1. Introduction
Platinum-based catalysts have long been at the forefront of heterogeneous catalysis due to their exceptional activity, selectivity, and stability in a wide range of chemical reactions, including hydrogenation, oxidation, and environmental remediation processes [
1]. In particular, platinum nanoparticles (PtNPs) on metal oxides provide a versatile platform for catalytic applications, as the nanoscale dispersion of Pt enhances the availability of active sites, while the support can influence the electronic properties and stability [
2,
3]. In addition to metallic PtNPs, recent studies have also investigated ionic platinum species as active catalytic centers, particularly in systems in which Pt is predominantly present in ionic states (e.g., Pt
2+ or Pt
4+). Okumura et al. [
4] synthesized platinum catalysts supported on metal oxides, in particular CeO
2, which contain highly dispersed ionic Pt species, mainly Pt
4+ and Pt
2+, and play a key role in catalytic performance. CeO
2 was effective in stabilizing Pt
4+, with some catalysts retaining over 70% Pt
4+ even after repeated use. These ionic species were catalytically active and could be reversibly reduced to Pt
0 under reaction conditions, forming dynamic active sites. The strong metal–support interaction and high dispersion of Pt
4+ were crucial for high activity, stability and reusability in hydrosilylation reactions. Mukri et al. [
5] presented the use of Pt-doped ionic catalysts in which Pt
2+ contributes significantly to the catalytic activity of titanium(IV) oxide (TiO
2). These ions, in conjunction with weakly bonded lattice oxygen, promote electron transfer and oxygen activation, which are essential for catalytic oxidation reactions. Pt
4+ was significantly less active, as these ions occupied octahedral sites and, thus, lacked the structural and electronic effects necessary for catalysis. These ionic Pt catalysts have demonstrated unique reactivity and stability, opening new avenues for catalyst design. Barsan et al. [
6] demonstrated that platinum can also to be present in its oxidized form. The oxidized platinum species, primarily as Pt
4+—in PtO
2-like clusters—are highly active and stable on the surface of SnO
2. These species are not reduced under operating conditions and remain catalytically accessible. Their strong interaction with the SnO
2 surface, probably via Pt–O–Sn linkages, enhances electron transfer and gas activation. Unlike metallic Pt, these oxidized Pt forms retain their activity even in humid environments, making them effective and durable catalytic centers for CO sensing in the presence of water vapor.
Bimetallic Pt-based catalysts offer enhanced catalytic activity, selectivity, and stability compared to pure Pt, while reducing the platinum content. Lai et al. [
7] have reported on carefully engineered nanostructures, such as Pt-Pd or Pt-Ni alloys in core–shell or dendritic form, which show excellent performance in reactions, such as fuel cell operation and hydrogen evolution. A major drawback remains the difficulty in scaling up reproducible, controllable synthesis methods on a larger scale, which limits their wider application. Xu et al. [
8] synthesized Pt-based ultrafine one-dimensional nanowires with superior catalytic activity and structural stability, making them suitable for advanced energy applications. Their elongated shape enhances transport properties and active site exposure, but challenges remain in scalable synthesis and precise control of their structure and composition, especially when scalable and reproducible methods are required. Recent advancements in Pt-based catalysts supported by carbon and conductive polymers have shown promise for direct methanol fuel cell anodes, especially with novel carbon materials, like nanotubes and graphene. Ramli and Kamarudin [
9] report that alloying Pt with metals, like Fe, Co, and Ni, and designing advanced nanostructures, such as nanowires or hollow particles, significantly improves catalytic activity and stability while reducing Pt usage. However, challenges remain in scalable synthesis, high metal loading, and uniform dispersion on various carbon supports, particularly with high surface area materials. Zhan et al. [
10] investigated non-carbon supports, such as metal oxides. Understanding and optimizing the strong metal–support interaction can significantly improve ORR performance. Single-atom catalysts offer high Pt utilization and catalytic efficiency, but future efforts need to focus on improving their stability and scalability for commercial applications. Among the various oxide supports, tin(IV) oxide (SnO
2) has received increasing attention due to its high thermal stability, chemical stability, and ability to synergistically interact with Pt species, which can improve catalytic performance through strong metal–support interactions. Specifically, SnO
2 offers the additional advantage of facilitating electron transfer and creating oxygen vacancies, which can enhance catalytic activity [
11]. When used as a support for Pt nanoparticles, SnO
2 has been shown to improve catalyst dispersion and stability, potentially leading to enhanced catalytic performance in redox reactions [
12]. The catalytic efficiency of Pt/SnO
2 systems is strongly influenced by the synthesis method and the purity of the final catalyst.
