3.1.1. EDS analysis of the CrAlN layers
We have plotted the evolution of the chemical composition of the CrAlN layer as a function of the polarization voltage of the aluminum target (
Figure 2) with an average uncertainty margin of 0.1%. It is observed that logically, the chromium content decreases while the aluminum content increases. Following the results obtained (
Figure 2-a), it turns out that it is the variation in the voltage on the chromium target which is the predominant parameter. This is because the sputtering rate of chromium is higher than that of aluminum, as confirmed by the concentration of Al which remains higher than that of Al for the same voltage applied to the two targets (-900 V). On the other hand, by increasing the power applied to the aluminum and chromium targets (
Figure 2a-b) to 2800 W and 1500W, respectively, we obtain a comparable sputtering rate on the two targets (Cr and Al), and thus almost identical percentages of the two elements in the CrAlN layer. Indeed, the sputtering yield deposition rate of Al is 1.08 atoms / ions and that of Cr is 1.36 atoms/ions [
34,
35,
36,
37]. The nitrogen percentage decreases slightly with the increasing power of the Al target up to 2800 W, indicating a slight substoichiometry for aluminum rich films. Similar results have been shown by Romero et al. [
38] and López et al. [
39].
Moreover, the overall content (Cr +Al) of the metallic elements remains substantially constant, because of the nitrogen content in the films. However, we note that the constant sensitivity is more regulated on CrAlNR than on CrAlNS films, which proves the impact of the rotation of the substrate on the homogeneity of the film. The results confirm that the structure of the CrAlN films synthesized in this work forms a solid solution allowing the substitution of Cr atoms by Al atoms because the radius of the covalent bond of Al (about 0.125 nm) is smaller than its counterpart of Cr (about 0.139 nm). The positions of Al atoms in the basic unit cell of the microstructure of a Cr lattice provide greater homogeneity to the CrAlNR layer. Due to the deposition temperature and deposition conditions meeting the thickness constraint of 2-3 µm, the nitrogen (N) atoms and their connection with the substrate atoms appear to be negligible. However, to achieve constructive results for further studies, investigations into the Cr-Al-N ternary phase diagram and the interactions with the substrate at the interface are being considered for both stationary and rotational modes.
In rotation mode, the impact of density and plasma fluxes on the incorporation of reactive species has been investigated in deposition experiments, and the preferred incorporation of Nitrogen (N) occurs when the growing coating surface faces the ion source. Thus, the growing surface is positioned in front of a region of high plasma density and characterized by large fluxes of film-forming species. The preferred incorporation of N takes place in a region of low plasma density where small fluxes are present. As the growing surface approaches the ion source under the effect of the rotational movement, low modulations in composition are observed (
Figure 3). This is caused by the rotation of the substrate since the growing coating surface is periodically exposed to regions of high plasma density and large fluxes of film-forming species and regions of low plasma density and small fluxes. These findings are highly relevant to all reactive industrial processes of plasma-assisted PVD coatings using substrate rotation.
Nitrogen (N) is easily incorporated in the zone under high plasma density and relevant flux of film-forming species (substrate in front of the ionization source) [
12]. It is logical to estimate that the reactions of non-metallic species in areas with high and low plasma densities are significantly different. CrAlN films are synthesized in abundance of nitrogen and in regions subject to high plasma density. Hence, the activation of nitrogen can build up by N
2+ and N
+ [
12]. However, for the growth conditions employed in our study, the modulations seem sensitive to variations caused by the rotation of the substrate, in the density of the plasma and related to changes in the flux of film-forming species. It is reasonable to think that the modulations of the different chemical species have implications on the performance of the coating (homogeneous, mechanical and tribological properties…). Therefore, these results are very relevant to the correct conduct of industrial processes of working under vacuum (plasma) with a mixture of reactive gases and under a rotation/stationary substrate. This is because the active surface (growth surface) will regularly encounter regions with fluctuating plasma density (high and low) that feature varying fluxes (significant and minor). As a result, the modulations of the chemical composition identified in this study at the nanoscale are influenced by periodic variations in plasma density and the fluxes of the film-forming species. These variations are induced by the rotation of the substrate. In addition, chemical modulations can be expected to occur in all plasma-assisted reactive industrial PVD processes using substrate rotation and may have drastic implications on the properties of the functional coatings. This modulation behavior is attenuated under the effect of the optimization of the speed values (Eq. 1 and Eq. 2). This relationship makes it possible to reduce the amplitude and the period of the modulations, and therefore the homogeneous growth of the CrAlN layer.
3.1.2. Morphology of the CrAlN layers
Transmission electron microscopy is a very precise technique which gives clearer information on the growth mode of thin films, their morphology, and their structure (
Figure 4a). To better compare, we chose the sample coated with 5% Aluminum, which exhibits better tribological performance, as shown in the work of Tlili et al. [
6]. We noticed that the layer is nanostructured over 25 nm at the substrate/deposition interface, with a period of 3 to 3.5 nm. Growth then continues with a columnar structure in the form of grains surrounded by an amorphous matrix (
Figure 4a). These observations correlate with the X-Ray Diffraction DRX analyzes (
Figure 6) which showed a large peak for the CrAlN layer (5% Al) proving that it is amorphous. Wang et al. [
36] have shown that the addition of Al allows the formation of amorphous/nanocrystalline nanocomposites and observed the presence of several nanocrystalline particles, in the form of super-lattices, of more than 20 nm in grain size and 0.20 nm period, surrounded by an amorphous structure. This crystalline phase is attributed to CrN (111) while Al is implanted in the amorphous phase. Further away from the substrate/coating interface, the CrAlN layer has a two-phase structure, formed of CrN and AlN. This phenomenon attests that the doping of Al in the CrN matrix strongly promotes the development of a mixed amorphous /crystalline nanocomposite microstructure. Our results are in line with those of Bobzin et al. [
22] where only a crystal phase attributed to a metastable structure of CrAlN (200) is observed.
Figure 4b represents an electron diffraction image of the CrAlN
s layer (5% Al). The rings obtained are discontinuous which confirms the inhomogeneity of the structure of the obtained film. Two inter-reticular distances could be measured, the first of 2.45 nm and the second of 2.04 nm. It is difficult to identify the first distance, while the second is measured according to Bobzin et al. [
22], relating to a new cubic structure of (Cr,Al)N having an inter-planar distance of 2.03 nm. Bobzin et al. [
22] carried out a Gaussian study of the diffraction peak obtained for a CrAlN layer of 29.6% Cr, 12.5% Al, 54.5% nitrogen and traces of carbon and oxygen. They showed that the DRX peak of this new phase is at an angle of 43 °, between that of Cr
2N (111) and CrN (200), using a copper anode (λ
Cu Kα = 325 nm). The comparison of the results obtained by transmission electron microscopy (TEM) (
Figure 4b) with those of DRX (
Figure 5), leads us to believe that the layer of CrAlN at 5% Al is a multiphase composite (Cr
2N, CrN and AlN).
The surface topography, given by atomic force microscopy (AFM) in
Figure 4, is globally uniform with some dimes and tiny craters spread all over the area (
Figure 4c). The dimensional measurements show that the domes have an average diameter (dm) around 30 nm, the craters have a maximum depth (P
max) of 73 nm, and the roughness (RMS) is about 10nm (CrAlNs at 5% Al). The dimensions of the CrAlN layers increase with the bias voltage (or Al%), which is justified on the same coating at 28% Al, with dm measuring approximately 55 nm, and P
max is of the order of 112 nm. The root-mean-square (RMS) surface roughness values of the CrAlN coating with the proportions of aluminum are quite different. Apparently, the surface roughness is proportional to the ratio of aluminum. The CrAlN coating is very smooth with an RMS value of about 9.8 nm for 5% Al, while the surface roughness increases to 16.5 nm for 28% Al (
Figure 4c).
Consequently, the particle size of the CrAiN
S film increases as the proportion of aluminum increases (about 10 nm for 5% Al and 20 nm for 28% Al). This change in particle size is accompanied by a phase change in the layer and an overall alteration in the coating’s structure. These findings align with the general observations of sputter deposition, wherein the deposition rate increases as the substrate-target distance decreases [
36].
In the case of stationary substrates, the substrate-target distance was the shortest and kept constant, and therefore, the substrates are more exposed to the denser and constant flux of adatoms from the sputtering target. Further, during the rotation of the substrate, the substrate-target distance varies from 70 mm to 180 mm with a weaker ion cloud of the flow of adatoms, and the deposition rate decreases. Thus, the structure is columnar with low density (
Figure 5a). In addition, the shadow effect is maximum in the case of a rotating substrate, which promotes the porosity of the CrAlN
R film (
Figure 5b). For the lowest percentage of Al (3.5%), AFM analysis shows small grains covering the entire surface of the substrate. These particles of tiny craters and domes of different sizes are distributed randomly. By doping more Al (24%), we observe that domes and tiny craters still exist all over the area, but the surface is a little denser, and the grain size is higher. This coincides with
Figure 4b and with other papers [
6]. This alteration in the film surface can be attributed to the high ionization rate and ion projection enhanced by the high bias voltage applied to the substrate. Thus, the growth rate of the CrAlN
R layer is very high, which results in a 3D growth on the top surface of the coating (
Figure 5b).
Based on the analysis of TEM images for the ternary CrAlN
R layer (
Figure 5), we find that this surface layer is composed of a well-developed crystalline phase following relatively arranged grain geometries, where the grain size measures about 20 nm. In addition, we notice that the large grains of this film are superimposed on small grains of the order of a few nanometers, and on a larger scale, the crystalline grains of average size of around 5 nm (black area) are developed in the part of the amorphous structure (white area) (
Figure 5c). The System Architecture Diagram (SAD) of the coating shows and specifies the presence of the diffraction peaks (111), (200) and (220) associated with a cubic phase structure centered on the face. This could be attributed to the nanocrystalline Cr-Al-N solid solution integrated in an amorphous matrix. Subsequently, the enlargement of the image of this structure (
Figure 5d), shows nanoparticles distributed along an interplanar crystal spacing of 0.206 nm, which is consistent with the orientation of the typical crystal plane of CrN (111). As a result, Al is mainly present in the amorphous structure. The High-Resolution TEM (HRTEM) image of
Figure 4c shows that the doping of Al in the CrN matrix strongly fosters the development of a mixed amorphous/crystalline nanocomposite micro-structure.
Overall, during this study by AFM, the variability of the RMS roughness as a function of the bias voltage of the substrate (i.e., the power applied to the target) shows a proportionality. This increase in surface roughness is due to an increase in atomic motion and coating densification due to the increased flow and ions energy. Moreover, structural densification and grain development may increase the surface roughness [
40]. In fact, the increase in the polarization voltage generates a high mobility of the adatoms and subsequently a nucleation density (
Figure 4c,e). Additionally, under high-energy ion bombardment, highly mobile adatoms can diffuse into inter-grain voids, making the film more compact and denser [
42]. High levels of Argon ion bombardment at a higher bias voltage also tend to improve etching and result in a smoother surface [
43,
44]. This has been demonstrated for high proportions of aluminum, such as CrAlNs with 28% of Al and CrAlNR with 24% of Al, which can benefit from high-energy bombardment [
45].
The previous descriptions are based on thermodynamic equilibrium assumptions. To better understand the growth of the coating, one must look at the elementary processes at the atomic level. The species resulting from the incident flow are adsorbed and can also be desorbed. The desorption being thermodynamically activated is negligible for the metals which are deposited at moderate room temperature. If the atoms remained at the impact site, the fluctuation of the incident flux would result in faster increasing roughness.
We could model this growth as non-interacting atomic columns subjected to a Poisson-type particle flow. The roughness would be the standard deviation of this distribution, which is proportional to the square root of the thickness [
38]. However, surface atomic diffusion (adsorbed atoms) ensures the redistribution of atoms. The diffusion of adsorbed atoms is much easier than volume diffusion (lower activation energy), and even at room temperature, it is active for metals. However, there are mechanisms that hinder the surface-smoothing effect of diffusion for adsorbed atoms. These hindrances are associated with the inherent characteristics of the crystal lattice, such as the adsorption of atoms in crystal lattice gaps or the formation of stable islands of a certain size. The movement of adsorbed atoms can occur through upward or downward diffusion, establishing mechanisms for their mobility.
3.1.3. XRD analyses of the CrAlN layers
Figure 6 shows the diffractograms of CrAlN
s films for different Al percentages as well as of the silicon substrate. The latter has been added to facilitate the identification of peaks by focusing only on those corresponding to the layers and not on the substrate. On the diffractogram of the CrN layer (0% Al), there is a large peak exhibiting almost amorphous layer. This peak is perhaps the result of the contribution of several others such as: the Cr
2N (111) observed at 50.42 °, Cr
2N (200) observed at 50.58° from the hexagonal phase Cr
2N (hcp), and CrN (200) observed at 51.20° from the face-centered cubic phase of CrN (fcc). The EDS analyzes of this film give a Cr/N ratio of 2.5, which suggests that under these deposition conditions Cr
2N is obtained instead of CrN or a two-phase mixture.
According to the previous results, Al addition improves the crystallization of the CrAlN
s layers and promotes the formation of different crystalline phases. Indeed, several peaks were observed in the case of cubic CrN (c-CrN (110) at 43.58° and (200) at 51.20°), hexagonal Cr
2N (h-Cr
2N (110) at 44.15°, (111) at 50.42° and (200) at 50.58°), cubic AlN (c-AlN (111) at 45.05°, (200) at 52.49 ° and (400) at 53.71°) and hexagonal AlN (h-AlN (002) at 42.10° and (101) at 44.45°). Comparable results have been obtained in other studies [
42]. A translation of these peaks towards the large angles has been visualized, which shows that CrAlNs films contain compressive stresses, a result already found in other studies [
43]. At 24% Al, a single AlN (101) peak is present at 44.45°. This peak disappears at 51% Al to give way to a large AlN (002) peak detected at 42.10°, and the CrAlN films become again amorphous. These last two peaks relate to the hexagonal AlN phase (hcp), which shows that there has been a crystal change from a cubic phase (C-AlN) to a hexagonal phase (H-AlN).
Moreover, equilibrium diagrams of Cr-N and Al-N [
42] showed that the formation of C-CrN and C-AlN is only possible at 50% nitrogen. However, according to AFM analyzes (
Figure 5), the nitrogen content in the CrAlN layers varies between 28 and 33%. These low nitrogen contents would allow the formation of hexagonal Cr
2N and AlN instead of cubic CrN and AlN. It is consequently probable that our layers are in the form of multiphases where we have the formation of cubic CrN and AlN when the stoichiometric ratios Cr/N and Al/N are close to 1. But if these ratios are close to 2, the hexagonal Cr
2N phases and AlN form.
Figure 6 shows the diffractograms of CrAlN
R films with different Al contents (Variation of the target power). All films show the AlN (101), CrN (200) and CrN (202) peaks. For a small amount of Al (3.5%), the cubic phase peak of AlN (101) is identified at the angle of 43.8°. The CrN (200) and CrN (202) phases are also detected at the angles 51.5° and 75.04°. The AlN (101) peak is broad with low intensity. This may be due to the low amount of Al and therefore low crystallinity. The CrN peak (200) is shifted towards the large angles. This shows that CrAlN
R films exhibit residual compressive stresses. By increasing the percentage of Al to 7%, the intensity of the AlN (101) phases increases. This good crystallinity is probably attributed to the J
ion/J
atom ratio with J
ion the ionic flux density and J
atom the flux density of atoms during layer growth [
43]. According to Wang [
45], this ratio is proportional to the I
b/R
b ratio where I
b is the bias current of the substrate during deposition and R
b is the deposition rate. They have shown that this ratio increases with the I
Al/I
Cr ratio. This increase in ion bombardment and the resulting improvement in ad-atom mobility are beneficial for the development of the crystalline phase of CrN (200). In addition, with the increasing low-energy ion bombardment, defects are reduced, and a denser microstructure is developed. Therefore, the crystallinity of CrAlN films improves with increasing Al content (or target Al flux) at low Al/Cr ratio values. However, as the Al content increases, the distortion of the lattice becomes larger due to the substitution of the Cr atoms by the smaller Al atoms. This results in a decrease in crystallinity. As the percentage of Al in the layer increases from 3.5% to 24%, so the intensity of the AlN (101) phases increases. This improvement in the crystallinity of CrAlN
R films with an Al content of 19 and 24% is explained by the increase in the crystallization rate [
44]. Indeed, the particle size of the CrAlN
R film with an Al content of 11% is about 38 nm. For an Al level of 24% in the layer, the size increases to 77 nm (
Figure 6). This result of grain size increase for a high bias voltage is already confirmed by AFM analysis (
Figure 5).
Following various analyzes carried out by DRX on CrAlNs and CrAlNR, we deduce that neither the growth nor the phases are similar. In fact, the impact of the rotation affects the morphology, the structure, the phases and even the composition of the coating. The increase in the proportion of aluminum does not lead to a further increase in the N content. At this stage, the CrAlN films are not saturated (N/(Al + Cr) <1); that is, the material properties of CrAlNR films are like those of a partially nitride coated film. The Al/(Al+Cr) ratio is maintained at values below 0.5, which favor the formation of the h-AlN (101) phase in the CrAlNs films. It should be noted that the oxygen content of all films is less than 0.5%.
When Cr is mixed with AlN, the larger Cr atoms partially occupy the Al sites to form a solid solution of CrAlN, thus changing the lattice parameters as follows: CrN (0.41480 nm) and AlN (0.4045 nm) [
45]. It should be noted that the hexagonal structure characteristic of AlN crystals appears in the two CrAlN films (rotating and stable substrates), due to the low ratio (Al / (Al + Cr)) <0.5; that is, below the critical point at which AlN would form. At 30% Al (CrAlNs) content, a broad peak appears due to the appearance of c-CrN (200) and c-AlN (200), with the mixture of a dense and well crystallized structure.