Next Article in Journal
Template–Free–Induced Synthesis of an Fe–N–C Electrocatalyst with Porous Yolk–Shell Structure Towards Oxygen Reduction Reaction
Previous Article in Journal
Geometric Matching Effect Induced High Dispersion of Na2WO4 Nanocluster on Cristobalite Support for Efficient Methyl Chloride-to-Vinyl Chloride Conversion
Previous Article in Special Issue
Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy

by
Matteo Tommasi
1,2,3,
Alice Gramegna
1,2,
Simge Naz Degerli
2,
Federico Galli
3,* and
Ilenia Rossetti
1,2
1
Chemical Plants and Industrial Chemistry Group, Dipartimento di Chimica, Università degli Studi di Milano, Via C. Golgi 19, 20133 Milan, Italy
2
INSTM Unit, Milano-Università, Via C. Golgi 19, 20133 Milan, Italy
3
Département de Génie Chimique et de Génie Biotechnologique, Université de Sherbrooke, 2500 Boul. de l’Université, Sherbrooke, QC J1K 2R1, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(4), 383; https://doi.org/10.3390/catal15040383
Submission received: 21 February 2025 / Revised: 3 April 2025 / Accepted: 10 April 2025 / Published: 16 April 2025
(This article belongs to the Special Issue Advances in Catalysis for a Sustainable Future)

Abstract

:
In this work, three catalysts, TiO2, WO3 and TiO2/WO3, have been synthesized through flame spray pyrolysis synthesis (FSP) and have been tested for CO2 photoreduction. The catalysts were fully characterized by XRD, DRS UV–Vis, N2 physisorption and SEM. Experimental tests were performed in a one-of-a-kind high-pressure reactor at 18 bar. TiO2 P25 was used as a benchmark to compare the productivities of the newly synthetized catalysts. The two single oxides showed comparable productivities, both slightly lower than the P25 reference value (ca. 17 mol/kgcat·h). The mixed oxide, TiO2/WO3, instead showed an impressive productivity of formic acid with 36 mol/kgcat·h, which is around 2.5 times higher than both of the single oxides alone. The formation of a type-II heterojunction has been confirmed through DRS analysis. The remarkable productivity demonstrates how FSP synthesis can be a crucial tool to obtain highly active and stable photocatalysts. This approach has already been successfully scaled up for the industrial production of various catalysts, showcasing its versatility and efficiency.

1. Introduction

Energy is indispensable, fulfilling fundamental needs and sustaining a comfortable and enhanced standard of living [1]. Over recent decades, energy consumption has escalated due to global population expansion and uncontrolled industrialization [2]. This has led to significant consumption of fossil fuels, resulting in energy shortages and elevated CO2 emissions. Emissions of CO2 and other greenhouse gases from carbon-based energy sources have caused numerous environmental and social challenges. CO2 emissions are responsible for more than 60% of global warming, with annual emissions exceeding 30 gigatons, primarily from fossil fuels [3]. The adoption of renewable energy is essential for reducing CO2 emissions, alleviating climate change and preserving a clean environment through sustainable energy solutions. Various strategies have been developed to mitigate the impact of CO2, such as CO2 capture, storage and reduction and its conversion into value-added chemicals. Among these, the conversion of CO2 into value-added chemicals is particularly appealing, as it also offers the potential to decrease fossil fuel consumption. CO2 utilization can be accomplished through several approaches, including chemical conversions, electrochemical conversions, biological conversions and reforming processes. However, each of these technologies faces some challenges, including requirements for high temperatures or electrical voltages to break the stable CO2 molecule, limitations in raw materials, high operational costs and issues related to sustainability [4]. Given these factors, one of the most promising approaches for the reutilization of CO2 is artificial photosynthesis. Despite the intermittency and variability in its intensity, sunlight is an inexhaustible energy source that is freely accessible. Natural photosynthesis is a process that transforms atmospheric CO2 into sugar-based biomass. Artificial photosynthesis emulates this process as a model to attempt the storing of solar energy in chemical form. The generation of solar fuels under sunlight can be achieved through photocatalysis. In heterogeneous photosynthetic processes, solid semiconductors are employed, activated by the absorption of photons with energy exceeding the bandgap [5]. When irradiation reaches the semiconductor surface, an electron–hole pair is generated; the excited electron moves from the valence band to the conduction band, leaving a positive hole that facilitates oxidation and reduction with the redox pairs present in the reaction medium. The efficiency of the reaction is negatively impacted by the possibility of electron–hole recombination. In recent years, a variety of materials have been adopted as efficient and environmentally sustainable photocatalysts for CO2 photoreduction, including TiO2, WO3, ZnO, CdS, CuO and g-C3N4. Among these, TiO2 stands out as one of the most promising due to its abundance, high chemical stability, low price and robust catalytic properties [6,7,8,9,10,11]. However, it has a limitation due to its wide bandgap of approximately 3.2 eV, making it responsive only to ultraviolet (UV) light, which constitutes approximately 3% of the solar spectrum [12,13]. Over recent decades, numerous studies have been published aiming to reduce the bandgap, enhance the light absorption and decrease the charge transfer and recombination rates of TiO2 by coupling it with other semiconductors that have narrower bandgap values, such as WO3, CdS, MoO2 [8,14,15,16,17,18]. WO3 has received considerable attention due to its relatively narrow bandgap of approximately 2.6–2.8 eV, non-toxicity, resistance to photo-corrosion, chemical stability and capability to exist in a wide range of oxidation states (2+, 3+, 4+, 5+ and 6+), which facilitates the storage of photogenerated electrons [19].
It is challenging to meet all of the needs, using only a semiconductor, for photocatalysis, such as having a wide visible-light response, long-term stability and a strong redox capacity, as well as preventing fast electron–hole recombination. A beneficial means that can be applied to enhance the properties of semiconductors is the type-II heterojunction, shown in Figure 1 [20]. The latter is formed when two semiconductors capable of absorbing light with different energy bands are put in contact. In this arrangement, the valence band (VB) of photocatalyst-I (PC-I) is more positive than the VB of PC-II, while the conduction band (CB) of PC-II is more negative compared to the CB of PC-I, as reported in Figure 1. When light irradiates the semiconductor surfaces, each generates an excited electron that transitions from the VB to the CB. In type-II heterojunctions, photo-promoted electrons tend to accumulate in the material with a more positive CB potential. On the other hand, holes tend to accumulate in the semiconductor with a more negative VB [21]. This kind of heterojunction differs from the so-called Z-scheme ones, where after two excited electrons and holes are generated, the electrons from the lower conduction band can recombine with the holes in the highest energy band, following the Z-scheme and making the potential differences between the holes and electrons higher than each of the materials alone [20,22].
Flame spray pyrolysis (FSP) is a promising route, but other well-established techniques have been widely explored in the literature to synthetize type-II heterojunctions with tunable properties as well. For the sake of completeness, some papers on different techniques that appeared of interest to the authors are given below. Lei et al. reported the synthesis of ZnxInySz by the hydrothermal method, and the relative ZnxInySz/ZnO composites with internal type-II heterojunctions prepared by a solvent-assisted interface, highlighting how the latter significantly enhanced the separation of the photocarriers [23]. Lately, electrospinning has gained attention as a method for fabricating one-dimensional (1D) nanostructured photocatalysts. This technique allows for the formation of nanofibers with high ratios, as well with large surface areas, for catalytic reactions. Experiments performed on TiO2/CdS nanofiber heterojunctions have reported significant improvements as H2 production photocatalysts, due to the formation of type-II heterojunctions that facilitate charge separation [24].
Despite the higher redox potential of the Z-scheme, type-II heterojunctions are able to improve photocatalytic efficiency when two materials are placed in intimate contact. It is well known that TiO2’s performance as a photocatalyst is highly affected by its extremely short e lifetime [25]. In this sense, the WO3/TiO2 type-II heterojunction can play a crucial role in enhancing charge separation and supporting CO2 reduction. De Castro et al. investigated WO3/TiO2 heterostructures towards organic contaminant degradation, reporting a particular improvement in the catalytic activity of the material for the 40 wt% WO3/TiO2 sample due to the larger interface and contact between the two phases [26]. In this study, we present a novel TiO2/WO3 hybrid material, synthetized though the flame spray pyrolysis (FSP) technique, that has been characterized in-depth and tested for CO2 photoreduction in a high-pressure photoreactor. The FSP technique allows us to place different materials in intimate contact as a strategy to favor the electron–hole transfer, improving the performances of the single photocatalysts.

