high hardness values of 42.6 and 40.6 GPa, high elastic modulus values of 459 and 439 GPa, high H/E ratios of 0.093 and 0.092, and high H3/E2 ratios of 0.367 and 0.347 GPa, respectively, accompanied with low wear rates of 8.5 and 6.3 × 10−6 mm3/(N·m), respectively. The decrease in crystallite sizes of the nanocomposite films at a higher TS of 400 °C was attributed to the homogeneous and random segregation of the SiNx tissue phase surrounding the CrN/W2N crystallites, resulting in higher nucleation rates in film deposition. The CrWSiN films with near-equal Cr and W contents revealed oxidation resistance at 800 °C in ambient air, whereas the Cr-enriched CrWSiN films (Cr37W4Si10N49 and Cr37W5Si10N48) exhibited higher oxidation resistance up to 900 °C. The surficial Cr2O3 layer restricted the progression of oxidation.

Author Contributions

Conceptualization, Y.-I.C.; validation, L.-C.C., C.-H.T. and T.-Y.O.; formal analysis, C.-H.T. and T.-Y.O.; investigation, C.-H.T. and T.-Y.O.; resources, L.-C.C. and Y.-I.C.; writing—original draft preparation, Y.-I.C.; supervision, Y.-I.C.; project administration, Y.-I.C.; funding acquisition, L.-C.C. and Y.-I.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science and Technology Council, Taiwan, grant numbers 111-2221-E-019-064 and 111-2221-E-131-028. The APC was funded by National Taiwan Ocean University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The analysis support from the Joint Center for High Valued Instruments at NSYSU for the FIB is acknowledged. The analysis support from the Instrumentation Center at National Tsing Hua University for the EPMA is appreciated.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Diyatmika, W.; Cheng, C.Y.; Lee, J.W. Fabrication of Cr-Si-N coatings using a hybrid high-power impulse and radio-frequency magnetron co-sputtering: The role of Si incorporation and duty cycle. Surf. Coat. Technol. 2020, 403, 126378. [Google Scholar] [CrossRef]
  2. Chang, L.C.; Liu, Y.H.; Chen, Y.I. Mechanical properties and oxidation behavior of Cr–Si–N coatings. Coatings 2019, 9, 528. [Google Scholar] [CrossRef]
  3. Ju, H.; He, S.; Yu, L.; Asempah, I.; Xu, J. The improvement of oxidation resistance, mechanical and tribological properties of W2N films by doping silicon. Surf. Coat. Technol. 2017, 317, 158–165. [Google Scholar] [CrossRef]
  4. Liu, Y.H.; Chang, L.C.; Liu, B.W.; Chen, Y.I. Mechanical properties and oxidation behavior of W–Si–N Coatings. Surf. Coat. Technol. 2019, 375, 727–738. [Google Scholar] [CrossRef]
  5. Chen, H.W.; Chan, Y.C.; Lee, J.W.; Duh, J.G. Oxidation behavior of Si-doped nanocomposite CrAlSiN coatings. Surf. Coat. Technol. 2010, 205, 1189–1194. [Google Scholar] [CrossRef]
  6. Dong, Y.; Zhu, H.; Ge, F.; Zhao, G.; Huang, F. Microstructural effects on the high-temperature steam oxidation resistance of magnetron sputtered Cr-Al-Si-N quaternary coatings on zirconium coupons. Surf. Coat. Technol. 2019, 374, 393–401. [Google Scholar] [CrossRef]
  7. Benkahoul, M.; Robin, P.; Gujrathi, S.C.; Martinu, L.; Klemberg-Sapieha, J.E. Microstructure and mechanical properties of Cr–Si–N coatings prepared by pulsed reactive dual magnetron sputtering. Surf. Coat. Technol. 2008, 202, 3975–3980. [Google Scholar] [CrossRef]
  8. Lee, J.W.; Chang, Y.C. A study on the microstructures and mechanical properties of pulsed DC reactive magnetron sputtered Cr–Si–N nanocomposite coatings. Surf. Coat. Technol. 2007, 202, 831–836. [Google Scholar] [CrossRef]
