3.1 Surface morphologies and microstructure
Figure 1. shows the cross-sectional SEM morphology of bilayer multi-element alloy (AlCrNbSiTiV) doped DLC samples produced with different DC power settings. The buffer layer (interlayer) is about 163 nm thick. The DLC films grow outwards from the metallic interlayer interface and throughout the layer, which possessed an intense flux, indicating the energy distribution of impinging ions was favorable. The bilayer film structure is uniform, having a dense morphology without micro-cracks and adhering perfectly. The metallic (AlCrNbSiTiV) interlayer shows a featureless cross-sectional composition structure, increasing the adhesion strength between the film coating and substrate. Moreover, as the DC power increased from 80 to 160 watts, the growth rate of the multi-element alloy doped DLC film increased from 9.40 nm/min (thickness of 188.0 nm) to 25.82 nm/min (thickness of 516.4 nm). Universally, the depositing film should be primarily bombarded by sputtering target atoms, gas ions and plasma gas atoms, and so on, which were closely associated with the sputtering power. The increased energy received from the greater sputtering power should lead to greater particle energy, along with greater bombardment efficiency of the target materials [21]. Therefore, the sputtering energy should significantly affect the deposition rate, along with the increased thickness of the multi-element alloy doped DLC films if the DC power were monotonically increased from 80 to 160 watts.
The effect of the concentrations of C, O and multi-element alloy on the doped DLC sample at various DC power values was examined using EDS. The results are shown in Fig. 2. As the DC power was raised from 80 to 160 watts, the contents of the carbon in the DLC film samples decreased from 57.2 to 43.1 at.%. In contrast, the doping concentrations of the multi-element alloy increased almost linearly. The results indicated the concentrations of carbon and doped multi-element alloy were associated with the doping efficiency due to the varying DC power. A similar trend had been reported that increased sputtering power should decrease the concentration ratio of carbon/metal [22]. A little of oxygen (smaller than 4.2 at.%) was observed for the DLC films, which perhaps were attributed to exposure of the sample in air. Moreover, it should be associated with the residual oxygen in the vacuum chamber because of the relatively high base pressure. The elements Si and Ti were easily bonding due to their low enthalpy to form Si compounds. While the elements Al, Cr, Nb and V are another group to form Al compounds. The deposition rate of Si compoundswas slightly higher than the Al compounds group. So, the metal composition percentage of Ti was slightly higher than the Al. The metal percentages and deposition rate were due to the fact that sputtering power helps to increase the atomic mobility on the growing surface. Different DC power values should result in different degrees of doping with multi-element alloy. However, the relative chemical concentration of the DLC films were similar to that of the multi-element alloy target with an identical DC power due to the similar deposition rate. The concentration of the chemical elements is approximately equiatomic.
The micrographs and phases of the multi-element alloy doped DLC sample were determined using HR-TEM. Figure 3 demonstrates the TEM images of the DLC film corresponding to the selected area electron diffraction (SAED) patterns. It can be observed that a typical nanocomposite structure was formed in the amorphous carbon phases matrix incorporated with metal carbide or metal nanocrystalline. This differs from the DLC coating samples without doping, which showed an amorphous structure without other nano-particulates or crystals [23]. Pauschitz et al. [24] presented the TEM pattern of the composite DLC films, showing a polycrystalline structure. However, as shown in Fig. 3, randomly oriented small grains form multi-element alloy carbide nanocomposites in the amorphous DLC carbon matrix and diffraction rings are also found, which corresponding to the (111), (200), (220) and (311) planes of the face-centered cubic (FCC) structure phase. The results are also verified by XRD measurement below. The metal doped DLC composite films displayed the FCC crystalline phase, which was in agreement with the results of Li et al. [25]. The crystalline structure and the polycrystalline FCC phase should be associated with the competitive and developmental growth procedure of the multi-element alloy doped DLC coatings. [26]. Figure 4 shows the TEM micrographs and the corresponding mapping of the multi-element alloy doped DLC sample, confirming the doped films have a homogeneous chemical distribution.
Figure 5 shows the AFM and SEM images for the multi-element alloy doped DLC films grown using different DC power values. All DLC films samples showed a relatively smooth nano-particulate structure without delamination, which was homogeneous and had grain-like features on the substrate surface. These results should be relative to the typical carbon-metal structure phase composition clusters in the DLC film [27]. As shown in Fig. 5, DLC films prepared with a low DC power (80 ~ 140 watts) have relatively low surface roughness (Ra = 1.23 ~ 3.02 nm). The surface roughness of film deposited at high DC power (160 watts) has a maximum value of Ra = 4.13 nm. With increased sputtering power, the surface of the films become rougher, which was associated with the agglomeration of nanoparticles due to the increased number of target atoms reaching the substrate within the same time using the greater sputtering power. A similar result was observed in films deposited with sputtering power from 100 to 250 watts [28].