Building on our previous work, where we developed a microwave-assisted synthesis of Pt/SnO
2 catalysts with high activity and reusability [
13], we now take this a step further by improving this green synthesis approach. Specifically, we have developed a near room-temperature method for synthesizing both the SnO
2 support and dispersing platinum, virtually without energy consumption. This advancement was made possible by removing chloride ions from the precursor solutions. By eliminating chloride during synthesis, we were able to obtain well-crystallized cassiterite-phase SnO
2 at room temperature. Chloride removal is particularly important, as it is believed that chloride ions can hinder catalyst performance by poisoning active sites, disrupting nanoparticle dispersion, or altering the surface chemistry of the support [
14]. This issue is especially pronounced in surface-sensitive reactions [
15] or in aqueous environments where chloride can leach or promote undesirable side reactions [
16]. To investigate how support properties influence platinum dispersion and catalytic behavior, we synthesized three distinct chloride-free SnO
2 supports, as follows:
SnA, obtained via room-temperature precipitation following ion exchange;
SnB, produced by hydrothermal treatment of SnA at 180 °C, resulting in larger SnO2 particles;
SnC, generated by annealing SnB at 600 °C to further increase crystallinity and particle size.
These supports were then used to prepare Pt/SnO2 catalysts with identical platinum content. Notably, the platinum precursor (H2PtCl6) was dispersed at a temperature of only 40 °C, in contrast to the conventional methods that require temperatures around 300–400 °C. We were aware that such a low-temperature treatment would not result in the formation of metallic platinum nanoparticles. This assumption was confirmed by XPS analysis, which revealed the presence of oxidized Pt species (Pt2+ and Pt4+) on the surface of the SnO2 supports, forming PtOx/SnO2 catalysts. The catalytic performance of these PtOx/SnO2 catalysts was evaluated in the reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP). The results were correlated with the structural and textural properties of the catalysts, such as surface area and pore size distribution. By comparing materials that differ only in SnO2 particle size and crystallinity, this study aims to clarify the influence of support structure on platinum dispersion and overall catalytic efficiency.
2. Materials and Methods
2.1. Chemicals
Tin(IV) tetrachloride (Product No. 244678), manufactured by Sigma Aldrich (Steinheim, Germany), and hexachloroplatinic acid (CAS 18497-13-7), manufactured by ThermoFischer (Karlsruhe, Germany), were used as received. AmberLite HPR 550 ion exchange resin is supplied by Sigma Aldrich (St. Louis, MO, USA) in the form of translucent orange spheres with a particle size of 590 ± 50 μm. It was used after being rehydrated in deionized Milli-Q (MQ) water. The chemicals used for the catalytic experiments were as follows: 4-nitrophenol—Sigma Aldrich (Steinheim, Germany), Reagent Plus, ≥99%, CAS: 100-02-7, Product No.: 241.326 and sodium borohydride (NaBH4)—Alfa Aesar (Karlsruhe, Germany), min.98%, CAS: 16940-66-2, Product No.: 88983. Deionized Milli-Q water was used for catalytic experiments.
2.2. Stock Solution Preparation
For the SnCl4 stock solution, 35.06 g of the SnCl4·5 H2O powder was weighed out and dissolved in 50 mL of deionized Milli-Q water. For the H2PtCl6 stock solution, 5 g of the solid H2PtCl6·6 H2O was mixed with 4.93 mL of deionized Milli-Q water. The calculated concentrations of the tin and the platinum stock solutions were 2.0 mol dm–3 each (2M SnCl4 and 2M H2PtCl6).
2.3. Anion Exchange Resin Preparation
The anion exchange resin, which is supplied in dehydrated form, must be rehydrated to restore its functionality. This is achieved by immersing 160 g of resin in deionized MQ water, so that the volume of the resin submerged in MQ water is 400 mL (performed in a plastic 500 mL beaker). After 20 min of stirring (with a magnetic stirrer) to allow the resin beads to swell and fully hydrate, the resin is treated with a sodium hydroxide solution. In this step, the resin is converted to its hydroxide form, which increases its anion exchange capacity. The process involves mixing the resin with a 2M NaOH solution, rinsing the resin with MQ water and repeating the treatment with 2M NaOH. The resin is then thoroughly rinsed with MQ water to remove all residual NaOH and displaced Na+ and OH– ions. This step is repeated until the rinse water reaches a neutral pH, indicating that no excess NaOH is present and confirming that the resin is in its desired hydroxide form.
