Next Article in Journal
Bibliometric Analysis of Coating Protection from 2015 to 2025
Previous Article in Journal
BBW YOLO: Intelligent Detection Algorithms for Aluminium Profile Material Surface Defects
Previous Article in Special Issue
Etching and Precursor Effects on Plasma-Modified Waste Polyethylene Terephthalate (PET) to Laccase Immobilization Applied in Catechol Biodegradation for Water Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Water Content in Pigment-Grade TiO2 Particles Coated with Al2O3 and SiO2, and Their Effect on Polypropylene Photodegradation

by
Edgar F. Armendáriz-Alonso
1,
Nancy Rivera-García
2,
J. Antonio Moreno-Razo
3,
Luis Octavio Meza-Espinoza
4,
Miguel A. Waldo-Mendoza
1,2,* and
Elías Pérez
5
1
Doctorado Institucional en Ingeniería y Ciencia de Materiales (DICIM), Universidad Autónoma de San Luis Potosí, San Luis Potosí 78210, Mexico
2
Tecnología Sustentable Greennova S.A. de C.V., San Luis Potosí 78395, Mexico
3
Departamento de Física, Universidad Autónoma Metropolitana, Ciudad de Mexico 09310, Mexico
4
Instituto de Física, Ing. Luís Rivera Terrazas, Benemérita Universidad Autónoma de Puebla, Puebla 72570, Mexico
5
Instituto de Física, Universidad Autónoma de San Luis Potosí (IF-UASLP), San Luis Potosí 78000, Mexico
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(6), 685; https://doi.org/10.3390/coatings15060685
Submission received: 26 April 2025 / Revised: 31 May 2025 / Accepted: 2 June 2025 / Published: 6 June 2025

Abstract

:
The influence of structural water in alumina (Al2O3) and silica (SiO2) coated titanium dioxide (TiO2) pigments on the photodegradation behavior of polypropylene (PP) composites was investigated. Four commercial rutile TiO2 pigments with varying surface inorganic coatings were incorporated into PP plaques and subjected to accelerated UV weathering to simulate outdoor exposure. Photodegradation was assessed through gloss retention measurements, the carbonyl index (CI), and stress at break retention, while pigment morphology and composition were analyzed using transmission electron microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS). Surface charge and water content were determined through the zeta potential (ζ), Karl Fischer titration, thermogravimetric analysis (TGA), and Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS). The results showed that low-alumina coating alone led to the lowest photodegradation resistance, the highest CI, and the lowest stress at break retention. In contrast, increasing alumina content enhanced photostability, reaching its maximum for combined alumina–silica coatings, which mitigated electron–hole pair migration. PP composites with high alumina–silica-coated TiO2 exhibited higher gloss retention (36%) compared to low-alumina samples (21%). Furthermore, statistical analysis using ANOVA revealed significant differences in coating content and ζ potential among the pigment grades. These findings provide novel insights into oxide-water interactions and the impact of structural water on the photodegradation of polymer composites.

1. Introduction

Titanium dioxide (titania) is one of the most widely used pigments across various industries due to its low toxicity, chemical stability, high refractive index, and minimal light absorption [1,2,3]. These properties make it suitable for applications in painting, papermaking, plastics, cosmetics, health care, and pharmaceuticals [4,5,6]. However, its use in polymeric systems can be limited by its high photocatalytic activity [7,8]. Under UV irradiation, TiO2 exhibits significant photocatalytic behavior, leading to the generation of electron–hole pairs. These charge carriers can react with adsorbed molecules, forming hydroxide ions (OH), which are highly reactive radicals capable of degrading organic material [9].
The mechanism of polymer degradation mediated by a photocatalytic process has been extensively studied [10,11]. Research has shown that photogenerated holes migrate to the crystal facets of TiO2 and react with adsorbed water molecules, leading to the formation of reactive oxygen species (ROS). These ROS have a strong affinity for polymer chains and promote chain scission, resulting in the generation of carboxylic groups at the ends of polymer molecules. Additionally, humidity can more easily permeate the amorphous regions of the polymer, altering its original properties [12]. Consequently, the presence of water near the titanium dioxide surface is a primary factor in the breakdown of polymers through photocatalysis [13,14,15]. Despite this understanding, few studies have examined the combined effect of adsorbed and structural water in the photodegradation of polymers, resins, or coatings, leaving the role of structural water unclear [16,17,18].
The modification of TiO2 particle surfaces to reduce the photodegradation of polymeric materials is an important issue, referred to as passivation. To achieve this, organic molecules and inorganic oxides have been used to decrease the photoactivity of TiO2 particles. Among these, alumina and silica are the most used and are typically incorporated during TiO2 pigment fabrication [19,20]. The selection and proportion of inorganic surface coatings applied to TiO2 pigments are determined by the end-use performance requirements, particularly concerning photostability and environmental exposure. Commercial TiO2 grades are generally classified into three durability categories: non-durable, semi-durable, and durable, based on their resistance to photocatalytic activity under environmental exposure conditions [21,22].
Non-durable TiO2 pigments are characterized by minimal surface passivation, typically containing less than 2 wt.% of alumina. These grades are primarily used in indoor applications, such as short-term industrial packaging films for food, where exposure to UV light is negligible. Semi-durable TiO2 pigments incorporate either additional silica into the low-alumina formulation or may contain only alumina in proportions exceeding 2 wt.%. These pigments are commonly used in interior architectural paints and temporary exterior coatings, but they have limited UV exposure resistance. In contrast, durable TiO2 pigments are designed for long-term performance in outdoor environments. These grades typically have high combined alumina and silica content, with surface treatments reaching approximately 11 wt.%. This formulation suppresses photocatalytic activity, enhancing pigment stability and minimizing polymer degradation. This grade is extensively used in polyolefin-based systems for automotive components, building materials, agricultural films, outdoor furniture, and automotive coatings, where sustained UV resistance is essential [23].
This study aimed to investigate the role of structural water in the photodegradation behavior of polypropylene plaques containing alumina- and silica-coated TiO2. To achieve this, PP composites were fabricated and exposed to UV irradiation in a controlled chamber, simulating outdoor weathering. The results indicated a direct correlation between PP gloss retention and TiO2 coating content. It has been demonstrated that these coatings reduce the reactivity of TiO2, hindering the photo-oxidation reaction. Since alumina and silica are insulators, it is suggested that any coating blocks the interaction of photogenerated charge carriers with reactant molecules. However, alumina and silica can also promote water adsorption through hydroxylation, which in turn could enhance ROS generation and polymer photodegradation. These opposing mechanisms (i.e., water adsorption and stabilization) compete with each other. This study found a direct correlation between PP gloss retention and the water content of TiO2 pigments. This was explained by identifying two types of water: adsorbed water and structural water. The former can be incorporated into the crystal lattice through chemical bonds, and the latter is physically attached to the surface of particles by weak interaction forces. Among the coatings studied, the low-alumina formulation exhibited the lowest structural water content, resulting in the weakest photodegradation resistance. In contrast, increasing the coating content effectively enhanced water retention, acting as a protective barrier against moisture loss while limiting the migration of electron–hole pairs or ROS. These observations suggest a synergistic mechanism in which alumina enhances water adsorption, while silica impedes electron–hole pair migration and the subsequent ROS formation. Overall, this study provides essential insights into the role of alumina and silica coatings on TiO2 pigments, concerning water adsorption dynamics and their impact on the photodegradation process in polymeric systems.

2. Materials and Methods

Four different grades of commercial rutile TiO2 samples were used, designated based on their formulated Al2O3-SiO2 coating content as follows: low alumina (LA, 1.7 wt.%), low alumina with added silica (LA 1.7 wt.% + S 3 wt.%), high alumina (HA, 3.2 wt.%), and high alumina with added silica (HA 3.2 wt.% + S 3.5 wt.%). Commercial polypropylene (PP) with a melt flow index (MFI) of 4 g/10 min, under a load of 2.16 kg at 190 °C, was used for plaque preparation to evaluate photodegradation. The PP was supplied by Greennova (San Luis Potosí, Mexico), and the TiO2 samples were obtained from Chemours (Altamira, Mexico).

2.1. Polypropylene/TiO2 Plaques Fabrication, Degradation, and Mechanical Tests

Polypropylene (PP) was compounded with TiO2 pigment at 10 wt.% load using hot melt extrusion in a co-rotating twin-screw extruder ZSK-30 (Werner & Pfleiderer, Stuttgart, Germany). The barrel and die section temperatures were maintained at 210 °C and 220 °C, respectively, with an extruder speed of 300 rpm and a length-to-diameter ratio of 24:1. Using this compounded material, PP plaques were fabricated in an injection molding machine (Battenfeld LCMI, Kottingbrunn, Austria) where the barrel temperature was set at 220 °C, the mold temperature 27 °C, and the injection pressure 8.27 MPa, with a shot volume of 43 cm3. The plaques were produced in dimensions of 60 × 60 × 2 mm, as depicted in Figure 1.
Following fabrication, the plaques were exposed to UV radiation in an aging chamber (Q-Panel model with a UVA lamp at 60 °C). The exposure cycle included condensation at 50 °C with deionized water spray and a radiation intensity of 0.8 W/m2/nm at 340 nm. The plaques were subjected to UV exposure for 0, 300, and 500 h. Durability was assessed by evaluating changes in optical properties, carbonyl index, and stress at break retention.
Gloss retention was measured using a TRI-gloss 4430 gloss meter (Byk Gardner, Geretsried, Germany) following ASTM D2457 standards. The carbonyl index was determined using Fourier transform infrared spectroscopy with attenuated total reflectance (FTIR-ATR) on a Cary 600 spectrometer (Agilent, Santa Clara, CA, USA) within a spectral range of 1200 to 2000 cm⁻1 and 32 scans per sample. Mechanical properties were evaluated by measuring stress at break retention at different UV weathering times, following ASTM D638 standards, using a universal testing machine (Instron, Norwood, MA, USA) equipped with load cells of 0.1, 5, and 30 kN.

2.2. Characterization of TiO2 Pigment

The crystalline phase of pigments was determined using X-ray powder diffraction (XRD) with an AXS D8 Advance diffractometer (Bruker, Karlsruhe, Germany) equipped with a Cu tube emitting Kα radiation at 1.5405 Å. The scanning range was set from 20° to 90° (2θ) with a step size of 0.02°/s.
Pigment particle size and chemical composition were analyzed using transmission electron microscopy (TEM) and energy-dispersive spectroscopy (EDS) in a JEM–1230 (JEOL, Tokyo, Japan). The zeta potential (ζ) was measured using an analyzer Zen3600 (Malvern Panalytical, Malvern, UK). The sample moisture content was determined using the Karl Fischer titration method with a DL39 titrator (Mettler Toledo, Greifensee, Switzerland).
Thermogravimetric analysis (TGA) was performed using a TGA Q600 (TA Instruments, New Castle, DE, USA) with thermograms recorded at a heating rate of 10 °C/min over a temperature range of 25–450 °C. DRIFTS spectra were obtained using a Nicolet 6700 infrared spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). Samples were placed in a stainless-steel sample cup (10 mm diameter × 3 mm deep), leveled at the top, and positioned in a Praying Mantis (Harrick Scientific, Pleasantville, NJ, USA). The experimental conditions were set from room temperature to 300 °C, with a heating rate of 10 °C/min.

2.3. ANOVA Statistical Analysis and Mathematical Modeling

An analysis of variance (ANOVA) was conducted to identify the sources of variance in one or more factors affecting TiO2 particles. This study aimed to determine whether differences arise from variability between pigment grades or within different samples of a single TiO2 type. ANOVA evaluates the sources of variation linked to independent variables and examines how these variables interact to influence the properties of the four titania samples.
The analysis tests the null hypothesis ( H 0 ), which states that all population means ( μ i ) are equal, against the alternative hypothesis ( H 0 ), which suggests differences among the four TiO2 samples.
H 0 : μ 1 = μ 2 = μ 3 = μ 4
H 1 : a t   l e a s t   o n e   i s   d i f f e r e n t   f r o m   t h e   o t h e r s
The ANOVA table decomposes the variance of the data into two components: the between-group component and the within-group component. The p-value defines the probability that, if the null hypothesis were true, the sampling variation would produce an estimate further from the hypothesized value than the observed estimate. If the p-value is less than 0.05, there is a statistically significant difference between the means of the four variables with a 95.0% confidence level.
To determine which means are significantly different, Fisher’s Least Significant Difference (LSD) test is used. This test is applied in the context of analysis of variance when the p-value suggests rejection of the null hypothesis ( H 0 ), indicating a significant difference between population means. The LSD test helps to identify populations whose means are statistically distinct.
Additionally, the experimental data were fitted to a polynomial mathematical model, correlating various sample properties such as diameter, zeta potential, Al2O3, and SiO2 composition on the TiO2 surface, with an adjustment exceeding a 95.0% confidence level. Before conducting ANOVA, the assumption of homogeneity of variances was evaluated using Levene’s test. A p-value greater than 0.05 indicates that variances are not significantly different, validating the use of ANOVA. The determined p-values from Levene’s test for chemical composition were 0.22 and 0.68 for TiO2 and Al2O3 across the four TiO2 grades, while for SiO2 in the LA+S and HA+S samples, the p-value was 0.84. For particle size across the four grades, the p-value from Levene’s test was 0.51.

3. Results and Discussion

3.1. Effect of Inorganic Coatings on Photodegradation of Polypropylene/TiO2 Plaques

PP plaques extruded with different grades of TiO2 were irradiated in a UV light aging chamber. Figure 2 illustrates the photodegradative effect on each sample in polypropylene plaques. The optical property was assessed by measuring gloss retention at three different exposure times, with this factor evaluated exclusively for pigmented composites to facilitate comparison.
The surface gloss retention graph indicates that the PP/HA+S plaque maintained the highest gloss retention after 500 h of UV exposure, while the PP/LA plaque exhibited the lowest, with values of 36% and 21%, respectively. A clear trend is observed in gloss retention as a function of Al2O3 and SiO2 surface content. The combination of both oxides appears to exert a synergistic stabilizing effect by inhibiting electron–hole pair separation and suppressing hydroxyl radical formation, which are key drivers of polymer degradation.
Additionally, a direct relationship is noted between gloss retention and moisture content, with the following order: PP/LA > PP/LA+S > PP/HA > PP/HA+S. To explain this effect, an ANOVA analysis was conducted to evaluate the TiO2 pigment’s capability of capturing hydroxyl groups as a function of Al2O3 and SiO2 content. This trend is also observed in the evaluation of mechanical properties, specifically stress at break retention measurements, as a function of TiO2 coating content.
Stress at break retention of PP/TiO2 was used as an indicator of polymeric composite degradation due to UV weathering exposure. Figure 3 illustrates that PP without TiO2 exhibited the greatest loss of mechanical properties across all three exposure times. In contrast, incorporating TiO2 into PP enhanced stress at break retention, stabilizing its mechanical properties. The influence of TiO2 coating on stabilization follows this sequence: PP/LA > PP/LA+S > PP/HA > PP/HA+S.
The carbonyl index (CI), measured as an indicator of photodegradation in polymeric materials, was determined by calculating the ratio of the carbonyl band (C=O) at approximately 1700–1740 cm⁻1 to the CH3 stretching mode band at around 1455 cm⁻1 as a reference. This method has been established as a reliable indicator of photodegradation in polyolefins [24,25].
Figure 4 presents the carbonyl index as a function of inorganic coating content and exposure time. The results indicate that the incorporation of TiO2 pigment reduces the photodegradation of polypropylene (PP) to varying degrees, with the extent of stabilization strongly influenced by the surface coating composition. The samples containing LA and LA+S coatings demonstrated enhanced resistance to photodegradation, while the HA-coated pigment provided superior stabilization. The most pronounced protective effect was observed in the pigment with the HA+S surface coating.

3.2. Characterization of TiO2 Pigment and Statistical Analysis

3.2.1. Structural Analysis by X-Ray Diffraction

Figure 5 presents the X-ray diffraction patterns of TiO2 pigments as a function of inorganic coating content. These diffraction patterns were compared with ICDD PDF 21-1276 and PDF 21-1272, which correspond to rutile and anatase phases, respectively. All samples exhibited the primary peaks of the rutile phase, with Miller indices (110), (101), (200), (111), (210), (211), (220), (002), (310), (112), and (301). No additional peaks associated with other phases were detected.
Although commercial rutile TiO2 samples often contain anatase impurities, these impurities may be below the detection limit of this technique. Consequently, no significant differences in crystalline structure were observed among the TiO2 samples, with the surface content of alumina and silica remaining the primary variable factor.

3.2.2. Coating of TiO2 Surface

The composition of pigment particles was determined using EDS coupled with TEM during image acquisition. EDS microanalysis spectra were obtained from representative regions of pigment particles, and the elemental concentrations (wt.%) of titanium (Ti), aluminum (Al), and silicon (Si) were converted to their respective oxide forms (i.e., TiO2, Al2O3, and SiO2) based on their stoichiometric ratios and molar masses. The results, displayed in Table 1, provide a detailed breakdown of the metallic oxide composition of each particle. Figure 6 presents a boxplot illustrating the distribution of three principal oxides identified in the sample particles: TiO2 (a), alumina (Al2O3) (b), and silica (SiO2) (c). According to the composition analysis, the LA sample exhibited the highest TiO2 content, suggesting that it may demonstrate greater reactivity compared to the other samples. The statistical significance of these differences was confirmed through ANOVA analysis, as shown in Table 2.
Since the p-value is less than 0.05, there is a statistically significant difference between the means of the four variables with a 95.0% confidence level. The LA sample exhibits the lowest alumina content at 1.5 ± 0.3 wt.%, while the LA+S sample contains 2.5 ± 0.5 wt.% of Al2O3. The HA and HA+S samples have alumina concentrations of 3.6 ± 0.3 wt.% and 3.8 ± 0.3 wt.%, respectively.
Regarding silica content, the p-value is also less than 0.05, indicating that at least one sample has a different concentration. Statistically, the LA+S and HA+S samples show distinct silica content, with compositions of 3.4 ± 0.3 wt.% and 2.5 ± 0.3 wt.%, respectively. These findings align with the estimated order of magnitude required for silica or alumina to coat the surface [20,26,27].

3.2.3. Particle Size

The morphology of the pigment particles was analyzed using TEM, with Figure 7 displaying micrographs as a function of inorganic coating content. In general, faceted particles were observed regardless of the coating composition. Additionally, the average particle size was determined from TEM images for each pigment grade. The HA+S sample exhibited the largest size, with a diameter of 233 ± 64 nm, while HA had the smallest size at 165 ± 27 nm. The LA and LA+S samples had diameters of 203 ± 40 nm and 199 ± 32 nm, respectively.
Figure 8 presents the LSD test and boxplot of particle size analyses, while Table 3 provides the ANOVA summary. The letters above each boxplot indicate statistically homogeneous groups where treatments sharing a letter are not significantly different. The reported p-value in the table exceeds 0.05, indicating that there is no statistically significant difference in the size parameter [28].

3.2.4. Zeta Potential (ζ)

Zeta potential measurements were conducted on aqueous suspensions of TiO2 at the as-obtained pH of 6.6, achieved by dispersing pigment particles in deionized water. This pH value represents the equilibrium condition of TiO2 in water as a function of inorganic coatings content. Figure 9 presents the LSD test for zeta potential in TiO2 particles, while Table 4 displays the ANOVA test results for these zeta potential values. Since the p-value is less than 0.05, at least one sample exhibited a statistically different zeta potential value. The LSD test assigned different letters (A, B, C, and D) to each TiO2 pigment formulation, indicating statistically significant differences among formulations.
All samples exhibit statistically distinct zeta potential values. The zeta potential measurements for LA, LA+S, HA, and HA+S were −28.7 ± 0.7 mV, −20.2 ± 0.3 mV, −30.5 ± 0.5 mV, and −24.6 ± 1.9 mV, respectively. Negative values are characteristic of metallic oxides due to the presence of retained -OH groups on their surfaces [29,30].
According to the mathematical model from ANOVA, different average values of zeta potential were fitted to the amount of aluminum and silica. The relationship is expressed by the following equation:
Z p = 26.20 1.24 C A l + 3.56 C S i
where C A l and C S i represent the percentage compositions of Al2O3 and SiO2, respectively. A linear model was adjusted with an R-squared value of 0.99 (Figure 10), indicating a strong correlation. These findings suggest that zeta potential is modulated by the composition of inorganic oxides incorporated into TiO2 particles, which in turn influences the effective charge of the particle surface.
The density of surface hydroxyl groups correlates with zeta potential, and these groups play a crucial role in promoting water adsorption. In this case, the zeta potential, as a function of the composition of inorganic coatings and the measured values, confirms the presence of hydroxyl groups on alumina- and silica-coated TiO2 surfaces. Silica modifies these surface hydroxyl groups, leading to a less negative surface charge.

3.2.5. Water Content

Water content plays a crucial role in photodegradation performance, since photocatalytic processes occur in an aqueous environment. Adsorbed water molecules interact with bridging oxygen atoms on the pigment surface through multiple hydrogen bonds, as depicted in Figure 11, enhancing proton transfer and photo-oxidative reactions [16,31]. This interaction facilitates the generation of reactive oxygen species, such as hydroxyl radicals (●OH), which subsequently attack the polymer backbone. These free radicals lead to the formation of carboxyl groups in the polymer chains, reducing molecular size and resulting in degradation.
In alumina- and silica-coated TiO2, two distinct types of water can exist: physically adsorbed water and structural water [32]. Physically adsorbed water is easily removed upon heating, whereas structural water is more strongly bound. However, the role of structural water in photodegradation remains unestablished. Due to the thermal desorption overlap of physically adsorbed and structural water within the 100–300 °C range, the water determination was not separated.
To quantify total water content regardless of its binding state, the Karl Fischer titration method was employed. The results, presented in Table 5, indicate that the lowest water content was observed in the LA sample (0.11 wt.%), followed by LA+S (0.19 wt.%) and HA (0.45 wt.%). The highest value was recorded for HA+S (0.54 wt.%). Based on these findings, the sequence of overall water loss was as follows: LA < LA+S < HA < HA+S.
These results suggest a direct proportional effect between the percentage of coating on the TiO2 surface and water content. Furthermore, gloss retention in polypropylene plaques directly correlates with water content. An analysis of TGA supported these findings, and the results are presented as thermograms in Figure 12. In both thermograms shown in Figure 12, the highest weight loss was observed for the HA+S sample. Here, an important point to consider is the presence of hydrous alumina particles, which lowers van der Waals forces between TiO2 pigment particles, decreasing particle-to-particle attractions by several orders of magnitude. Thus, hydrous aluminum oxide (Al(OH)3) phases are used to improve dispersibility more effectively than other hydroxides and oxides [33]. Accordingly, the observed weight loss at around 300 °C can be attributed to the dehydroxylation of aluminum hydroxide, resulting in the formation of alumina and water [34,35,36]. The initial step in the thermal decomposition of aluminum hydroxide can be assigned to the diffusion of protons and the reaction with hydroxyl ions (OH) to form water [37,38]. This process removes the binding force between subsequent units in the aluminum hydroxide structure. The HA+S sample exhibits the highest weight loss at higher temperatures or prolonged exposure times, consistent with its high alumina and silica content. In contrast, LA shows the lowest weight loss. The TGA results up to 450 °C reflect both physically adsorbed and structural water. Compared with the Karl Fischer values, the HA+S sample contains the most significant structural water content. This finding is further corroborated by DRIFTS analysis.
DRIFTS analysis enabled the complementary evaluation of pigments under thermal conditions [34]. The infrared spectra of the four grades of TiO2 (Figure 13) revealed broad absorption bands between 3700 and 3260 cm−1 (O-H stretching) and 1627–1640 cm−1 (H-O-H bending), attributed to surface hydroxyl and adsorbed water, respectively. These bands were more intense in the LA and HA samples. In silica-containing samples (LA+S and HA+S), absorption bands between 1047 and 1055 cm−1 and between 1230 and 1245 cm−1 were observed, suggesting chemical interaction between hydrous silica and TiO2 surfaces [26,30]. The disruption of Si-O tetrahedral symmetry by Ti atoms leads to IR band splitting or activation of otherwise inactive modes [39]. It is worth noting that the persistence of O-H bands after heating to 300 °C confirms the presence of structural water. Upon heating, band intensity decreased significantly for LA, partially for LA+S, and remained strong in HA+S, suggesting a stabilizing effect of alumina–silica coatings. These results indicate that LA primarily retains physically adsorbed water, while HA and HA+S retain more structural water. In addition, silica enhances water retention, as shown by stronger O-H bands in DRIFTS. Structural water does not participate directly in surface reaction, but hydroxyl coverage governs interfacial interactions, clearly evidenced in the LA sample.
Based on these findings, structural water on the TiO2 surface, particularly that associated with the inorganic coatings of Al2O3 and SiO2, appears to contribute to increased UV protection efficiency in polypropylene matrices. The composition and ratio of coating materials determine the density of hydroxyl groups and water retention behavior. This correlation was validated through DRIFTS, Karl Fischer titration, and zeta potential measurements. ANOVA analyses confirmed statistically significant differences between pigment grades. Despite differences in water content and surface properties, the optical performance of the polymer composites indicates a synergistic effect of alumina and silica, enhancing surface passivation and improving resistance to UV-induced degradation, as measured by gloss retention, stress at break retention, and the carbonyl index.

3.3. Water Stabilization Mechanism by Silica on TiO2 Surfaces

The results of this study provide significant evidence regarding the role of silica in stabilizing adsorbed water on the TiO2 surface. DRIFTS analyses revealed that silica-containing samples (LA+S and HA+S) exhibit characteristic absorption bands at 1047–1055 cm⁻1, indicating a chemical interaction between hydrous silica and TiO2 surfaces [39,40]. This molecular interaction is fundamental to understanding the protection mechanism.
Silica acts as a water stabilizer through several interrelated mechanisms, including hydrogen bond formation, modification of water mobility, physical barrier effects, and passivation of active sites. These mechanisms contribute to enhanced photostability by reducing the likelihood of photocatalytic degradation in polymeric systems.
  • Hydrogen Bond Formation: The tetrahedral structure of silica, with its surface silanol (Si-OH) groups, offers numerous sites for hydrogen bond formation with water molecules [16]. These bonds are stronger than water-TiO2 interactions, leading to more stable adsorption.
  • Modification of Water Mobility: Water adsorbed on the silica surface exhibits reduced mobility compared to water on the alumina surface or pure TiO2 [17,31]. This decreased mobility restricts the ability of water molecules to participate in photocatalytic reactions.
  • Physical Barrier for Reactive Species Migration: The silica layer functions as a physical barrier, preventing the migration of photogenerated electron–hole pairs to the surface [19]. This reduction in migration significantly decreases the likelihood of these pairs interacting with water molecules to form reactive oxygen species (ROS).
  • Passivation of Active Sites: Silica effectively passivates the catalytically active sites on the TiO2 surface, thereby reducing the pigment’s ability to initiate photocatalytic reactions that lead to polymer degradation [8,41].
Zeta potential analysis confirms this mechanism, demonstrating that samples containing silica (LA+S and HA+S) exhibit less negative zeta potential values (−20.2 ± 0.3 mV and −24.6 ± 1.9 mV) compared to samples without silica (LA and HA, with −28.7 ± 0.7 mV and −30.5 ± 0.5 mV, respectively). This shift in zeta potential indicates a significant modification in surface charge, which directly influences the interaction with water molecules [29,30].
The equation derived from the ANOVA model ( Z p = 26.20 1.24 C A l + 3.56 C S i ) quantifies this relationship, revealing that silica’s contribution to the zeta potential is approximately 2.3 times greater than that of alumina. This finding suggests that silica has a more pronounced effect on modifying the surface properties of TiO2, which explains its superior effectiveness in stabilizing adsorbed water [34].

3.4. Practical Implications for Polymer Applications

The findings of this study have significant practical implications for the formulation and application of polypropylene compounds containing TiO2, providing valuable insights into optimizing their stability and performance.
  • Coating optimization for outdoor applications: For polypropylene products designed for outdoor applications, including garden furniture, exterior automotive components, and construction materials, the use of TiO2 pigments with combined alumina and silica coatings (like the HA+S sample) is recommended to enhance gloss retention and durability [10,27].
  • Processing considerations: The presence of silica coatings on TiO2 particles may influence the rheological properties of the molten polymer during processing [7]. To account for these effects, manufacturers should consider adjusting processing parameters, including melting temperature and extrusion speed.
  • Cost–benefit balance: Although TiO2 pigments with alumina and silica coatings may have a higher initial cost, their significant improvement in durability and gloss retention can lead to lower long-term costs due to extended product life [4].
  • Reduction in additional additives: The enhanced UV stability achieved with properly coated TiO2 pigments may enable a reduction in the amount of additional UV stabilizers in the polymer formulation, potentially streamlining the formulation process and lowering costs [9,12].
  • Specific Applications: For applications where gloss retention is essential, such as consumer packaging or decorative pieces, selecting TiO2 pigments with an optimal balance of alumina and silica coatings can significantly enhance product performance [6].

3.5. Comparison with Previous Studies on TiO2 Coatings

Our results on the role of alumina and silica coatings on TiO2 expand the current understanding of UV protection mechanisms in polymeric systems. Previous studies, such as those by Egerton and colleagues, have demonstrated that inorganic coatings reduce the photoactivity of TiO2 [39]. However, these studies focus primarily on coatings acting as physical barriers that prevent charge carrier migration. Our work adds a new dimension to this understanding by highlighting the critical role of structural and adsorbed water in the photodegradation process.
Unlike earlier research that mainly examined the effect of alumina and silica on TiO2 particle dispersion and opacity, our study directly links coating composition to water retention and, ultimately, UV protection performance. Guo et al.’s studies on thin Al2O3 films for suppressing the photocatalytic activity of TiO2 showed similar results in reducing photocatalytic activity but did not address the specific mechanism related to water retention [20]. Our findings complement these studies by providing a mechanistic model.
Wang and colleagues investigated TiO2@SiO2 nanocomposites for printing pigment applications, demonstrating improvements in dispersibility and optical performance [26]. However, they did not explore the effect on UV durability in polymeric matrices. Our study extends those findings to the context of polypropylene photodegradation.
Perhaps most significantly, our results challenge some conventional assumptions in the literature. While previous studies have suggested that alumina alone is sufficient for passivation, our data indicate that a combination of alumina and silica provides a superior synergistic effect for UV stabilization. This observation has important implications for the formulation of TiO2 pigments in polymer applications where long-term durability is critical [32,42].
In summary, our research provides a more comprehensive and nuanced understanding of the role of inorganic coatings in modifying the TiO2 surface for polymer applications, specifically highlighting the importance of water–oxide interactions in the photodegradation mechanism [15].

4. Conclusions

This study establishes a direct correlation between structural water content in TiO2 particles and photodegradation resistance of polypropylene composites, yielding several significant findings with important industrial implications. The investigation of four commercial TiO2 variants with different Al2O3 and SiO2 surface treatments revealed that water–oxide interactions play a crucial and previously underappreciated role in polymer photostability.
Comprehensive characterization through TEM, EDS, zeta potential analysis, and DRIFTS demonstrates that alumina coatings significantly increase water adsorption on TiO2 surfaces. However, this water becomes a liability for photostability unless stabilized by silica. The sample with low alumina content (LA) exhibited the most severe photodegradation despite showing minimal weight loss during thermal analysis, indicating that its surface hydroxyl groups were particularly susceptible to participating in photocatalytic degradation processes.
Statistical analysis using ANOVA confirmed significant differences in surface properties across the four TiO2 variants, with zeta potential measurements providing a quantitative indicator of hydroxyl group density. The mathematical model developed ( Z p = 26.20 1.24 C A l + 3.56 C S i ) demonstrates that silica has approximately 2.3 times greater impact on surface charge properties than alumina, which correlates strongly with improved UV protection performance.
The synergistic mechanism revealed in this study operates as follows: alumina coating provides sites for water adsorption through its hydroxyl-rich surface, while silica stabilizes this adsorbed water through stronger hydrogen bonding and simultaneously forms a physical barrier against electron–hole pair migration. This dual protective action results in significantly enhanced gloss and stress at break retention in PP composites containing alumina- and silica-coated TiO2 (HA+S). This last composite reached maximum properties retention after 500 h of accelerated UV exposure, with 36% gloss retention, the highest stress at break retention, and the lowest carbonyl index.
These findings provide valuable guidance for optimizing TiO2 surface treatments in polymer applications requiring enhanced weather resistance. For exterior applications, where UV stability is critical, TiO2 pigments should ideally incorporate balanced aluminum and silicon oxide coatings to maximize photostability. The insights gained about water–oxide interactions may also inform the development of novel stabilization strategies beyond conventional approaches focused solely on electron–hole pair suppression.
Future research directions should explore the long-term stability of these water–oxide interactions under varied environmental conditions, the potential for further optimizing the Al2O3:SiO2 ratio for specific polymer systems, and the extension of these findings to other photocatalytically active pigments and polymeric matrices.

Author Contributions

Conceptualization, M.A.W.-M. and E.F.A.-A.; methodology, J.A.M.-R.; software, L.O.M.-E.; validation, E.P., N.R.-G. and J.A.M.-R.; formal analysis, L.O.M.-E.; investigation, E.F.A.-A.; resources, E.P.; data curation, E.F.A.-A.; writing—original draft preparation, E.F.A.-A.; writing—review and editing, M.A.W.-M.; visualization, N.R.-G.; supervision, M.A.W.-M.; project administration, M.A.W.-M.; funding acquisition, E.P. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by Greennova.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The dataset is available on request from the authors.

Acknowledgments

The authors thank Greennova for the technical and funding support during this research. Edgar F. Armendáriz-Alonso acknowledges Secretaría de Ciencia, Humanidades, Tecnología e Innovación (SECIHTI) for the Ph.D. fellowship (CVU 733640). The authors acknowledge E. Sambriski for his valuable review of the manuscript and M.M.I.M. Rosa Lina Tovar Tovar from Instituto de Metalurgia, UASLP, for technical support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Arun, J.; Nachiappan, S.; Rangarajan, G.; Alagappan, R.P.; Gopinath, K.P.; Lichtfouse, E. Synthesis and Application of Titanium Dioxide Photocatalysis for Energy, Decontamination and Viral Disinfection: A Review. Environ. Chem. Lett. 2023, 21, 339–362. [Google Scholar] [CrossRef] [PubMed]
  2. Nawaz, H.; Umar, M.; Maryam, R.; Nawaz, I.; Razzaq, H.; Malik, T.; Liu, X. Polymer Nanocomposites Based on TiO2 as a Reinforcing Agent: An Overview. Adv. Eng. Mater. 2022, 24, 2200844. [Google Scholar] [CrossRef]
  3. Samoylov, A.M.; Popov, V.N. Titanium Dioxide (TiO₂) and Its Applications; Elsevier: Amsterdam, The Netherlands, 2021; ISBN 9780128199602. [Google Scholar]
  4. Yang, C.; Liu, C.; Xu, L.; Nie, M.; Wang, Q. Flow-Induced Alignment of Titanium Dioxide Nanorods in Polyolefin Tubes for Use in Catheters. Mater. Des. 2023, 233, 112277. [Google Scholar] [CrossRef]
  5. Wu, X. Applications of Titanium Dioxide Materials. In Titanium Dioxide—Advances and Applications; IntechOpen: London, UK, 2022. [Google Scholar]
  6. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M.H.D.A.; Jiangyong, H. A Review on the Synthesis of the Various Types of Anatase TiO2 Facets and Their Applications for Photocatalysis. Chem. Eng. J. 2020, 384, 123384. [Google Scholar] [CrossRef]
  7. Wu, F.; Misra, M.; Mohanty, A.K. Challenges and New Opportunities on Barrier Performance of Biodegradable Polymers for Sustainable Packaging. Prog. Polym. Sci. 2021, 117, 101395. [Google Scholar] [CrossRef]
  8. Zapata-Tello, D.L.; Escobar-Barrios, V.; Gonzalez-Calderon, J.A.; Pérez, E. Chemical Modification of Titanium Dioxide Nanoparticles with Dicarboxylic Acids to Mediate the UV Degradation in Polyethylene Films. Polym. Bull. 2020, 77, 6409–6431. [Google Scholar] [CrossRef]
  9. Parvate, S.; Singh, J.; Dixit, P.; Vennapusa, J.R.; Maiti, T.K.; Chattopadhyay, S. Titanium Dioxide Nanoparticle-Decorated Polymer Microcapsules Enclosing Phase Change Material for Thermal Energy Storage and Photocatalysis. ACS Appl. Polym. Mater. 2021, 3, 1866–1879. [Google Scholar] [CrossRef]
  10. Andrady, A.L.; Barnes, P.W.; Bornman, J.F.; Gouin, T.; Madronich, S.; White, C.C.; Zepp, R.G.; Jansen, M.A.K. Oxidation and Fragmentation of Plastics in a Changing Environment; from UV-Radiation to Biological Degradation. Sci. Total Environ. 2022, 851, 158022. [Google Scholar] [CrossRef]
  11. Bourgogne, D.; Therias, S.; Gardette, J.-L. Wavelength Effect on Polymer Photooxidation under LED Weathering Conditions. Polym. Degrad. Stab. 2022, 202, 110021. [Google Scholar] [CrossRef]
  12. Alavian Petroody, S.S.; Hashemi, S.H.; Škrlep, L.; Mušič, B.; van Gestel, C.A.M.; Sever Škapin, A. UV Light Causes Structural Changes in Microplastics Exposed in Bio-Solids. Polymers 2023, 15, 4322. [Google Scholar] [CrossRef]
  13. Vital-Grappin, A.D.; Ariza-Tarazona, M.C.; Luna-Hernández, V.M.; Villarreal-Chiu, J.F.; Hernández-López, J.M.; Siligardi, C.; Cedillo-González, E.I. The Role of the Reactive Species Involved in the Photocatalytic Degradation of HDPE Microplastics Using C,N-TiO2 Powders. Polymers 2021, 13, 999. [Google Scholar] [CrossRef] [PubMed]
  14. Sharara, A.; Samy, M.; Mossad, M.; Gar Alalm, M. Photodegradation of Polyethylene Debris in Water by Sulfur-Doped TiO2: System Optimization, Degradation Mechanism, and Reusability. Environ. Sci. Pollut. Res. 2023, 31, 3951–3963. [Google Scholar] [CrossRef]
  15. Bussiere, P.-O.; Gardette, J.-L.; Rapp, G.; Masson, C.; Therias, S. New Insights into the Mechanism of Photodegradation of Chitosan. Carbohydr. Polym. 2021, 259, 117715. [Google Scholar] [CrossRef]
  16. Hosseinpour, S.; Tang, F.; Wang, F.; Livingstone, R.A.; Schlegel, S.J.; Ohto, T.; Bonn, M.; Nagata, Y.; Backus, E.H.G. Chemisorbed and Physisorbed Water at the TiO2/Water Interface. J. Phys. Chem. Lett. 2017, 8, 2195–2199. [Google Scholar] [CrossRef] [PubMed]
  17. Litke, A.; Su, Y.; Tranca, I.; Weber, T.; Hensen, E.J.M.; Hofmann, J.P. Role of Adsorbed Water on Charge Carrier Dynamics in Photoexcited TiO2. J. Phys. Chem. C 2017, 121, 7514–7524. [Google Scholar] [CrossRef]
  18. Chapa-Rodríguez, R.; Avila-de la Rosa, G.; Pérez, E. Thermal Stability and Ageing Properties of PP–PE Film Modulated by Nano-Silica Particles: Comparison between Dry and Moist Particles. Polym. Bull. 2021, 78, 3071–3088. [Google Scholar] [CrossRef]
  19. Burak, D.; Han, J.H.; Han, J.S.; Kim, I.S.; Rahman, M.A.; Yang, J.K.W.; Cho, S.-H. Controlling TiO2 Photocatalytic Behaviour via Perhydropolysilazane-Derived SiO 2 Ultrathin Shell. Nanoscale 2024, 16, 21960–21969. [Google Scholar] [CrossRef]
  20. Guo, J.; Van Bui, H.; Valdesueiro, D.; Yuan, S.; Liang, B.; Van Ommen, J. Suppressing the Photocatalytic Activity of TiO2 Nanoparticles by Extremely Thin Al2O3 Films Grown by Gas-Phase Deposition at Ambient Conditions. Nanomaterials 2018, 8, 61. [Google Scholar] [CrossRef]
  21. Ti-PureTM TiO2 by Chemours Titanium Dioxide Solutions for Plastics. Available online: https://www.tipure.com/en/applications/plastics (accessed on 22 May 2025).
  22. Titanium Dioxide—Venator. Available online: https://www.venatorcorp.com/products-and-applications/products/titanium-dioxide (accessed on 22 May 2025).
  23. Ti-PureTM Solutions for Coatings: Applications. Available online: https://www.tipure.com/en/applications/coatings (accessed on 22 May 2025).
  24. Campanale, C.; Savino, I.; Massarelli, C.; Uricchio, V.F. Fourier Transform Infrared Spectroscopy to Assess the Degree of Alteration of Artificially Aged and Environmentally Weathered Microplastics. Polymers 2023, 15, 911. [Google Scholar] [CrossRef]
  25. Zhou, Q.; Liu, X.; Lu, Y.; Dao, X.; Qiu, L. Accelerated Laboratory Weathering of Polypropylene/Poly (Lactic Acid) Blends. Polymers 2022, 15, 17. [Google Scholar] [CrossRef]
  26. Wang, L.; Xie, G.; Mi, X.; Zhang, B.; Du, Y.; Zhu, Q.; Yu, Z. Surface-Modified TiO2 @SiO2 Nanocomposites for Enhanced Dispersibility and Optical Performance to Apply in the Printing Process as a Pigment. ACS Omega 2023, 8, 20116–20124. [Google Scholar] [CrossRef] [PubMed]
  27. Shi, Y.; Zhao, Y.; Ma, M.; Chen, S.; He, H.; Zhu, Y.; Wang, X. LLDPE/TiO2 Composites with High UV Aging Resistance and Mechanical Properties by Controlling the Alumina Coating on TiO2. Polym. Compos. 2025, 46, 1383–1399. [Google Scholar] [CrossRef]
  28. Al-Fahham, A.A. Development of New LSD Formula When Numbers of Observations Are Unequal. Open J. Stat. 2018, 8, 258–263. [Google Scholar] [CrossRef]
  29. Mui, J.; Ngo, J.; Kim, B. Aggregation and Colloidal Stability of Commercially Available Al2O3 Nanoparticles in Aqueous Environments. Nanomaterials 2016, 6, 90. [Google Scholar] [CrossRef]
  30. Ceballos-Chuc, M.C.; Ramos-Castillo, C.M.; Rodríguez-Pérez, M.; Ruiz-Gómez, M.Á.; Rodríguez-Gattorno, G.; Villanueva-Cab, J. Synergistic Correlation in the Colloidal Properties of TiO2 Nanoparticles and Its Impact on the Photocatalytic Activity. Inorganics 2022, 10, 125. [Google Scholar] [CrossRef]
  31. Kang, X.; Chatzitakis, A.; Aarholt, T.; Sun, X.; Negri, C.; Norby, T. Facet-Engineered TiO2 Nanomaterials Reveal the Role of Water–Oxide Interactions in Surface Protonic Conduction. J. Mater. Chem. A Mater. 2022, 10, 218–227. [Google Scholar] [CrossRef]
  32. Kunc, F.; Gallerneault, M.; Kodra, O.; Brinkmann, A.; Lopinski, G.P.; Johnston, L.J. Surface Chemistry of Metal Oxide Nanoparticles: NMR and TGA Quantification. Anal. Bioanal. Chem. 2022, 414, 4409–4425. [Google Scholar] [CrossRef]
  33. Li, W.; Franco, D.C.; Yang, M.S.; Zhu, X.; Zhang, H.; Shao, Y.; Zhang, H.; Zhu, J. Comparative Study of the Performances of Al(OH)3 and BaSO4 in Ultrafine Powder Coatings. Processes 2019, 7, 316. [Google Scholar] [CrossRef]
  34. Liu, X. DRIFTS Study of Surface of γ-Alumina and Its Dehydroxylation. J. Phys. Chem. C 2008, 112, 5066–5073. [Google Scholar] [CrossRef]
  35. Milligan, L.H. The Mechanism of the Dehydration of Crystalline Aluminum Hydroxide and of the Adsorption of Water by the Resulting Alumina. J. Phys. Chem. 1922, 26, 247–255. [Google Scholar] [CrossRef]
  36. Vasile, B.S.; Dobra, G.; Iliev, S.; Cotet, L.; Neacsu, I.A.; Surdu, V.A.; Nicoara, A.I.; Boiangiu, A.; Filipescu, L. Thermally Activated Al(OH)3 Part II—Effect of Different Thermal Treatments. Ceramics 2021, 4, 564–575. [Google Scholar] [CrossRef]
  37. Zheng, X.; Yuan, F.; Ma, A.; Tian, S. Experimental Study on Microwave Drying Aluminum Hydroxide. Coatings 2024, 14, 687. [Google Scholar] [CrossRef]
  38. Wang, X.; Li, G.; Eckhoff, R.K. Kinetics Study of Hydration Reaction between Aluminum Powder and Water Based on an Improved Multi-Stage Shrinking Core Model. Int. J. Hydrogen Energy 2021, 46, 33635–33655. [Google Scholar] [CrossRef]
  39. Egerton, T. The Influence of Surface Alumina and Silica on the Photocatalytic Degradation of Organic Pollutants. Catalysts 2013, 3, 338–362. [Google Scholar] [CrossRef]
  40. Li, J.; Liu, Y.; Wang, Y.; Wang, W.; Wang, D.; Qi, T. Hydrous Alumina/Silica Double-Layer Surface Coating of TiO2 Pigment. Colloids Surf. A Physicochem. Eng. Asp. 2012, 407, 77–84. [Google Scholar] [CrossRef]
  41. Oguma, J.; Kakuma, Y.; Murayama, S.; Nosaka, Y. Effects of Silica Coating on Photocatalytic Reactions of Anatase Titanium Dioxide Studied by Quantitative Detection of Reactive Oxygen Species. Appl. Catal. B 2013, 129, 282–286. [Google Scholar] [CrossRef]
  42. Navidpour, A.H.; Abbasi, S.; Li, D.; Mojiri, A.; Zhou, J.L. Investigation of Advanced Oxidation Process in the Presence of TiO2 Semiconductor as Photocatalyst: Property, Principle, Kinetic Analysis, and Photocatalytic Activity. Catalysts 2023, 13, 232. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of PP/TiO2 plaques fabrication.
Figure 1. Schematic representation of PP/TiO2 plaques fabrication.
Coatings 15 00685 g001
Figure 2. PP plaques gloss retention under UV irradiation with 10 wt.% of TiO2 as a function of inorganic coatings content.
Figure 2. PP plaques gloss retention under UV irradiation with 10 wt.% of TiO2 as a function of inorganic coatings content.
Coatings 15 00685 g002
Figure 3. Stress at break retention of PP and PP/TiO2 composite as a function of inorganic coating of TiO2.
Figure 3. Stress at break retention of PP and PP/TiO2 composite as a function of inorganic coating of TiO2.
Coatings 15 00685 g003
Figure 4. Carbonyl index (CI) as a function of inorganic coatings content at different exposure times.
Figure 4. Carbonyl index (CI) as a function of inorganic coatings content at different exposure times.
Coatings 15 00685 g004
Figure 5. X-ray diffraction patterns of TiO2 as a function of inorganic coating content. The main diffraction peaks of rutile and anatase are included for comparison.
Figure 5. X-ray diffraction patterns of TiO2 as a function of inorganic coating content. The main diffraction peaks of rutile and anatase are included for comparison.
Coatings 15 00685 g005
Figure 6. Boxplot of statistical analyses, with chemical composition by EDS, for the three principal oxides in TiO2 particles: (a) TiO2 composition; (b) Al2O3 composition; (c) SiO2 composition.
Figure 6. Boxplot of statistical analyses, with chemical composition by EDS, for the three principal oxides in TiO2 particles: (a) TiO2 composition; (b) Al2O3 composition; (c) SiO2 composition.
Coatings 15 00685 g006
Figure 7. Transmission electron micrographs of TiO2 particles: (a) LA; (b) LA+S; (c) HA; (d) HA+S.
Figure 7. Transmission electron micrographs of TiO2 particles: (a) LA; (b) LA+S; (c) HA; (d) HA+S.
Coatings 15 00685 g007
Figure 8. LSD test and boxplot of TiO2 particle sizes. The letters A, B, and C correspond to model interactions.
Figure 8. LSD test and boxplot of TiO2 particle sizes. The letters A, B, and C correspond to model interactions.
Coatings 15 00685 g008
Figure 9. Boxplot of zeta potential and LSD test of TiO2 samples, at pH 6.6.
Figure 9. Boxplot of zeta potential and LSD test of TiO2 samples, at pH 6.6.
Coatings 15 00685 g009
Figure 10. Zeta potential as a function of Al2O3 and SiO2 composition. Experimental data (●) and model fitting (□).
Figure 10. Zeta potential as a function of Al2O3 and SiO2 composition. Experimental data (●) and model fitting (□).
Coatings 15 00685 g010
Figure 11. Schematic representation of hydroxyl ions formation on the rutile TiO2 surface, titanium (blue spheres), oxygen (white spheres).
Figure 11. Schematic representation of hydroxyl ions formation on the rutile TiO2 surface, titanium (blue spheres), oxygen (white spheres).
Coatings 15 00685 g011
Figure 12. TGA thermograms of TiO2 pigment of different grades: (a) dynamic analysis at 10 °C/min of heating rate; (b) isothermal analysis at 300 °C.
Figure 12. TGA thermograms of TiO2 pigment of different grades: (a) dynamic analysis at 10 °C/min of heating rate; (b) isothermal analysis at 300 °C.
Coatings 15 00685 g012
Figure 13. DRIFTS spectrum of each TiO2 sample at a heating rate of 10 °C/min: (a) LA; (b) LA+S; (c) HA; (d) HA+S.
Figure 13. DRIFTS spectrum of each TiO2 sample at a heating rate of 10 °C/min: (a) LA; (b) LA+S; (c) HA; (d) HA+S.
Coatings 15 00685 g013
Table 1. Experimental composition of TiO2 pigment.
Table 1. Experimental composition of TiO2 pigment.
Sample IDInorganic Coatings Content
Experimental (wt.%)
TiO2Al2O3SiO2
LA98.5 ± 0.31.5 ± 0.3/
LA+S95.0 ± 0.32.5 ± 0.52.5 ± 0.3
HA96.4 ± 0.33.6 ± 0.3/
HA+S92.7 ± 0.63.8 ± 0.33.4 ± 0.3
/ not reported.
Table 2. ANOVA summary with chemical analyses for the principal components: TiO2, Al2O3, and SiO2 (a degree of freedom, b Fisher statistic, c probability).
Table 2. ANOVA summary with chemical analyses for the principal components: TiO2, Al2O3, and SiO2 (a degree of freedom, b Fisher statistic, c probability).
SourceSum of Squaresd.f. aMean SquaresF bP c
TiO2Between52.30317.43110.980.0000
Within (error)1.2680.16
Total53.5511
Al2O3Between10.3533.4526.580.0000
Within (error)1.0480.13
Total11.3911
SiO2Between1.1311.1314.050.0199
Within (error)0.3240.08
Total1.465
Table 3. ANOVA summary table for particle diameters. Here, ANOVA analysis decomposes the data variance into two components: between-group component and within-group component (a degree of freedom, b Fisher statistic, c probability).
Table 3. ANOVA summary table for particle diameters. Here, ANOVA analysis decomposes the data variance into two components: between-group component and within-group component (a degree of freedom, b Fisher statistic, c probability).
SourceSum of Squaresd.f. aMean SquaresF bP c
Between9409.4333136.481.670.226
Within (error)22,579.92121881.66
Total31,989.3515
Table 4. ANOVA summary table for zeta potential at pH 6.6 (a degree of freedom, b Fisher statistic, c probability).
Table 4. ANOVA summary table for zeta potential at pH 6.6 (a degree of freedom, b Fisher statistic, c probability).
SourceSum of Squaresd.f. aMean SquaresF bP c
Between411.473137.16110.480.0000
Within (error)29.80241.24
Total441.2627
Table 5. Total water values obtained by the Karl Fischer titration method at T = 190 °C.
Table 5. Total water values obtained by the Karl Fischer titration method at T = 190 °C.
IDMoisture
(wt.%)
HA+S0.54
HA0.45
LA+S0.19
LA0.11
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

Armendáriz-Alonso, E.F.; Rivera-García, N.; Moreno-Razo, J.A.; Meza-Espinoza, L.O.; Waldo-Mendoza, M.A.; Pérez, E. Structural Water Content in Pigment-Grade TiO2 Particles Coated with Al2O3 and SiO2, and Their Effect on Polypropylene Photodegradation. Coatings 2025, 15, 685. https://doi.org/10.3390/coatings15060685

AMA Style

Armendáriz-Alonso EF, Rivera-García N, Moreno-Razo JA, Meza-Espinoza LO, Waldo-Mendoza MA, Pérez E. Structural Water Content in Pigment-Grade TiO2 Particles Coated with Al2O3 and SiO2, and Their Effect on Polypropylene Photodegradation. Coatings. 2025; 15(6):685. https://doi.org/10.3390/coatings15060685

Chicago/Turabian Style

Armendáriz-Alonso, Edgar F., Nancy Rivera-García, J. Antonio Moreno-Razo, Luis Octavio Meza-Espinoza, Miguel A. Waldo-Mendoza, and Elías Pérez. 2025. "Structural Water Content in Pigment-Grade TiO2 Particles Coated with Al2O3 and SiO2, and Their Effect on Polypropylene Photodegradation" Coatings 15, no. 6: 685. https://doi.org/10.3390/coatings15060685

APA Style

Armendáriz-Alonso, E. F., Rivera-García, N., Moreno-Razo, J. A., Meza-Espinoza, L. O., Waldo-Mendoza, M. A., & Pérez, E. (2025). Structural Water Content in Pigment-Grade TiO2 Particles Coated with Al2O3 and SiO2, and Their Effect on Polypropylene Photodegradation. Coatings, 15(6), 685. https://doi.org/10.3390/coatings15060685

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