Blog

Innovative microwave in situ approach for crystallizing TiO2 nanoparticles with enhanced activity in photocatalytic and photovoltaic applications | Scientific Reports

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Scientific Reports volume  14, Article number: 12617 (2024 ) Cite this article Sulfate Titanium Dioxide

Innovative microwave in situ approach for crystallizing TiO2 nanoparticles with enhanced activity in photocatalytic and photovoltaic applications | Scientific Reports

This investigation introduces an innovative approach to microwave-assisted crystallization of titania nanoparticles, leveraging an in situ process to expedite anatase crystallization during microwave treatment. Notably, this technique enables the attainment of crystalline material at temperatures below 100 °C. The physicochemical properties, including crystallinity, morphology, and textural properties, of the synthesized TiO2 nanomaterials show a clear dependence on the microwave crystallization temperature. The presented microwave crystallization methodology is environmentally sustainable, owing to heightened energy efficiency and remarkably brief processing durations. The synthesized TiO2 nanoparticles exhibit significant effectiveness in removing formic acid, confirming their practical utility. The highest efficiency of formic acid photodegradation was demonstrated by the T_200 material, reaching almost 100% efficiency after 30 min of irradiation. Furthermore, these materials find impactful application in dye-sensitized solar cells, illustrating a secondary avenue for the utilization of the synthesized nanomaterials. Photovoltaic characterization of assembled DSSC devices reveals that the T_100 material, synthesized at a higher temperature, exhibits the highest photoconversion efficiency attributed to its outstanding photocurrent density. This study underscores the critical importance of environmental sustainability in the realm of materials science, highlighting that through judicious management of the synthesis method, it becomes feasible to advance towards the creation of multifunctional materials.

In the face of escalating environmental concerns and the imperative to embrace sustainable practices, pursuing innovative technologies becomes increasingly critical1,2. This publication explores a novel avenue of material synthesis, focusing on the crystallization of TiO2 nanoparticles through the pioneering approach of microwave in situ synthesis.

The urgency to transition towards green chemistry is underscored by the adverse effects of escalating pollutant emissions on our environment and the pressing need for clean energy solutions3,4,5. While significant strides are being made in altering top-down approaches to energy production, a complementary bottom-up approach becomes indispensable. Yang et al.6 emphasize the critical need to explore groundbreaking technologies within materials science, particularly nanomaterials exhibiting precisely defined morphology, crystallinity, and textural properties7,8. These materials hold the key to addressing the challenges posed by sustainable development and zero-emission strategies across scientific and industrial domains9,10.

Titanium dioxide emerges as a crucial material in the domain of photocatalysis, owing to its exceptional properties such as robust oxidizing abilities, superhydrophilicity, and chemical stability11,12. The photocatalytic activity of TiO2 hinges on generating charge carriers upon UV light absorption, facilitating the degradation of organic pollutants. However, optimizing efficiency necessitates meticulous control over particle size, surface area, pore structure, and crystalline phase13,14. Therefore, developing performance improvements by adjusting these factors remains the focus of photocatalysis research.

In materials science, the microwave-assisted hydrothermal method has emerged as a pioneering approach for synthesizing TiO2 nanoparticles, presenting a transformative pathway for enhancing their efficiency in photocatalytic applications15,16. Integrating microwaves into the hydrothermal reactor accelerates heating and overcomes the limitations of slow convection and heat conduction associated with traditional methods. Microwave synthesis reduces energy costs and enables rapid reaction mechanisms, resulting in high-yield, pure products in significantly shorter timescales. Dar et al.17 reported that a microwave-based approach achieves ~ 7 nm and ∼ 100–400 nm nanostructures of anatase titania. On the other hand, an important conclusion was presented by Kubiak et al.18, who obtained TiO2 in the microwave process supported by surfactants. The research team confirmed that the obtained materials were characterized by a reduction in the bandgap energy concerning the literature values while maintaining only the anatase structure, which may suggest defects on the TiO2 surface (e.g. formation of Ti3+ species). The application of microwaves was an efficient way to improve powder quality by enhancing crystallinity and decreasing processing times. Notably, the widely used microwave process borrowed from hydrothermal techniques often involves several stages, where heat treatment concludes the technological process19,20. However, microwave heating fundamentally differs from conventional processes such as oil-bath or hydrothermal heating regarding rapid reaction mechanisms, nucleation, and development of the formed material's nanostructure21,22. Therefore, to fully harness the potential of the microwave pathway, an innovative and comprehensive approach to the formation of nanomaterials is necessary.

Presently, an intriguing avenue in advancing the microwave route lies in utilizing the in situ process, a methodology hailed as "future-proof" by Kumar et al.23, enabling the continuous evolution of microwave phenomena. This innovative approach involves introducing a promoter into the reactor, which undergoes decomposition under the influence of rising temperatures during processing, thereby initiating the desired material's microwave-driven precipitation. The scientific literature reveals the successful application of hydrothermal in situ routes for synthesizing ZnO, notably employed as a component in concrete composites with heightened antibacterial activity24. Additionally, researchers have posited that in situ materials obtained through microwave means may serve as a versatile platform for the deposition of lanthanides25. The paramount advantage of in situ pathways is evident in their ability to facilitate one-step reactions, eliminating the need for intermediate product synthesis and preventing particle agglomeration while maintaining an advantageous particle size distribution26,27. These in situ routes have demonstrated remarkable efficiency by promoting crystalline growth and achieving desired crystallinity while also earning praise for their environmentally friendly attributes28,29,30. Significantly, this methodology allows precise control over the size and morphology of the formed materials by adjusting synthesis parameters, such as precursor concentration and the type and quantity of the precipitation agent31.

In this context, the application of an in situ microwave approach emerges as an innovative technique for synthesizing TiO2 NPs. This process involves using TiCl4 and NH4F as precursor materials. This unique synthesis pathway aims to produce multifunctional TiO2 nanoparticles with outstanding photocatalytic and photovoltaic performance. Additionally, our investigation delves into the influence of microwave treatment temperature on various physicochemical parameters. This technique enables the achievement of crystalline material even at temperatures below 100 °C. The TiO2 NPs synthesized through this approach exhibit significant effectiveness in removing formic acid, substantiating their practical applicability. Furthermore, our nanomaterials demonstrate impactful potential in dye-sensitized solar cells (DSSCs), providing an additional avenue for their utilization. This study underscores the paramount importance of environmental sustainability in materials science. It emphasizes that thoughtful management of the synthesis method makes it possible to progress toward developing multifunctional materials.

Titanium(IV) chloride (TiCl4, 97%, Merck, Germany), ammonium fluoride (NH4F, p.a., S Merck, Germany), formic acid (FA, 99%, Merck, Germany), ethylcellulose (p.a., Sigma-Aldrich, USA), α-terpineol (96%, Merck, Germany), 1-propyl-3-methyl-imidazole iodide (97%, Sigma-Aldrich, USA), 4-tert-butylpiridine (98%, Sigma-Aldrich, USA), potassium dihydrogen phosphate (p.a., Sigma-Aldrich, USA), acetonitrile (LC–MS ultra-pure, Honeywell, USA), acetic acid (p.a., POCh, Poland), anhydrous ethanol (99.8%, POCh, Poland), iodine (p., POCh, Poland), N3 dye (p., Ruthenizer 535, Solaronix, Switzerland), hexachloroplatinic acid hexahydrate (99.9%, Merck, Germany), hydrofluoric acid (40%, Suprapur, Merck, Germany), guanidine thiocyanate (97%, Fluka, Switzerland), ionomeric foil Meltonix (Solaronix, Switzerland), were used. All reagents were of analytical grade and used without any further purification. The water used in all experiments was deionized.