  9. Chang, L.C.; Sung, M.C.; Chen, Y.I.; Tseng, C.H. Mechanical properties and oxidation behavior of CrWSiN films. Surf. Coat. Technol. 2022, 437, 128368. [Google Scholar] [CrossRef]
  10. de Boer, F.R.; Boom, R.; Mattens, W.C.M.; Miedema, A.R.; Niessen, A.K. Transition Metal Alloys; North-Holland: Amsterdam, The Netherlands, 1988. [Google Scholar]
  11. Barin, I. Thermochemical Data of Pure Substances, 3rd ed.; VCH: New York, NY, USA, 1995. [Google Scholar]
  12. Telu, S.; Patra, A.; Sankaranarayana, M.; Mitra, R.; Pabi, S.K. Microstructure and cyclic oxidation behavior of W-Cr alloys prepared by sintering of mechanically alloyed nanocrystalline powders. Int. J. Refract. Met. Hard Mater. 2013, 36, 191–203. [Google Scholar] [CrossRef]
  13. Shin, S.H.; Kim, M.W.; Kang, M.C.; Kim, K.H.; Kwon, D.H.; Kim, J.S. Cutting performance of CrN and Cr-Si–N coated end-mill deposited by hybrid coating system for ultra-high speed micro machining. Surf. Coat. Technol. 2008, 202, 5613–5616. [Google Scholar] [CrossRef]
  14. Kim, C.; Kang, M.C.; Kim, J.S.; Kim, K.H.; Shin, B.S.; Je, T.J. Mechanical properties and cutting performance of nanocomposite Cr–Si–N coated tool for green machining. Curr. Appl. Phys. 2009, 9, S145–S148. [Google Scholar] [CrossRef]
  15. Xiang, Y.; Huang, L.; Zou, C. Effects of bias voltages on the structural, mechanical and oxidation resistance properties of Cr–Si–N nanocomposite coatings. Coatings 2020, 10, 796. [Google Scholar] [CrossRef]
  16. Chang, L.C.; Sung, M.C.; Chen, Y.I. Effects of bias voltage and substrate temperature on the mechanical properties and oxidation behavior of CrSiN films. Vacuum 2021, 194, 110580. [Google Scholar] [CrossRef]
  17. Tan, S.; Zhang, X.; Wu, X.; Fang, F.; Jiang, J. Effect of substrate bias and temperature on magnetron sputtered CrSiN films. Appl. Surf. Sci. 2011, 257, 1850–1853. [Google Scholar] [CrossRef]
  18. Subramanian, B.; Ananthakumar, R.; Vidhya, V.S.; Jayachandran, M. Influence of substrate temperature on the materials properties of reactive DC magnetron sputtered Ti/TiN multilayered thin films. Mater. Sci. Eng. B 2011, 176, 1–7. [Google Scholar] [CrossRef]
  19. Le, V.V.; Nguyen, T.T.; Kim, S.K. The influence of nitrogen pressure and substrate temperature on the structure and mechanical properties of CrAlBN thin films. Thin Solid Films 2013, 548, 377–384. [Google Scholar] [CrossRef]
  20. Ran, Y.; Lu, H.; Zhao, S.; Jia, L.; Li, Y.; Jiang, Z.; Wang, Z. Effects of substrate bias and temperature on the structure and dielectric properties of TixZr1−xNy ternary nitride thin films. Surf. Coat. Technol. 2019, 359, 258–264. [Google Scholar] [CrossRef]
  21. Chang, L.C.; Tzeng, C.H.; Chen, Y.I. Effects of W content on structural and mechanical properties of TaWN films. Coatings 2022, 12, 700. [Google Scholar] [CrossRef]
  22. Cullity, B.D.; Stock, S.R. Elements of X-ray Diffraction, 3rd ed.; Prentice-Hall: Hoboken, NJ, USA, 2001; pp. 169–170. [Google Scholar]
  23. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  24. Simmons, G.; Wang, H. Single Crystal Young’s Constants and Calculated Aggregate Properties: A Handbook; MIT Press: Cambridge, MA, USA, 1971. [Google Scholar]
  25. Olaya, J.J.; Wei, G.; Rodil, S.E.; Muhl, S.; Bhushan, B. Influence of the ion–atom flux ratio on the mechanical properties of chromium nitride thin films. Vacuum 2007, 81, 610–618. [Google Scholar] [CrossRef]
  26. Polcar, T.; Parreira, N.M.G.; Cavaleiro, A. Structural and tribological characterization of tungsten nitride coatings at elevated temperature. Wear 2008, 265, 319–326. [Google Scholar] [CrossRef]