The phase structure of the multi-element alloy doped DLC composite films coated at different DC sputtering power settings was determined by means of X-ray diffraction and shown in Fig. 6. No diffraction peaks relating to the carbon phase are observed, indicating the formation of amorphous carbon matrix in the DLC film. However, the (111), (200), (220) and (311) crystal planes of a nano-grain FCC crystal phase structure can be found in the multi-element alloy doped DLC films sample, which is also confirmed by the TEM topographic images in Fig. 3. Meanwhile, the relative intensity of these diffraction peaks gradually increases with an increase in DC power from 100 to 160 watts. The DLC film samples prepared using a lower DC power exhibited a relatively low XRD pattern and poor crystallinity. It has been reported that a metal-carbon crystalline phase was formed in Ti-doped DLC samples with a high concentration of metal doping, and the fraction of the metal-carbon structure was positively correlated with the content of the metal [29]. Therefore, the low XRD pattern should be attributed to the limited amount of metal atomic of nanoparticles, which would be difficult to uniformly disperse into the amorphous DLC structure due to the lower DC power.
To investigate the chemical composition and bond structure information of the multi-element alloy doped DLC films, analysis of XPS spectra was employed. The high resolution XPS analysis of the Ti 2p, Nb 3d, V 2p, Si 2p, Cr 2p, Al 2p and C 1s regions for the multi-element alloy doped DLC films coated with DC sputtering power of 120 watts, as demonstrated in Fig. 7 (a)–(g). In addition, owing to the signal intensity from the XPS measurement is too low, the hydrogen concentration in the films was neglected, which was also examined by Dai et al. [23]. In Fig. 7(a), the main peak and the others located at ~ 74.8 eV and ~ 75.4 eV corresponding to Al in oxide and Al-O-H bonding situation, respectively, revealed the atom of Al did not bond with C to format carbide. This is also similar to the results of Dai et al. [27]. In Fig. 7(b), binding energy located at ~ 573.8 eV and ~ 575.4 eV were associated with Cr-C/Cr and Cr-O bonding situation, respectively. Because of the spectra for Cr metallic and Cr in carbide had the same bind energy, resulting in difficult to discern the chemical structures between Cr metallic and Cr carbide [27]. In Fig. 7(c), the different binding energy located at ~ 207.5 eV, ~ 210.3 eV and ~ 208 eV represented Nb-O, Nb-C and NbCO/Nb-O structures bonds, respectively, showed that metallic Nb served as compound of carbide in the DLC samples. As shown in Fig. 7(d), the bonds of Si-C and Si-O showed that the Si elements principally about bonded with C to constitute the Si-C sp3 structures [23]. Moreover, bonds of Ti-C and Ti-O was located at ~ 459.1 eV and ~ 459.9 eV, respectively, as shown in Fig. 7(e). This result indicated that Ti-doped mainly acted as compound of carbide in the DLC films [30]. As shown in Fig. 7(f), a binding energy at ~ 517.7 eV was related to the oxide structure of V elements. Additionally, from Fig. 7(g), the C1s spectra exhibited C-C/C-H structures and C-O bonds, was located at ~ 284.8 eV and ~ 286.1 eV, respectively.
Raman spectroscopy is used to determine the carbon atomic bonded characteristics of multi-element alloy doped DLC films. The results are shown in Fig. 8. The spectra display an asymmetric dispersion curve with a range of 1000 to 2000 cm− 1, which was deconvoluted using Gaussian fitting and had the typical features of a DLC structure [29, 31]. Two resulting deconvoluted sub-peaks were used to express the Raman spectra, implying a remarkable transformation of the carbon bond structure. The D-peak (‘disorder’) related to the carbon disordered structure (lower wave number at 1335 ~ 1455 cm− 1) corresponded to the breathing vibration mode of the sp2 sites in aromatic rings. The G-peak (‘graphite’) at a higher wave number (1552 ~ 1580 cm− 1) of graphitic carbon was assigned to the bond stretching vibration for all pairs of sp2 carbon elements in both the aromatic rings and carbon atomic chains [27, 32]. A higher ratio of metallic doping resulted in decreased scattering intensity of the DLC films, which should cause increased sp2 and less sp3 content in the metal-doped DLC film [33].