The synthesis of TiO2 NPs was achieved through the in situ microwave route. The precursor was titanium(IV) chloride, the solution of which was prepared in distilled water in an ice-water bath, according to the procedure described previously. The concentration of titanium(IV) chloride was adjusted to 1%. Subsequently, 1 g of ammonium fluoride was introduced into the 100 cm3 TiCl4 solution and stirred using an IKA magnetic stirrer (IKA Werke GmbH, Germany). The resulting clear solution was then transferred to a microwave reactor (CEM, Discover 2.0, USA) and heated with a maximum power of 300 W to temperatures ranging from 60 to 200 °C (in 20 °C increments). The heating process was conducted at various temperatures, where samples, e.g., T_60, were heated to 60 °C, and T_80 was heated to 80 °C. The solution was promptly cooled down to room temperature upon reaching the set temperature. Subsequent nanomaterials were obtained following the procedure above. Finally, the resulting nanomaterials were filtered, thoroughly washed, and dried at 60 °C for 6 h.

The crystal structure analysis was done using a D8 Advance diffractometer (Bruker, Germany). The measured material was placed in a measuring cuvette and subjected to analysis with CuKα radiation (λ = 1.5406 Å) within the 2θ range of 20°–80° at a scan speed of 1°/min. XRD data were analyzed using Rietveld refinement method and applying the Fullprof program32. The crystallite size of the synthesized materials was determined using Williamson-Hall method33, represented by the equation:

where β—is the line broadening at half the maximum intensity (FWHM), θ—is the Bragg angle, K—is a shape factor (0.891), D—size of crystallinity, λ—is the X-ray wavelength, and ε—lattice strain.

For the characterization of the porous structure, including BET surface area, pore volume, and pore size, a Quantachrome Autosorb iQ surface characterization analyzer (Quantachrome, USA) was employed. The Brunauer–Emmett–Teller (BET) method, based on low-temperature N2 sorption, was used for analysis. Surface area determination utilized the multipoint BET method with adsorption data in a relative pressure p/p0 range of 0.05–0.30.

Transmission electron microscopy (TEM) analysis was conducted using the FEI TECNAI G2 F20 electron microscope operating at 200 kV, featuring a Gatan CCD camera for high-resolution imaging. Specimens for high-resolution TEM (HR-TEM) analysis were prepared by sonication of a small amount of material in 2-propanol, followed by suspension onto a copper grid with a holey carbon film. Micrographs were captured after solvent evaporation, ensuring a comprehensive and statistically representative mapping of the studied materials.

X-ray Photoelectron Spectroscopy (XPS) experiments were recorded using a Specs UHV spectrometer (SPECS, Germany) with a charge neutralizer. The C 1s peak at 284.8 eV was a reference for rectifying the binding energies.

The light-absorption properties were measured through diffuse reflectance spectroscopy (DRS) in the 200–800 nm range. The bandgap energy of the samples was calculated from the (F(R)·E)0.5 against the E graph, where E is photon energy and F(R) is the Kubelka–Munk function proportional to the radiation's absorption. Thermo Scientific Evolution 220 spectrophotometer (Waltham, USA), equipped with a PIN-757 integrating sphere using BaSO4 as a reference, was employed for these measurements.

Photoluminescence (PL) measurements were conducted using a spectrofluorometer (Fluorolog version-3 Horiba, Japan) with a 450 W high-pressure xenon arc lamp as an excitation source. Photoluminescence excitation (λ = 320 nm) and emission spectra were acquired at room temperature with a spectral resolution of 2 nm and a slit width of 2 mm.

In the standard procedure, a magnetically stirred 60 mL cylindrical quartz reactor was utilized and placed within a custom-made housing consisting of a black box mounted on an optical bench. The light source for irradiation was a convenient LED photosystem previously detailed elsewhere34, emitting ultraviolet light within the range of 360–410 nm. The entire setup was maintained at ambient temperature through a continuous air stream. All aqueous suspensions subjected to irradiation contained photocatalysts at a concentration of 0.1 g L−1. The photocatalysts, dispersed in pure water, underwent preliminary sonication in a Eurosonic, Model 22, apparatus for 30 min. Subsequently, the required volume of formic acid solution was added to achieve an initial formic acid concentration of 50 ppm. The suspension was then magnetically stirred in the dark for 15 min to reach the adsorption equilibrium of the substrate on the photocatalyst's surface before initiating irradiation. Stirring was maintained throughout the experimental runs. The lamp was switched on, and at various time intervals during the runs, 2 mL samples of the suspension were withdrawn from the reactor and centrifuged using an EBA-20 Hettich centrifuge (Hettich Group, Germany). The supernatant was analyzed for residual FA content through ion chromatography with conductivity detection, utilizing a Metrohm 761 Compact IC instrument (Metrohm AG, Switzerland), after calibration for formate ion concentration in the 0–50 ppm range.

Formic acid (FA) kinetics were assessed using a pseudo-first-order kinetic model. This model postulates that the degradation rate is directly proportional to the surface coverage (θ) of FA, expressed as follows:

Here, k represents the reaction rate constant, 'θ' denotes the surface coverage by FA, K and Ks are the adsorption coefficients for FA and water, respectively, C0 stands for the initial concentration of formic acid, and Cs represents the concentration of water. The concentration of water, Cs remains nearly constant and is significantly higher than the concentration of formic acid. Consequently, we can express Eq. (3) in the following form:

In Eq. (3), k1 signifies the first-order rate constant, and t is the time of irradiation.

Working electrodes have been prepared by spreading titania paste on the FTO substrate using the doctor blade technique. The pastes containing investigated materials were obtained using the well-known literature method of mixing the titania powders with ethanol, alpha-terpineol, acetic acid, and ethylcellulose in the mass ratio 1:9.8:3.3:0.18:0.5, respectively35. After the paste deposition, working electrodes have been annealed at 450 °C in the air for 2 h. Cooled-down electrodes have been immersed in a 40 mM TiCl4 water solution for 1 h at 70 °C to prepare an electron recombination-preventing layer. Afterward, the electrodes were washed with distilled water and ethanol, dried, and again annealed at 450 °C for 1 h in the air. After cooling to approximately 80 °C, electrodes were immersed in 10–4 M N3 dye solution in absolute ethanol and left overnight. Counter electrodes have been prepared using our standard procedure by spreading H2PtCl6 ethanolic solution on FTO substrates followed by annealing at 450 °C in the air for 1 h36. Photovoltaic devices were assembled in the sandwich mode using a 25 µm thick ionomeric hot-melted foil as a spacer and sealing agent. Afterward, the liquid electrolyte, containing 0.6 M 1-propyl-3-methyl-imidazolium iodide, 0.03 M iodine, 0.1 M guanidine thiocyanate, and 0.5 M 4-tert-butylpiridine in acetonitrile, have been injected into the cell by two holes predrilled in the counter electrode and the final device seal have been made using an ionomeric foil and microscope cover slide. The typical active area of the investigated cells was 0.125 cm2. Additional working electrodes with an active area of about 2.5 cm2 have been prepared for the XRD and DRS measurements.