  27. Dai, C.L.; Chang, Y.M. A resonant method for determining mechanical properties of Si3N4 and SiO2 thin films. Mater. Lett. 2007, 61, 3089–3092. [Google Scholar] [CrossRef]
  28. Janssen, G.C.A.M.; Abdalla, M.M.; van Keulen, F.; Pujada, B.R.; van Venrooy, B. Celebrating the 100th anniversary of the Stoney equation for film stress: Developments from polycrystalline steel strips to single crystal silicon wafers. Thin Solid Films 2009, 517, 1858–1867. [Google Scholar] [CrossRef]
  29. Wang, Q.M.; Kim, K.H. Microstructural control of Cr–Si–N films by a hybrid arc ion plating and magnetron sputtering process. Acta Mater. 2009, 57, 4974–4987. [Google Scholar] [CrossRef]
  30. Martinez, E.; Sanjinés, R.; Banakh, O.; Lévy, F. Electrical, optical and mechanical properties of sputtered CrNy and Cr1−xSixN1.02 thin films. Thin Solid Films 2004, 447–448, 332–336. [Google Scholar] [CrossRef]
  31. Bielawski, M. Residual stress control in TiN/Si coatings deposited by unbalanced magnetron sputtering. Surf. Coat. Technol. 2006, 200, 3987–3995. [Google Scholar] [CrossRef]
  32. Pogrebnjak, A.D.; Beresnev, V.M.; Bondar, O.V.; Postolnyi, B.O.; Zaleski, K.; Coy, E.; Jurga, S.; Lisovenko, M.O.; Konarski, P.; Rebouta, L.; et al. Superhard CrN/MoN coatings with multilayer architecture. Mater. Des. 2018, 153, 47–59. [Google Scholar] [CrossRef]
  33. Tsui, T.Y.; Pharr, G.M.; Oliver, W.C.; Bhatia, C.S.; White, R.L.; Anders, S.; Anders, A.; Brown, I.G. Nanoindentation and nanoscratching of hard carbon coatings for magnetic disks. Mater. Res. Soc. Symp. Proc. 1995, 383, 447–452. [Google Scholar] [CrossRef]
  34. Chen, X.; Du, Y.; Chung, Y.W. Commentary on using H/E and H3/E2 as proxies for fracture toughness of hard coatings. Thin Solid Films 2019, 688, 137265. [Google Scholar] [CrossRef]
  35. Chang, L.C.; Sung, M.C.; Chu, L.H.; Chen, Y.I. Effects of the nitrogen flow ratio and substrate bias on the mechanical properties of W–N and W–Si–N films. Coatings 2020, 10, 1252. [Google Scholar] [CrossRef]
  36. Vierneusel, B.; Benker, L.; Tremmel, S.; Göken, M.; Merle, B. Isolating the effect of residual stresses on coating wear by a mechanical stress relaxation technique. Thin Solid Films 2017, 638, 159–166. [Google Scholar] [CrossRef]
Figure 1. GIXRD patterns of (a) Batch 50 and (b) Batch 90 samples.
Figure 1. GIXRD patterns of (a) Batch 50 and (b) Batch 90 samples.
Coatings 13 01672 g001
Figure 2. BBXRD patterns of (a) Batch 50 and (b) Batch 90 samples.
Figure 2. BBXRD patterns of (a) Batch 50 and (b) Batch 90 samples.
Coatings 13 01672 g002
Figure 3. (a) XTEM image and (b) SAED pattern of the as-deposited A50T(400) sample, and (c) HRTEM image of the area indicated in (a).
Figure 3. (a) XTEM image and (b) SAED pattern of the as-deposited A50T(400) sample, and (c) HRTEM image of the area indicated in (a).
Coatings 13 01672 g003
Figure 4. (a) XTEM image and (b) SAED pattern of the as-deposited A90B(100) sample, and (c) HRTEM image of the area indicated in (a).
Figure 4. (a) XTEM image and (b) SAED pattern of the as-deposited A90B(100) sample, and (c) HRTEM image of the area indicated in (a).
Coatings 13 01672 g004
Figure 5. Relationships between H and (a) Tc(200) and (b) residual stress of the CrWSiN films.
Figure 5. Relationships between H and (a) Tc(200) and (b) residual stress of the CrWSiN films.
Coatings 13 01672 g005
Figure 6. Relationships between (a) H and (b) E and crystallite size of the CrWSiN films.
Figure 6. Relationships between (a) H and (b) E and crystallite size of the CrWSiN films.
Coatings 13 01672 g006
Figure 7. Relationship between H and E for (Cr,W)(Si)N films [2,4,9,16,35].
