Next Article in Journal
Degradation of HVOF-MCrAlY + APS-Nanostructured YSZ Thermal Barrier Coatings
Previous Article in Journal
Review on Metal (-Oxide, -Nitride, -Oxy-Nitride) Thin Films: Fabrication Methods, Applications, and Future Characterization Methods
Previous Article in Special Issue
Controlled Compositions in Zn–Ni Coatings by Anode Material Selection for Replacing Cadmium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Fe Ions on the Surface, Microstructural and Optical Properties of Solution Precursor Plasma-Sprayed TiO2 Coatings

1
Center for Nanoscience Research, Premier Research Institute of Science and Mathematics (PRISM), Mindanao State University–Iligan Institute of Technology, Andres Bonifacio Ave., Tibanga, Iligan City 9200, Philippines
2
Department of Physics, Mindanao State University–Iligan Institute of Technology, A. Bonifacio Avenue, Iligan City 9200, Philippines
3
IT and Physics Department, College of Natural Sciences and Mathematics, Mindanao State University–General Santos, Brgy. Fatima, General Santos City 9500, Philippines
4
Faculty of Mechanical Engineering, Wrocław University of Science and Technology, Lukasiewicza 5, 50-371 Wrocław, Poland
5
Faculty of Electronics, Photonics and Microsystems, Wrocław University of Science and Technology, Janiszewskiego 11/17, 50-372 Wrocław, Poland
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(8), 870; https://doi.org/10.3390/coatings15080870
Submission received: 23 June 2025 / Revised: 19 July 2025 / Accepted: 21 July 2025 / Published: 24 July 2025

Abstract

This work investigates on how Fe incorporation influences the surface, microstructural, and optical properties of solution precursor plasma-sprayed TiO2 coatings. The Fe-TiO2 coatings were prepared using titanium isopropoxide and iron acetylacetonate as precursors, with ethanol as the solvent. X-ray diffraction analysis revealed the existence of both anatase and rutile TiO2 phases, with a predominant rutile phase, also confirmed by Raman spectroscopy. There was an increase in the anatase crystals upon the addition of Fe ions. A longer spray distance further enhanced the anatase content and reduced the average TiO2 crystallite sizes present in the Fe-added coatings. SEM cross-sectional images displayed finely grained, densely packed deposits in the Fe-added coatings. UV-Vis spectroscopy showed visible-light absorption by the Fe-TiO2 coatings, with reduced band gap energies ranging from 2.846 ± 0.002 eV to 2.936 ± 0.003 eV. Photoluminescence analysis showed reduced emission intensity at 356 nm (3.48 eV) for the Fe-TiO2 coatings. These findings confirm solution precursor plasma spray to be an effective method for developing Fe-TiO2 coatings with potential application as visible-light-active photocatalysts.

1. Introduction

Hydrogen production through photocatalytic water splitting utilizes semiconductor materials like titanium dioxide (TiO2) to absorb light energy and catalyze the splitting of water into hydrogen and oxygen, offering a promising avenue for sustainable and environmentally friendly hydrogen generation [1,2,3]. While TiO2 is widely used in photocatalysis, its efficiency is limited by a high optical band gap (~3.2 eV) and a high recombination rate of electron–hole pairs (EHPs) [4,5]. To enhance its photocatalytic activity, it is necessary to extend the light absorption of TiO2 to the visible spectrum, which can be achieved by modifying its optical properties through incorporation of other elements and compounds such as transition metals and metal oxides [6,7,8]. Metal/semiconductor (M/S) composites enhance photocatalytic performance by improving the charge separation at the semiconductor interface and increasing the visible light absorption [9]. Numerous metal-added TiO2 nanocomposites with various structural variations have been developed for hydrogen production, such as Cu-TiO2, Au-TiO2, Pt-TiO2, and others [10,11,12,13]. The introduction of foreign elements (dopants) into the surface of a material alters its electronic structure, which could result in increased light absorption, improved charge carrier separation, and enhanced overall efficiency. This approach makes doped coatings a promising approach for advancing the performance of photocatalysts.
The transition metal iron (Fe) is a suitable dopant for TiO2 due to the similar ionic radii of Fe3+ (0.64 Å) and Ti4+ (0.68 Å) [14,15]. The addition of Fe narrows the band gap of TiO2 by introducing intermediate energy levels, enabling visible light absorption and a photo-response [16,17]. Moreover, Fe3+ ions can act as electron–hole traps, reducing the recombination rates as the energy levels of Fe2+/Fe3+ align close with those of Ti3+/Ti4+ [14]. This improves the ability of TiO2 to absorb visible light effectively. Several methods, including the sol-gel method [18,19,20], metal organic chemical vapor deposition [21,22], and reactive magnetron sputtering [18], have been used to deposit Fe-added TiO2 coatings, with studies reporting enhanced photocatalytic activity of TiO2 under visible-light irradiation, as reported elsewhere [19,23,24,25].
The group led by Dholam [18] demonstrated that Fe-doped TiO2 coatings derived via the sol-gel method achieved high H2 production rates during photocatalytic water splitting. Despite this result, this coating fabrication method is time consuming, which limits its scalability. Furthermore, it results in fine porosity [26], which limits its efficiency. High porosity is important in order to expose the holes to a high surface area in the smallest volume possible in order to react with water [27]. Another work also presented enhanced H2 production using Fe(III)-doped TiO2 nanoparticles compared to pure TiO2 nanoparticles [28]. However, nanopowders pose health risks [29], so their usage should be limited.
The top layer of the metal-oxide-based hybrid photocatalyst plays a critical role in their overall functionality. They are ideal for adsorption due to their high surface area and microstructural features [30]. The active sites and high surface area enable influential adsorption events with excellent chemical stability [31,32]. During light excitation, processes such as electron-phonon scattering, irreversible electron transition, involvement of holes during oxidation, and electrons undergoing reduction reaction significantly influence charge recombination and, consequently, photocatalytic efficiency. Thus, understanding the processing technique and precursor composition and properties is essential for enhancing the performance of photocatalytic coatings.
The present work is focused on the investigation of the influence of Fe ions on the surface, microstructural, and optical properties of TiO2 coatings developed using the solution precursor plasma spray (SPPS) process. This innovative method allows for the production of homogenous large-area coatings and strong substrate adhesion over conventional methods [33,34]. While TiO2 coatings developed using SPPS have been studied [35,36,37,38,39,40], there are only a few reports regarding SPPS-deposited TiO2 incorporated with metal ions and analysis of their properties for photocatalytic applications [40,41,42]. This study provides a rare insight into how Fe incorporation affects the light-responsive behavior of SPPS-deposited TiO2, which remains largely unexplored in the prior literature.

2. Experimental Section

2.1. Solution Precursor Preparation

A 1.0 L 5 mol% Fe-added TiO2 solution was used in the plasma spray process. The TiO2 solution precursors were prepared by dissolving titanium isopropoxide (TTIP, Sigma-Aldrich, St. Louis, MO, USA, 97%) in an absolute ethanol (Scharlau, Barcelona, Spain, 99%), achieving a concentration of 0.6 M, followed by continuous stirring at 300 rpm using a magnetic stirrer. Simultaneously, the required amount of iron(II) acetylacetonate (Sigma-Aldrich, 97%) was thoroughly dissolved in ethanol by vigorous stirring at 500 rpm while heating to 60 °C. The TTIP and ethanol solution was then gradually diluted by adding the stirred Fe+–ethanol solution dropwise. To avoid gelation and ensure thorough mixing, a final stirring step of 30 min was performed.

2.2. Operational Plasma Spray Parameters

A direct current SG-100 plasma torch (Praxair S.T., Indianapolis, IN, USA) was used to deposit the TiO2 and Fe-TiO2 coatings on grit-blasted stainless-steel substrates with a thickness of 2.0 mm. Argon (Ar) and hydrogen (H2) were used as the primary and secondary plasma working gases, respectively. The plasma spray experiments were conducted at stand-off distances of 40, 50, and 60 mm. The plasma torch was controlled at a speed of 500 mm/s to ensure successful deposition of individual particles on the temperature-controlled substrates. The direct deposition of TiO2 and Fe-TiO2 coatings via SPPS has been previously reported [43,44]. The detailed spray parameters are listed in Table 1 based on the reference work [44], while Table 2 presents the samples along with the specific parameter values used during spraying.