Photovoltaic parameters (I–V curves) measurements have been performed under AM 1.5G illumination using the Abet Sun 2000 solar simulator (Abet Technologies, USA). The light intensity was calibrated to 1 sun using a ReRa Solutions reference cell with a KG5 filter (ReRa Solutions, The Netherlands). Electrochemical impedance spectroscopy (EIS) has been performed in the 1 sun illumination over a frequency range from 0.1 Hz to 100 kHz at VOC bias potential, with sinusoidal VAC = 10 mV. The I–V curves and EIS spectra have been registered on a Gamry Interface 1010 E potentiostat–galvanostat (Gamry Instruments, USA). EIS data have been fitted using ZView 3.2 software. Incident photon to current conversion efficiency (IPCE) was investigated on Bentham PVE300 EQE/IPCE apparatus (Bentham Instruments Ltd., UK) using a chopped mode with an external light bias.

In the initial phase of the physicochemical analysis, X-ray diffraction (XRD) was conducted to determine the capability of proposed microwave in situ pathways for crystal growth and achieving the desired crystallinity. The resulting XRD patterns are illustrated in Fig. 1. The distinct peaks identified in the XRD patterns of the TiO2 nanomaterials, occurring at 2θ values of 25.28°, 36.9°, 37.8°, 47.9°, 53.8°, 55°, 62.6°, 68.7°, 70°, and 75.05°, closely correspond to the characteristic peaks of the anatase crystalline structure (as per crystallographic database card no. 9009086) in space group I41/amd (no. 141)37,38,39. Furthermore, the XRD analysis disclosed the presence of crystalline planes, specifically (101), (103), (004), (200), (105), (211), (213), (116), (220), and (215) in the analyzed nanomaterials40. It should be emphasized, however, that in the case of materials obtained within the temperature range of 60–100 °C, additional diffraction peaks were observed, suggesting the presence of other crystalline phases, particularly titanium fluoride, with its characteristic peak observed at 2θ = 28.6°41,42. Nevertheless, irrespective of the temperature applied in the microwave process, no discernible peaks characteristic of other titanium dioxide phases, such as rutile and brookite, were identified. This observation underscores that the proposed methodology exclusively promotes the growth of crystalline anatase.

Results of (a) XRD patterns, and (b) low-temperature N2 sorption isotherms for TiO2 NPs fabricated by microwave in situ route.

Rietveld refinement was performed to obtain a thorough insight into the crystallinity of the acquired materials, and the corresponding results are summarized in Figure S1 of the Supplementary Materials. The determined lattice parameters and average size of crystallites are detailed in Table S1 (see Supplementary Materials). Figure S2 compares the anatase peak location at 2θ of about 25°. Analysis of the collected data reveals a noticeable shift in the anatase peak towards larger values with the increasing microwave processing temperature. This shift likely indicates a reduction in the lattice distance (d), causing the peak to shift towards higher 2θ values. As per the Bragg equation, the decrease in the distance between crystal planes results in an increased 2θ angle, implying a contraction in the crystal dimensions along a specific direction43. This observation is supported by the determined parameters of the anatase crystal lattice. However, it is crucial to emphasize that the shift in the peak towards higher 2θ values may also be attributed to stresses within the crystal lattice. Stress-induced changes can affect the separation between crystal planes, thereby influencing X-ray diffraction patterns. In the analyzed samples, stress–strain (ε) values were quantified, revealing a proportional increase with the rise in microwave processing temperature44. The above underscores the dynamic nature of the in situ microwave process, leading to heightened stresses within the crystal structure as the process temperature increases. Moreover, the analysis of the determined parameters indicates that an increase in microwave processing temperature leads to crystalline growth, a phenomenon we have previously validated in our earlier studies. However, the obtained XRD data show that the in situ microwave process causes the crystallization of anatase with surface defects, linked to the contraction of the crystal lattice in the c-direction and the presence of stress–strain.

The N2 isotherms depicted in Fig. 1b exhibit characteristic features of type IV shape, indicating reversible mono- and multilayer adsorption in the lower range of p/p0, followed by a hysteresis loop at higher pressure values45. Across all synthesized materials, a discernible reduction in the adsorbed volume is evident with an increase in microwave temperature, particularly in the lower section of the isotherm, aligning with the observed improvement in material crystallinity. Furthermore, two specific types of hysteresis loops were identified. Materials synthesized at 60–80 °C and 160–200 °C displayed an H3 hysteresis loop, indicating aggregates of plate-like particles giving rise to slit-shaped pores46,47. In contrast, materials obtained at temperatures of 100–140 °C exhibited an H1 hysteresis loop, often associated with porous materials consisting of agglomerates or compacts of approximately uniform spheres in a fairly regular array, resulting in narrow pore size distributions45. These distinctions are closely tied to the crystallinity of the materials. At a temperature around 100 °C, the transformation of amorphous TiO2 or titanium fluoride into nanocrystalline anatase occurs, forming agglomerates composed of TiO2 nanoparticles. This transformation significantly contributes to the observed augmentation in specific surface area and total pore volume. With a further increase in temperature (above 140 °C), notable crystalline growth occurs, leading to the formation of tetragonal TiO2 nanoparticles. This results in decreased specific surface area and the emergence of slit-shaped pores between nanocrystalline anatase particles. A comprehensive set of sorption measurements, encompassing calculated BET surface area, total pore volume, and average pore size, is presented in Table S1 in Supplementary Materials.

For a comprehensive investigation of TiO2 NPs, both TEM, HRTEM and FFT analysis were carried out. The outcomes are illustrated in Fig. 2.

The TEM, HR-TEM, and FFT images for: (a) T_60, (b) T_100, (c) T_140, (d) T_200.

In the case of the T_60 material (Fig. 2a), a distinct structure is evident in the high-contrast image, indicating a high contribution of the amorphous phase. High-resolution imaging shows crystallographic spacings indicative of partial crystallization, consistent with XRD data. Additionally, only single spots are observed in the obtained FFT images, which may indicate poor crystallinity of the sample and is consistent with previous reports by Bielan et al.48 and Dozzi et al.49,50. Increasing the microwave treatment temperature to 100 °C (sample T_100, Fig. 2b) results in an elevated share of the crystalline phase. Additionally, single crystalline particles, including cubic ones, are observable in lower magnification images. High-resolution images confirm the presence of lattice spacings characteristic of anatase. The number of spots in the FFT image indicates that the rapid microwave crystallization resulted in the growth of TiO2 crystals in different directions, which is confirmed by the varied morphology forms and wide particle size distribution51. For sample T_140 (Fig. 2c), only nanocrystalline particles are observed, without any amorphous areas, indicating that 140 °C is the temperature limit, allowing for nearly complete crystallization of anatase in the in situ microwave procedure. In the detailed examination of the T_140 sample, an ordered morphology was revealed, characterized by the presence of octahedral anatase TiO2 bipyramids with well-exposed {001}/{101} facets, aligning with previous reports by Kubiak et al.52. This morphological evolution is attributed primarily to two factors: termination by fluoride ions, promoting crystal growth in specific directions, and the extension of the microwave process. Fluoride ion termination was found to stabilize (001) surfaces over (101) due to the balancing of O–O/O–F repulsions and Ti–O/Ti–F attractions, stabilizing Ti and O atoms on the surface. Further raising the microwave treatment temperature to 200 °C (sample T_200, Fig. 2d) leads to crystal growth, increasing the average particle size to 40–60 nm. Additionally, a visible change in the shape of the obtained nanoparticles is noted. In contrast to the T_140 material, the nanoparticles obtained for T_200 are characterized by a predominance of octahedral shape. These considerations clearly confirm previous XRD results, demonstrating that the in situ microwave processing temperature decisively influences the morphology of the obtained materials.