Figure 9 shows the change in the intensity ratio for the D band to G band (ID/IG) and the G-peak position of DLC samples with various DC power values. It is seen that the intensity ratio (ID/IG) and the G-peak position moderately reduced when the DC power increased from 80 to 120 watts. At a DC power of 120 watts, the minimum values of ID/IG and the G-peak position are 0.59 and 1552.2 cm− 1, respectively. However, ID/IG and the G-peak position value increase as the DC power increases by more than 120 watts. The position of G-peak and the intensity ratios, ID/IG, provided qualitative information about the structure of the aromatic rings and carbon chains, which were used to determine the sp2/sp3 bonding ratio [34, 35]. The results indicated the increase in DC power from 80 to 120 watts should lead to a decreasing wave numbers shift and decreasing ID/IG, which should be associated with the increased sp3 content along with the reduced (sp2-C) fraction [36]. At this phase, the consumed sp2-C in the carbide should be related to the formation of longer metallic-C bonds, leading to increased degree of disorder for the bond arrangement which was linked with higher sp3 fraction [27]. Inversely, if the DC power was more than 120 watts, the increased ID/IG should be attributed to the transform from the sp3 to sp2-C fraction. Normally, the increased fraction of distorted bond lengths should result in the increased residual stress. However, the increase of the adatom mobility induced by metal ions bombardment probably relaxed the internal stress due to the multi-element alloys doping [41]. The results also indicated that the higher DC power would not be conducive to the formation of longer metallic-C bonds.
3.2 Mechanical behavior
Nano indentation was used to determine the surface nano-mechanical behavior of the DLC film samples. The elastic modulus (E, GPa), hardness (H, GPa) and elastic recovery rate (%Re) were determined using a loading and unloading procedure [37]. The Re value represents the capability to resist plastic deformation, which is related to the toughness of the specimen, as defined by Eq. (1):
(1)
where hmax is the indentation depth at the peak load and hr is the residual depth when the load is removed [37].
Figure 10 shows the load versus displacement curves of nano-indentation tests for bilayer multi-element alloy doped DLC samples grown using different DC power settings. The films show a multiple pop-in event (abrupt displacement burst) due to the interplay between dislocations during the process of elastic-plastic deformation. The pop-ins event was complex and occurred randomly, implying nanoindentation-induced dislocation propagation [38, 39]. In Fig. 10, as the DC power increased, the maximum loading and the Re value increased, reaching a maximum value of 1.60 mN and 62.9%, respectively, at a DC power of 120 watts, and then decreasing as the DC power was increased further. For comparison, the pure DLC and AlCr co-doped DLC sample [27], reported a maximum Re value about 39% and 58%, respectively. The higher Re value should increase the hardness and the adhesion of the films, which was attributed to a change in the crystallite size and relaxation of the elastic strain [40]. Combining with the Raman spectroscopy tests at a DC power of 120 watts, a greater Re value corresponds to a lower ID/IG intensity ratio, implying the presence of more sp3 bonds, which should enhance the hardness property of DLC films.
Qiang et al. [41] presented the stress of the pure DLC thin films was up to about 0.9 GPa. For the low concentration Ti-doped DLC films (Ti concentration was about 0.21 at.%), the stress decreased significantly to 0.3 GPa (almost three times). This significant decrease in internal stress by Ti atom doping should be attributed to the increase of the sp2-C fraction and distortion of the atomic bond. Compared to the pure DLC, the elastic modulus (E) of the Ti-doped DLC films was decreased from 117 to 97 GPa, and the hardness (H) reduced from 10.5 to 10.3 GPa. Li et al. [42] reported that residual stress, H and E of films were strongly dependent on the concentrations of the co-doped Ti/Al atoms. The doped Ti mainly bonded with carbon and the doped Al existed in oxidation state. As the increase in Ti/Al concentrations, the hard Ti carbide particles were generated following the reduce in sp2 content, leading to the increase of films hardness, which showed the higher H of 16.4 GPa and E of 178 GPa. Dai et al. [30] proposed the AlTiSi multi-doped DLC films produced with high power impulse magnetron sputtering. The doped AlTiSi elements concentration was controlled by regulate the Ar fraction in the sputter gas admixture of C2H2 and Ar. As a result, the mechanical properties of the coatings enhanced as the Ar fraction increased. It can be seen that the highest E and H value for the AlTiSi co-doped DLC films were 200 GPa and 22 GPa, respectively.
Table 1 summarizes the H, E, ID/IG ratio and coefficient of friction (COF) for bilayer multi-element alloy doped DLC samples, which are grown using different DC power settings. Clearly, as the DC power increases from 80 to 120 watts, the H and E values increase to a maximum (H = 17.62 GPa, E = 193.79 GPa) and the ID/IG intensity ratio decreases to a minimum, and then decrease (the H and E) and slightly increase (the ID/IG ratio), respectively, as the DC power was further increased to 160 watts. Compared with previous reports [41–43], for DLC films sputtered with DC power of 120 watts, a decent mechanical property was obtained using metal doping. The mechanical characteristics of the DLC sample should be associated with the concentration of carbon and doped multi-element alloy, which depended on the DC power. In particular, the mechanical performance of DLC samples was attributed to the sp3 carbon structure bond inter-link matrix nanocomposite structure. Moreover, a greater sp3 concentration would lead to higher H and E values [43]. The H3/E2 value (resistance to plastic deformation) and the H/E value (resistance against fracture toughness) illustrated the elastic-plastic behavior of hard film [44] and were used to determine the effect of bilayer multi-element alloy doped DLC on the mechanical properties. As shown in Table 1, the H/E ratio is nearly unchanged, but the H3/E2 ratio increases from 0.119 to 0.146, if the DC power increased from 80 to 120 watts. However, a different trend was found if the DC power increased further. The results indicated the toughness of the DLC film was enhanced along with reasonable ID/IG intensity ratio using a proper DC power. Therefore, a proper DC power enables the coating to better accommodate deformation without cracking. Further, the mechanical characteristics of the DLC film could be optimized by properly adjusting the DC power.
The tribological features of the bilayer multi-element alloy doped DLC films sample were also investigated using a ball-on-disk tribometer. The relation between the slide distance and the COF for DLC samples coated with different DC power settings is shown in Fig. 11. Clearly, the COF changes in the range of 0.38 to 0.61 when applying various DC power values. The DLC specimen coated with a DC power of 120 watts has the lowest COF of approximate 0.38. The results are consistent with the nano-indentation tests, and the small friction coefficient should be associated with greater resistance to plastic deformation due to a proper DC power of 120 watts.
Table 1
Hardness (H), elastic modulus (E), ID/IG ratio and coefficient of friction (COF) of bilayer multi-element alloy doped DLC samples coated using different DC power settings.
DC power (watt)
|
H (GPa)
|
E (GPa)
|
H/E
|
H3/E2
|
COF
|
ID/IG
ratio
|
80
|
11.57 ± 0.19
|
114.22 ± 1.84
|
0.101
|
0.119
|
0.61
|
0.79
|
100
|
12.85 ± 0.16
|
131.56 ± 2.92
|
0.098
|
0.123
|
0.59
|
0.69
|
120
|
17.62 ± 0.13
|
193.78 ± 3.28
|
0.091
|
0.146
|
0.38
|
0.59
|
140
|
15.82 ± 0.19
|
177.95 ± 2.49
|
0.089
|
0.125
|
0.49
|
0.73
|
160
|
15.62 ± 0.12
|
173.41 ± 2.71
|
0.090
|
0.127
|
0.45
|
0.84
|
The Rockwell-C adhesion test was adopted to measure the adhesion of the coated films. Jin et al. [45] evaluated six classes (HF1–HF6) of adhesion strength of the coated films. HF1–HF2 and HF3–HF6 denoted sufficient and insufficient adhesion, respectively. In our work, Rockwell C indentation tests were performed using a total load of 1470 N to measure the adhesion of the coated films and the results are shown in Fig. 12 (a). No obvious cracking or peeling around the circumference of the indentation symbol is observed, which is characteristic of the class HF1 adhesive strength of the DLC film grown with a DC power of 120 watts.
Moreover, scratch testing is applied to determine the deformation behavior and adhesion strength of film coatings [46]. The critical scratch load LC1, LC2, and LC3 were generally used, which could be identified from the scratch crack patterns [28]. LC1 is the critical normal load to characterize the initial crack appearing in the coated specimen. Lc2 is defined when the film is initial adhesive failure from the substrate surface [47]. Critical load LC3 is determined to the loaded at which there is complete delamination of the film. Figure 12 (b) denotes the results of the scratch tracks on the multi-element alloy doped DLC samples. Clearly, cracks were only examined in the scratch track. The LC1 load was about 20.1 N and LC2 load was about 42.9 N. However, the LC3 load was not observed, demonstrating the DLC film had the acceptable adhesion. As reported by Guo et al. [3], the adhesion relation between film layer and substrate, film thickness and coefficient of friction affected the value of critical load. For (Cr, N)-DLC/DLC multilayer films, the respective LC1 and LC2 were 10.2 N and about 50.0 N, denote that well adhesion strength between coated film and substrates.
Milling operations are used to produce various components. Dry machining did not employ a cutting fluid, thus it has less environment pollution [48]. For coated and uncoated samples, the SEM images of the surface roughness of the machined workpiece and the flank wear of the cutter inserts are illustrated in Fig. 13. Clearly, for cutter inserts with a bilayer DLC coated specimen, the flank wear was significantly reduced from 52.41 ± 1.97 µm to 31.68 ± 2.16 µm (the improvement rate in flank wear is 39.6%) and the surface roughness was reduced from Ra = 3.85 ± 0.29 to Ra = 2.54 ± 0.22 µm (the improvement rate in surface roughness was 34.0%). The decreased flank wear was attributed to the lower surface roughness and the improved mechanical characteristics because of the helpful multi-element alloy interlayer. The results were consistent with the tribological features and the nano indentation test. This indicates cutting tool inserts coated with multi-element alloy doped DLC film exhibit better machining performance.