2. Results and Discussion

2.1. Material Characterization

In the WO3 sample (Figure 2a,b), a petal-like structure was observed. The latter was due to the non-isotropic growth of the WO3 crystals, where certain crystal facets grew faster than others, leading to the formation of these morphologies. The whole surface of the sample resulted in small spherical particles of WO3. In contrast, on the TiO2 FSP sample (Figure 2e,f), this petal-like structure was not observed, but instead, a structure characterized by a well-defined porous network and particle aggregation was noted. Fewer spherical TiO2 particles can be observed on the surface of the sample. On the WO3/TiO2 hybrid material (Figure 2c,d), as reported in Figure 2d, it can be seen how the petal-like structure of the WO3 and the more classical porous structure due to the TiO2 joined with each other. The non-isotropic growth was still present, but to a lower extent. The WO3 deposited on top of the titania in the hybrid sample did not adversely affect the surface area of the latter, as reported in Table 1, allowing for excellent contact between the phases, as shown by the DRS analysis.

2.2. XRD Analysis

The XRD diffractograms are reported in Figure 3. In both the FSP-TiO2 and FSP- TiO2/WO3 60/40 samples, one can observe the main reflections at 25.3° (2θ) and 27.5° (2θ) due to the anatase (JCPDS 21-1272—anatase) and rutile (JCPDS 21-1276—rutile) phases, respectively. The FSP-WO3 diffractogram showed correspondence with the monoclinic WO3 across all ranges of measurement. These results are in accordance with the fact that this phase is the most stable at room temperature. Furthermore, authors have reported that WO3 is generally unable to retain alternative phases such as β-WO3 (orthorhombic) and α-WO3 (tetragonal) that are more stable at high temperatures [27]. The main reflection of the WO3 phase was still visible at 23° (2θ) in the FSP-TiO2/WO3 60/40, though much less crystalline and possibly dispersed into the titania phase, whose crystallinity was also decreased with respect to the bare titania sample. The crystallite sizes of the different catalysts are reported in Table 1, together with the calculated anatase/rutile compositions of the titania-containing compounds. The TiO2 prepared through FSP showed a significant difference with commercial P25 regarding the anatase/rutile composition. The FSP sample showed an anatase/rutile ratio of 48/52, compared to the 78/22 ratio of the commercial P25 sample. As reported by some authors, a higher rutile content can improve charge separation due to the formation of a homojunction with anatase [28].

2.3. N2 Physisorption Analysis

The results from the N2 physisorption are reported in Figure 4. All of the samples displayed a type-IV isotherm. The BET SSA, total pore volume and BJH adsorption pore width results obtained from the N2 physisorption curve elaboration are reported in Table 1. The BET SSA area of the FSP-TiO2 was one order of magnitude higher than that of the pure WO3, with values of 29.7 and 3.4 m2/g, respectively. The commercial P25 showed a higher surface area of 52.7 m2/g. The pore volume and pore width also varied greatly between the WO3 and TiO2, with the latter showing roughly half of the total pore volume compared to the sample of P25 (0.109 cm3/g vs. 0.225 cm3/g). The composite catalyst showed slight decreases in the SSA and the total pore volume compared to the FSP-TiO2. The BJH calculated pore width displayed was also lower (9.8 nm).

2.4. DRS UV–Vis Analysis

Both the DR UV–Vis spectra of all of the samples and the Tauc plot elaboration are shown in Figure 5. The calculation of the bandgap was obtained starting from the Tauc equation and implementing the Kubelka–Munk function. Both TiO2 and WO3, as explained, are semiconductors with an indirect transition. The values of the bandgaps can be obtained from the graph, plotting (F(R∞)hν)γ against the energy. The results obtained are reported in Table 1. The P25 and FSP-TiO2 bandgaps were estimated, respectively, as 3.22 and 3.10 eV. The titanium dioxide and tungsten trioxide bandgaps obtained by FSP were in agreement with the literature, which attributes a value of 3.2 eV for the former [29] and 2.7 eV for the latter [30]. The bandgap of the TiO2 FSP sample was slightly lower compared to that of the P25, which can be attributed to the higher rutile content. Anatase and rutile have bandgaps of, respectively, 3.2 eV and 3.0 eV [31], and the TiO2 FSP bandgap of 3.10 eV is in line with the anatase/rutile composition that was calculated through XRD. All of these characteristics align with the higher flame temperature and longer residence time, which enhance the crystallinities of both the anatase and rutile phases while reducing defects. This is advantageous in photocatalysis, as a lower defect density improves charge separation, reduces recombination and, combined with a high surface area for reactant adsorption and efficient light exposure, enhances the photocatalytic performance.
The bandgap of the mixed oxide is interesting because its value, 2.70 eV, was lower than that of both of the pure oxides synthetized, and resembles that of the pure WO3. This behavior is typical of type-II heterojunctions, where a small reduction in the measured bandgap is observed by DRS. This means that better sunlight harvesting can be exploited by this substance: a lower bandgap can be translated into a wider range of wavelengths capable of being absorbed by the substance considered. Being that the sunlight spectrum shifted to visible light, rather than UV radiation, this improvement is relevant for photocatalyst application in real sunlight.

2.5. Photoreduction of CO2

Tests were performed at a pH of 14, 18 bar and 6 h of reaction time, with 1.66 g/L of sodium sulfite as a hole scavenger (HS) and 31 mg/L of catalyst concentration on the prepared catalysts, as previously optimized. The productivity results and HS conversions are reported in Figure 6. Starting from the productivity, the total stored energy was calculated considering the (Lower Heating Value) values of the primary reduction products [32], using Equations (1) and (2). Both the productivity and the calculated total stored energy are reported in Table 2.
n i m o l = p r o d u c t i v i t y i m o l k g c a t × h × t h × m c a t k g
T o t a l   s t o r e d   e n e r g y J = n i m o l × L H V i J m o l
The performance of our benchmark, namely, a commercial TiO2 P25 produced by Evonik (formerly Degussa) and supplied by EIGENMANN & VERONELLI S.p.A, was ca. 15% better than that of the titania sample prepared by FSP. This can be attributed to the combined effects of the higher anatase content and higher surface area of the P25 sample. The performances of the two bare oxides prepared by FSP were comparable, with the FSP titania showing a slightly lower productivity of HCOOH (13.9 mol/kgcat·h) compared to that of the FSP WO3 (14.7 mol/kgcat·h). This result is attributed to the lower bandgap of the WO3, which suggests a better utilization capacity of the absorbed light when provided the use of the same UV lamp. The WO3/TiO2 40/60 showed a two-fold increase in HCOOH productivity compared to the P25 with an astonishing value of 36.5 mol/kgcat·h, among the highest reported in the literature, as reported in Table 3.
The hole scavenger conversion followed the productivity of the different catalysts, with the mixed oxide reporting a conversion of 77% compared to the 43% of the WO3 and the 40% of the TiO2. The greater HS conversion is in line with the higher productivity recorded. It should be underlined that HS conversion should be maintained well below 90 percent to avoid the photoreforming of the obtained products [33].
As shown by the DRS measurements, the TiO2 FSP sample showed a slightly lower bandgap compared to the P25, which can be attributed to the higher rutile content. When TiO2 and WO3 are put in intimate contact, such as through FSP synthesis, a type-II heterojunction is formed, as confirmed by the DRS measurements. The latter showed a bandgap slightly inferior even to that of the pure WO3. In type-II heterojunctions, the charge recombination is reduced by spatially separating the photogenerated electrons and holes, with the electrons migrating to the lower conduction band (WO3) and the holes to the higher valence band (TiO2). In this case, a small reduction in the measured bandgap was observed, as reported in Figure 1. Some authors have also reported that the greater acidic character of WO3 should interact more with the carbonate ions present in the solution, due to its greater acidity [34].
As previously mentioned, CO2 photoreduction is now a well-known and growing field. The photocatalytic reduction of CO2 to formic acid represents a small fraction of the research carried out, despite the high added value of the formic acid produced. Recently, an interesting review was published on this topic [35]. A quick overview of the relevant productivities recorded in the literature is shown in Table 3.
Table 3. Literature comparison—catalyst, reaction medium, light source and productivity.
Table 3. Literature comparison—catalyst, reaction medium, light source and productivity.
CatalystReaction MediumLight SourceProductivity (mol/kgcat h)Reference
Cu2OAqueous suspension in 0.3 M NaHCO3, 2 M glycerol as HSArc Lamp
1000 W Xenon
2.76[36]
Ag/TiO2Aqueous suspension in 0.2 M NaHCO3Arc Lamp
1000 W Xenon
3.89[37]
Co3O4Aqueous suspension in waterLED lamp
21 W, λ = 510
HCOOH
4.53 (mmol/kgcat·h)
HCHO
0.62 (mmol/kgcat·h)
[38]
ZnO-Cu2O
NPs
Aqueous suspension in water and
ZnO-Cu2O
300 W Xe lamp
0.2 M Na2CO3
300 W Xe lampCH4
1080 (mmol/kgcat·h)
CO
1.4 (mmol/kgcat·h)
[39]
TiO2 FSPAqueous suspension in water, HS 1.66 g/L of Na2SO3Medium Hg-pressure UV lamp 250 W13.9This work
WO3/TiO2 (40/60) FSPAqueous suspension in water, HS 1.66 g/L of Na2SO3Medium Hg-pressure UV lamp 250 W36.5This work
The performances of both of the bare oxides prepared by FSP were comparable, with high productivity compared to what has been reported in the literature, although it is worth noticing that different conditions, such as lamps and hole scavengers, were used. Authors have reported a Ag/TiO2 catalyst which showed a remarkable productivity of 3.89 mol/kgcat·h when tested in an aqueous suspension (0.2 M NaHCO3) under irradiation by a 1000 W Xenon arc lamp [37]. As previously said, the hybrid oxide being prepared through FSP synthesis allowed it to obtain an astonishing productivity of 36.5 mol/kgcat·h, emphasizing that FSP represents a fundamental tool for the preparation of oxide-based catalysts, and all the more so in the field of photocatalysis, where the intimate contact obtained through this technique allows for simple and fast preparations of type-II heterojunctions.

3. Materials and Methods

3.1. Materials Used

Commercial TiO2 P25, produced by Evonik (formerly Degussa; Essen, Germany) and supplied by EIGENMANN & VERONELLI S.p.A. (Rho, Italy), was used as a benchmark. Carbon dioxide (CO2), methane (CH4), oxygen (O2) and helium (He), all with a purity >99.99%, were purchased from AirLiquide (Milano, Italy). Ammonium metatungstate hydrate (AMT) (Sigma Aldrich, St. Louis, MO, USA, >99.99% purity) and titanium tetraisopropoxide (TTIP) (Sigma Aldrich, >97% purity) were used as precursors for the preparation of the metal oxides through flame spray pyrolysis. Dimetilformamide (DMF) (Sigma Aldrich, >99.9% purity) was used as a solvent for the preparation of the precursor solution. Sodium sulfite (Na2SO3, Sigma Aldrich, >98% purity) was used as a hole scavenger (HS) during the photocatalytic test. NaOH (Sigma Aldrich, 98% purity) was used to achieve a basic pH (pH = 14) during the experiments. The operating conditions, including the use of a basic pH, were optimized in previous work to maximize the catalyst productivity [40].