X-ray Photoelectron Spectroscopy analysis was employed to investigate the surface composition and oxidation states of elements in the TiO2 nanoparticles. The examination of XPS data substantiated the presence of three crucial elements integral to the synthesis process, namely Ti, O, and F. The high-resolution XPS spectra are presented in Fig. 3, and Table S2 in the Supplementary Materials provides a comprehensive summary of the atomic percentages of each element. Noteworthy is the observed variation in fluoride atomic content across the analyzed systems, demonstrating a decline as the temperature of microwave treatment for the samples increases.

XPS spectra of Ti 2p, O 1s, and F 1s for TiO2 NPs.

The Ti 2p region reveals two distinct peaks at 459.3 eV and 465 eV, attributed to Ti 2p3/2 and Ti 2p1/2, respectively. The 5.7 eV separation between Ti 2p1/2 and Ti 2p3/2 signifies a standard Ti4+ state53,54,55. In the O 1s region, TiO2 spectra exhibit a double peak that can be resolved into two oxygen atom types: lattice oxygen (530.5 eV) and adsorbed surface −OH groups (532.1 eV), with a spacing consistent with literature values, such as the 1.5 eV difference between lattice O2− and −OH56,57. The F 1s region displays a peak at 685 eV, corresponding to F atoms adsorbed on TiO2. This F 1s peak originates from surface fluoride formed by ligand exchange between F− and surface hydroxyl groups on TiO2. The F 1s peak is precisely located at 685 eV, and no evidence of F− ions in the lattice was detected, aligning with existing literature14,58.

Analysis of DRS data reveals the presence of a well-established absorption band in the 250–400 nm range for the synthesized TiO2 nanomaterials (Fig. 4a). It is noteworthy that an increase in the microwave processing temperature leads to a gradual enhancement of absorbance, particularly towards longer wavelengths. This phenomenon can be attributed to alterations in the quantity of O 2p bonding orbitals, specifically those within the Ti3O plane51. The energy bandgap of the synthesized samples was determined using the Kubelka–Munk equation. Tauc plots for selected materials are shown in Figure S3 in the Supplementary Materials. An absorption peak at around 3.0–3.1 eV is consistently observed across all samples, denoting electron transfer from the valence band to the conductivity band in the pure anatase structure, aligning with existing literature. It is essential to highlight that introducing fluorine ions during the synthesis of TiO2 nanoparticles often leads to the formation of surface Ti–F bonds, absorbing light around 500 nm, as reported by Dudziak et al.59. However, this effect is not observed in our materials, indicating the absence of Ti-F bonds on the surface of the obtained anatase nanocrystals. Furthermore, the lack of a band gap shift is correlated with the absence of surface defects, as confirmed by the XPS results.

The (a) DRS and (b) emission spectra for TiO2 NPs.

The observed photoluminescence emission (Fig. 4b) primarily results from the recombination of excited electrons and holes, offering crucial insights into carrier separation efficiency in photo-induced processes. Across all examined materials, a consistently noted broad luminescence band around 450 nm was observed. In contrast, the typical literature on TiO2 photoluminescence spectra highlights two primary emission peaks at approximately 396 and 462 nm, corresponding to energies of 3.13 and 2.68 eV, respectively40,60. However, broader luminescence bands encompassing these narrow peaks have also been documented, as reported by Meng et al.61. These peaks are attributed to bandgap transition emission with light energy roughly equal to the anatase bandgap energy (387.5 nm) and an emission signal arising from the charge-transfer transition from Ti3+ to oxygen anions in a TiO68− complex62,63.

Notably, the obtained PL spectra can be categorized into two groups based on the synthesis temperature. In the microwave processing temperature range of 60–120 °C, lower luminescence efficiency and a peak shift towards longer wavelengths, peaking at 430 nm, were observed. Conversely, for materials synthesized at temperatures ranging from 140 to 200 °C, an increase in luminescence efficiency was evident, with a peak at approximately 420 nm. These findings suggest that elevating the in situ microwave processing temperature enhances luminescence and increases charge carrier recombination.

One crucial aspect of the research was assessing the photooxidation capabilities of the TiO2 NPs in the formic acid degradation. This evaluation was conducted utilizing a convenient LED photoreactor, and the findings are illustrated in Fig. 5.

(a) Photocatalytic degradation efficiency and (b) first-order plots for the photooxidation of formic acid.

The influence of microwave treatment temperature on the photocatalytic activity of TiO2 is evident (Fig. 5a). The T_60 sample exhibited the lowest degradation efficiency due to its reduced crystallinity, displaying only a 25% maximum efficiency within the 60–140 °C temperature range. However, increasing the temperature beyond 140 °C significantly enhanced the pollutant removal efficiency, with the TiO2_200 sample achieving the highest degradation rate of 99% after 30 min. Interestingly, the T_200 material showed the highest charge carrier recombination, as indicated by photoluminescence data, although this did not negatively impact formic acid degradation. The superior performance of the TiO2_200 sample is primarily attributed to its highly crystalline anatase structure with minimal surface defects, as revealed by X-ray diffraction (XRD) analysis. These minimal defects reduce electron–hole recombination, thereby enhancing photocatalytic efficiency. Additionally, the morphology of this sample, characterized by uniform octahedral shapes with well-defined facets, provides numerous active sites for the photocatalytic reaction. TEM analyses further supported this beneficial morphology, showing a predominant crystalline phase with larger particle sizes that promote improved charge carrier dynamics. Moreover, DRS measurements indicated a slight blue shift in the absorption edge for the TiO2_200 sample, which enhances UV light utilization and potentially extends the light absorption spectrum. Collectively, these physicochemical properties contribute to the enhanced photocatalytic activity of the TiO2_200 sample, clearly demonstrating the impact of synthesis conditions on material properties.

The sorption process was intentionally excluded from the calculations to establish the rate constant parameter, as shown in Fig. 5b. Using Eqs. (2) and (3), the apparent values of parameter k1 for each catalyst were determined by analyzing the slope of the ln Ct/C0 vs. time plot. The computed values are detailed in Table S3 and available in the Supplementary Materials for reference and comparison. The outcomes underscore the significant impact of microwave treatment temperature on the photooxidative capacity of the synthesized materials. Among the studied materials, the T_200 material exhibited the highest reaction rate constant at 0.095 min−1. Despite increased charge carrier recombination, as evident in the photoluminescence spectra, the reaction rate constantly rose with elevated microwave treatment temperatures. The above further proves that the formic acid degradation process with TiO2 nanoparticles is not solely contingent on charge carrier recombination but is intricately linked to physicochemical parameters such as crystal structure and morphology.

Nanostructured titanium dioxide's synthesis methods and physicochemical properties significantly influence its photocatalytic performance. For instance, research conducted by Dorosheva et al.65 focused on preparing nanostructured TiO2 via a sol–gel process, resulting in photocatalytically active materials for organic reactions under visible light exposure. The photocatalytic activity of these TiO2 nanoparticles was notably enhanced by controlling the crystallite size, ranging from 10 to 85 nm. This size modification was achieved through annealing processes, inducing the transformation of the material's phase from amorphous to anatase under treatment in a hydrogen atmosphere at specific temperatures. Similarly, the study conducted by Parida et al.66 on zinc oxide revealed significant influences of physicochemical properties such as surface area, surface acidity, and crystallite sizes of ZnO particles on photocatalytic activity towards the oxidation of 4-nitrophenol and the reduction of Cr(VI). ZnO samples prepared by microwave irradiation and calcined at 300 °C exhibited the highest surface area, acid sites, and the lowest crystallite sizes, leading to superior activity in photocatalytic reactions, this underscores the pivotal role of nanoparticle morphology in influencing photocatalytic efficiency.

These findings underscore the crucial role of the morphological characteristics of photocatalysts in determining their efficiency67,68. By adjusting synthesis conditions and post-synthesis treatments, researchers can tailor nanoparticles' crystal structure, surface properties, and morphology to optimize their photocatalytic performance. The correlation between nanoparticle morphology and photocatalytic activity highlights the importance of detailed physicochemical characterization in developing efficient photocatalytic materials.

Similar conclusions were drawn based on obtained SEM images (see Figure S4 in the Supplementary Materials). An increase in microwave processing temperature resulted in a more uniform particle size distribution, facilitating the production of material with uniform morphology, which critically impacted the efficiency achieved in the formic acid degradation process.

Table 1 shows the current state of knowledge in the elimination of formic acid.

The T_200 and T_180 photocatalysts were chosen for the reusability test. Five consecutive cycles of formic acid photodegradation were conducted to assess photocatalytic reusability, as depicted in Figure S5 in Supplementary Materials. After each cycle, filtration separated the photocatalyst from the reaction suspension. Subsequently, the separated photocatalyst was reused without any further treatment. The photocatalytic degradation efficiency decreased by approximately 5% after the fifth cycle compared to the first. This slight decrease in activity after each irradiation cycle could be attributed to photocatalyst losses during the separation process.

Studied materials have also been tested as semiconductor layers in DSSCs. It should be emphasized that during the working electrode preparation, the materials were annealed at 450 °C (the detailed procedures described in the Experimental section of this manuscript), significantly influencing their structure. However, their properties were still strongly related to the nature of the bare samples. The XRD analysis of prepared electrodes, see Figure S6 in Supplementary materials, shows that all the materials present well-formed anatase structures with no post-synthetic fluorine moieties residues. The additional reflexes observed on each diffractogram originated from the SnO (JPDS no. 2-1337), the conductive layer of FTO substrates. The reduced intensities of the main anatase reflexes observed for the samples synthesized at lower temperatures compared to those obtained at higher ones prove the influence of primary preparation temperature on the material's properties even after high-temperature post-treatment. The electron properties of the investigated materials were also changed after annealing during the electrode preparation. As the main DRS spectra shape has been preserved (see Figure S7 in Supplementary materials), the hypochromic effect of the absorption edges of particular materials is observed (see Table 2). Still, higher Eg values are observed for the materials prepared at lower synthesis temperatures. Several authors have previously described such effects—the preservation of some properties tendencies of the TiO2 materials originally prepared in different temperatures75,76. This type of behavior may be understood as a kind of synthesis-conditions-fingerprint of the material. The above observations lead to the expectation of significant differences in investigated materials' photovoltaic parameters.

Assembled DSSC devices have been characterized with J–V curve measurements under simulated solar light, and the results are presented in Fig. 6 and Table 2. The performances of the DSSCs present somewhat different tendencies than those observed in our photocatalytic results. Surprisingly, the best photoconversion efficiency (η) has been observed for the T_100 device. The materials prepared in lower temperatures, T_60 and T_80, were the least efficient as electrode materials in DSSCs, similar to photocatalysis. Moreover, following our photocatalytic results, the materials obtained at 140 °C and above were more efficient. The more profound analysis of the photovoltaic parameters shows that the reason for the T-100 cell's best efficiency is the outstanding photocurrent density (JSC) of 12.2 mA/cm2, 10% better than the second-best registered for T_160 material (11.0 mA/cm2) and over three times better than the worst result registered for T_60 cell (3.96 mA/cm2). Such results suggest that the T_100 working electrodes have relatively good electron transportation properties and/or poor electron recombination process efficiency. The ease in electron transportation may also be the reason for such good T-100 cell efficiency, even though the open circuit potential (VOC) value, which is the driving force of electron transportation in the device77, is the lowest among investigated cells. At this stage, the fill factor (FF) values, mainly understood as energy loss related to inherent resistance in the photovoltaic device78,79, seem worth considering. In this case, the FF values do not vary as much as the other photovoltaic parameters of the presented cells, but still, the result registered for the T_100 cell is one of the best, yielding only the one obtained for the T_160 cell. The relatively high FF value may suggest that the internal resistances in the T_100 cell are most probably optimal for good electron transportation and to avoid the recombination effects.

J–V curves registered for DSSCs studied.

Electrochemical impedance spectroscopic measurements were performed for deeper insight into the resistance properties of investigated cells. The Nyquist plots obtained for investigated cells are presented in Figure S8 in Supplementary materials, and the calculated values of the resistances (R1, R2, and R3, obtained by fitting the experimental results) and estimated electron lifetimes are collected in Table S4 in Supplementary materials. As R1 and R2 values, which are the serial resistance of the FTO substrate or wires connection and the charge transfer resistance at the counter electrode/electrolyte interface, respectively, are mainly considered to have no significant influence on DSSCs performance80,81, we focus on the R3 value interpretation. The R3 value, the second semicircle on the Nyquist plot, represents resistances in the TiO2/dye/electrolyte interface and directly influences the cell performance. One may see that the R3 values are the almost exact representation of the JSC values according to dependence; the lower the R3, the higher the JSC. This observation strongly supports our abovementioned hypotheses that the ease in electron transportation through T_100 electrode material is the most probable reason for its outstanding performance. On the other hand, the analysis of estimated electron lifetimes (τ) suggests that the stability of the electrons injected into the T_100 semiconductor is not as unusual when compared to the other materials tested. The latter observation may suggest that the lower efficiency of PL does not simply translate to the cell performance.

The IPCE measurements were performed to determine the efficiency of the conversion incident photon into the cell's internal photocurrent, and the results obtained are presented in Figure S9 in Supplementary Materials and Table 2. In short, the shape of the IPCE curves approximately reflects the sensitizer's (N3 dye) absorption spectrum. IPCE values determined about the N3 dye adsorption maximum (~ 540 nm) refer to the JSC values supporting the above discussion about the crucial role of the ease of electron injection and transportation in the performance of presented cells81,82.