2.3. Characterization of the Coatings

Scanning electron microscopy–energy dispersive X-ray (SEM-EDX) spectroscopy was used to examine the microstructure of the plasma-sprayed TiO2 and Fe-TiO2 coatings at low and high magnifications. Cross-sectional images of the coatings were obtained at 5000× magnification using a Tescan Vega3 SEM (Brno, Czech Republic) equipped with a backscattered electron (BSE) detector. Imaging was conducted under the following conditions: accelerating voltage of 20.0 kV, view field of 60.7 µm, and working distance of 14.19 mm. Moreover, the internal coating structure was analyzed on polished cross-sections at 80,000× magnification using SEM/Xe-PFIB FEI Helios G4 PFIB CXe microscope (Hillsboro, OR, USA) under a 2.00 kV accelerating voltage, utilizing a 2.59 µm view field and a 4.2 mm working distance. Additionally, elemental mapping analysis was conducted to assess the spatial distribution of the elements within the coating. The crystallinity and crystal structure of all the samples were characterized using X-ray diffraction (XRD) using an Ultima IV (Rigaku, Tokyo, Japan) system utilizing Cu Kα radiation with a wavelength of 1.54056 Å. The XRD patterns were recorded over a 2θ range of 10° to 70° with a scanning rate of 2°/min at room temperature. This was also used to determine the crystalline phase content and average crystallite sizes of all the SPPS coatings. The Raman scattering (RS) spectra were obtained using a LabRam HR800 (HORIBA/Jobin-Yvon, Kyoto, Japan) spectrometer fitted with an Ar+ laser at 514.55 nm. The laser power was kept at 50 mW while data was being collected within the spectral range from 50 to 4000 cm−1. Photoluminescence (PL) spectroscopy was used to analyze the spontaneous light emission from a photo-excited material, providing insights into the molecular adsorption and surface reactions in photocatalytic material. The PL spectra were detected using a HORIBA FluoroMax Plus spectrofluorometer at room temperature with an excitation wavelength of 300 nm. The ultraviolet–visible (UV-Vis) light absorption of the coatings was evaluated on the basis of the reflectance spectra. The UV-Vis spectra were acquired using a QE6500 spectrometer (Ocean Optics, Orlando, FL, USA) with a coupled halogen–deuterium lamp as a light source in the wavelength range of 250–800 nm. The diffused reflectance spectra were used as another approach to estimate the band gap of the coatings. The Tauc plot ( F ( R ) × E ) 1 / 2 was plotted as a function of the incident photon energy. The band gap of each coating was determined by fitting the reflectance spectra data according to the following equation:
F R × E 1 / 2   =   A h ν E g ,
where F R = K / S is the Kubelka–Munk function ( K and S are the absorption and scattering coefficients, respectively), h ν is the photon energy, A is a constant number, and E g is the band gap energy.
As a supporting analysis, the solution viscosity was measured prior to the spraying experiments. Additionally, under the experimental conditions described, white nanometric powders were obtained, dried, and subsequently analyzed using differential thermal analysis–thermogravimetric analysis (DTA-TGA) using STA Regulus 2500 (NETZSCH, Waldkraiburg, Germany) to assess their thermal behavior. This analysis was also used to determine the phase transition temperature of TiO2. The powder samples were heated from room temperature to 1000 °C at a rate of 10 °C/min in a nitrogen atmosphere. XRD analysis was also conducted on powders heated at 450 °C.

3. Results

3.1. SEM-EDX

Figure 1 and Figure 2 depict typical polished cross-sections of the SPPS TiO2 and Fe-TiO2 coatings. Unlike conventional atmospheric plasma-sprayed coatings, which typically exhibit coarse splat boundaries, these features are absent in the SPPS TiO2 and Fe-TiO2 coatings obtained in the current work. Notably, no horizontal or vertical fissures are observed in the coatings. All the coatings can be compared to a two-zone microstructure, characterized by grainy particles and compact regions, as previously described in the reference paper [45].
The magnified cross-sectional images of TiO2 and Fe-TiO2 reveal a homogeneous microstructure with good adhesion to the stainless-steel substrate. The dense TiO2 coating’s internal structure, shown in Figure 1b, consists of finely grained particles with clearly defined grain boundaries, considered to be one of the zones mentioned. On the other hand, the internal structure of the Fe-TiO2 coatings, shown in Figure 2b, features significantly finer and more tightly packed granules compared to the pure TiO2 coatings. This difference can be attributed to the changes in the thermal behavior and particle formation of TiO2 and Fe-TiO2 solution droplets when introduced into the plasma jet. Furthermore, the thickness of all the coatings ranges from 5 to 12 μm, showing the versatility of SPPS in depositing thin and thick coatings. The SEM images of the other TiO2 and Fe-TiO2 coatings sprayed at 50 mm and 60 mm are provided in the Supplementary Information.
Additionally, the appearance of lighter spots indicated by arrows within the TiO2 coatings is described as one of the two zones, which is also noticeable in Figure 2a. These spots are likely associated with processed TiO2 agglomerates. This observation is supported by the EDS images in Figure 3 and Figure 4, which indicate a significant presence of Ti and O elements in Figure 3c,d and Figure 4b. Figure 3 illustrates the EDS mapping of the 60TiO2 coating’s cross-section. The BSE micrograph in Figure 3a is the region selected for the elemental analysis boxed in Figure 1a. The Ti mapping, shown in Figure 3c, reveals its strong and widespread presence in the coating, which confirms that Ti is the main constituent of the coating. Figure 3d displays O uniformly distributed and co-localized with Ti, which signifies the formation of a pure TiO2, especially as the C in Figure 3e only surrounds that specific dense region where T and O dominate. Fe, shown in Figure 3f, is barely detected in this sample, supporting the suggestion that it is Fe-free, in contrast to Figure 4. The EDS images in Figure 4a present the elemental mapping of the 60Fe-TiO2 coating’s cross-section. Figure 4a shows the selected EDS mapping area. The coating appears as the top, relatively denser, clearly distinct from the substrate. The Ti mapping in Figure 4b shows an intense detection, which also confirms its majority in the coating. Figure 4c displays the elemental distribution of Fe, showing a slightly lower intensity in the coating region, which suggests co-deposition with Ti, and its presence may have influenced the morphology of the Fe-TiO2 coatings. The low visible Fe signal may be attributed to the low concentration of deposited Fe ions, which is likely below the detection limit of the equipment. Similar to the O mapping in Figure 4e, it is distributed throughout both the coating and the substrate, which limits the ability to confirm TiO2 formation solely from this data and necessitates verification through other characterization techniques.

3.2. XRD Results

The X-ray diffractograms of all the coatings were analyzed using the ICSD patterns for the anatase and rutile phases of tetragonal TiO2. Figure 5 displays the XRD patterns of the TiO2 and Fe-TiO2 coatings at various spray distances, revealing that the coatings are polycrystalline and consist of both rutile and anatase phases. This bi-phasic composition is beneficial for photocatalysis, as TiO2 photocatalysts containing both phases often exhibit significantly higher catalytic activity than either phase alone [46].
The addition of Fe ions influenced the peak intensities of the stainless-steel substrates, with increased austenite (111) and (200) peaks observed at 2θ = 43° and 2θ = 54°, respectively. The XRD spectra of the Fe-TiO2 coatings support the EDS mapping results, showing minimal detectable traces of Fe and that Fe3+ ions may have substituted Ti4+ ions at some TiO2 lattice sites. The absence of additional peaks corresponding to other TiO2- or Fe-related phases suggests that no secondary crystalline phases are present when the high-temperature SPPS process (>900 °C) is used, in contrast to reports in the literature [47] on coatings produced by lower-temperature processes (<900 °C).
The percentages of the rutile and anatase phases in the coatings sprayed at varied stand-off distances were calculated from the XRD data. The following equation was used to calculate the anatase phase content ( f A ) of the as-sprayed coatings using the peak intensities of anatase (101) and rutile (110) TiO2 denoted by ( I A ) and ( I A ), respectively [44]:
f A = I A I A + 26 I R
For the pure TiO2 coatings, the anatase-to-rutile ratios were 11.3% anatase and 88.7% rutile for 40TiO2, 21.6% anatase and 78.4% rutile for 50TiO2, and 26.4% anatase and 73.6% rutile for the as-sprayed 60TiO2 coating. The incorporation of Fe further increased the anatase phase content: 40Fe-TiO2 showed 20.5% anatase and 79.5% rutile, 50Fe-TiO2 had 32.4% anatase and 67.6% rutile, and 60Fe-TiO2 exhibited 39.4% anatase and 60.6% rutile. As shown in Figure 6, Fe addition to the TiO2 coatings consistently reduced the rutile content while enhancing the anatase phase. Likewise, increasing the stand-off distance led to a further decrease in the rutile content and a corresponding increase in the anatase content.
The Scherrer formula below was used to estimate the crystallite sizes from the measured width of the crystal planes’ diffraction curves [48]:
t h k l = 0.9 λ β cos θ B
where t h k l is the crystallite size, λ = 0.15045 nm is the wavelength of X-ray radiation, β is the peak intensity full width at half-maximum (FWHM), and θ B is the Bragg angle of reflection of a specific crystal plane (hkl). Figure 7 shows the calculated average crystallite sizes for anatase and rutile, all of which are below 20 nm. Without Fe, the rutile crystals have comparable sizes for all the stand-off distances, which are around 14 nm, whereas the anatase crystallite sizes exhibit a broader distribution. In contrast, the Fe-TiO2 coatings display a more uniform anatase crystallite size, while the rutile crystallite sizes vary depending on the stand-off distance.
Figure 7 also shows that adding Fe to the solution affects the average growth of TiO2 crystallites as the stand-off distance changes. For the pure TiO2 coatings, the average crystallite size increases with an increasing stand-off distance. However, in the presence of Fe, the trend is reversed, with the average crystallite size decreasing as the stand-off distance increases.

3.3. Raman Spectroscopy

Raman spectroscopy provides detailed structural information about TiO2 coatings, making it an essential tool for characterizing their crystalline properties and allowing extended analysis of their phase composition. Figure 8 presents the Raman spectra of the TiO2 and Fe-TiO2 coatings, revealing sharp and well-defined peaks indicative of high crystallinity coating material. All the TiO2-based coatings exhibited the prominent characteristic Raman peaks of rutile, corresponding to the symmetries of B1g, Eg, and A1g, confirming the presence of the rutile TiO2 phase [49,50]. The B1g mode at 180 cm1 is associated with bending vibrations involving Ti and O atoms within the coatings. The peaks at 400 cm1 and 600 cm1 correspond to the Eg and A1g stretching and bending vibration modes of the oxygen atoms in the TiO2 lattice, respectively [51]. These findings support the XRD results, further confirming that rutile TiO2 is more abundant than anatase TiO2 in the coatings. In addition, a broad vibrational peak between 235 and 270 cm1, arising from multiple-phonon scattering processes, is distinctly observed in all the samples [49].

3.4. Photoluminescence

Figure 9 shows the PL spectra of the TiO2 and Fe-TiO2 coatings excited at 300 nm. All the samples exhibit broad emission bands within the 360–520 nm range, with two notable regions: emissions around 340–375 nm, attributed to direct band-to-band recombination [52,53], and broader emissions between 380 nm and 500 nm, centered around 439 nm, associated with oxygen vacancies and surface defect states [53,54,55,56].
The strong emission at 356 nm (3.48 eV) for TiO2 corresponds to direct band-to-band recombination [52]. In contrast, the Fe-TiO2 coatings exhibit significantly lower luminescence intensity in this wavelength range, indicating that Fe incorporation effectively suppresses direct e-h+ recombination. This quenching effect is attributed to Fe ions acting as electron-trapping centers, thereby enhancing the charge separation. The shifting of this peak observed at 360–370 nm in the PL spectra of Fe-TiO2 in Figure 9d–f may prove the latter, reflecting the energy state of excited electrons prior to recombination, which suggests intermediate energy levels introduced by Fe ions. Additionally, the emission peaks observed between 380 nm (3.26 eV) and 500 nm (2.48 eV) in the TiO2 samples, such as the emissions of wavelengths 382–384 nm, 403–405 nm, 425 nm, 431–442 nm, 466–468 nm, and 489–492 nm, are attributed to surface defects and e--h+ recombination at oxygen vacancy sites within the coating structure [53,54,55,56]. The presence of Fe ions increases the number of oxygen vacancies and surface defects, as evidenced by the broad emission bands spanning this region, as illustrated by the 400–401 nm, 438–439 nm, 468 nm, and 495–496 nm emissions wavelengths. These key spectral features provide qualitative insights into the impact of Fe addition on the recombination behavior of EHPs, ultimately supporting the role of Fe in improving the charge separation efficiency within the TiO2 matrix.

3.5. UV-Vis Results

The absorption curves of the pure TiO2 and Fe-TiO2 coatings are shown in Figure 10. As shown in the spectra, the pure TiO2 coatings exhibited a strong broad absorption in the UV region, attributed to the optical band gap of TiO2. The onsets of the absorption edges for the pure TiO2 coatings are at 375 nm, 391 nm, and 399 nm for increasing spray distances. The Fe-TiO2 coatings exhibit UV absorption peaks but also feature an absorption tail extending into the visible region. The localized spectra indicate that the onset of absorption occurs at 433 nm, 427 nm, and 424 nm for the Fe-TiO2 coatings sprayed at 40, 50, and 60 mm distances, respectively.
Using the following equation, λ = 1240 / E g , the calculated band gaps of the TiO2 coatings are determined to be 3.310 ± 0.326 eV, 3.171 ± 0.434 eV, and 3.108 ± 0.414 eV for increasing spray distances. On the other hand, the band gap energies of the 40Fe-TiO2, 50Fe-TiO2, and 60Fe-TiO2 coatings are 2.864 ± 0.377 eV, 2.904 ± 0.314 eV, and 2.924 ± 0.284 eV, respectively.
The band gap energies were obtained directly from the plot with the use of the baseline approach [57,58,59,60,61,62]. The band gap energies of the Fe-TiO2 coatings are 2.846 ± 0.002 eV, 2.920 ± 0.003 eV, and 2.936 ± 0.003 eV for the 40, 50, and 60 mm stand-off distances, respectively, which align well with the results presented in Figure 10. The fitted band gap energies from the absorption spectra and the Tauc calculation are summarized in Table 3. It was observed that the band gap energies of pure TiO2 decreased with an increasing spray distance, while at the same time, the band gap energies increased for the Fe-TiO2 coatings. This result can be related to the average TiO2 crystallite size deposited on the substrate with and without the addition of Fe. As the spray distance increases, the average crystallite size of TiO2 increases, while for the Fe-TiO2 coatings, the crystallite size decreases.

4. Discussion

4.1. Influence of Fe Ions on the Microstructural Properties of TiO2 Coatings

Addressing the grainy morphology of the coatings involves identifying the specific cause, which may be linked to the formulation, application process, or environmental conditions. A difference in the morphology between TiO2 and Fe-TiO2 coatings is observed. As illustrated in Figure 1, the Fe-TiO2 coatings exhibit a fine-grained structure compared to the coarser morphology of the TiO2 coatings. Since both coatings were obtained using the same spray parameters, the variation in morphology is likely due to the addition of Fe ions to the feedstock solution, which influenced the plasma environment during the spraying process.
The modification of the properties of the plasma vicinity altered the chemical and physical changes that the solution underwent. As reported in [63], there was an increase in the plasma ability of heating, which is caused by the Fe ions as additional mobile species present in the solution. This modification allows Fe-TiO2 particle droplets to remain within the plasma for a longer duration, exposing them to higher temperatures than the TiO2 droplets. As a result, the Fe-TiO2 coatings consist of smaller deposited grainy particles. Additionally, this phenomenon can be inferred from the recorded substrate temperature during the spray process. Using thermocouples, the substrate temperature was measured, revealing a maximum of ~340 °C for TiO2 and ~525 °C for Fe-TiO2. The temperature of the substrate for TiO2 remained relatively stable, whereas the Fe-TiO2 substrate temperature increased progressively throughout the deposition process. A higher substrate temperature correlated with a higher plasma jet temperature, further explaining the formation of finer particles due to the rapid vaporization and fragmentation of the solution droplets during the spray process. Despite the morphological differences, the coating thicknesses for each spray distance remain comparable, as both coatings were deposited under the same spray conditions.
This claim is further supported by the viscosity measurements of the TiO2 and Fe-TiO2 solution precursors. Two flow curves were measured for each sample before the spray process, covering a shear rate range from 10 to 130 s−1. The viscosity values were initially taken using an increasing shear rate ramp, followed by a subsequent decreasing ramp from 130 to 10 s−1. Figure 11 presents the viscosity measurements of the prepared solutions. Both solutions showed similar average viscosity measurements for both ramp tests. However, a slight difference was observed, with the Fe-TiO2 solution displaying a lower viscosity compared to pure TiO2. Since the Fe-TiO2 solution has lower viscosity than TiO2, it has a greater tendency to form tiny droplets under primary atomization and evaporate during spraying [64]. This leads to the faster breaking up of droplets and the formation of fine molten deposits. Indeed, the size of the deposited particles is dependent on the viscosity and the evaporation dynamics of the precursor solution.
Additionally, the XRD results confirmed the presence of both rutile and anatase phases, with rutile being the dominant phase. This is advantageous for photocatalysis, as mixed TiO2 phases generally exhibit higher catalytic activity than either of the component phases [46,65,66]. Both coatings displayed a more pronounced rutile (110) peak compared to the anatase (100) peak, indicating a higher fraction of rutile TiO2 in the sprayed coatings. However, the phase percentage calculations at different stand-off distances revealed an increase in the anatase content and a drop in the rutile content with the addition of Fe ions. These observations can be explained by the thermal behavior of the powders used as coating precursors.
In Figure 12, the precursors undergo sequential processes of solvent vaporization, decomposition, crystallization, and phase transformation. The crystallization of amorphous TiO2 into anatase TiO2 is marked by an exothermic peak at a temperature of ~423 °C, as verified in its first derivative plot, and is comparable with the temperatures previously reported [34,67]. When Fe was added, the crystallization of TiO2 shifted to a higher temperature of ~454 °C, indicating that Fe ions significantly influence the crystallization of TiO2. The presence of Fe prolonged the formation of the anatase phase, probably due to Fe ions slowing down the growth of crystalline TiO2 [68]. Despite this, the anatase-to–rutile phase transition temperature for TiO2 and Fe-TiO2 was at 812–824 °C. This also explains the increased anatase content when Fe was added, as a lower percentage of anatase was fully transformed into rutile TiO2. In many cases, an increased anatase content tends to result in better TiO2 photocatalytic activity, and its combination with rutile further enhances the catalytic activity [66,69,70]. From the literature, the synergistic interaction between the anatase (011) and rutile (110) planes enhances the charge separation and suppresses the electron–hole recombination, resulting in good photocatalytic activity [71].
Moreover, no characteristic peaks corresponding to iron oxide phases were observed in the XRD spectra of any of the coating samples. This can be attributed to the low concentration of Fe ions, and hence, could not be detected by XRD. This is further supported by the XRD spectra of the heated Fe-TiO2 precursor powders shown in Figure 13, which also did not exhibit any diffraction peaks for Fe. This implies that Fe is below the detection limit and that the Fe3+ ions may have substituted Ti4+ in some of the TiO2 lattice sites [72,73]. The incorporation of Fe ions into the TiO2 lattice structure likely contributed to the prolonged TiO2 crystallization temperature, thereby influencing the anatase and rutile phase content of the Fe-TiO2 coatings. Moreover, the anatase peaks in Figure 13 are dominant upon the addition of Fe, which conveys an increased anatase content in the sprayed Fe-TiO2 coating.
For both coating types, increasing the stand-off distance resulted in a higher anatase phase content, as particles experienced longer dwell times in the plasma jet. Conversely, a shorter stand-off distance caused a faster temperature rise, exceeding 600 °C, which promoted the transformation of nanocrystalline anatase into rutile TiO2. This explains why the coatings sprayed at 40 mm developed larger rutile crystals, due to the increased heat exposure and anatase’s high surface energy, and also, the higher percentage of rutile formation. Thus, the farther the coating from the solution injection point, the fewer the anatase content that transforms into rutile TiO2.
Regarding the crystallite size, the rutile phase exhibited uniform crystal sizes across all the stand-off distances in the TiO2 coatings. However, the anatase crystals in the Fe-TiO2 coatings maintained a consistent average size at all the stand-off distances. This implies that the incorporation of Fe atoms into some TiO2 crystals effectively controlled their growth, resulting in a wider range of anatase crystal sizes.

4.2. Influence of Fe Ions on the Optical Properties of TiO2 Coatings

All the TiO2-based coatings exhibited the characteristic stretching peaks of rutile without any trace of anatase TiO2. Since Raman spectroscopy is a point-to-point characterization technique, it is likely that rutile TiO2 crystals were primarily detected during the measurement. This confirms the XRD results, which indicate higher rutile TiO2 content than anatase in the coatings. Furthermore, there are no significant changes in the Raman spectra upon the addition of Fe to TiO2, suggesting that the phase content and average crystal sizes of the TiO2 and Fe-TiO2 coatings are comparable. It is noteworthy to mention the appearance of multi-phonon scattering in the Raman spectra. This happens especially when the coatings are being excited by a laser pulse, generating an EHP, suggesting an electron upconversion via intervalley transition, particularly in nanostructured TiO2 [74]. The excited electron, after relaxation, is prohibited from recombining in a hole that occupies a different valley [75]. As soon as the electron is delivered by upconversion, it can immediately recombine with the hole in the valence band, releasing a photon that the PL intensity detector can measure.
PL reflects charge carrier recombination near the semiconductor surface, making its intensity and spectrum sensitive to surface reactions and indicative of photocatalytic efficiency. Evidently, in Figure 9, recombination of the charge carriers is evident in both the TiO2 and Fe-TiO2 coatings. The differences in the PL spectra between these samples indicate that the Fe ion introduces new light-emitting phenomena. These observations are summarized in the proposed PL mechanism for Fe-TiO2 coatings illustrated in Figure 14.
The process starts with electron excitation from the valence band to the conduction band and shallow states [76]. The relaxed photogenerated electrons in the conduction band can then follow multiple recombination paths [77]. In (path a), a direct recombination happened, emitting 3.48 eV of energy (356 nm), as observed in the TiO2 coatings. The reduced PL intensity at this energy when Fe ions were introduced is attributed to the enhanced metal–metal (Fe-Fe) interactions, acting as luminescence quenchers [78].
Moreover, shallow states, which are initially empty electron traps, can be populated from band-tail states. These electrons can subsequently relax by settling into trap states, caused by Fe ions, located significantly below the conduction band edge. These trapped electrons will recombine to a hole in the valence band (path b) [79,80], like the PL emission with wavelength 361–370 nm (3.43–3.35 eV). Another dominant pathway for above-band gap excitation involves the recombination of conduction band electrons with deep empty states created by OH or trapped holes (path c) [76,81]. This can be presented by 468 nm light emission of 2.65 eV energy. Also, defects can cause PL emissions [82]. Several PL emission peaks, such as 400 nm (3.08 eV) and 439 nm (2.82 eV), are attributed to the presence of various defect states and impurities in the coatings. Defect emissions are the results of oxygen vacancies. These oxygen vacancies serve as recombination and/or carrier trap sites, offering alternative pathways for energy transfer and contributing to the additional emission features in the PL spectrum [79]. The emission at 2.52 eV results from the recombination of an e- and an h+, where the electron is situated from an oxygen vacancy located in the sub-bandgap as a dense trap state (path d).
The presence of oxygen vacancies confirms the dominance of the TiO2 rutile phase in all the coatings. Aside from the very high temperature of the plasma jet that helped attain the phase transformation, oxygen vacancies offer space for atomic arrangement and promote anatase–rutile transformation [83]. This also accounts for the higher PL peak intensities observed in the Fe-TiO2 coatings compared to pure TiO2. The substitution of Fe3+ ions into the TiO2 lattice introduces oxygen vacancies as a result of charge compensation, which increases the PL emission [84]. It can be supported by the possible emergence of impurity energy levels within the band gap, primarily associated with Fe states. The Fe ions act as trapping sites that inhibit direct recombination. The PL emission energy of 3.48 eV, corresponding to the direct recombination of e- and h+, aligns with the band gap transition energy. This is comparable to the calculated band gap energy of 3.33 eV from the UV-Vis spectroscopy via the Tauc plot.
The addition of Fe to catalyst-free TiO2 precursors produces denser coatings composed of ultrafine-grained particles, which is beneficial for photocatalytic applications. Nanostructured coatings have a high surface atom fraction, improving the light absorption. In pure TiO2, light absorption primarily occurs via interband electron transitions, which are direct but limited due to crystal symmetry constraints. Minimal absorption results, as direct transitions are typically disallowed. However, when absorption occurs at the crystal boundaries, such as surfaces or interfaces, it allows indirect electron transitions. These transitions significantly enhance the light absorption, which will contribute to improving the photocatalytic efficiency in Fe-TiO2 coatings [85].
This means that when a material is made up of nanocrystals, just like the SPPS Fe-TiO2 coatings, a reasonable enhancement of the absorption can be observed. As shown in Figure 10, the SPPS Fe-TiO2 coatings exhibit significantly enhanced light absorption in the visible region. This enhancement can be attributed to the nanoscale dimensions of the particles, which result in a high surface-to-volume ratio and a substantial fraction of surface atoms. Notably, the increase in interfacial absorption becomes particularly pronounced when the particle size approaches ~20 nm or smaller [85,86]. The XRD results revealed that the TiO2 crystalline sizes are less than 20 nm, which justifies the absorption of Fe-TiO2 in the visible region.
The Fe-TiO2 coatings have lower calculated band gap values than TiO2. This confirms the successful incorporation of Fe into TiO2 because of the potential appearance of certain impurity energy levels within the band gap that are primarily made of Fe. The valence and conduction bands gradually broaden in the presence of Fe, which causes the band gap to decrease. Thus, the addition of Fe ions greatly improves the activity of TiO2 in absorbing visible light.
Plasma spraying of a catalyst-free Fe-TiO2 solution precursors proves to be an effective method for producing nanostructured, crystalline Fe-TiO2 coatings with excellent deposition quality and precisely controlled anatase content. Furthermore, the emergence of traps because of the shallow states caused by Fe ions and oxygen vacancies, the multi-phonon scattering, and the large surface-to-volume ratio of the SPPS Fe-TiO2 coating creates the possibility of well-timed utilization of photogenerated carriers in interfacial processes for photocatalysis.

5. Conclusions

This study presents the development of Fe-TiO2 coatings via the innovative SPPS process and explores the influence of Fe ion incorporation on the surface, microstructural, and optical properties of TiO2. The properties of the TiO2 coatings for photocatalytic application were found to improve upon the addition of Fe ions. Cross-sectional images revealed dense deposits with fine grains in the Fe-TiO2 coatings, indicating the successful incorporation of Fe ions into the TiO2 structure through substitution, facilitated by the close radii of Fe3+ and Ti4+. Structural analysis revealed the presence of both anatase and rutile TiO2 phases, with the increased rutile content confirmed by both XRD and Raman spectroscopy. Interestingly, Fe incorporation also promoted anatase crystal formation, especially at longer spray distances, which additionally contributed to a reduction in the average crystallite size. These structural changes were accompanied by a significant enhancement in visible-light absorption and a decrease in band gap energy, from 2.936 ± 0.003 eV to 2.846 ± 0.002 eV, indicating improved photocatalytic potential under visible light. PL analysis revealed a decrease in the direct recombination of EHPs, as indicated by the reduced emission intensity at 356 nm (3.48 eV) in the Fe-TiO2 coatings, along with the appearance of radiative emission at 361 nm, suggesting enhanced charge separation prior to recombination. Overall, the results highlight SPPS as a viable technique for depositing Fe-TiO2 coatings with enhanced physico-chemical and optical properties, making them promising candidates for photocatalytic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/coatings15080870/s1, Figure S1: Cross-section of 50TiO2 coating; Figure S2: Cross-section of 60TiO2 coating; Figure S3. Cross-section of 50Fe-TiO2 coating; Figure S4: Cross-section of 60Fe-TiO2 coating; Figure S5. Internal structure of 50Fe-TiO2 coating; Figure S6. Internal structure of 60Fe-TiO2 coating.

Author Contributions

Conceptualization, K.S. and R.C.J.; methodology, K.S., P.S. and R.C.J.; validation, R.U., A.G., M.M., P.S. and R.C.J.; formal analysis, K.S.; investigation, K.S., P.S., and R.C.J.; resources, A.G., M.M., P.S.; writing—original draft preparation, K.S. and R.C.J.; writing—review and editing, K.S., R.U. and R.C.J.; visualization, K.S.; supervision, P.S. and R.C.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Acknowledgments

Key Simfroso acknowledges the Department of Science and Technology–Accelerated Science and Technology Human Resource Development Program (DOST-ASTHRDP) for the scholarship grant. The DOST–Philippine Council or Industry, Energy and Emerging Technology Research and Development (PCIEERD) is also acknowledged for supporting this work. Special thanks are extended to Adam Sajbura, Tomasz Kiełczawa and Martyna Adach for their invaluable assistance with the spraying and sample preparation, and to Tomas Tesar for the viscosity measurements and for his insightful comments and suggestions on the paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tang, J.; Durrant, J.R.; Klug, D.R. Mechanism of Photocatalytic Water Splitting in TiO2. Reaction of Water with Photoholes, Importance of Charge Carrier Dynamics, and Evidence for Four-Hole Chemistry. J. Am. Chem. Soc. 2008, 130, 13885–13891. [Google Scholar] [CrossRef] [PubMed]
  2. Eidsvåg, H.; Bentouba, S.; Vajeeston, P.; Yohi, S.; Velauthapillai, D. TiO2 as a Photocatalyst for Water Splitting—An Experimental and Theoretical Review. Molecules 2021, 26, 1687. [Google Scholar] [CrossRef] [PubMed]
  3. Moridon, S.N.F.; Arifin, K.; Yunus, R.M.; Minggu, L.J.; Kassim, M.B. Photocatalytic water splitting performance of TiO2 sensitized by metal chalcogenides: A review. Ceram. Int. 2022, 48, 5892–5907. [Google Scholar] [CrossRef]
  4. ul Haq, A.; Saeed, M.; Khan, S.G.; Ibrahim, M. Photocatalytic Applications of Titanium Dioxide (TiO2). In Titanium Dioxide—Advances and Applications; IntechOpen: London, UK, 2021. [Google Scholar]
  5. Gupta, S.; Tripathi, M. A review of TiO2 nanoparticles. Chin. Sci. Bull. 2011, 56, 1639–1657. [Google Scholar] [CrossRef]
  6. Li, R.; Weng, Y.; Zhou, X.; Wang, X.; Mi, Y.; Chong, R.; Han, H.; Li, C. Achieving overall water splitting using titanium dioxide-based photocatalysts of different phases. Energy Environ. Sci. 2015, 8, 2377–2382. [Google Scholar] [CrossRef]
  7. Li, Y.; Peng, Y.-K.; Hu, L.; Zheng, J.; Prabhakaran, D.; Wu, S.; Puchtler, T.J.; Li, M.; Wong, K.-Y.; Taylor, R.A.; et al. Photocatalytic water splitting by N-TiO2 on MgO (111) with exceptional quantum efficiencies at elevated temperatures. Nat. Commun. 2019, 10, 4421. [Google Scholar] [CrossRef]
  8. Rusinque, B.; Escobedo, S.; de Lasa, H. Hydrogen Production via Pd-TiO2 Photocatalytic Water Splitting under Near-UV and Visible Light: Analysis of the Reaction Mechanism. Catalysts 2021, 11, 405. [Google Scholar] [CrossRef]
  9. Fu, Y.-S.; Li, J.; Li, J. Metal/Semiconductor Nanocomposites for Photocatalysis: Fundamentals, Structures, Applications and Properties. Nanomaterials 2019, 9, 359. [Google Scholar] [CrossRef]
  10. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  11. Sulaiman, S.N.A.; Noh, M.Z.; Adnan, N.N.; Bidin, N.; Razak, S.N.A. Effects of photocatalytic activity of metal and non-metal doped TiO2 for Hydrogen production enhancement—A Review. IOP Conf. Ser. J. Phys. Conf. Ser. 2017, 1027, 012006. [Google Scholar]
  12. Lv, C.; Lan, X.; Wang, L.; Yu, Q.; Zhang, M.; Sun, H.; Shi, J. Alkaline-earth-metal-doped TiO2 for enhanced photodegradation and H2 evolution: Insights into the mechanisms. Catal. Sci. Technol. 2019, 9, 6124–6135. [Google Scholar] [CrossRef]
  13. Rafique, M.; Hajra, S.; Irshad, M.; Usman, M.; Imran, M.; Assiri, M.A.; Ashraf, W.M. Hydrogen Production Using TiO2-Based Photocatalysts: A Comprehensive Review. ACS Omega 2023, 8, 25640–25648. [Google Scholar] [CrossRef] [PubMed]
  14. Nair, P.B.; Justinvictor, V.B.; Daniel, G.P.; Joy, K.; Ramakrishnan, V.; Kumar, D.D.; Thomas, P.V. Structural, optical, photoluminescence and photocatalytic investigations on Fe doped TiO2 thin films. Thin Solid Films 2014, 550, 121–127. [Google Scholar] [CrossRef]
  15. Safari, M.; Talebi, R.; Rostami, M.H.; Nikazar, M.; Dadvar, M. Synthesis of iron-doped TiO2 for degradation of reactive Orange16. J. Environ. Health Sci. Eng. 2014, 12, 19. [Google Scholar] [CrossRef] [PubMed]
  16. Zahid, R.; Manzoor, M.; Rafiq, A.; Ikram, M.; Nafees, M.; Butt, A.R.; Hussain, S.G.; Ali, S. Influence of Iron Doping on Structural, Optical and Magnetic Properties of TiO2 Nanoparticles. Electron. Mater. Lett. 2018, 14, 587–593. [Google Scholar] [CrossRef]
  17. Yamashita, H.; Harada, M.; Misaka, J.; Takeuchi, M.; Neppolian, B.; Anpo, M. Photocatalytic degradation of organic compounds diluted in water using visible light-responsive metal ion-implanted TiO2 catalysts: Fe ion-implanted TiO2. Catal. Today 2003, 84, 191–196. [Google Scholar] [CrossRef]
  18. Dholam, R.; Patel, N.; Adami, M.; Miotello, A. Hydrogen production by photocatalytic water-splitting using Cr- or Fe-doped TiO2 composite thin films photocatalyst. Int. J. Hydrogen Energy 2009, 34, 5337–5346. [Google Scholar] [CrossRef]
  19. Moradi, H.; Eshaghi, A.; Hosseini, S.R.; Ghani, K. Fabrication of Fe-doped TiO2 nanoparticles and investigation of photocatalytic decolorization of reactive red 198 under visible light irradiation. Ultrason. Sonochem. 2016, 32, 314–319. [Google Scholar] [CrossRef]
  20. Kim, H.-J.; Jeong, K.-J.; Bae, D.-S. Synthesis and Characterization of Fe Doped TiO2 Nanoparticles by a Sol-Gel and Hydrothermal Process. Korean J. Mater. Res. 2012, 22, 249–252. [Google Scholar]
  21. Zhang, X.; Zhou, M.; Lei, L. Co-deposition of photocatalytic Fe doped TiO2 coatings by MOCVD. Catal. Commun. 2006, 7, 427–431. [Google Scholar] [CrossRef]
  22. Othman, S.H.; Rashid, S.A.; Ghazi, T.I.M.; Abdullah, N. Fe-Doped TiO2 Nanoparticles Produced via MOCVD: Synthesis, Characterization, and Photocatalytic Activity. J. Nanomater. 2011, 2011, 571601. [Google Scholar] [CrossRef]
  23. Kim, Y.; Yang, S.; Jeon, E.H.; Baik, J.; Kim, N.; Kim, H.S.; Lee, H. Enhancement of Photo-Oxidation Activities Depending on Structural Distortion of Fe-Doped TiO2 Nanoparticles. Nanoscale Res. Lett. 2016, 11, 41. [Google Scholar] [CrossRef]
  24. El Mragui, A.; Logvina, Y.; da Silva, L.P.; Zegaoui, O.; da Silva, J.C.G.E. Synthesis of Fe- and Co-Doped TiO2 with Improved Photocatalytic Activity Under Visible Irradiation Toward Carbamazepine Degradation. Materials 2019, 12, 3874. [Google Scholar] [CrossRef]
  25. Wahyuni, E.T.; Lestari, N.D.; Cinjana, I.R.; Annur, S.; Natsir, T.A.; Mudasir, M. Doping TiO2 with Fe from iron rusty waste for enhancing its activity under visible light in the Congo red dye photodegradation. J. Eng. Appl. Sci. 2023, 70, 9. [Google Scholar] [CrossRef]
  26. Modan, E.M.; Plaiasu, A.G. Advantages and Disadvantages of Chemical Methods in the Elaboration of Nanomaterials. Ann. Dun. Jos. Univ. Galati, Fasc. IX Metall. Mater. Sci. 2020, 43, 1. [Google Scholar] [CrossRef]
  27. Mondal, K.; Malode, S.J.; Shetti, N.P.; Alqarni, S.A.; Pandiaraj, S.; Alodhayb, A. Porous nanostructures for hydrogen generation and storage. J. Energy Storage 2024, 76, 109719. [Google Scholar] [CrossRef]
  28. Ismael, M. Enhanced photocatalytic hydrogen production and degradation of organic pollutants from Fe (III) doped TiO2 nanoparticles. J. Environ. Chem. Eng. 2020, 8, 103676. [Google Scholar] [CrossRef]
  29. Skocaj, M.; Filipic, M.; Petkovic, J.; Novak, S. Titanium dioxide in our everyday life; is it safe? Radiol. Oncol. 2011, 45, 227–247. [Google Scholar] [CrossRef]
  30. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  31. Yang, L.; Wang, J.; Lv, H.; Ji, X.-M.; Liu, J.-M.; Wang, S. Hollow-Structured Microporous Organic Networks Adsorbents Enabled Specific and Sensitive Identification and Determination of Aflatoxins. Toxins 2022, 14, 137. [Google Scholar] [CrossRef]
  32. Mendez, M.S.; Lemarchand, A.; Traore, M.; Perruchot, C.; Sassoye, C.; Selmane, M.; Nikravech, M.; Amar, M.B.; Kanaev, A. Photocatalytic Activity of Nanocoatings Based on Mixed Oxide V-TiO2 Nanoparticles with Controlled Composition and Size. Catalysts 2021, 11, 1457. [Google Scholar]
  33. Chien, L.K. Solution Precursor Plasma Spray Deposition of Porous Tin Oxide Coatings for Gas Sensors; Library and Archives Canada = Bibliothèque et Archives Canada: Ottawa, ON, Canada, 2006. [Google Scholar]
  34. Mittal, G.; Paul, S. Suspension and Solution Precursor Plasma and HVOF Spray: A Review. J. Therm. Spray Technol. 2022, 31, 1443–1475. [Google Scholar] [CrossRef]
  35. Du, L.; Coyle, T.W.; Chien, K.; Pershin, L.; Li, T.; Golozar, M. Titanium Dioxide Coating Prepared by Use of a Suspension-Solution Plasma-Spray Process. J. Therm. Spray Technol. 2015, 24, 915–924. [Google Scholar] [CrossRef]
  36. Aruna, S.T.; Vismaya, A.; Balaji, N. Photocatalytic behavior of titania coatings fabricated by suspension and solution precursor plasma spray processes. Mater. Manuf. Process. 2021, 36, 868–875. [Google Scholar] [CrossRef]
  37. Chen, D.; Jordan, E.H.; Gell, M.; Ma, X. Dense TiO2 Coating Using the Solution Precursor Plasma Spray Process. J. Am. Ceram. Soc. 2008, 91, 865–872. [Google Scholar] [CrossRef]
  38. Chen, D.; Jordan, E.H.; Gell, M. Porous TiO2 coating using the solution precursor plasma spray process. Surf. Coat. Technol. 2008, 202, 6113–6119. [Google Scholar] [CrossRef]
  39. Adán, C.; Marugán, J.; van Grieken, R.; Chien, K.; Pershin, L.; Coyle, T.; Mostaghimi, J. Effect of Liquid Feed-Stock Composition on the Morphology of Titanium Dioxide Films Deposited by Thermal Plasma Spray. J. Nanosci. Nanotechnol. 2015, 15, 6651–6662. [Google Scholar] [CrossRef] [PubMed]
  40. Li, D.-Y.; Zhang, X.; Li, J.-W.; Zhang, Y.-J.; Zhang, Y.; Zhang, C. Effects of Si Doping on Structure and Photocatalytic Performance of TiO2 Coatings Deposited by Solution Plasma Spraying. Surf. Technol. 2018, 5, 220–226. [Google Scholar]
  41. Mittal, G.; Bhamji, I.; Fanicchia, F.; Paul, S. Development of Solution Precursor Plasma Spray TiO2/Al2O3 Composite Coatings for Heat Exchanger Application. In Proceedings of the Corrosion 2021, Virtual, 19–30 April 2021. NACE-2021-16304. [Google Scholar]
  42. Kumar, R.; Govindarajan, S.; Janardhana, R.K.S.K.; Rao, T.N.; Joshi, S.V.; Anandan, S. Facile One-Step Route for the Development of In-Situ Co-Catalyst Modified Ti3+ Self-Doped TiO2 for Improved Visible-Light Photocatalytic Activity. ACS Appl. Mater. Interfaces 2016, 8, 27642–27653. [Google Scholar] [CrossRef]
  43. Unabia, R.; Candidato, R.T., Jr.; Pawlowski, L. Current Progress in Solution Precursor Plasma Spraying of Cermets: A Review. Metals 2018, 8, 420. [Google Scholar] [CrossRef]
  44. Simfroso, K.T.; Cabo, S.R.S.; Unabia, R.B.; Britos, A.; Sokołowski, P.; Candidato, R.T., Jr. Solution Precursor Plasma Spraying of TiO2 Coatings Using a Catalyst-Free Precursor. Materials 2023, 16, 1515. [Google Scholar] [CrossRef] [PubMed]
  45. Kozerski, S.; Pawłowski, L.; Jaworski, R.; Roudet, F.; Petit, F. Two zones microstructure of suspension plasma sprayed hydroxyapatite coatings. Surf. Coat. Technol. 2010, 204, 1380–1387. [Google Scholar] [CrossRef]
  46. Yang, D. An Efficient Photocatalyst Structure: TiO2(B) Nanofibers with a Shell of Anatase Nanocrystals. J. Am. Chem. Soc. 2009, 131, 17885–17893. [Google Scholar] [CrossRef] [PubMed]
  47. Ferrara, M.C.; Montecchi, M.; Mittiga, A.; Schioppa, M.; Mazzarelli, S.; Tapfer, L.; Lovergine, N.; Prete, P. Synthesis and annealing effects on microstructure and optical properties of wide-bandgap polycrystalline ferro-pseudobrookite FeTi2O5 sol-gel layers. Ceram. Int. 2025, 51, 9669–9676. [Google Scholar] [CrossRef]
  48. Cullity, B.D. Elements of X-ray Diffraction; Addison-Wesley Metallurgy Series; Addison-Wesley Publishing: Boston, MA, USA, 1956. [Google Scholar]
  49. Challagulla, S.; Tarafder, K.; Ganesan, R.; Roy, S. Structure sensitive photocatalytic reduction of nitroarenes over TiO2. Sci. Rep. 2017, 7, 8783. [Google Scholar] [CrossRef]
  50. Wu, Z.; Zhang, C.; Liu, J.; Wen, F.; Cao, H.; Pei, Y. The Investigation of Microstructure, Photocatalysis and Corrosion Resistance of C-Doped Ti–O Films Fabricated by Reactive Magnetron Sputtering Deposition with CO2 Gas. Coatings 2021, 11, 881. [Google Scholar] [CrossRef]
  51. Loan, T.T.; Huong, V.H.; Huyen, N.T.; Quyet, L.V.; Bang, N.A.; Long, N.N. Anatase to rutile phase transformation of iron-doped titanium dioxide nanoparticles: The role of iron content. Opt. Mater. 2021, 111, 110651. [Google Scholar] [CrossRef]
  52. Khan, M.I.; Bhatti, K.A.; Qindeel, R.; Althobaiti, H.S.; Alonizan, N. Structural, electrical and optical properties of multilayer TiO2 thin films deposited by sol–gel spin coating. Results Phys. 2017, 7, 1437–1439. [Google Scholar] [CrossRef]
  53. Chetibi, L.; Busko, T.; Kulish, N.P.; Hamana, D.; Chaeib, S.; Achour, S. Photoluminescence properties of TiO2 nanofibers. J. Nanopart. Res. 2017, 19, 129. [Google Scholar] [CrossRef]
  54. Horti, N.C.; Kamatagi, M.D.; Patil, N.R.; Nataraj, S.K.; Sannaikar, M.S.; Inamdar, S.R. Synthesis and photoluminescence properties of titanium oxide (TiO2) nanoparticles: Effect of calcination temperature. Optik 2019, 194, 163070. [Google Scholar] [CrossRef]
  55. Shyniya, C.R.; Amarsingh Bhabu, K.; Rajasekaran, T.R. Enhanced electrochemical behavior of novel acceptor doped titanium dioxide catalysts for photocatalytic applications. J. Mater. Sci. Mater. Electron. 2017, 28, 6959–6970. [Google Scholar] [CrossRef]
  56. Saha, A.; Moya, A.; Kahnt, A.; Iglesias, D.; Marchesan, S.; Wannemacher, R.; Prato, M.; Vilatela, J.J.; Guldi, D.M. Interfacial Charge Transfer in Functionalized Multi-walled Carbon Nanotube@TiO2 nanofibres. Nanoscale 2017, 9, 7911–7921. [Google Scholar] [CrossRef]
  57. Liboon, A., Jr.; Cabo, S.R.; Unabia, R.B.; Bagsican, F.R.G.; Candidato, R.T., Jr. Phase Composition, Microstructure, and Optical Characteristics of Spin-Coated La-TiO2 and Fe-TiO2. Phys. Status Solidi A 2023, 220, 2200756. [Google Scholar] [CrossRef]
  58. Dev, P.R.; David, T.M.; Justin, S.J.M.; Wilson, P.; Palaniappan, A. A plausible impact on the role of pulses in anodized TiO2 nanotube arrays enhancing Ti3+ defects. J. Nanopart. Res. 2020, 22, 56. [Google Scholar] [CrossRef]
  59. Makuła, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [PubMed]
  60. Suram, S.K.; Newhouse, P.F.; Gregoire, J.M. High Throughput Light Absorber Discovery, Part 1: An Algorithm for Automated Tauc Analysis. ACS Comb. Sci. 2016, 18, 673–681. [Google Scholar] [CrossRef] [PubMed]
  61. Gesesse, G.D.; Gomis-Berenguer, A.; Barthe, M.-F.; Ania, C.O. On the analysis of diffuse reflectance measurements to estimate the optical properties of amorphous porous carbons and semiconductor/carbon catalysts. J. Photochem. Photobiol. A Chem. 2010, 398, 112622. [Google Scholar] [CrossRef]
  62. Khurram, R.; Wang, Z.; Ehsan, M.F.; Peng, S.; Shafiq, M.; Khan, B. Synthesis and characterization of an α-Fe2O3/ZnTe heterostructure for photocatalytic degradation of Congo red, methyl orange and methylene blue. RSC Adv. 2020, 10, 44997–45007. [Google Scholar] [CrossRef]
  63. Simfroso, K.T.; Liboon, A.Q., Jr.; Candidato, R.T., Jr. A Numerical Investigation of the Thermal Transport Properties of Argon + Hydrogen Plasma Working Gases in the Presence of Various TiO2 Precursor Solutions. Phys. Chem. Res. 2023, 11, 897–912. [Google Scholar]
  64. Zhang, Y.; Zhang, Z.; Yang, J.; Yue, Y.; Zhang, H. Evaporation characteristics of viscous droplets on stainless steel superhydrophobic surface. Int. J. Therm. Sci. 2023, 18, 107843. [Google Scholar] [CrossRef]
  65. Rosa, D.; Abbasova, N.; Di Palma, L. Titanium Dioxide Nanoparticles Doped with Iron for Water Treatment via Photocatalysis: A Review. Nanomaterials 2024, 14, 293. [Google Scholar] [CrossRef] [PubMed]
  66. Eddy, D.R.; Permana, M.D.; Sakti, L.K.; Sheha, G.A.N.; Solihudin; Hidayat, S.; Takei, T.; Kumada, N.; Rahayu, I. Heterophase Polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for Efficient Photocatalyst: Fabrication and Activity. Nanomaterials 2023, 13, 704. [Google Scholar] [CrossRef] [PubMed]
  67. Baszczuk, A.; Jasiorski, M.; Winnicki, M. Low-Temperature Transformation of Amorphous Sol–Gel TiO2 Powder to Anatase During Cold Spray Deposition. J. Therm. Spray Technol. 2018, 27, 1551–1562. [Google Scholar] [CrossRef]
  68. Othman, S.H.; Abdul Rashid, S.; Mohd Ghazi, T.I.; Abdullah, N. Effect of Fe doping on Phase Transition of TiO2 Nanoparticles Synthesized by MOVCD. J. Appl. Sci. 2010, 10, 1044–1051. [Google Scholar] [CrossRef]
  69. Bojinova, A.; Kralchevska, R.; Poulios, I.; Dushkin, C. Anatase/rutile TiO2 composites: Influence of the mixing ratio on the photocatalytic degradation of Malachite Green and Orange II in slurry. Mater. Chem. Phys. 2007, 106, 187–192. [Google Scholar] [CrossRef]
  70. Tian, B.; Li, C.; Zhang, J. One-step preparation, characterization and visible-light photocatalytic activity of Cr-doped TiO2 with anatase and rutile bicrystalline phases. Chem. Eng. J. 2012, 191, 402–409. [Google Scholar] [CrossRef]
  71. He, J.; Du, Y.; Bai, Y.; An, J.; Cai, X.; Chen, Y.; Wang, P.; Yang, X.; Feng, Q. Facile Formation of Anatase/Rutile TiO2 Nanocomposites with Enhanced Photocatalytic Activity. Molecules 2019, 24, 2996. [Google Scholar] [CrossRef]
  72. Fan, X.; Fan, J.; Hu, X.; Liu, E.; Kang, L.; Tang, C.; Ma, Y.; Wu, H.; Li, Y. Preparation and characterization of Ag deposited and Fe doped TiO2 nanotube arrays for photocatalytic hydrogen production by water splitting. Ceram. Int. 2014, 40, 15907–15917. [Google Scholar] [CrossRef]
  73. Kusumawardani, L.J.; Syahputri, Y. Study of structural and optical properties of Fe(III)-doped TiO2 prepared by sol-gel method. IOP Conf. Ser. Earth Environ. Sci. 2019, 299, 012066. [Google Scholar] [CrossRef]
  74. Dash, P.; Thirumurugan, S.; Tseng, C.-L.; Lin, Y.-C.; Chen, S.-L.; Dhawan, U.; Chung, R.-J. Synthesis of Methotrexate-Loaded Dumbbell-Shaped Titanium Dioxide/Gold Nanorods Coated with Mesoporous Silica and Decorated with Upconversion Nanoparticles for Near-Infrared-Driven Trimodal Cancer Treatment. ACS Appl. Mater. Interfaces 2023, 15, 33335–33347. [Google Scholar] [CrossRef]
  75. Král, K.; Menšík, M. Photoluminescence of Nanostructures with Indirect Band Gap. In Proceedings of the International Conference on Transparent Optical Networks ICTON 2014, IEEE Xplore, Graz, Austria, 6–10 July 2014. [Google Scholar]
  76. Pallotti, D.K.; Passoni, L.; Maddalena, P.; Di Fonzo, F.; Letteiri, S. Photoluminescence Mechanisms in Anatase and Rutile TiO2. J. Phys. Chem. C 2017, 121, 9011–9021. [Google Scholar] [CrossRef]
  77. Tizei, L.H.G.; Kociak, M. Chapter Four—Quantum Nanooptics in the Electron Microscope. Adv. Imaging Electron Phys. 2017, 199, 185–235. [Google Scholar]
  78. Nasralla, N.H.S.; Yeganeh, M.; Šiller, L. Photoluminescence study of anatase and rutile structures of Fe-doped TiO2 nanoparticles at different dopant concentrations. Appl. Phys. A 2020, 126, 192. [Google Scholar] [CrossRef]
  79. Miriyala, N.; Prashanthi, K.; Thundat, T. Oxygen vacancy dominant strong visible photoluminescence from BiFeO3 nanotubes. Phys. Status Solidi RRL 2013, 7, 668–671. [Google Scholar] [CrossRef]
  80. Nguyen, V.N.; Nguyen, N.K.T.; Nguyen, P.H. Hydrothermal synthesis of Fe-doped TiO2 nanostructure photocatalyst. Adv. Nat. Sci. Nanosci. Nanotechnol. 2011, 2, 035014. [Google Scholar] [CrossRef]
  81. Sood, S.; Umar, A.; Mehta, S.K.; Kansal, S.K. Highly effective Fe-doped TiO2 nanoparticles photocatalysts for visible-light driven photocatalytic degradation of toxic organic compounds. J. Colloid Interface Sci. 2015, 450, 213–223. [Google Scholar] [CrossRef]
  82. Fu, Q.; Wang, S.; Zhou, B.; Xia, W.; Liu, X.; Han, X.; Duan, Z.; Liu, T.; Sun, X.; Yuan, X.; et al. Defect-Mediated Efficient and Tunable Emission in van der Waals Integrated Light Sources at Room Temperature. Adv. Funct. Mater. 2024, 35, 2414062. [Google Scholar] [CrossRef]
  83. Gouma, P.I.; Mills, M.J. Anatase-to-rutile transformation in titania powders. J. Am. Ceram. Soc. 2001, 84, 619–622. [Google Scholar] [CrossRef]
  84. Gao, Q.; Wu, X.; Fan, Y. The effect of iron ions on the anatase–rutile phase transformation of titania (TiO2) in mica–titania pigments. Dyes. Pigm. 2012, 95, 96–101. [Google Scholar] [CrossRef]
  85. Banerjee, A.N. The design, fabrication, and photocatalytic utility of nanostructured semiconductors: Focus on TiO2-based nanostructures. Nanotechnol. Sci. Appl. 2011, 4, 35–65. [Google Scholar] [CrossRef]
  86. Smijs, T.G.; Pavel, S. Titanium dioxide and zinc oxide nanoparticles in sunscreens: Focus on their safety and effectiveness. Nanotechnol. Sci. Appl. 2011, 4, 95–112. [Google Scholar] [CrossRef]
Figure 1. (a) Cross-sectional images of the TiO2 coating sprayed at 40 mm, scale bar = 10 μm; and (b) its magnified internal structure with a scale bar of 500 nm. The arrows indicate the TiO2 agglomerates, which represent one of the two identified zones. The boxed area is the region selected for the elemental analysis.
Figure 1. (a) Cross-sectional images of the TiO2 coating sprayed at 40 mm, scale bar = 10 μm; and (b) its magnified internal structure with a scale bar of 500 nm. The arrows indicate the TiO2 agglomerates, which represent one of the two identified zones. The boxed area is the region selected for the elemental analysis.
Coatings 15 00870 g001
Figure 2. (a) Cross-sectional image and (b) magnified internal structure of the Fe-TiO2 coating sprayed at 40 mm. The scale bars for (a,b) are 10 μm and 500 nm, respectively. The arrow points to TiO2 agglomerates, representing one of the two identified zones.
Figure 2. (a) Cross-sectional image and (b) magnified internal structure of the Fe-TiO2 coating sprayed at 40 mm. The scale bars for (a,b) are 10 μm and 500 nm, respectively. The arrow points to TiO2 agglomerates, representing one of the two identified zones.
Coatings 15 00870 g002
Figure 3. EDS analysis and elemental mapping images of the 60TiO2 coating sample: (a) EDS area; (b) gold; (c) titanium, (d) oxygen, (e) carbon, (f) iron, and (g) combined (be) elements.
Figure 3. EDS analysis and elemental mapping images of the 60TiO2 coating sample: (a) EDS area; (b) gold; (c) titanium, (d) oxygen, (e) carbon, (f) iron, and (g) combined (be) elements.
Coatings 15 00870 g003
Figure 4. EDS analysis and elemental mapping images of the 60Fe-TiO2 coating sample: (a) EDS area; (b) titanium; (c) iron, (d) carbon, (e) oxygen, (f) nickel, and (g) combined (be) elements. The scale bar represents 10 μm.
Figure 4. EDS analysis and elemental mapping images of the 60Fe-TiO2 coating sample: (a) EDS area; (b) titanium; (c) iron, (d) carbon, (e) oxygen, (f) nickel, and (g) combined (be) elements. The scale bar represents 10 μm.
Coatings 15 00870 g004
Figure 5. SPPS TiO2 and Fe-TiO2 coatings from full-pass sprayed solutions at various stand-off distances, as shown by XRD. The red and blue lines, each with their respective labels, represent the Miller indices of anatase and rutile TiO2, respectively.
Figure 5. SPPS TiO2 and Fe-TiO2 coatings from full-pass sprayed solutions at various stand-off distances, as shown by XRD. The red and blue lines, each with their respective labels, represent the Miller indices of anatase and rutile TiO2, respectively.
Coatings 15 00870 g005
Figure 6. Anatase and rutile phase content of (a) TiO2 and (b) Fe-TiO2 coatings.
Figure 6. Anatase and rutile phase content of (a) TiO2 and (b) Fe-TiO2 coatings.
Coatings 15 00870 g006
Figure 7. Calculated anatase (A) and rutile (R) crystallite sizes for (a) TiO2 and (b) Fe-TiO2 coatings.
Figure 7. Calculated anatase (A) and rutile (R) crystallite sizes for (a) TiO2 and (b) Fe-TiO2 coatings.
Coatings 15 00870 g007
Figure 8. Raman spectra of TiO2 and Fe-TiO2 SPPS coatings.
Figure 8. Raman spectra of TiO2 and Fe-TiO2 SPPS coatings.
Coatings 15 00870 g008
Figure 9. Photoluminescence emission spectra of (ac) TiO2 and (df) Fe-TiO2 coatings excited at 300 nm, including the Gaussian fitting of the emission bands corresponding to the contribution of direct band-to-band recombination (red), trapped electrons from defects and O vacancies (yellow, green, brown, blue, magenta) and trapped holes (cyan).
Figure 9. Photoluminescence emission spectra of (ac) TiO2 and (df) Fe-TiO2 coatings excited at 300 nm, including the Gaussian fitting of the emission bands corresponding to the contribution of direct band-to-band recombination (red), trapped electrons from defects and O vacancies (yellow, green, brown, blue, magenta) and trapped holes (cyan).
Coatings 15 00870 g009
Figure 10. UV-Vis absorption spectra of SPPS TiO2 and Fe-TiO2 coatings.
Figure 10. UV-Vis absorption spectra of SPPS TiO2 and Fe-TiO2 coatings.
Coatings 15 00870 g010
Figure 11. Viscosity–shear rate curves illustrating the flow behavior of TiO2 and Fe-TiO2 precursor solutions.
Figure 11. Viscosity–shear rate curves illustrating the flow behavior of TiO2 and Fe-TiO2 precursor solutions.
Coatings 15 00870 g011
Figure 12. DTA-TGA curves of dried (a) TiO2 and (b) Fe-TiO2 precursor powders at a heating rate of 10 °C/min. Graphs of the first derivation of the DTA results are included. Vertical dotted lines indicate the crystallization and phase transformation temperatures.
Figure 12. DTA-TGA curves of dried (a) TiO2 and (b) Fe-TiO2 precursor powders at a heating rate of 10 °C/min. Graphs of the first derivation of the DTA results are included. Vertical dotted lines indicate the crystallization and phase transformation temperatures.
Coatings 15 00870 g012
Figure 13. XRD patterns of TiO2 and Fe-TiO2 powders heated at 450 °C.
Figure 13. XRD patterns of TiO2 and Fe-TiO2 powders heated at 450 °C.
Coatings 15 00870 g013
Figure 14. Schematic diagram illustrating the physical origin of the PL emissions in the SPPS Fe-TiO2 nanostructure surface.
Figure 14. Schematic diagram illustrating the physical origin of the PL emissions in the SPPS Fe-TiO2 nanostructure surface.
Coatings 15 00870 g014
Table 1. Spray parameters for the solution precursor plasma spraying of TiO2 and Fe-TiO2 coatings.
Table 1. Spray parameters for the solution precursor plasma spraying of TiO2 and Fe-TiO2 coatings.
ParametersValues
Plasma power28 kW
Ar flow rate45 slpm
H2 flow rate5 slpm
Plasma torch velocity500 mm/s
Solution feed rate35 g/min
Stand-off distance40, 50, 60 mm
SubstrateStainless steel
Nozzle diameter0.2 mm
Injection pressure1 bar
Number of spray passes10 cycles
Table 2. Sample labeling of the sprayed TiO2 and Fe-TiO2 coatings.
Table 2. Sample labeling of the sprayed TiO2 and Fe-TiO2 coatings.
Stand-Off DistanceCoating Label
TiO2Fe-TiO2
40 mm40TiO240Fe-TiO2
50 mm50TiO250Fe-TiO2
60 mm60TiO260Fe-TiO2
Table 3. Comparison of the calculated band gap energies from the absorption spectra and Tauc calculations.
Table 3. Comparison of the calculated band gap energies from the absorption spectra and Tauc calculations.
Stand-Off DistanceTiO2Fe-TiO2
E g   =   1240 / λ Tauc Plot E g   =   1240 / λ TAUC PLOT
40 mm3.310 ± 0.326 eV3.328 ± 0.002 eV2.864 ± 0.377 eV2.846 ± 0.002 eV
50 mm3.171 ± 0.434 eV3.168 ± 0.007 eV2.904 ± 0.314 eV2.920 ± 0.003 eV
60 mm3.108 ± 0.414 eV3.139 ± 0.001 eV2.924 ± 0.284 eV2.936 ± 0.003 eV
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

Simfroso, K.; Unabia, R.; Gibas, A.; Mazur, M.; Sokołowski, P.; Candidato, R., Jr. Influence of Fe Ions on the Surface, Microstructural and Optical Properties of Solution Precursor Plasma-Sprayed TiO2 Coatings. Coatings 2025, 15, 870. https://doi.org/10.3390/coatings15080870

AMA Style

Simfroso K, Unabia R, Gibas A, Mazur M, Sokołowski P, Candidato R Jr. Influence of Fe Ions on the Surface, Microstructural and Optical Properties of Solution Precursor Plasma-Sprayed TiO2 Coatings. Coatings. 2025; 15(8):870. https://doi.org/10.3390/coatings15080870

Chicago/Turabian Style

Simfroso, Key, Romnick Unabia, Anna Gibas, Michał Mazur, Paweł Sokołowski, and Rolando Candidato, Jr. 2025. "Influence of Fe Ions on the Surface, Microstructural and Optical Properties of Solution Precursor Plasma-Sprayed TiO2 Coatings" Coatings 15, no. 8: 870. https://doi.org/10.3390/coatings15080870

APA Style

Simfroso, K., Unabia, R., Gibas, A., Mazur, M., Sokołowski, P., & Candidato, R., Jr. (2025). Influence of Fe Ions on the Surface, Microstructural and Optical Properties of Solution Precursor Plasma-Sprayed TiO2 Coatings. Coatings, 15(8), 870. https://doi.org/10.3390/coatings15080870

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