Figure 7. Relationship between H and E for (Cr,W)(Si)N films [2,4,9,16,35].
Coatings 13 01672 g007
Figure 8. Wear scars of the CrWSiN films.
Figure 8. Wear scars of the CrWSiN films.
Coatings 13 01672 g008
Figure 9. COFs of the CrWSiN films during the wear test.
Figure 9. COFs of the CrWSiN films during the wear test.
Coatings 13 01672 g009
Figure 10. GIXRD patterns of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at (a) 800 and (b) 900 °C.
Figure 10. GIXRD patterns of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at (a) 800 and (b) 900 °C.
Coatings 13 01672 g010
Figure 11. Surficial SEM morphologies of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at 800 °C.
Figure 11. Surficial SEM morphologies of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at 800 °C.
Coatings 13 01672 g011
Figure 12. XSEM images of the 800 °C annealed A50T(400), A50BT, A90T(400), and A90BT samples.
Figure 12. XSEM images of the 800 °C annealed A50T(400), A50BT, A90T(400), and A90BT samples.
Coatings 13 01672 g012
Figure 13. (a,b) XTEM and (cf) HRTEM images of the A50T(400) sample after annealing at 800 °C for 2 h.
Figure 13. (a,b) XTEM and (cf) HRTEM images of the A50T(400) sample after annealing at 800 °C for 2 h.
Coatings 13 01672 g013aCoatings 13 01672 g013b
Figure 14. (a) XTEM image, (b,c) SAED patterns, and (d,e) HRTEM images of the 800 °C and 2 h annealed A90T(400) sample.
Figure 14. (a) XTEM image, (b,c) SAED patterns, and (d,e) HRTEM images of the 800 °C and 2 h annealed A90T(400) sample.
Coatings 13 01672 g014
Figure 15. XSEM images of the 900 °C annealed A50T(400), A50BT, A90T(400), and A90BT samples.
Figure 15. XSEM images of the 900 °C annealed A50T(400), A50BT, A90T(400), and A90BT samples.
Coatings 13 01672 g015
Figure 16. Surficial SEM morphologies of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at 900 °C.
Figure 16. Surficial SEM morphologies of the A50T(400), A50BT, A90T(400), and A90BT samples after 2 h of annealing at 900 °C.
Coatings 13 01672 g016
Figure 17. (a) XTEM image, (b,e) SAED patterns, and (c,d) HRTEM images of the A90T(400) sample after annealing at 900 °C for 2 h.
Figure 17. (a) XTEM image, (b,e) SAED patterns, and (c,d) HRTEM images of the A90T(400) sample after annealing at 900 °C for 2 h.
Coatings 13 01672 g017
Table 1. Chemical compositions, deposition rate, and thicknesses of the CrWSiN films.
Table 1. Chemical compositions, deposition rate, and thicknesses of the CrWSiN films.
Sample Chemical Composition (at.%)R 1T 2
CrWSiNO(nm/min)(nm)
A50Cr21W28Si9N4220.9 ± 0.327.6 ± 0.29.4 ± 0.141.5 ± 0.40.6 ± 0.25.9705
A50B(50)Cr26W20Si7N4725.5 ± 0.420.3 ± 0.46.8 ± 0.346.6 ± 1.20.8 ± 0.26.4831
A50B(100)Cr24W25Si9N4224.3 ± 0.025.0 ± 0.28.5 ± 0.142.0 ± 0.10.2 ± 0.26.1794
A50B(150)Cr22W24Si8N4622.0 ± 0.524.0 ± 0.27.6 ± 0.246.1 ± 0.80.3 ± 0.25.2679
A50T(300)Cr23W25Si7N4522.5 ± 0.125.3 ± 0.26.5 ± 0.445.2 ± 0.70.5 ± 0.26.0778
A50T(400)Cr20W28Si9N4320.2 ± 0.327.6 ± 0.49.4 ± 0.142.8 ± 0.70.0 ± 0.05.4700
A50BTCr21W28Si10N4220.9 ± 0.428.1 ± 0.49.5 ± 0.441.5 ± 1.00.0 ± 0.05.5719
A90Cr39W5Si9N4738.2 ± 0.44.2 ± 0.18.9 ± 0.047.0 ± 0.31.7 ± 0.16.8880
A90B(100)Cr41W4Si8N4740.3 ± 0.33.9 ± 0.27.9 ± 0.046.0 ± 0.51.9 ± 0.27.2937
A90T(400)Cr37W4Si10N4935.4 ± 0.24.3 ± 0.09.3 ± 0.147.6 ± 0.43.4 ± 0.25.4701
A90BTCr37W5Si10N4835.9 ± 0.34.4 ± 0.29.7 ± 0.147.0 ± 0.43.0 ± 0.27.2937
1 R: deposition rate. 2 T: thickness.
Table 2. Lattice constants, texture coefficients, and average crystallite sizes of the CrWSiN films.