The anion exchange resin used to remove chloride anions from the solution becomes saturated with chloride ions over time, so regeneration is required to restore its capacity for effective ion exchange. Regeneration converts the resin from the chloride form back to the hydroxide form, reactivating its anion exchange function. First, the regenerating NaOH solution must be prepared at a concentration of 6 to 8% by weight. The regenerating solution is then stirred with the resin for 15 min and then thoroughly rinsed twice with MQ water in a sieve to remove all residual NaOH and displaced chloride ions [
17]. This process is repeated five or more times, depending on the condition of the anion exchange resin, until all chloride ions are removed from the resin beads, which is checked using silver nitrate (AgNO
3) to exclude the presence of chloride ions. Once there is no more reaction, the resin is ready for reuse.
2.4. Synthesis of the Supports and Samples
For all three types of support, i.e., SnA, SnB, and SnC, 18 mL of 2 M SnCl
4 was diluted with 102 mL of MQ water to prepare a 0.3 M SnCl
4 solution. This solution was stirred for 20 min with a preconditioned anion exchange resin (the aforementioned 160 g of resin) to facilitate the removal of chloride ions by replacing them with hydroxide ions [
17]. The resulting milky white suspension appeared after mixing for 20 min and reaching pH 3.5. It was carefully decanted, taking care not to entrain any resin beads. The entire support synthesis process is shown in
Figure 1, while the precise steps are given in the next three paragraphs.
For the SnA support, the suspension was quantitatively transferred to a glass beaker and stirred with a magnetic stirrer at 140 rpm and 40 °C for 72 h until complete evaporation of the water. The precipitate was dried overnight at 60 °C and then scraped from the beaker and homogenized using a mortar and pestle. Then, 1 g of the resulting white powder was stirred in 20 mL of MQ water with 35 µL of 2M H
2PtCl
6 to obtain a Pt loading of 1 mol% (
Figure 2, identical for all three supports). This Pt-loaded SP1a sample was stirred with a magnetic stirrer at 140 rpm and 40 °C for 48 h, and then dried overnight at 60 °C. In the same manner, a sample with 10 mol% Pt was synthesized by adding 221 µL of 2M H
2PtCl
6 to 0.6 g of SnA support to obtain SP10a.
For the SnB support, the suspension was subjected to hydrothermal treatment in an autoclave at 180 °C for 24 h. After autoclaving, the suspension was transferred to a glass beaker with a magnetic stirrer and stirred at 140 rpm at 40 °C for 72 h to evaporate the water. The resulting precipitate (1 g) was dried overnight at 60 °C, collected, homogenized in a mortar and pestle, and used for platinum precipitation with 35 µL of 2M H2PtCl6 to reach a final Pt content of 1 mol%. The SP1b sample was stirred and dried, like the SP1a sample.
For the SnC support, the suspension was hydrothermally treated in an autoclave at 180 °C for 24 h. After autoclaving, the suspension was transferred to a glass beaker with a magnetic stirrer and stirred at 140 rpm at 40 °C for 72 h to evaporate the water. The powder was dried overnight at 60 °C and then annealed in a tube furnace at 600 °C for 2 h to enhance crystallinity. The annealed SnO2 (1 g) was then combined with 35 µL of 2M H2PtCl6 to achieve a Pt loading of 1 mol%. This SP1c sample was stirred and dried, like the SP1a and SP1b samples.
2.5. Instrumental Analysis
X-ray diffraction (XRD) measurements were carried out at room temperature using an APD 2000 diffractometer (CuKα radiation, graphite monochromator, NaI-Tl detector) from ITALSTRUCTURES, Riva Del Garda, Italy.
Scanning electron microscopy (SEM) was performed with a Jeol Ltd. (Tokyo, Japan) 700F field emission SEM coupled to the EDS/INCA 350 system for energy-dispersive X-ray spectrometry (EDXS), constructed by Oxford Instruments Ltd. (Abingdon, UK).
An atomic resolution scanning transmission electron microscope (AR STEM), namely the Jeol ARM 200 CF model (JEOL Ltd., Tokyo, Japan) operating at 200 kV, was used for this study. This instrument was coupled to the Gatan Quantum ER system, which includes electron energy loss spectroscopy and energy-dispersive X-ray spectrometry capabilities using the Jeol Centurio 100 module.
For the nitrogen adsorption analysis performed at 77 K, the Quantachrome Autosorb iQ3 system (from Quantachrome Instruments, Boynton Beach, FL, USA) was used, which utilizes the Brunauer–Emmett–Teller (BET) technique to evaluate material properties. Prior to testing, a controlled heating process up to 250 °C was carried out under vacuum conditions to remove residual gases and moisture. The evacuation process was continued until the pressure fluctuations stopped increasing rapidly and reached a value below 50 millitorr per minute. Isothermal adsorption and desorption measurements were then performed at 77 K over a relative pressure range of about 10–5 up to almost 0.99.
X-ray photoelectron spectroscopy (XPS) was used to investigate the oxidation state of Sn and Pt in the oxidized Pt species on SnO
2 supports. The analysis was carried out under ultra-high vacuum (UHV) conditions using a SPECS instrument (SPECS Surface Nano Analysis GmbH, Berlin, Germany). The experimental setup used an excitation energy of 1486.74 eV derived from the Al Kα X-ray emission and the Phoibos100 electron energy analyzer. To neutralize the charge accumulation in non-conductive oxide samples, a 5 eV electron flooding method was applied during the XPS analysis. A pass energy of 50 eV was chosen for the evaluation of the Pt 4f core levels, while a pass energy of 10 eV was used for the spectra around the Sn 3d levels. The experimental curves were fitted using a combination of Gaussian and Lorentz functions via the Unifit software 2024 (R. R. Hesse–UNIFIT Software, Leipzig, Germany) [
18]. All photoemission spectra were calibrated using the C 1s peak, which was set at a binding energy (BE) of 284.5 eV.
Thermogravimetric analysis (TGA) was performed with a Mettler Toledo TGA/DSC 3+ instrument (Mettler Toledo, Schwerzenbach, Switzerland). In a typical experiment, 10-12 mg of the SnO2 support was placed in an aluminum oxide crucible and inserted into a furnace. The supports experienced non-isothermal heating from 35 °C to 1000 °C at a heating rate of 10 °C/min and a constant nitrogen (N2) gas flow at a rate of 50 mL/min. The thermogravimetric analysis and differential scanning calorimetry (DSC) data were recorded using a computer synchronized with the furnace. The differential thermogravimetry (DTG) data, generated from the first derivative of the TG, depicted the mass loss rate of the materials with increasing time or temperature. The thermal behavior and characteristic parameters of the supports in question were concluded from the TG and DTG data.
Then, 119Sn Mössbauer spectroscopy measurements were performed with a WissEl Mössbauer spectrometer (WissEl GmbH, Starnberg, Germany) setup operated in transmission geometry, with the samples and the source being kept at room temperature. A 119mSn(CaSnO3) radioactive source (RITVERC JSC) with an activity of ~0.35 mCi provided the γ-rays. The source was driven by a sinusoidal velocity signal, with velocity extrema of approximately ±6 mm s−1. The raw spectra consisted of 2048 channels, which were subsequently folded into 1024 channels for further processing. The 119Sn isomer shift (δ) values are given relative to that of a SnO2 reference powder (Merck, Budapest, Hungary) having an isomer shift equal to that of the CaSnO3 source matrix. Circular absorbers with a diameter of 15.5 mm were prepared by uniformly mixing approximately 100 mg of cellulose (as a filler) with either 14.2 mg of SP1a powder or 15 mg of SP1b and SP1c powders. The MossWinn 4.0 program (Institute for Nuclear Research (Atomki), Debrecen, Hungary) was used to analyze the spectra, assuming the thin absorber approximation.
UV–Vis reflectance spectra were recorded with a Shimadzu UV/VIS/NIR spectrometer, model UV-3600 (Shimadzu Corporation, Kyoto, Japan). The wavelength range used was from 600 to 200 nm.
2.6. Catalytic Measurements
The catalytic reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP) was investigated using UV–visible spectrophotometry in the presence of NaBH
4 and the synthesized Pt-loaded samples. Before each measurement, the solutions containing 4-NP and NaBH
4 were not purged with nitrogen. Immediately before each experiment, a fresh aqueous solution of NaBH
4 was prepared to maintain the reducing efficiency. For the SP1a sample, 0.3 μmol of 4-NP (20 μL of a 0.015 M solution) was diluted with 2.7 mL of ultrapure water in a quartz cuvette, followed by the addition of 79.3 μmol of NaBH
4 (20 μL of a 0.793 M solution) [
19]. Samples SP1b and SP1c, because of their size, settled down before being able to act as catalysts. For better dispersion, samples SP1b and SP1c were prepared for catalytic measurements in the following manner: 0.3 μmol of 4-NP (20 μL of a 0.015 M solution) was diluted with 2.7 mL of 1% polyvinylpyrrolidone (PVP) solution in a quartz cuvette, followed by the addition of 79.3 μmol of NaBH
4 (20 μL of a 0.793 M solution). Subsequently, 20 μL of a catalyst suspension (3 mg/mL in ultrapure water) was added and rapidly mixed using a micropipette. UV–visible spectra were recorded immediately after catalyst addition and monitored over time until the characteristic absorbance of nitrophenolate ions at 400 nm disappeared. The progress of the reaction and the formation of 4-AP were tracked by the emergence of a new absorption maximum at 300 nm.
The catalyst reusability was tested under the same experimental conditions in ten consecutive cycles. After each cycle, an additional 0.3 μmol of 4-NP was added to the cuvette and thoroughly mixed, followed by spectral recording, as described. To maintain a constant excess of reducing agent, 79.3 μmol of NaBH4 was also added before the fourth and seventh cycles. This approach ensured sufficient NaBH4 concentration throughout the experiment, allowing a reliable evaluation of catalyst stability and durability.
4. Discussion
In previous studies, platinum-based catalysts were synthesized on various metal oxides, such as Fe
2O
3, SnO
2, and MnO
2, by thermal or mechanochemical methods, often using organometallic precursors and high-temperature post-treatments in controlled atmospheres. Platinum-doped SnO
2, for example, was produced by wet impregnation of Pt(acac)
2 in toluene, followed by ball-milling and annealing in argon and oxygen atmospheres at 400 °C [
32]. Another approach was to synthesize MnO
2 nanostructures by microwave-assisted hydrothermal treatment, followed by solvent evaporation and annealing after mixing with Pt(acac)
2 in toluene to achieve different Pt loadings [
33]. While these methods are effective, they require either high temperatures, long processing times, or organic solvents, which can limit scalability and environmental compatibility.
The method we used in our previous work [
13] was the microwave-assisted hydrothermal synthesis of Pt-doped SnO
2 materials by direct coprecipitation of H
2PtCl
6 and SnCl
4 at very low pH, followed by ammonia-induced precipitation, microwave treatment at 230 °C, and calcination at 400 °C. This process provided better control over the Pt doping concentration with molar ratios ranging from 0 to 15 mol% and aimed to produce well-dispersed Pt species within the SnO
2 matrix. Although this route is more aqueous in nature, it still required high-temperature microwave synthesis and post-synthetic annealing to ensure oxide formation and phase purity.
In contrast, the present work presents an aqueous low-temperature synthesis of Pt/SnO2 catalysts based on a chloride-free precursor approach, in which SnCl4 is first purified by anion exchange and platinum is introduced as H2PtCl6 under mild conditions without the use of external reducing agents or organic solvents. This synthesis strategy eliminates the need for the high-temperature decomposition of organometallic complexes and enables the formation of catalytically active PtOx species (Pt2+ and Pt4+) at room temperature. Moreover, the approach emphasizes strong metal–support interactions on hydroxyl-rich SnO2 surfaces, which not only stabilize the platinum species but also contribute to their catalytic activity, as demonstrated in the reduction of 4-nitrophenol. This work, therefore, provides a more sustainable and controllable route to platinum-based catalysts compared to previous methods, while expanding the mechanistic understanding of oxidized Pt species in heterogeneous catalysis.
The structural, morphological and surface chemical properties of the supports and samples were systematically analyzed by XRD with Rietveld refinement, STEM imaging, TGA/DSC/DTG analysis, BET surface analysis, XPS, Raman spectroscopy, and 119Sn Mössbauer spectroscopy. The catalytic activity of the samples was evaluated using the model reduction of 4-nitrophenol to 4-aminophenol in aqueous NaBH4 under ambient conditions and monitored by UV–Vis spectroscopy.
The Rietveld refinements (
Figure 3 and
Table 1) provide detailed structural comparisons between the three SnO
2-based samples—SP1a, SP1b, and SP1c—which were each subjected to increasingly intense thermal treatment. SP1a, which was synthesized only by removing chloride ions through anion exchange, shows broad and low-intensity diffraction peaks, indicating low crystallinity and a nanocrystalline structure with high lattice disorder. This is to be expected for a material that has not undergone hydrothermal or thermal processing. SP1b, which was hydrothermally treated at 180 °C for 24 h, shows narrower and more intense diffraction peaks, indicating a significant improvement in crystal quality and order in the SnO
2 lattice [
34]. The most pronounced improvement in crystallinity is observed in SP1c, which was subjected to both hydrothermal treatment and subsequent annealing at 600 °C. This sample exhibits the sharpest and most intense peaks as well as the narrowest full width at half maximum (FWHM), indicating larger crystallite sizes and minimal microstrain [
35]. These results are in excellent agreement with complementary data from XPS and Raman spectroscopy, which also indicate a higher degree of crystallinity and fewer surface defects. The degree of SnO
2 crystallinity influences the local environment of Sn
4+ ions, as evidenced by variations in the isomer shift and quadrupole splitting values observed in the
119Sn Mössbauer spectroscopy. Taken together, these structural differences help to explain the different catalytic activity and thermal stability of the three samples. The quality of the Rietveld fit, as indicated by the
Rwp values, remains acceptable for all samples with values between 0.053 and 0.061, confirming the reliability of the refinements. These results show that crystallinity and structural coherence gradually improve with increasing thermal treatment.
STEM-EDXS mapping and high-resolution imaging (
Figure 4 and
Figure 5) confirm the homogeneous distribution of the Pt species on the SnO
2 support. In particular, the Pt particles are molecularly dispersed, with no signs of agglomeration or large clusters, supporting the hypothesis that platinum is present in an oxidized and highly dispersed form.
We chose two platinum loadings—1 mol% and 10 mol%—to achieve a balance between catalytic relevance and structural characterization. The 1 mol% Pt loading was chosen as a representative concentration commonly used in catalysis to ensure sufficient dispersion of the active sites while minimizing platinum consumption. However, due to the high dispersion and low atomic contrast of the PtOx nanoparticles at this concentration, direct imaging and confirmation of their presence using STEM proved to be difficult. To overcome this limitation and clearly confirm the localization and morphology of the Pt species on the SnA support, a second sample was synthesized with 10 mol% Pt. The histogram in
Figure 6 displays the size distribution of PtOx NPs obtained from STEM micrographs with fits based on normal and lognormal distribution models. The normal distribution fit yields a mean particle size of 0.89 nm and a standard deviation of 0.25 nm, indicating a narrow and uniform distribution of particle size. The lognormal distribution, which better reflects skewed distributions common in nanoparticle systems, yields a mode of 0.84 nm, a median of 0.90 nm, and a mean of 0.98 nm. The slight right skew of the histogram supports the appropriateness of the lognormal model and reflects the presence of some larger particles. Overall, the data confirm that the PtNPs have a narrow and well-controlled size distribution centered below 1 nm, consistent with high dispersion on the support, which is advantageous for catalytic applications due to the larger surface area and number of accessible active sites [
23].
SnA exhibits the highest nitrogen uptake and BET surface area (134.4 m
2/g), indicating a well-developed mesoporous structure with a narrow pore size distribution centered around 3.2 nm. This high mesoporosity results from the interparticle spaces between uniform 5 nm nanocrystalline cassiterite particles formed via a low-temperature, chloride-free aqueous precipitation route. Specifically, SnA was prepared via an anion exchange route, wherein an aqueous solution of SnCl
4 was subjected to chloride removal using an anion exchange resin. This process leads to the immediate (in 20-30 min) appearance of a milky-white colloidal suspension, indicative of the rapid formation of Sn(IV)-oxo or hydroxide clusters due to extensive hydrolysis and subsequent olation/oxolation reactions, especially in the absence of chloride, which normally stabilizes Sn
4+ in solution. The primary hydrolysis step can be represented with Reaction (1), followed by the condensation reactions in Reaction (2):
This leads to the formation of small, highly hydrated tin–oxo clusters, which then aggregate into nanocrystalline cassiterite (SnO2), as confirmed by XRD and STEM. Due to the low temperature and aqueous environment, crystal growth is kinetically limited, resulting in crystallite sizes of ~5 nm. Importantly, these nanocrystals do not coarsen significantly due to the absence of thermal treatment, preserving high dispersion and interparticle voids, thus forming a mesoporous network without the need for templating agents. The H2-type hysteresis suggests constricted or “bottleneck” pores due to the random packing of nanocrystals. SnB also displays a mesoporous profile with a slightly reduced surface area of 117.4 m2/g and its pore size distribution peaking around 3.4 nm, indicating a similar but marginally wider pore network after autoclaving. In contrast, SnC, which was annealed at 600 °C, exhibits significantly lower nitrogen uptake and a reduced BET surface area of 35.1 m2/g, reflecting a much less developed mesostructure. Its pore size distribution is broader and shifted toward larger diameters (~8.5–9 nm), indicating a less uniform and partially collapsed or sintered porous structure due to the thermal treatment.
The introduction of Pt, through hydrolysis of H2PtCl6 at 40 °C, generally leads to a decrease in specific surface areas and altered pore volumes across all supports, while maintaining their type IV/H2 characteristics. SP1a (Pt on SnA) displays a BET surface area of 122.6 m2/g, slightly lower than SnA (134.4 m2/g), indicating partial pore blockage or surface coverage by highly dispersed 0.85 nm PtOx clusters (“faint stains” or patches) as observed by STEM. The isotherm maintains the H2-type hysteresis, with a subtle narrowing in the desorption branch suggesting a more constrained mesopore network, while the pore size distribution remains centered around 3.2 nm. The Pt precursor (H2PtCl6) was added without chloride removal, introducing residual Cl− into the system. The strong interaction of H2PtCl6 with the hydrated SnO2 surface, combined with ambient oxidation conditions, likely leads to the formation of highly dispersed hydrated PtOx species via hydrolysis reactions on the surface. For SP1b (Pt on autoclaved SnB), the surface area reduces to 103.1 m2/g compared to SnB (117.4 m2/g), consistent with Pt-induced partial pore coverage. The H2-type hysteresis remains visible but is narrower and shifted toward lower relative pressures, indicating smaller or more constricted pores, possibly due to structural rearrangement during hydrothermal treatment and subsequent Pt deposition, with the pore size distribution still centered around 3.2–3.4 nm. The strong interaction with the support promotes well-dispersed PtOx even with residual chloride. SP1c (Pt on annealed SnC) shows the most pronounced changes, with its surface area dropping significantly to 28.5 m2/g (from 35.1 m2/g for SnC). The isotherm reveals a broader hysteresis loop (H1-type for SnC and SP1c) at higher relative pressures (P/P0 > 0.6), indicating a more disordered and less uniform mesoporous system due to crystallite sintering and grain growth during annealing at 600 °C. The pore size distribution shifts dramatically to larger diameters (10–15 nm), reflecting the collapse of smaller pores and partial destruction of mesoporosity. Consequently, the PtOx dispersion occurs primarily at the external surface due to the largely lost internal mesopores. In summary, the synthetic conditions critically influence the mesoporous texture, with SnA and SnB retaining high surface areas and uniform mesoporosity post-Pt incorporation, making them suitable for active species dispersion. The observed textural changes upon Pt incorporation, including the subtle reduction in surface area and potential pore narrowing, reflect the effective dispersion of ultrasmall PtOx domains on the SnO2 supports.
From the combined characterization results, we conclude that platinum in the SP10a sample is predominantly present in the form of surface-anchored PtOx species with the presence of Sn-O-Sn, Pt-O-Pt, Pt–Cl, Pt–OH, and, most importantly, Pt-O-Sn bonds on the surface of SnO
2 catalyst. Although the SP10a sample was synthesized with a significantly higher Pt loading (10 mol%) compared to SP1a (1 mol%), STEM-EDS analysis (
Figure S3) shows no increase in chloride content, suggesting that the chloride is not freely dispersed but rather is coordinated to the Pt species. Thermal analysis of SP10a (
Figure S4) shows a more gradual and prolonged mass loss profile than the chloride-free SnA support, indicating the presence of strongly bound hydrolyzed Pt complexes. This is confirmed by Raman spectroscopy (
Figure S8), where characteristic bands are observed in the 300–330 cm
−1 range, and are typically associated with the Pt–Cl vibrational modes. The Raman band at 710 cm
−1 strongly suggested the presence of Pt-O-Sn surface bonds, thus confirming strong SnO
2/PtOx interactions. In addition, XPS analysis (
Figure 9 and
Figure S5) confirms the predominance of PtOx platinum species, particularly Pt
2+ and Pt
4+, and shows a strong signal corresponding to hydroxyl groups, consistent with Pt–OH surface coordination. Taken together, these results suggest a system in which platinum is immobilized on SnO
2 by stable PtOx species instead of forming mostly metallic nanoparticles.
Importantly, despite the mild synthesis conditions—room temperature and no external reducing agents—the XPS analysis revealed the presence of metallic platinum (Pt
0) on the surface of the Pt/SnO
2 catalysts. In the sample with 1 mol% Pt loading, about 25% of the Pt was present in metallic state, while in the sample with 10 mol% Pt, the Pt
0 content increased to about 35% (
Figure S5). Although this observation may seem counterintuitive since H
2PtCl
6 is normally thermally reduced at ≥300 °C, several factors may explain the spontaneous partial reduction of Pt
4+ under our synthesis conditions. First, the SnO
2 support (SnA) synthesized at room temperature contains a considerable number of surface hydroxyl groups (OH form) due to the use of an anion exchange resin. These surface –OH groups, which are coordinated to Sn
4+ centers, can act as mild reducing agents. Their involvement in redox reactions is already known, especially in systems with strong metal–support interactions [
36]. The following simplified redox reaction illustrates the possible mechanism:
The higher concentrations of starting materials appear to favor partial autoreduction, likely due to increased local concentration and enrichment of Pt species, which may lead to local reduction through disproportionation or surface-mediated pathways during mixing and drying. Furthermore, the minor contribution of the photoreduction of Pt4+ cannot be excluded.
There is a direct correlation between the degree of thermal treatment and Raman spectral features (
Figure 11) and XPS oxygen speciation (
Figure 9, right panel). These two spectroscopies provide complementary insight into the defect structure of metal oxide materials. SP1a exhibits a broad, weak A
1g peak near 630 cm
−1, indicating low crystallinity and high structural disorder, consistent with a SnO
2 phase that has not been thermally treated. The presence of additional low-frequency scattering likely reflects a high density of surface defects and oxygen vacancies formed during ion exchange and precipitation without a subsequent ordering step. This was confirmed by the XPS spectrum measured around the O1s core levels, as the sample has a significant amount of water (H
2O) and surface hydroxyl groups (O–H). SP1b shows a better defined A
1g Raman mode with reduced baseline noise, which is expected from a hydrothermal step facilitating the restructuring of the SnO
2 framework [
25]. Improved structural order with maintained surface reactivity is evident from the increase in lattice oxygen (O
L) and the slight decrease in H
2O and O–H in the XPS spectrum. SP1c displays the sharpest and most intense A
1g peak, confirming that annealing at 600 °C promotes high crystallinity and the removal of residual hydroxyl groups.
While both SnB and SnC were subjected to hydrothermal treatment, the additional thermal annealing at 600 °C in SnC promotes grain growth and surface tension relaxation, which may reduce the efficiency of Raman scattering for certain vibrational modes and broaden the bands due to increased phonon–phonon interactions or reduced phonon confinement. In contrast, SnB has a relatively high degree of crystallinity with moderate crystallite size and higher defect density, e.g., oxygen vacancies and hydroxyl groups on the surface, which can localize vibrational modes and increase Raman activity. This explains why SnB has sharper and more intense bands, especially for the A1g and B2g modes. The shift in Raman band positions between samples is primarily attributed to differences in crystallite size, lattice strain, and local bonding environment. These shifts are consistent with the phonon confinement model, where smaller crystallites (as in SnA and SnB) lead to slight blue or red shifts in peak positions due to the relaxation of selection rules.
The detection of Sn
2+ and Sn
0 species via XPS (
Figure 9, left panel), based on Sn 3d core level analysis, contrasts with the absence of their signal in the
119Sn Mössbauer spectra, indicating that these species are confined to the surface of the SnO
2 particles. Consequently, the average tin oxidation states reported in
Table 2 primarily reflect the surface composition of SnO
2. The interaction between Pt
2+/Pt
4+ species and the SnO
2 support, a reducible n-type semiconductor, may facilitate electron transfer during the reaction, enabling catalytic function without the need for metallic nanoparticles.
Among the three catalysts, SP1a, prepared without thermal treatment, exhibited the highest specific surface area and the smallest average pore diameter (
Figure 8), as well as the best catalytic activity (
kapp = 1.27 × 10
−2 s
−1) and reusability, as it maintained a conversion of over 84% after ten cycles (
Figure 13 and
Figure 14). The observed catalytic activity despite the absence of metallic Pt
0 underlines the functionality of the oxidized Pt species. The Pt
2+ and Pt
4+ species anchored to SnO
2 appear to facilitate electron transfer from the reducing agent (NaBH
4) to 4-NP via the support. This is likely due to the reducible, n-type semiconducting nature of SnO
2, which can mediate charge transfer through oxygen vacancies or surface hydroxyls. The results are consistent with previous work [
4,
5], emphasizing the catalytic importance of oxidized Pt species, particularly Pt
2+, in redox reactions. The increased activity of SP1a suggests that the combination of high surface hydroxylation, a defect-rich structure, and molecularly dispersed Pt species creates an ideal environment for efficient catalysis. While the SP1c catalyst is the slowest of the three catalysts, it exhibits exceptional structural stability and robustness, making it well suited for operation under extreme conditions. The improved crystallinity of the SnO
2 support combined with the thermal removal of the hydroxyl groups on the surface and the improved lattice ordering contribute to its resistance to structural degradation at high temperatures. In addition, the strong metal–support interaction between oxidized Pt species (Pt
2+/Pt
4+) and the crystalline SnO
2 support improves the chemical resistance of the catalyst, allowing it to maintain its catalytic activity even in harsh pH environments and under high-pressure conditions [
36]. These properties make SP1a promising candidate for catalytic processes where longevity and performance stability are required in demanding industrial or environmental environments [
37].
In summary, the combination of surface-sensitive and bulk characterization techniques shows a consistent picture: highly dispersed oxidized Pt species anchored on defect-rich SnO2 supports—especially those synthesized under mild conditions—exhibit strong catalytic activity for the reduction of nitroaromatics. The role of synthesis temperature, support purity (almost chloride-free), and surface hydroxyl content is critical in tuning both structural integrity and catalytic performance. This work not only demonstrates the viability of PtOx species as active sites but also provides a scalable and environmentally friendly synthetic strategy for the development of reusable metal oxide catalysts. These findings offer a promising route to the development of cost-effective and environmentally friendly noble metal catalysts for the reduction of nitroaromatic pollutants.