3.2. Flame Spray Pyrolysis

Flame spray pyrolysis, often called FSP, is a method that allows us to obtain single or mixed metal oxides which are thermally stable, with a high degree of purity and an interesting specific surface area. The metal or metals whose oxide is to be obtained are contained in organometallic salts, called precursors. The precursors are dissolved in a combustible organic solvent, and the solution thus obtained is fed by means of a syringe pump through a needle in a vertical nozzle placed in the center of the burner. A scheme of the FSP apparatus is reported in Figure 7. The end of the capillary tube is coaxially lapped by a flow of O2, which acts as a comburent and at the same time has the function of dispersing the solution into very fine droplets. The heterogeneous O2/solution mixture is ignited by twelve CH4/O2 flames, arranged in a cone around the nozzle. The instantaneous dispersion and vaporization of the solution droplets and the decomposition (pyrolysis) of the precursor take place in the flame, giving rise to the first oxide nuclei, which increase by coalescence and condensation. The primary nanoparticles thus obtained aggregate in the final part of the flame, forming the desired powders. The powders are collected from the dust collection glass chamber on which they are deposited. The size of the formed particles depends on the temperature of the flame, on the residence time and on the concentration of particles in the flame. When the values of these parameters increase, the probability that sintering phenomena will occur increases as well, leading to the formation of larger particles with lower surface areas. The FSP process is affected by many parameters, including the nature of the solvent, the concentration and flowrate of the precursor solution, the flow of the oxygen and methane and the overpressure of the oxygen used for the dispersion of the solution. The latter can play a significant role in the final surface area of the catalyst, and for this reason, a manometer is used to constantly control this value.
FSP can be defined as one of the so-called non-conventional synthesis techniques. Several advantages over traditional synthesis methods like thermal annealing, sol-gel and wet impregnation are present, but it also comes with some challenges. One of its main strengths is the ability to produce nanoparticles in a single step, without the need for multiple drying and calcination stages. This translates as well to better control over the particle size and dispersion. The FSP process is highly scalable, making it suitable for industrial applications as well, where large amounts of material need to be synthetized. The main downsides for this technology when compared to more conventional ones such as thermal annealing include the use of specialized equipment, including a high-temperature flame and oxygen supply, which cannot be carried out with conventional lab equipment. In comparison, thermal annealing is the most straightforward method, but typically results in larger particles and lower surface areas, which can negatively impact catalytic performance. The powder collection efficiency varies greatly based on the filters and collection systems that are put in place. In the case of setups such as the one used in the present work, the dust collection is about 25% of the theoretical maximum amount. This occurs because only the powder deposited on the glass wall of the collection chamber is collected. When considering the synthesis of catalysts on a pilot or industrial scale, the powder collection efficiency improves greatly due to the combination of baghouse filters and HEPA porous stainless steel filters, reaching values close to 100%.
Figure 7. FSP schematic representation. Adapted from [41].
Figure 7. FSP schematic representation. Adapted from [41].
Catalysts 15 00383 g007

3.3. Catalyst Preparation

All of the catalysts tested in this work, except for the commercial P25 that was used as a benchmark, were synthesized via FSP. In order to synthetize the pure WO3 and TiO2 and the mixed oxide WO3/TiO2, ammonium metatungstate hydrate (AMT) and titanium tetraisopropoxide (TTIP) were used as precursors, respectively, for WO3 and TiO2. AMT is a water-soluble compound which cannot be dissolved in alcohols. On the other side, TTIP is strongly hygroscopic, and requires the use of an anhydrous organic solvent to avoid hydrolysis directly in the flask. DMF (dimetilformamide) was used as a solvent, since it is an anhydrous organic solvent able to solubilize AMT and due to its boiling point (153 °C). Both a low-boiling and a very high-boiling solvent must be avoided to prevent, respectively, the sintering of the produced particles or the formation of egg-shell phenomena, which lead to the formation of hollow particles. Solutions with a concentration of the organic precursor equal to 3% wt. compared to the solvent were used for injection in the FSP apparatus. This concentration was taken as a reference from a previous work [40] to avoid the vaporization of the solvent and the precursor directly into the tip of the needle, which can lead to clogging and to a twisted flame. Furthermore, increasing the concentration above a certain value again incurs the phenomenon of particle sintering, due to the higher probability of collision of the formed particles. All of the instrumentation was well dried in the oven to eliminate any moisture. The prepared solutions were injected into the FSP apparatus with a flow of 2.7 mL/min. The flowrates for the methane, external and the internal oxygen flows were, respectively, 0.5 L/min, 1 L/min and 5 L/min. During all of our syntheses, an overpressure of 1.5 bar of oxygen used for dispersion was maintained. After synthesis, all of the powder obtained was collected from the wall of the glass chamber. The composite catalyst obtained had final mass and molar ratios of the two oxides equal, respectively, to TiO2/WO3 = 60/40 (wt/wt) and TiO2/WO3 = 81.3/18.7 (mol/mol).

3.4. Catalyst Characterization

Images from the scanning electron microscopy (SEM) were captured using a Hitachi TM 1000 tabletop scanning electron microscope (Marunouchi, Japan). Carbon tape was used as a substrate for the SEM analysis to prevent charging effects without requiring metallization.
The X-ray diffraction (XRD) analyses were performed with a Rigaku Miniflex-600 horizontal-scan powder diffractometer (Tokyo, Japan) using Cu-Kα radiation, with a graphite monochromator on the diffracted beam. Scherrer’s Equation (3) was used to calculate the size of the crystallites.
D = (K∙λ)/(β∙cosθ)
where D is the crystal size, λ is the X-ray wavelength (0.154 nm with the Cu Kα generator), K is the shape factor (0.9), β is the width at the half maximum of the peak (i.e., FWHM) and θ is the Bragg angle.
The N2 physisorption analyses were performed with an ASAP 2020 instrument by Micromeritics (Norcross, GA, USA). The pretreatment before the actual analysis included the degassing of the samples, performed under a vacuum at 150 °C for 4 h. This pretreatment allowed us to remove any contaminants or traces of water present in the catalysts. This non-destructive analysis was performed on 200 mg of nano powder at a temperature of −196 °C. The data obtained were processed following the BET theory, BJH theory and t-plot analysis method.
The diffuse reflectance (DR) UV–Vis spectra of the catalysts were recorded on a Shimadzu UV−3600 Plus (Kyoto, Japan) in the range of 200–800 nm, using an integrating sphere and BaSO4 as reference standards. Using the reflectance spectra as the input data, a Kubelka–Munk elaboration was performed, according to Equation (4) [42].
F(R∞) = (1 − R∞)2/2R
The calculated (F(R)hν)1/r (with r = 2 or ½ for the direct and indirect bandgaps) was plotted versus hν to obtain the bandgap of each sample [43]. To prevent the saturation of the samples that absorb too much, all of the photocatalysts were diluted one-to-one or one-to-two with BaSO4, which was the reflectance standard. The shapes of the curves were not affected by blending with the BaSO4, so a comparison of the absolute value of the reflectance could not be performed. The final form of the equation is reported in Equation (5).
(F(R∞)hν)γ = A(hν − Eg)
with the parameter F(R∞) corresponding to the Kubelka–Munk function, A being a constant, Eg being the bandgap energy of the semiconductor and hν being the energy of the incident radiation calculated using Planck’s constant and the frequency of the radiation itself. The exponent γ can assume two values: 2 or ½. In the specific case of the synthetized FSP catalysts, the value of γ considered was ½, because both TiO2 and WO3 (monoclinic) are semiconductors with an indirect transition [44].

3.5. Photoreduction Setup

The tests with the synthetized catalysts were performed in a special high-pressure photoreactor previously developed by our group and described elsewhere [45]. This reactor allows us to perform experiments in the liquid phase up to 20 bar and 100 °C. An AISI 316 stainless steel jacketed reactor (cylinder shaped) with an internal capacity of 1.7 L was used. A quartz sleeve able to withstand pressures up to more than 20 bar was placed in the middle of the reactor. Housed inside the quartz jacket was the UV lamp, consisting of two medium Hg-pressure UV bulbs, each one with a power of 125 W. The lamp irradiation was in the range of 254–365 nm, with the peak emission located at 365 nm. The irradiance was checked periodically using a photoradiometer (Delta OHM HD2102.2 equipped with a LP 471A-UVeff probe, producer Senseca Italy Srl (ex Delta OHM Srl), Caselle di Selvazzano (PD), Italy), active in the range of 250–400 nm and resulting in an average of 120 W/m2. Sodium sulfite (Na2SO3) was used as a hole scavenger (HS). To test the catalysts, 1.2 L of water were loaded into the reactor, together with 1.66 g/L of Na2SO3 and 31 mg/L of the catalyst dispersed in it [25,46,47,48]. All of the photocatalytic tests were performed at a basic pH (pH = 14), 80 °C and 18 bar under intense and constant stirring to ensure the homogeneous dispersion of the catalyst. NaOH was used to obtain a basic environment. Each experiment lasted 6 h.
The gaseous products were analyzed with a GC Agilent 7890A (producer Agilent Technologies, Santa Clara, CA, USA) equipped with two different columns, HP PlotQ and MS, respectively, used for the analyses of the gaseous product and permanent gases. The GC analysis was performed at 70 °C, He was used as a carrier gas and the quantification was performed by means of a TCD. The two columns were placed in a series, with the second (MS) being able to be bypassed and sent into isolation. At the beginning of the analysis, the two columns were connected in a series, which allowed the permanent gases (H2, N2, O2, CH4 and CO), which were the first to leave the HP PlotQ, to flow directly into the MS column. The MS column was then set into bypass mode, and the CO2 coming out from the HP PlotQ was directly detected. After this, the bypass of the MS column was removed, and the permanent gases were quantified.
The liquid products were evaluated through an HPLC analysis (Jasco, Tokyo, Japan), equipped with a PS-DVB sulphonated gel column (SepaChrom Benson Polymeric BP-OA BL0053 2000–0, Rho, Italy), a UV–Vis (Jasco 4070) set at 210 nm and a Refractive Index (RI) (Jasco 4030) detector. An aqueous H2SO4 solution (0.05 mM) was used as an eluent, with a flowrate of 0.4 mL/min. The column was held in a bath at 45.5 °C. An iodometric titration was performed after the tests to evaluate the sodium sulfite conversion.

4. Conclusions

The present study demonstrates that FSP is a highly effective synthesis method for producing photocatalysts with well-defined properties to enhance photocatalytic performance. Three semiconductors were synthetized: TiO2, WO3 and a type-II heterojunction, WO3/TiO2 (40/60 weight ratio). These three semiconductors were subsequently tested for their efficacy in CO2 photoreduction. TiO2 is one of the most common photocatalysts used in the world. Of all commercial products, the TiO2 (P25) produced by Evonik (formerly Degussa) is not only one of the best known in the world, but is also produced industrially in a flame with a similar technique to FSP, using titanium tetrachloride (TiCl4) in an oxygen-rich flame. For these reasons, TiO2 P25 (Evonik) was therefore used as a reference catalyst for benchmarking purposes.
All catalysts tested were characterized by X-ray diffraction (XRD), nitrogen physisorption, scanning electron microscopy (SEM) and diffuse reflectance spectroscopy (DRS) UV–Vis. The latter returned values that align with those of the literature. The slightly lower bandgap of the TiO2 FSP compared to P25 can be attributed to the higher rutile content, as confirmed by the XRD analysis.
The experimental tests were conducted using a unique high-pressure reactor operating at 18 bar. The results showed that the two single oxides, TiO2 and WO3, despite their significant difference in bandgap energy (3.10 eV vs. 2.73 eV), exhibited comparable productivities in CO2 photoreduction, with each slightly underperforming relative to the P25 benchmark, which had a productivity of 16.9 mol/kgcat·h.
A key advantage of FSP emerged in the synthesis of the mixed oxide WO3/TiO2, where intimate contact between the phases was obtained, leading to the formation of a type-II heterojunction. This heterojunction is evidenced by the slight reduction in the bandgap compared to the pure WO3, as observed in the DRS measurements. As shown in Figure 1, the type-II alignment facilitated charge separation by directing the photogenerated electrons to the conduction band of the WO3 and the holes to the valence band of the TiO2, suppressing recombination and improving the photocatalytic efficiency. The WO3/TiO2 (40/60 weight ratio) catalyst exhibited a remarkable two-fold increase in formic acid productivity compared to P25, reaching 36.5 mol/kgcat·h, which is among the highest values reported in the literature.
To further reinforce the advantages of FSP, a comparative benchmark with sol-gel or other synthesis techniques would be beneficial. However, the results presented herein already indicate that FSP enables intimate contact between semiconductor phases, leading to superior photocatalytic performance.
By achieving an outstanding productivity of 36 mol/kgcat·h, this method leveraged the synergy between flame spray pyrolysis (FSP), high-pressure photoreduction and the type-II heterojunction. FSP proved to be an efficient and scalable synthesis technique for the synthesis of type-II heterojunctions, where phase interaction plays a crucial role in enhancing activity. The integration of these techniques not only allows us to optimize catalytic performance, but also paves the way for innovative applications in industrial catalysis.