Table 2. Lattice constants, texture coefficients, and average crystallite sizes of the CrWSiN films.
SampleLattice ConstantsTc(111) 1Tc(200) 1Crystallite Size
(nm) (nm)
A500.41881.960.0444
A50B(50)0.41861.990.0151
A50B(100)0.41841.990.0151
A50B(150)0.41931.950.0543
A50T(300)0.41821.800.2047
A50T(400)0.41861.710.2937
A50BT0.41821.440.5625
A900.41461.940.0662
A90B(100)0.41432.000.0077
A90T(400)0.41391.870.1337
A90BT0.41391.910.0940
1 Tc(111), Tc(200): texture coefficients of reflections (111) and (200).
Table 3. Mechanical properties, residual stresses, and crystallite sizes of CrWSiN films.
Table 3. Mechanical properties, residual stresses, and crystallite sizes of CrWSiN films.
SampleH 1 (GPa)E 2 (GPa)H/EH 3/E 2 (GPa)We 3 (%)Stress (GPa)D 4 (nm)
A5025.0 ± 0.6323 ± 80.0780.15361−0.84 ± 0.2244
A50B(50)20.8 ± 2.6267 ± 170.0780.12654−0.08 ± 0.0851
A50B(100)20.4 ± 1.0255 ± 60.0800.13154−0.05 ± 0.0151
A50B(150)25.3 ± 0.9278 ± 80.0910.21059−1.19 ± 0.1843
A50T(300)26.5 ± 1.4291 ± 90.0910.22059−0.21 ± 0.1847
A50T(400)42.6 ± 0.9459 ± 110.0930.36772−0.69 ± 0.0537
A50BT40.6 ± 2.2439 ± 210.0920.34771−0.42 ± 0.1625
A908.6 ± 1.1208 ± 170.0410.015350.23 ± 0.0562
A90B(100)6.6 ± 0.7179 ± 70.0370.009310.26 ± 0.0477
A90T(400)26.2 ± 2.3369 ± 260.0710.132590.58 ± 0.0737
A90BT24.4 ± 1.1355 ± 210.0690.115580.49 ± 0.1740
1 H: hardness. 2 E: elastic modulus. 3 We: elastic recovery. 4 D: crystallite size.
Table 4. Wear test results of CrWSiN films.
Table 4. Wear test results of CrWSiN films.
SampleT 1 (nm)Wear Depth (nm)Wear Width (μm)μ 2Wear Rate (mm3/(N·m))
A5070511201290.563.6 × 10−5
A50B(100)794433940.511.1 × 10−5
A50T(400)700497830.478.5 × 10−6
A50BT7193211640.466.3 × 10−6
A9088012322250.397.9 × 10−5
A90B(100)93710343240.452.5 × 10−5
A90T(400)70133930.501.2 × 10−7
A90BT707611080.632.6 × 10−7
1 T: film thickness. 2 μ: coefficient of friction.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, L.-C.; Tzeng, C.-H.; Ou, T.-Y.; Chen, Y.-I. Effects of Bias Voltage and Substrate Temperature on the Mechanical Properties and Oxidation Behavior of CrWSiN Films. Coatings 2023, 13, 1672. https://doi.org/10.3390/coatings13101672

AMA Style

Chang L-C, Tzeng C-H, Ou T-Y, Chen Y-I. Effects of Bias Voltage and Substrate Temperature on the Mechanical Properties and Oxidation Behavior of CrWSiN Films. Coatings. 2023; 13(10):1672. https://doi.org/10.3390/coatings13101672

Chicago/Turabian Style

Chang, Li-Chun, Chin-Han Tzeng, Tzu-Yu Ou, and Yung-I Chen. 2023. "Effects of Bias Voltage and Substrate Temperature on the Mechanical Properties and Oxidation Behavior of CrWSiN Films" Coatings 13, no. 10: 1672. https://doi.org/10.3390/coatings13101672

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop