3.1. Microstructural Characterisation
The top side of the deposits was characterised using optical microscopy and SEM, as seen in
Figure 1,
Figure 2 and
Figure 3. The figures show ovular primary carbides are surrounded by a eutectic matrix structure. Dark voids are present in the non-heat-treated Sample A and the 650HT Sample B. Fewer of these voids appear after heat treatment after 950 °C, as can be more easily seen in the optical microscopy pictures in
Figure 1,
Figure 2 and
Figure 3. The formation of such voids is noted in the literature to be more likely due to carbide particle pullout rather than “true” pores within the coating [
6].
The samples heat-treated at 950 °C show a marked carbide decomposition compared to the samples heat-treated at 650 °C.
The number and size of carbides were evaluated from the optical microscopy pictures. This was carried out using three equally sized images from each sample, with typical surfaces as shown previously in
Figure 1,
Figure 2 and
Figure 3.
From
Table 5, it can be seen that heat treatment decreased both the average size and number of carbides. This trend appears consistent as heat treatment temperature increases. For instance, the mean carbide size decreases from 154 µm
2 to 138 µm
2 to 98.7 µm
2 as heat treatment temperature increases from 650 °C to 950 °C. Similarly, the maximum carbide size found on the 950HT Sample C was almost half that of the maximum carbide found in the non-heat-treated original sample (322 µm
2 and 628 µm
2, respectively). Given this reduction in size and number of the carbides, this optical analysis indicates that the chromium carbides increasingly dissolve into the matrix as heat treatment temperature also increases.
Similar quantification of surface voids was undertaken using images such as those within
Figure 1,
Figure 2 and
Figure 3. As with quantification of carbides in imageJ, this was performed using three equally sized images from each sample. The results for these measurements are shown in
Table 6.
Quantification in this manner confirms the reduction in both average size and number of voids with increasing heat treatment. Interestingly, the average void size for the 650HT Sample B appears to remain close to that for the noHT Sample A (8 µm2 compared to 9 µm2), as does the maximum void size (170 vs. 189 µm2), despite having a significantly lower number of voids (75 compared to 121). This suggests that at this heat treatment temperature, the smaller voids were the most reduced, whilst the larger voids only began to shrink slightly.
SEM-EDS analysis of the uncorroded samples was also undertaken. As with the optical microscopy, the SEM imagery showed ovular-shaped chromium carbides surrounded by a matrix material. Elemental analysis by EDS was conducted on the matrix and carbides. Representative locations for the carbide and matrix EDS analysis are shown in
Figure 4.
From this, the comparison of elemental compositions of the carbides and the matrix between all samples in their as-sprayed state and post-heat treatments is shown in
Table 7 and
Table 8.
From
Table 7, the noHT Sample A had the highest nickel content in the matrix, at 89.3 wt%. This nickel content in the matrix decreases with increasing heat treatment temperature, from 87.6 to 77.6 wt% at 650 °C and 950 °C, respectively. Similarly, in the as-sprayed state, chromium was found to be 7.9 wt% in the matrix for the original noHT Sample A. This increased to 10.5%Cr in the 650HT Sample B and 20.6%Cr for the 950HT Sample C.
The composition of the carbides is provided in
Table 8. The composition of the carbides for the noHT Sample A showed the highest chromium content at 86.6 wt% and the lowest nickel content at 10.5%. With increasing heat treatment temperatures, there is a decrease in the wt% of chromium (85.4% to 81.6%) and an increase in the wt% of nickel (11.7% to 14.9%), respectively. This shows that it is likely that these carbides dissolve during heat treatment. This dissolution increases as heat treatment temperature increases. At 950HT, nickel content increased to 14.9% from 10.5 wt% in the noHT Sample A. This increase in nickel is most likely from the matrix. This is further supported by the optical microscopy results in
Table 5 showing a decrease in the size and number of carbides during heat treatment.
X-ray diffraction (XRD) was also used to gauge changes in microstructural composition during heat treatment. This confirmed the presence of Cr
3C
2 peaks at 2θ ≈ 53° and 77° [
26,
30] as well as Ni/NiCr at 2θ ≈ 44° [
26], as seen in
Figure 5. Closer analysis of the XRD spectrum over 35–55° also reveals chromium carbides present in the form Cr
23C
6, as seen in
Figure 6.
Rietveld analysis was conducted on the XRD spectra in order to gauge changes in the proportion of compounds using High Score software (version 5.2). The analysis was carried out in the range of 2θ = 35–55° where chromium carbide is expected. These results found that Cr23C6 content grew from 8.2% in the noHT Sample A to 13.9% in 650HT Sample B before reducing again to 6.4% in the 950HT Sample C, with nickel/nickel chromium being the remaining proportion to balance. This indicates that carbides were increasingly transformed with heat treatment into a more stable Cr23C6 state at 650 °C, whilst at 950 °C these carbides were being dissolved back into the matrix. It is important to note that although compositional percentage changes for Cr23C6 are given through this analysis, these are not the true compositional values due to the lack of Cr3C2 analysis.
XRD data can also be used to calculate other parameters such as crystallite size, microstrain, and dislocation density, as they provide insights into the crystallographic changes with the microstructural evolution due to heat treatment. Origin software (Version 10.1.5.132) was used to find the angle position (2
) and full width at half maximum (β) of the peaks. Using a wavelength λ of 1.54 (for the copper target) and Equations (1)–(3), crystallite size, microstrain, and dislocation density were calculated [
30]. The results are provided in
Table 9 and in
Figure 7,
Figure 8 and
Figure 9.
The microstructural evolution of Cr
3C
2-NiCr coatings subjected to heat treatment at varying temperatures reveals complex phase transformations and structural refinements that significantly impact the coating’s properties. In the as-sprayed state, the XRD pattern reveals two primary phases: Cr
3C
2 (chromium carbide) and the NiCr matrix. The Cr
3C
2 phase is identified by strong peaks at 2θ values of approximately 53° and 77°, confirming the presence of the intended hard carbide [
31]. The face-centred cubic (FCC) nickel/chromium solid solution matrix is evidenced by peaks at 2θ values around 44° [
26]. The broad, asymmetric nature of these peaks suggests some degree of alloying and potential nanocrystalline or amorphous structure in the as-sprayed matrix.
Heat treatment at 650 °C induces several notable changes in the coating’s microstructure. The relative intensity of the Cr
3C
2 peaks increases compared to the as-sprayed condition, suggesting potential precipitation of additional carbides from the supersaturated matrix. A small shift to higher 2θ angles is observed for the NiCr matrix peaks, indicating a reduction in lattice parameters due to the precipitation of dissolved carbide-forming elements. Interestingly, small peaks corresponding to Cr
23C
6 are detected, with quantitative analysis revealing approximately 13.9% Cr
23C
6 content. This indicates the onset of carbide transformation at 650 °C, consistent with the known phase transformations in the Cr-C system at this temperature [
19].
Further heat treatment at 950 °C leads to more pronounced microstructural changes. All peaks become more intense, indicating continued grain growth and stress relief. The relative intensity of the Cr3C2 peaks increases further, suggesting more extensive carbide precipitation and growth. Surprisingly, the Cr23C6 content decreases to 6.4% compared to the 650 °C sample, which may be due to either dissolution back into the matrix or transformation to other carbide phases not clearly resolved in the XRD pattern.
The XRD data also provide insights into the changes in crystallite size (D), microstrain (ε), and dislocation density for both the Cr3C2 and NiCr phases. For the Cr3C2 phase, the crystallite size decreases from 23.32 nm in the as-sprayed condition to 21.21 nm at 650 °C and further to 18.35 nm at 950 °C. Concurrently, the microstrain in the Cr3C2 phase increases from 0.00293 to 0.00323 at 650 °C and 0.00374 at 950 °C, indicating a rise in lattice distortions and defects. The dislocation density also shows an increasing trend, rising from 0.00184 × 10−9 m−2 to 0.00222 × 10−9 m−2 at 650 °C and 0.00297 × 10−9 m−2 at 950 °C.
The NiCr matrix also exhibits similar trends in its microstructural evolution. The crystallite size of the matrix phase decreases from 32.41 nm in the as-sprayed condition to 26.05 nm at 650 °C and 24.33 nm at 950 °C, indicating a refinement of the matrix crystal structure. The microstrain in the NiCr matrix increases from 0.00295 to 0.00370 at 650 °C and 0.00396 at 950 °C, suggesting a rise in lattice distortions. The dislocation density shows a significant increase from (9.52 × 10−4) × 10−9 m−2 to (1.47 × 10−3) × 10−9 m−2 at 650 °C and (1.69 × 10−3) × 10−9 m−2 at 950 °C.
These microstructural changes can be attributed to several mechanisms occurring during heat treatment. The partial transformation of Cr3C2 to Cr23C6 at higher temperatures introduces additional phase interfaces and stresses. The relaxation of residual stresses from the cold spray process is coupled with the generation of new thermal stresses during heat treatment. The refinement of the matrix structure occurs through the precipitation of dissolved elements and the formation of new interface boundaries. The observed microstructural evolution has important implications for the coating’s performance. The refinement of both carbide and matrix structures could potentially lead to changes in hardness. However, the increases in microstrain and dislocation density may affect the coating’s ability to accommodate deformation.
The XRD results correlate well with the microstructural characterisation data, which show a decrease in average carbide size from 154 μm
2 in the non-heat-treated condition to 98.7 μm
2 at 950 °C, consistent with the reduction in Cr
3C
2 peak intensities. The increase in matrix Cr content from 7.9 wt% to 20.6 wt% at 950 °C (shown in
Table 7) aligns with the observed peak shifts in the NiCr matrix, confirming significant chromium diffusion. The decrease in carbide Cr content from 86.6 wt% to 81.6 wt% at 950 °C supports the XRD evidence of carbide dissolution and transformation.
The microstructural evolution during heat treatment exhibits some intriguing and atypical behaviours. The decrease in crystallite size, coupled with increases in dislocation density and microstrain, suggests a complex interplay of mechanisms occurring within the alloy system. One possibility is the precipitation of fine secondary phases, such as carbides, within the NiCr matrix during heat treatment. It could be speculated that these precipitates may act as barriers to grain growth and even promote the formation of new, smaller grains. Additionally, if the heat treatment temperature falls within a specific range, it might trigger recrystallisation processes, leading to the formation of new, finer phases from the highly deformed cold-sprayed structure [
19].
The increase in dislocation density is equally unusual, as annealing typically reduces dislocation content. This behaviour might be explained by thermal mismatch between the NiCr matrix and CrC particles, generating new dislocations during cooling from the heat treatment temperature. Phase transformations or partial dissolution and reprecipitation of carbides could also introduce new dislocations due to volume changes and lattice mismatches [
32]. The presence of two phases with different thermal properties can lead to the development of internal stresses, further contributing to dislocation generation [
33].
The observed increase in microstrain is consistent with the rise in dislocation density, as microstrain is often directly related to the presence of lattice defects. The formation of new phases or changes in existing phases during heat treatment could lead to local lattice distortions, increasing overall microstrain. These observations suggest that the heat treatment is causing modifications to the alloy’s structure, possibly involving precipitation, phase transformations, and thermal stress effects. The simultaneous decrease in crystallite size and increases in dislocation density and microstrain indicate a delicate balance between competing mechanisms of microstructural evolution.
It is worth noting that similar complex behaviours have been observed in other alloy systems. For instance, in some high-entropy alloys, particle size effects have been shown to significantly influence dislocation density, microstructure, and phase transformations [
34]. In aluminium alloys, solution heat treatment can lead to significant changes in grain size, precipitate distribution, and consequently, mechanical properties [
35].
The heat treatment of Cr3C2-NiCr coatings results in a complex interplay of phase transformations, precipitation phenomena, and structural refinements. The formation of new carbide phases, such as Cr23C6, along with the observed changes in crystallite size, microstrain, and dislocation density, highlight the dynamic nature of these coatings under thermal exposure. The refinement of the microstructure and such temperature-induced changes are expected to impact the coating’s corrosion resistance, which is explored in the following sections.