Author Contributions

Conceptualization, I.R.; methodology, M.T. and A.G.; validation, F.G., I.R. and M.T.; formal analysis, S.N.D.; investigation, A.G. and M.T.; resources, I.R.; writing—original draft preparation, M.T.; writing—review and editing, M.T., I.R. and F.G.; visualization, M.T. and S.N.D.; supervision, I.R.; project administration, I.R.; funding acquisition, I.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fondazione Cariplo (Italy) through the grant “2021-0855 SCORE – Solar energy for circular CO2 photoconversion and chemicals regeneration”, in the frame of the 2021 call on circular economy (I. Rossetti and G. Ramis). I. Rossetti and G. Ramis gratefully acknowledge the financial contributions of Next Generation EU—PIANO NAZIONALE DI RIPRESA E RESILIENZA (PNRR), Missione 4 “Istruzione e Ricerca”—Componente C2 Investimento 1.1, “Fondo per il Programma Nazionale di Ricerca e Progetti di Rilevante Interesse Nazionale (PRIN2022PNRR)”, through the grant “P20227LB45—SCORE2—Solar-driven COnveRsion of CO2 with HP-HT 2156 photoReactor”.This research is also part of the project “One Health Action Hub: University Task Force for the resilience of territorial ecosystems”, supported by Università degli Studi di Milano—PSR 2021—GSA—Linea 6 (I. Rossetti). This study was carried out within the Agritech National Research Center and received funding from the European Union Next-GenerationEU (PIANO NAZIONALE DI RIPRESA E RESILIENZA (PNRR)—MISSIONE 4 COMPONENTE 2, INVESTIMENTO 1.4—D.D. 1032 17/06/2022, CN00000022). This manuscript reflects only the authors’ views and opinions, and neither the European Union nor the European Commission can be considered responsible for them. M. Tommasi, A. Gramegna and I. Rossetti acknowledge specifically the participation and funding of Tasks 8.2.3, 8.3.2 and 8.4.1.

Data Availability Statement

Data will be available on request.