In this study, TiO2 nanomaterials were synthesized using an innovative microwave in situ method, presenting a novel approach for their preparation. The investigation highlighted the significant impact of microwave treatment temperature on key physicochemical properties, such as crystal structure, morphology, and surface area development. Importantly, higher microwave processing temperatures were observed to stimulate crystallite growth, as evidenced by HR-TEM images, thereby influencing the content of the crystalline phase. The precise control over material properties achieved through in situ microwave processing temperature makes the resulting TiO2 nanoparticles suitable for diverse applications.

Notably, TiO2 nanomaterials synthesized at elevated microwave treatment temperatures exhibited improved performance in formic acid photo-oxidation. Despite an increase in luminescence intensity, indicating charge carrier recombination, the photocatalytic activity showed a positive correlation with the rise in microwave treatment temperature. This observation suggests that the photodegradation process of formic acid is not dependent on the rate of charge carrier recombination.

Moreover, the synthesized materials were employed in DSSC, with the T_100 material demonstrating the highest efficiency. This emphasizes the significance of customized material properties in achieving enhanced activity and reduced energy consumption during synthesis by utilizing lower process temperatures. Such an approach represents a crucial step towards incorporating sustainability into materials science, showcasing a conscientious effort to address environmental concerns in material synthesis.

The data that support the findings of this research are available from the corresponding author upon reasonable request.

Wunderling, N. et al. Global warming overshoots increase risks of climate tipping cascades in a network model. Nat. Clim. Chang. 13, 75–82 (2023).

Canadell, J. G. et al. Contributions to accelerating atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of natural sinks. Biol. Sci. 104, 18866–18870 (2007).

Donnelly, C. et al. Impacts of climate change on European hydrology at 1.5, 2 and 3 degrees mean global warming above preindustrial level. Clim. Change 143, 13–26 (2017).

Eckert, E. & Kovalevska, O. Sustainability in the European Union: Analyzing the discourse of the European Green Deal. J. Risk Financ. Manag. 14, 80 (2021).

Huang, X. et al. In-situ synthesis of efficient N-graphyne/Bi/BiOBr photocatalysts for contaminants removal and nitrogen fixation. J. Alloys Compd. 976, 173025 (2024).

Yang, X. et al. Recent advances in photodegradation of antibiotic residues in water. Chem. Eng. J. 405, 126806 (2021).

Article  CAS  PubMed  Google Scholar 

Zheng, Y. et al. Nitrogen-doped graphyne/BiOBr nanocomposites: In-situ sonochemical synthesis and boosted photocatalytic performance. Sep Purif Technol 301, 122062 (2022).

Liu, W. et al. Ball milling synthesis of porous g-C3N4 ultrathin nanosheets functionalized with alkynyl groups for strengthened photocatalytic activity. Sep Purif Technol 282, 120097 (2022).

Sacchi, R., Bauer, C., Cox, B. & Mutel, C. When, where and how can the electrification of passenger cars reduce greenhouse gas emissions?. Renew. Sust. Energ. Rev. 856, 112475 (2023).

Islam, A. et al. Progress in recent sustainable materials for greenhouse gas (NOx and SOx) emission mitigation. Prog. Mater. Sci. 132, 101033 (2023).

Kubiak, A., Siwińska-Ciesielczyk, K. & Jesionowski, T. Titania-based hybrid materials with ZnO, ZrO2 and MoS2: A review.Materials 11, 2295 (2018).

Article  ADS  PubMed  PubMed Central  Google Scholar 

Huang, X. et al. One-step hydrothermal formation of porous N-graphyne decorated TiO2/Ti3C2 composites with enhanced photocatalytic activity. Int. J. Hydrog. Energy 55, 581–591 (2024).

Dozzi, M. V., Montalbano, M., Marra, G., Mino, L. & Selli, E. Effects of anatase TiO2 morphology and surface fluorination on environmentally relevant photocatalytic reduction and oxidation reactions. Mater. Today Chem. 22, 100624 (2021).

Dozzi, M. V. et al. Effects of photodeposited gold vs platinum nanoparticles on N, F-doped TiO2 photoactivity: A time-resolved photoluminescence investigation. J. Phys. Chem. C 122, 14326–14335 (2018).

Tsuji, M., Hashimoto, M., Nishizawa, Y., Kubokawa, M. & Tsuji, T. Microwave-assisted synthesis of metallic nanostructures in solution. Chem. Eur. J. 11, 440–452 (2005).

Article  CAS  PubMed  Google Scholar 

Zhu, Y. J. & Chen, F. Microwave-assisted preparation of inorganic nanostructures in liquid phase. Chem. Rev. 114, 6462–6555 (2014).

Article  CAS  PubMed  Google Scholar 

Dar, M. I., Chandiran, A. K., Grätzel, M., Nazeeruddin, M. K. & Shivashankar, S. A. Controlled synthesis of TiO2 nanoparticles and nanospheres using a microwave assisted approach for their application in dye-sensitized solar cells. J. Mater. Chem. A Mater. 2, 1662–1667 (2014).

Kubiak, A. et al. Synthesis of titanium dioxide via surfactantassisted microwave method for photocatalytic and dye-sensitized solar cells applications. Catalysts 10, 586 (2020).

Lyu, J. et al. Electrochemical performance of hydrothermally synthesized rGO based electrodes. Mater. Today Energy 13, 277–284 (2019).

Pore, O. C., Fulari, A. V., Shejwal, R. V., Fulari, V. J. & Lohar, G. M. Review on recent progress in hydrothermally synthesized MCo2O4/rGO composite for energy storage devices. Chem. Eng. J. 426, 131544 (2021).

Moniz, S. J. A. & Tang, J. Charge transfer and photocatalytic activity in CuO/TiO2 nanoparticle heterojunctions synthesised through a rapid, one-pot, microwave solvothermal route. ChemCatChem 7, 1659–1667 (2015).

Zia, J., Farhat, S. M., Aazam, E. S. & Riaz, U. Highly efficient degradation of metronidazole drug using CaFe2O4/PNA nanohybrids as metal-organic catalysts under microwave irradiation. Environ. Sci. Pollut. 28, 4125–4135 (2021).

Kumar, A., Choudhary, P., Kumar, A., Camargo, P. H. C. & Krishnan, V. Recent advances in plasmonic photocatalysis based on TiO2 and noble metal nanoparticles for energy conversion, environmental remediation, and organic synthesis. Small 18, 2101638 (2022).

Klapiszewska, I., Kubiak, A., Parus, A., Janczarek, M. & Ślosarczyk, A. The in situ hydrothermal and microwave syntheses of zinc oxides for functional cement composites. Materials 15, 1069 (2022).

Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

Kubiak, A. et al. Design and microwave-assisted synthesis of TiO2-lanthanides systems and evaluation of photocatalytic activity under UV-LED light irradiation. Catalysts 12, 8 (2022).

Wu, Q. et al. Catalytic dehydration of carbohydrates on in situ exfoliatable layered niobic acid in an aqueous system under microwave irradiation. ChemSusChem 6, 820–825 (2013).

Article  CAS  PubMed  Google Scholar 

Ling, C. et al. Enhanced removal of sulfamethoxazole by a novel composite of TiO2 nanocrystals in situ wrapped-Bi2O4 microrods under simulated solar irradiation. Chem. Eng. J. 384, 123278 (2020).

Yu, J., Tao, H. & Cheng, B. In situ monitoring of heterogeneous catalytic reactions. ChemPhysChem 11, 1617–1618 (2010).

Article  CAS  PubMed  Google Scholar 

Liu, S. J., Han, S. D., Zhao, J. P., Xu, J. & Bu, X. H. In-situ synthesis of molecular magnetorefrigerant materials. Coord. Chem. Rev. 394, 39–52 (2019).

Zhao, H., Qu, Z. R., Ye, H. Y. & Xiong, R. G. In situ hydrothermal synthesis of tetrazole coordination polymers with interesting physical properties. Chem. Soc. Rev. 37, 84–100 (2008).

Zhang, H. et al. A facile synthesis of novel amorphous TiO2 nanorods decorated rgo hybrid composites with wide band microwave absorption. Nanomaterials 10, 1–17 (2020).

Rodríguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Physica B Condens. Matter. 192, 55–69 (1993).

Williamson, G. K. & Hall, W. H. X-ray line broadening from filed aluminium and wolfram. Acta Metallurgica 1, 22–31 (1953).

Kubiak, A., Fuks, H., Frankowski, M., Szymczyk, A. & Cegłowski, M. Removal of environmentally toxic sulfamethoxazole waste using TiO2-graphite systems and a tailor-made LED photoreactor: Unraveling the role of spectra-matching. Appl. Surf. Sci. 638, 158089 (2023).

Bartkowiak, A. et al. The importance of anchoring ligands of binuclear sensitizers on electron transfer processes and photovoltaic action in dye-sensitized solar cells. Sci. Rep. 13, 16808 (2023).

Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

Zalas, M. & Jelak, K. Optimization of platinum precursor concentration for new, fast and simple fabrication method of counter electrode for DSSC application. Optik 206, 164314 (2020).

Article  ADS  CAS  Google Scholar 

Dai, Y., Cobley, C. M., Zeng, J., Sun, Y. & Xia, Y. Synthesis of anatase TiO2 nanocrystals with exposed 001 facets. Nano Lett. 9, 2455–2459 (2009).

Article  ADS  CAS  PubMed  Google Scholar 

Yang, H. G. et al. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 453, 638–641 (2008).

Article  ADS  CAS  PubMed  Google Scholar 

Jiang, X. et al. Anatase and rutile in evonik aeroxide P25: Heterojunctioned or individual nanoparticles?. Catal. Today 300, 12–17 (2018).

Radha, E., Komaraiah, D., Sayanna, R. & Sivakumar, J. Photoluminescence and photocatalytic activity of rare earth ions doped anatase TiO2 thin films. J. Lumin. 244, 118727 (2022).

Lee, T. Y., Lee, C. Y. & Chiu, H. T. Enhanced photocatalysis from truncated octahedral bipyramids of anatase TiO2 with exposed 001}/{101 facets. ACS Omega 3, 10225–10232 (2018).

Article  CAS  PubMed  PubMed Central  Google Scholar 

Cheng, X., Yu, X., Xing, Z. & Wan, J. Enhanced photocatalytic activity of nitrogen doped TiO2 anatase nano-particle under simulated sunlight Iirradiation. Energy Procedia 16, 598–605 (2012).

Vorontsov, A. V. & Tsybulya, S. V. Influence of nanoparticles size on XRD patterns for small monodisperse nanoparticles of Cu0 and TiO2 anatase. Ind. Eng. Chem. Res. 57, 2526–2536 (2018).

Zhao, X. et al. Shape- and size-controlled synthesis of uniform anatase TiO2 nanocuboids enclosed by active 100 and 001 facets. Adv. Funct. Mater. 21, 3554–3563 (2011).

Sing, K. S. et al. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity. Pure Appl. Chem. 57, 603–619 (1985).

Schneider, P. Adsorption isotherms of microporous-mesoporous solids revisited. Appl. Catal. A Gen. 129, 157–165 (1995).

Muttakin, M., Mitra, S., Thu, K., Ito, K. & Saha, B. B. Theoretical framework to evaluate minimum desorption temperature for IUPAC classified adsorption isotherms. Int. J. Heat Mass. Transf. 122, 795–805 (2018).

Bielan, Z. et al. Mono- and bimetallic (Pt/Cu) titanium(IV) oxide core–shell photocatalysts with UV/Vis light activity and magnetic separability. Catal. Today 361, 198–209 (2021).

Dozzi, M. V., Saccomanni, A. & Selli, E. Cr(VI) photocatalytic reduction: Effects of simultaneous organics oxidation and of gold nanoparticles photodeposition on TiO2. J. Hazard. Mater. 211–212, 188–195 (2012).

Naldoni, A. et al. Influence of TiO2 electronic structure and strong metal-support interaction on plasmonic Au photocatalytic oxidations. Catal. Sci. Technol. 6, 3220–3229 (2016).

Dudziak, S. et al. Solvothermal growth of 0 0 1 exposed anatase nanosheets and their ability to mineralize organic pollutants. The effect of alcohol type and content on the nucleation and growth of TiO2 nanostructures. Appl. Surf. Sci. 563, 150360 (2021).

Kubiak, A. et al. Unraveling a novel microwave strategy to fabricate exposed 001}/{101 facets anatase nanocrystals: Potential for use to the elimination of environmentally toxic metronidazole waste. Mater. Res. Bull. 167, 112438 (2023).

Zouheir, M. et al. Effective photocatalytic conversion of formic acid using iron, copper and sulphate doped TiO2. J. Cent. South Univ. 29, 3592–3607 (2022).

Hurum, D. C., Agrios, A. G., Gray, K. A., Rajh, T. & Thurnauer, M. C. Explaining the enhanced photocatalytic activity of Degussa P25 mixed-phase TiO2 using EPR. J. Phys. Chem. B 107, 4545–4549 (2003).

Chi, M., Sun, X., Sujan, A., Davis, Z. & Tatarchuk, B. J. A quantitative XPS examination of UV induced surface modification of TiO2 sorbents for the increased saturation capacity of sulfur heterocycles. Fuel 238, 454–461 (2019).

Abdullah, S. A. et al. Neutron beam interaction with rutile TiO2 single crystal (1 1 1): Raman and XPS study on Ti3+-oxygen vacancy formation. Mater Lett. 263, 127143 (2020).

Chi, M., Sun, X., Lozano-Blanco, G. & Tatarchuk, B. J. XPS and FTIR investigations of the transient photocatalytic decomposition of surface carbon contaminants from anatase TiO2 in UHV starved water/oxygen environments. Appl. Surf. Sci. 570, 151147 (2021).

Dozzi, M. V., Prati, L., Canton, P. & Selli, E. Effects of gold nanoparticles deposition on the photocatalytic activity of titanium dioxide under visible light. Phys. Chem. Chem. Phys. 11, 7171–7180 (2009).

Article  CAS  PubMed  Google Scholar 

Dudziak, S., Bielan, Z., Kubica, P. & Zielinska-Jurek, A. Optimization of carbamazepine photodegradation on defective TiO2-based magnetic photocatalyst. J. Environ. Chem. Eng. 9, 105782 (2021).

Ola, O. & Maroto-Valer, M. M. Review of material design and reactor engineering on TiO2 photocatalysis for CO2 reduction. J. Photochem. Photobiol. C: Photochem. Rev. 24, 16–42 (2015).

Meng, H., Hou, W., Xu, X., Xu, J. & Zhang, X. TiO2-loaded activated carbon fiber: Hydrothermal synthesis, adsorption properties and photo catalytic activity under visible light irradiation. Particuology 14, 38–43 (2014).

Xue, P. et al. Synergistic manipulation of Zn2+ ion flux and nucleation induction effect enabled by 3D hollow SiO2/TiO2/carbon fiber for long-lifespan and dendrite-free Zn–metal composite anodes. Adv. Funct. Mater. 31, 2106417 (2021).

Rankin, S. M., Moody, M. K., Naskar, A. K. & Bowland, C. C. Enhancing functionalities in carbon fiber composites by titanium dioxide nanoparticles. Compos. Sci. Technol. 201, 108491 (2021).

Pellegrino, F. et al. Formic acid photoreforming for hydrogen production on shape-controlled anatase TiO2 nanoparticles: assessment of the role of fluorides, 101}/{001 surfaces ratio, and platinization. ACS Catal 9, 6692–6697 (2019).

Dorosheva, I. B. et al. Synthesis and physicochemical properties of nanostructured TiO2 with enhanced photocatalytic activity. Inorg. Mater. 57, 503–510 (2021).

Parida, K. M., Dash, S. S. & Das, D. P. Physico-chemical characterization and photocatalytic activity of zinc oxide prepared by various methods. J. Colloid. Interface Sci. 298, 787–793 (2006).

Article  ADS  CAS  PubMed  Google Scholar 

Smith, Y. R., Kar, A. & Subramanian, V. Investigation of physicochemical parameters that influence photocatalytic degradation of methyl orange over TiO2 nanotubes. Ind. Eng. Chem. Res. 48, 10268–10276 (2009).

Zhou, F., Yan, C., Liang, T., Sun, Q. & Wang, H. Photocatalytic degradation of Orange G using sepiolite-TiO2 nanocomposites: Optimization of physicochemical parameters and kinetics studies. Chem. Eng. Sci. 183, 231–239 (2018).

Tolosana-Moranchel, A., Faraldos, M. & Bahamonde, A. Assessment of an intrinsic kinetic model for TiO2-formic acid photodegradation using LEDs as a radiation source. Catal. Sci. Technol. 10, 6198–6211 (2020).

Hamandi, M., Berhault, G., Dappozze, F., Guillard, C. & Kochkar, H. Titanium dioxide nanotubes/polyhydroxyfullerene composites for formic acid photodegradation. Appl. Surf. Sci. 412, 306–318 (2017).

Article  ADS  CAS  Google Scholar 

Sadi, A. B., Al Bilali, R. K., Abubshait, S. A. & Kochkar, H. Low temperature design of titanium dioxide anatase materials decorated with cyanuric acid for formic acid photodegradation. J. Saudi Chem. Soc. 24, 351–363 (2020).

Alomair, N. A. The role of strontium on the enhancement of photocatalytic response of TiO2 nanotubes-application in methylene blue and formic acid photodegradation under visible light and UV-A. Arab. J. Basic Appl. Sci. 29, 162–174 (2022).

Holm, A., Hamandi, M., Sahel, K., Dappozze, F. & Guillard, C. Impact of H2O2 on the lactic and formic acid degradation in presence of TiO2 rutile and anatase phases under UV and visible light. Catalysts 10, 1–15 (2020).

Zouheir, M. et al. Bandgap optimization of sol–gel-derived TiO2 and its effect on the photodegradation of formic acid. Nano Futures 5, 025004 (2021).

Article  ADS  CAS  Google Scholar 

Zalas, M. & Schroeder, G. Template free synthesis of locally-ordered mesoporous titania and its application in dye-sensitized solar cells. Mater. Chem. Phys. 134, 170–176 (2012).

Hegazy, A. & Prouzet, E. Room temperature synthesis and thermal evolution of porous nanocrystalline TiO2 anatase. Chem. Mater. 24, 245–254 (2012).

Borbón, S. et al. Open-circuit voltage (VOC) enhancement in TiO2-based DSSCs: Incorporation of ZnO nanoflowers and Au nanoparticles. ACS Omega 5, 10977–10986 (2020).

Article  PubMed  PubMed Central  Google Scholar 

Pradhan, B., Batabyal, S. K. & Pal, A. J. Vertically aligned ZnO nanowire arrays in Rose Bengal-based dye-sensitized solar cells. Energy Mater Sol. Cells 91, 769–773 (2007).

Biswas, S., Hossain, M. F. & Takahashi, T. Fabrication of Grätzel solar cell with TiO2/CdS bilayered photoelectrode. Thin Solid Films 517, 1284–1288 (2008).

Article  ADS  CAS  Google Scholar 

Jiang, W., Yin, L., Liu, H. & Ding, Y. Nanograss-structured counter electrode for dye-sensitized solar cells. J. Power Sources 218, 405–411 (2012).

Article  ADS  CAS  Google Scholar 

Bartkowiak, A., Korolevych, O., Chiarello, G. L., Makowska-Janusik, M. & Zalas, M. Experimental and theoretical insight into DSSCs mechanism influenced by different doping metal ions. Appl. Surf. Sci. 597, 153607 (2022).

Subramanian, A., Ho, C. Y. & Wang, H. Investigation of various photoanode structures on dye-sensitized solar cell performance using mixed-phase TiO2. J. Alloys Compd. 572, 11–16 (2013).

This work was funded by the National Science Centre of Poland (2021/43/D/ST5/01190).

Faculty of Chemistry, Adam Mickiewicz University, Poznan, Uniwersytet Poznanskiego 8, 61614, Poznan, Poland

Adam Kubiak, Maciej Zalas & Michał Cegłowski

You can also search for this author in PubMed  Google Scholar

You can also search for this author in PubMed  Google Scholar

You can also search for this author in PubMed  Google Scholar

A.K.: Conceptualization, Methodology, Formal analysis, Investigation, Data curation, Resource, Visualization, Writing—Original Draft, Writing—Review & Editing; M.Z.: Methodology, Formal analysis, Investigation, Data curation, Writing—Original Draft, Writing—Review & Editing; M.C.: Supervision, Writing—Review & Editing, Writing—Original Draft, Funding acquisition, Project administrator.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Kubiak, A., Zalas, M. & Cegłowski, M. Innovative microwave in situ approach for crystallizing TiO2 nanoparticles with enhanced activity in photocatalytic and photovoltaic applications. Sci Rep 14, 12617 (2024). https://doi.org/10.1038/s41598-024-63614-7

DOI: https://doi.org/10.1038/s41598-024-63614-7

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Scientific Reports (Sci Rep) ISSN 2045-2322 (online)

Innovative microwave in situ approach for crystallizing TiO2 nanoparticles with enhanced activity in photocatalytic and photovoltaic applications | Scientific Reports

Sulfuric Titanium Dioxide Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.