3.2. Electrochemical Impedance Spectroscopy
Electrochemical impedance spectroscopy (EIS) was undertaken for each sample under a three-electrode setup, with artificial seawater as the solution medium. The results from the EIS are shown in
Figure 10 and
Figure 11.
Results from the EIS data can be modelled to equivalent circuit models in order to gauge their behaviour. For coatings, this is expected to follow the double layer model R
soln[C
c[R
po[C
corR
cor]]] as shown in
Figure 12 [
36].
In this model, R
soln indicates solution resistance, R
po represents porous resistance of the coating, C
c represents coating capacitance, C
cor represents the double layer capacitance, and R
cor represents charge transfer resistance [
36]. With this model, it was also found that the system was better represented by replacing the capacitors with constant phase elements (Q), which may be due to non-homogeneity within the system such as that due to imperfect coatings. As such, the representation parameters obtained from modelling this system are shown in
Table 10.
Chi-squared values found from these models were all less than 10−3, indicating a high quality of fit towards the data.
The electrochemical impedance spectroscopy (EIS) results for the heat-treated NiCr-CrC coatings reveal significant differences in corrosion behaviour among the samples, providing valuable insights into the effects of heat treatment on the coating’s microstructure and electrochemical properties. The Nyquist plot (
Figure 10) shows distinct semicircular arcs for each sample, with the 650 °C heat-treated sample (Sample B) exhibiting the largest arc diameter, followed by the non-heat-treated sample (Sample A), and the 950 °C heat-treated sample (Sample C) showing the smallest arc. This trend is consistent with the charge transfer resistance (Rcor) values extracted from the equivalent circuit modelling (
Table 10), where Sample B demonstrates the highest Rcor (1.37 × 10
6 Ω), indicating superior corrosion resistance compared to Samples A (4.61 × 10
5 Ω) and C (3.03 × 10
5 Ω). The observed differences in corrosion behaviour can be attributed to microstructural changes induced by heat treatment. At 650 °C, the coating likely undergoes beneficial transformations that enhance its protective properties. This temperature may promote the formation of a more compact and homogeneous microstructure, potentially through the reduction in porosity and the redistribution of chromium carbides. The increased Rcor for Sample B suggests the formation of a more stable passive layer, which acts as a barrier against corrosive species. Conversely, the 950 °C heat treatment (Sample C) appears to have a detrimental effect on the coating’s corrosion resistance.
The non-heat-treated sample (A) shows intermediate performance, suggesting that while the as-sprayed coating provides some corrosion protection, it can be further improved through optimised heat treatment. The capacitance values (Qc and Qcor) provide additional insights into the coating’s properties. Sample C exhibits the lowest Qc (4.88 × 10−6 Ω−1cm−2sn), which might indicate a thicker oxide layer with respect to specific phases or altered dielectric properties due to the high-temperature treatment. However, its higher Qcor (1.93 × 10−5 Ω−1cm−2sn) suggests increased surface area at the coating/electrolyte interface, consistent with the observed high porosity.
These EIS results align with findings from previous studies on heat-treated Cr
3C
2-NiCr coatings. For instance, Matthews et al. observed that heat treatment at moderate temperatures (500–700 °C) led to beneficial microstructural changes, including carbide precipitation and matrix densification, which improved corrosion resistance [
37]. However, treatments at higher temperatures (>800 °C) resulted in excessive carbide coarsening and potential chromium depletion in the matrix, compromising the coating’s protective properties [
28]. Primarily, excessive carbide dissolution occurs at this elevated temperature, where chromium carbides may dissolve into the matrix. This process can potentially create new pores or channels within the coating structure, leading to a more interconnected pore network. Such interconnectivity facilitates easier electrolyte penetration, despite the overall reduction in total porosity.
Figure 11a shows the Bode plot (phase angle vs. frequency) observed for the three samples. The behaviour in the low-frequency regime is particularly important as it relates to the electrochemical processes occurring at the electrode/electrolyte interface. In the low-frequency region, Sample B (650 °C heat-treated) exhibits a higher phase angle of 45° at 0.1 Hz, indicating superior corrosion resistance and suggesting formation of a more stable, protective passive film. This aligns with the optimised microstructure observed in the SEM analysis, where the balanced carbide size (138 μm
2) and increased Cr
23C
6 content (13.9%) potentially contribute to enhanced passivation behaviour.
Conversely, Sample C (950 °C heat-treated) shows the lowest phase angle of 30° at very low frequencies (0.1 Hz), implying that despite improved homogeneity, the 950 °C treatment may result in a less effective corrosion barrier. This could be attributed to the excessive carbide dissolution (carbide size reduced to 98.7 μm2). In the mid-frequency range at about 80–100 Hz, all samples show a continued increase in phase angle. Sample C exhibits the steepest rise, overtaking Sample A around 10 Hz and reaching the highest peak around 100 Hz. Sample B maintains a relatively steady increase, reaching its peak at a slightly higher frequency than Sample C. Sample A shows the least pronounced peak in this region.
The behaviour in the mid-frequency range often relates to the properties of the coating itself, including its porosity and overall integrity. The steeper rise and higher peak of Sample C could indicate that the high-temperature treatment at 950 °C likely led to the formation of new interfaces within the coating microstructure. These interfaces could be between different phases or at grain boundaries. The sharp peak at 100 Hz may represent the charge transfer processes occurring at these newly formed interfaces. According to recent research on Ni-Cr alloys, high-temperature treatments can lead to the formation of complex oxide layers with distinct interfaces between metal/inner oxide and inner oxide/outer oxide layers [
38]. The charge transfer across these interfaces could contribute to the observed peak. The peak also indicates a time constant typically associated with another specific electrochemical process [
39]. This peak suggests improved coating homogeneity, attributable to significant carbide size reduction (98.7 μm
2) and increased matrix Cr content (20.6 wt%) as observed in microstructural characterisation. These changes likely created new diffusion pathways within the coating. The peak at 100 Hz could represent the characteristic frequency of ion diffusion through these modified pathways, as due to the lowest pore resistance compared to other samples as observed in
Table 9. Sample B’s moderate but steady response indicates a balanced microstructure combining good coating integrity with optimal passive film properties. Sample B’s optimised carbide size and distribution likely contribute to its favourable electrochemical impedance response [
40]. Sample A lacks such a pronounced peak, suggesting a more homogeneous electrochemical response, potentially indicating a uniform passive film compared to Sample C, with higher overall corrosion resistance as observed in
Table 9.
In the high-frequency range (>1000 Hz), all samples converge towards lower phase angles, approaching purely resistive behaviour, typical in electrochemical systems and representing electrolyte resistance response [
39]. This convergence indicates similar bulk electrolyte properties for all samples. However, subtle differences in approaching this convergence point can provide information about coating pore resistance and capacitance.
Figure 11b shows the Bode plot displaying the modulus of impedance (|Z|) versus frequency on a logarithmic scale for all three samples. At low frequencies (0.1–1 Hz), the impedance values provide information about the overall corrosion resistance of the coating system. In this region, we observe that the 650 °C heat-treated sample (red line) exhibits the highest impedance, followed by the 950 °C heat-treated sample (blue line), and lastly, the non-heat-treated sample (black line). This trend suggests that heat treatment, particularly at 650 °C, enhanced the corrosion resistance of the coating. The higher impedance of the 650 °C heat-treated sample indicates a more effective barrier against corrosive species, which could be attributed to microstructural changes induced by the heat treatment. Previous studies have shown that moderate heat treatment temperatures can lead to the formation of beneficial phases and improved coating densification [
41]. The enhanced corrosion resistance at 650 °C may be due to the optimal balance between carbide dissolution and matrix phase transformation, resulting in a more homogeneous and protective coating structure. Furthermore, the behaviour in the low- to mid-frequency range provides insights into the properties of the electrochemical double layer at the coating/electrolyte interface. The higher impedance of the 650 °C sample indicates a more effective barrier against charge transfer processes, which is crucial for long-term corrosion protection [
39].