Acknowledgments

Daniele Marinotto, for the support given during the DRS analysis.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analysis or interpretation of the data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Soon, W.; Baliunas, S.L.; Robinson, A.B.; Robinson, Z.W. Environmental Effects of Increased Atmospheric Carbon Dioxide. Clim. Res. 1999, 13, 149–164. [Google Scholar] [CrossRef]
  2. Khan, A.A.; Tahir, M. Recent Advancements in Engineering Approach towards Design of Photo-Reactors for Selective Photocatalytic CO2 Reduction to Renewable Fuels. J. CO2 Util. 2019, 29, 205–239. [Google Scholar] [CrossRef]
  3. Albo, J.; Luis, P.; Irabien, A. Carbon Dioxide Capture from Flue Gases Using a Cross-Flow Membrane Contactor and the Ionic Liquid 1-Ethyl-3-Methylimidazolium Ethylsulfate. Ind. Eng. Chem. Res. 2010, 49, 11045–11051. [Google Scholar] [CrossRef]
  4. Das, S.; Wan Daud, W.M.A. A Review on Advances in Photocatalysts towards CO2 Conversion. RSC Adv. 2014, 4, 20856–20893. [Google Scholar] [CrossRef]
  5. Xie, S.; Zhang, Q.; Liu, G.; Wang, Y. Photocatalytic and Photoelectrocatalytic Reduction of CO2 Using Heterogeneous Catalysts with Controlled Nanostructures. Chem. Commun. 2016, 52, 35–59. [Google Scholar] [CrossRef]
  6. Sorcar, S.; Hwang, Y.; Grimes, C.A.; In, S.I. Highly Enhanced and Stable Activity of Defect-Induced Titania Nanoparticles for Solar Light-Driven CO2 Reduction into CH4. Mater. Today 2017, 20, 507–515. [Google Scholar] [CrossRef]
  7. Liao, F.; Zeng, Z.; Eley, C.; Lu, Q.; Hong, X.; Tsang, S.C.E. Electronic Modulation of a Copper/Zinc Oxide Catalyst by a Heterojunction for Selective Hydrogenation of Carbon Dioxide to Methanol. Angew. Chem. Int. Ed. 2012, 51, 5832–5836. [Google Scholar] [CrossRef] [PubMed]
  8. Ahmad Beigi, A.; Fatemi, S.; Salehi, Z. Synthesis of Nanocomposite CdS/TiO2 and Investigation of Its Photocatalytic Activity for CO2 Reduction to CO and CH4 under Visible Light Irradiation. J. CO2 Util. 2014, 7, 23–29. [Google Scholar] [CrossRef]
  9. Lan, H.; Tang, Y.; Zhang, X.; You, S.; Tang, Q.; An, X.; Liu, H.; Qu, J. Decomplexation of Cu(II)-EDTA over Oxygen-Doped g-C3N4: An Available Resource towards Environmental Sustainability. Chem. Eng. J. 2018, 345, 138–146. [Google Scholar] [CrossRef]
  10. Kohno, Y.; Tanaka, T.; Funabiki, T.; Yoshida, S. Reaction Mechanism in the Photoreduction of CO2 with CH4 over ZrO2. Phys. Chem. Chem. Phys. 2000, 2, 5302–5307. [Google Scholar] [CrossRef]
  11. Lu, X.; Xie, J.; Chen, X.; Li, X. Engineering MPx (M = Fe, Co or Ni) Interface Electron Transfer Channels for Boosting Photocatalytic H2 Evolution over g-C3N4/MoS2 Layered Heterojunctions. Appl. Catal. B Environ. 2019, 252, 250–259. [Google Scholar] [CrossRef]
  12. Raza, A.; Shen, H.; Haidry, A.A.; Sun, L.; Liu, R.; Cui, S. Studies of Z-Scheme WO3-TiO2/Cu2ZnSnS4 Ternary Nanocomposite with Enhanced CO2 Photoreduction under Visible Light Irradiation. J. CO2 Util. 2020, 37, 260–271. [Google Scholar] [CrossRef]
  13. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  14. El-Yazeed, W.S.A.; Ahmed, A.I. Photocatalytic Activity of Mesoporous WO3/TiO2 Nanocomposites for the Photodegradation of Methylene Blue. Inorg. Chem. Commun. 2019, 105, 102–111. [Google Scholar] [CrossRef]
  15. Gao, L.; Du, J.; Ma, T. Cysteine-Assisted Synthesis of CuS-TiO2 Composites with Enhanced Photocatalytic Activity. Ceram. Int. 2017, 43, 9559–9563. [Google Scholar] [CrossRef]
  16. Gan, T.; Li, Y.; Wang, X.Z.; Wang, X.T.; Wang, C.W. Cu2 ZnSnS4@TiO2 P-n Heterostructured Nanosheet Arrays: Controllable Hydrothermal Synthesis and Enhanced Visible Light-Driven Photocatalytic Activity. Appl. Surf. Sci. 2017, 408, 60–67. [Google Scholar] [CrossRef]
  17. Wei, R.B.; Huang, Z.L.; Gu, G.H.; Wang, Z.; Zeng, L.; Chen, Y.; Liu, Z.Q. Dual-Cocatalysts Decorated Rimous CdS Spheres Advancing Highly-Efficient Visible-Light Photocatalytic Hydrogen Production. Appl. Catal. B Environ. 2018, 231, 101–107. [Google Scholar] [CrossRef]
  18. Xu, X.; Wang, Y.; Ji, Y.; Xu, Y.; Xie, M. A Novel Phase Retrieval Method from Three-Wavelength in-Line Phase-Shifting Interferograms Based on Positive Negative 2$π$ Phase Shifts. J. Mod. Opt. 2018, 65, 8–15. [Google Scholar] [CrossRef]
  19. Khan, H.; Rigamonti, M.G.; Patience, G.S.; Boffito, D.C. Spray Dried TiO2/WO3 Heterostructure for Photocatalytic Applications with Residual Activity in the Dark. Appl. Catal. B Environ. 2018, 226, 311–323. [Google Scholar] [CrossRef]
  20. Zhang, G.; Wang, Z.; Wu, J. Construction of a Z-Scheme Heterojunction for High-Efficiency Visible-Light-Driven Photocatalytic CO2 Reduction. Nanoscale 2021, 13, 4359–4389. [Google Scholar] [CrossRef]
  21. Li, J.; Tu, P.; Yang, Q.; Cui, Y.; Gao, C.; Zhou, H.; Lu, J.; Bian, H. A Novel Type-II–II Heterojunction for Photocatalytic Degradation of LEV Based on the Built-in Electric Field: Carrier Transfer Mechanism and DFT Calculation. Sci. Rep. 2024, 14, 1–16. [Google Scholar] [CrossRef]
  22. Li, J.; Yuan, H.; Zhang, W.; Jin, B.; Feng, Q.; Huang, J.; Jiao, Z. Advances in Z-Scheme Semiconductor Photocatalysts for the Photoelectrochemical Applications: A Review. Carbon Energy 2022, 4, 294–331. [Google Scholar] [CrossRef]
  23. Lei, X.; Yin, X.; Meng, S.; Li; Wang, H.; Xi, H.; Yang, J.; Xu, X.; Yang, Z.; Lei, Z. Fabrication of Type II Heterojunction in ZnIn2S4@ZnO Photocatalyst for Efficient Oxidative Coupling of Benzylamine under Visible Light. Mol. Catal. 2022, 519. [Google Scholar] [CrossRef]
  24. Ge, H.; Xu, F.; Cheng, B.; Yu, J.; Ho, W. S-Scheme Heterojunction TiO2/CdS Nanocomposite Nanofiber as H2-Production Photocatalyst. ChemCatChem 2019, 11, 6301–6309. [Google Scholar] [CrossRef]
  25. Vequizo, J.J.M.; Matsunaga, H.; Ishiku, T.; Kamimura, S.; Ohno, T.; Yamakata, A. Trapping-Induced Enhancement of Photocatalytic Activity on Brookite TiO2 Powders: Comparison with Anatase and Rutile TiO2 Powders. ACS Catal. 2017, 7, 2644–2651. [Google Scholar] [CrossRef]
  26. De Castro, I.A.; Avansi, W.; Ribeiro, C. WO3/TiO2 Heterostructures Tailored by the Oriented Attachment Mechanism: Insights from Their Photocatalytic Properties. CrystEngComm 2014, 16, 1514–1524. [Google Scholar] [CrossRef]
  27. Zheng, H.; Ou, J.Z.; Strano, M.S.; Kaner, R.B.; Mitchell, A.; Kalantar-zadeh, K. Nanostructured Tungsten Oxide – Properties, Synthesis, and Applications. Adv. Funct. Mater. 2011, 21, 2175–2196. [Google Scholar] [CrossRef]
  28. Huang, X.; Zhang, R.; Gao, X.; Yu, B.; Gao, Y.; Han, Z. TiO2-Rutile/Anatase Homojunction with Enhanced Charge Separation for Photoelectrochemical Water Splitting. Int. J. Hydrogen Energy 2021, 46, 26358–26366. [Google Scholar] [CrossRef]
  29. Kutorglo, E.M.; Schwarze, M.; Nguyen, A.D.; Tameu, S.D.; Huseyinova, S.; Tasbihi, M.; Görke, O.; Primbs, M.; Šoóš, M.; Schomäcker, R. Efficient Full Solar Spectrum-Driven Photocatalytic Hydrogen Production on Low Bandgap TiO2/Conjugated Polymer Nanostructures. RSC Adv. 2023, 13, 24038–24052. [Google Scholar] [CrossRef]
  30. Liu, J.; Yang, L.; Li, C.; Chen, Y.; Zhang, Z. Optimal Monolayer WO3 Nanosheets/TiO2 Heterostructure and Its Photocatalytic Performance under Solar Light. Chem. Phys. Lett. 2022, 804, 139861. [Google Scholar] [CrossRef]
  31. Landmann, M.; Rauls, E.; Schmidt, W.G. The Electronic Structure and Optical Response of Rutile, Anatase and Brookite TiO2. J. Phys. Condens. Matter 2012, 24. [Google Scholar] [CrossRef] [PubMed]
  32. NIST. Formic Acid. Available online: https://webbook.nist.gov/cgi/cbook.cgi?ID=64-18-6 (accessed on 2 December 2024).
  33. Tommasi, M.; Conte, F.; Alam, M.I.; Ramis, G.; Rossetti, I. Highly Efficient and Effective Process Design for High-Pressure CO2 Photoreduction over Supported Catalysts. Energies 2023, 16, 4990. [Google Scholar] [CrossRef]
  34. Kumar, S.G.; Rao, K.S.R.K. Comparison of Modification Strategies towards Enhanced Charge Carrier Separation and Photocatalytic Degradation Activity of Metal Oxide Semiconductors (TiO2, WO3 and ZnO). Appl. Surf. Sci. 2017, 391, 124–148. [Google Scholar] [CrossRef]
  35. Pan, H.; Heagy, M.D. Photons to Formate—A Review on Photocatalytic Reduction of CO2 to Formic Acid. Nanomaterials 2020, 10, 2422. [Google Scholar] [CrossRef]
  36. Pan, H.; Chowdhury, S.; Premachandra, D.; Olguin, S.; Heagy, M.D. Semiconductor Photocatalysis of Bicarbonate to Solar Fuels: Formate Production from Copper(I) Oxide. ACS Sustain. Chem. Eng. 2018, 6, 1872–1880. [Google Scholar] [CrossRef]
  37. Pan, H.; Heagy, M.D. Plasmon-Enhanced Photocatalysis: Ag/TiO2 Nanocomposite for the Photochemical Reduction of Bicarbonate to Formic Acid. MRS Adv. 2019, 4, 425–433. [Google Scholar] [CrossRef]
  38. Mendoza, J.A.; Kim, H.K.; Park, H.K.; Park, K.Y. Photocatalytic Reduction of Carbon Dioxide Using Co3O4 Nanoparticles under Visible Light Irradiation. Korean J. Chem. Eng. 2012, 29, 1483–1486. [Google Scholar] [CrossRef]
  39. Bae, K.-L.; Kim, J.; Lim, C.K.; Nam, K.M.; Song, H. Colloidal Zinc Oxide-Copper(I) Oxide Nanocatalysts for Selective Aqueous Photocatalytic Carbon Dioxide Conversion into Methane. Nat. Commun. 2017, 8, 1156. [Google Scholar] [CrossRef]
  40. Bahadori, E.; Tripodi, A.; Villa, A.; Pirola, C.; Prati, L.; Ramis, G.; Rossetti, I. High Pressure Photoreduction of CO2: Effect of Catalyst Formulation, Hole Scavenger Addition and Operating Conditions. Catalysts 2018, 8, 430. [Google Scholar] [CrossRef]
  41. Chiarello, G.L.; Rossetti, I.; Forni, L. Flame-Spray Pyrolysis Preparation of Perovskites for Methane Catalytic Combustion. J. Catal. 2005, 236, 251–261. [Google Scholar] [CrossRef]
  42. Makuła, P.; Pacia Michałand Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV-Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [PubMed]
  43. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  44. Poongodi, S.; Kumar, P.S.; Masuda, Y.; Mangalaraj, D.; Ponpandian, N.; Viswanathan, C.; Ramakrishna, S. Synthesis of Hierarchical WO3 Nanostructured Thin Films with Enhanced Electrochromic Performance for Switchable Smart Windows. RSC Adv. 2015, 5, 96416–96427. [Google Scholar] [CrossRef]
  45. Rossetti, I.; Villa, A.; Pirola, C.; Prati, L.; Ramis, G. A Novel High-Pressure Photoreactor for CO2 Photoconversion to Fuels. RSC Adv. 2014, 4, 28883–28885. [Google Scholar] [CrossRef]
  46. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  47. Tan, H.L.; Abdi, F.F.; Ng, Y.H. Heterogeneous Photocatalysts: An Overview of Classic and Modern Approaches for Optical, Electronic, and Charge Dynamics Evaluation. Chem. Soc. Rev. 2019, 48, 1255–1271. [Google Scholar] [CrossRef]
  48. Yamada, Y.; Kanemitsu, Y. Determination of Electron and Hole Lifetimes of Rutile and Anatase TiO2 Single Crystals. Appl. Phys. Lett. 2012, 101, 133907. [Google Scholar] [CrossRef]
Figure 1. Type-II heterojunction between WO3 and TiO2.
Figure 1. Type-II heterojunction between WO3 and TiO2.
Catalysts 15 00383 g001
Figure 2. SEM images of (a,b) WO3, (c,d) WO3/TiO2 and (e,f) TiO2. Magnification: (left) ×5000, (right) ×10,000.
Figure 2. SEM images of (a,b) WO3, (c,d) WO3/TiO2 and (e,f) TiO2. Magnification: (left) ×5000, (right) ×10,000.
Catalysts 15 00383 g002
Figure 3. XRD of FSP catalysts with references (★ TiO2-anatase JCPDS no. 21-1272, ● TiO2-rutile JCPDS no. 21-1276, ■ WO3-monoclinic JCPDS no. 71-2141).
Figure 3. XRD of FSP catalysts with references (★ TiO2-anatase JCPDS no. 21-1272, ● TiO2-rutile JCPDS no. 21-1276, ■ WO3-monoclinic JCPDS no. 71-2141).
Catalysts 15 00383 g003
Figure 4. N2 physisorption curves of FSP catalysts.
Figure 4. N2 physisorption curves of FSP catalysts.
Catalysts 15 00383 g004
Figure 5. (Left) Diffuse reflectance (DR) UV–Vis spectra of FSP catalysts; (right) Tauc plots of FSP catalysts.
Figure 5. (Left) Diffuse reflectance (DR) UV–Vis spectra of FSP catalysts; (right) Tauc plots of FSP catalysts.
Catalysts 15 00383 g005
Figure 6. Productivity and HS conversion of tested catalysts.
Figure 6. Productivity and HS conversion of tested catalysts.
Catalysts 15 00383 g006
Table 1. Characterization data for tested catalysts.
Table 1. Characterization data for tested catalysts.
SampleBET SSA (m2/g)Total Pore Volume (cm3/g)BJH ads. Pore Width (nm)Crystallite Size (nm)Phase %BG (eV)
P2552.70.22518.715(A); 26(R)78(A); 22(R)3.22
FSP-TiO229.70.10914.520(A); 24(R)48(A); 52(R)3.10
FSP-WO33.40.01535.042/2.73
TiO2/WO3 60/40 wt29.20.0779.822(A); 16(R)45(A); 65(R)2.7
Table 2. Productivity and stored energy of tested catalysts—18 bar—pH 14.
Table 2. Productivity and stored energy of tested catalysts—18 bar—pH 14.
CatalystHS Conv.
(%)
HCOOH
(mol/kgcat h)
HCOOH
(g/kgcat h)
Total Stored
Energy (J) in 6 h
P254716.9778195
FSP TiO24013.9640160
FSP WO34314.7677169
FSP TiO2/WO3 60/40 7736.51680406
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tommasi, M.; Gramegna, A.; Degerli, S.N.; Galli, F.; Rossetti, I. High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy. Catalysts 2025, 15, 383. https://doi.org/10.3390/catal15040383

AMA Style

Tommasi M, Gramegna A, Degerli SN, Galli F, Rossetti I. High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy. Catalysts. 2025; 15(4):383. https://doi.org/10.3390/catal15040383

Chicago/Turabian Style

Tommasi, Matteo, Alice Gramegna, Simge Naz Degerli, Federico Galli, and Ilenia Rossetti. 2025. "High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy" Catalysts 15, no. 4: 383. https://doi.org/10.3390/catal15040383

APA Style

Tommasi, M., Gramegna, A., Degerli, S. N., Galli, F., & Rossetti, I. (2025). High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy. Catalysts, 15(4), 383. https://doi.org/10.3390/catal15040383

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop