Next Article in Journal
Broadening the Scope of Polyoxometalates as Artificial Proteases in Surfactant Solutions: Hydrolysis of Ovalbumin by Zr(IV)-Substituted Keggin Complex
Previous Article in Journal
Photoluminescent Coordination Polymers Based on Group 12 Metals and 1H-Indazole-6-Carboxylic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Activity of Multicompound TiO2/SiO2 Nanoparticles

1
Department of Chemistry, University of Eastern Finland, P.O. Box 111, FI-80101 Joensuu, Finland
2
Aerosol Physics Laboratory, Department of Physics, Tampere University, P.O. Box 692, FI-33101 Tampere, Finland
*
Author to whom correspondence should be addressed.
Submission received: 23 February 2021 / Revised: 19 March 2021 / Accepted: 21 March 2021 / Published: 25 March 2021
(This article belongs to the Section Inorganic Solid-State Chemistry)

Abstract

:
Multicompound TiO2/SiO2 nanoparticles with a diameter of 50–70 nm were generated using a liquid flame spray (LFS) nanoparticle deposition in a single flame. Here, we study the photocatalytic activity of deposited multicompound nanoparticles in gas-phase via oxidation of acetylene into carbon dioxide that gives new insight about the multicompound nanoparticle morphology. A small addition of SiO2 content of 0.5%, 1.0% and 3.0% significantly suppressed the photocatalytic activity by 33%, 44% and 70%, respectively, whereas 5.0% SiO2 addition completely removed the activity. This may be due to a formation of a thin passivating SiO2 layer on top of the of the TiO2 nanostructures during the LFS nanoparticle deposition. Surface wetting results support this hypothesis with a significant increase in water contact angle as the SiO2 content is increased.

1. Introduction

Accumulation of toxic compounds into ecosystem is a global and growing problem. Effective catalytic materials are needed to remove toxic, typically organic, compounds from wastewater to airborne pollution. Several technologies have been developed for solving this problem ranging from advanced oxidative technologies (AOT) [1] widely used in the wastewater management plants to solar-driven catalysis using novel multicompound metal and metal oxide based photocatalysts [2]. Photocatalysis is a promising approach for heterogeneous catalysis in which chemical reactions can be accelerated by the excitation of a catalyst by incident photons. The photocatalytic activity depends on the ability of the photocatalyst to form electron-hole pairs, which produce free radicals, which participate in secondary reactions [2].
Nowadays the most studied photocatalyst is titanium dioxide (TiO2). TiO2 is a well-known, non-toxic and widely used white pigment with good photostability [3]. In anatase crystalline form the bandgap between the valence and the conduction band is 3.2 eV that corresponds to a photon wavelength of 387 nm in the ultraviolet A (UVA) range. Photons with shorter wavelength (and thus higher energy) can excite electrons from the valence band to the conduction band: the generated electrons and holes can diffuse onto the surface and facilitate various reduction and oxidation reactions, respectively. In the seminal work by Fujishima and Honda in 1972 [4], water splitting to hydrogen and oxygen was demonstrated by irradiating TiO2 electrode with UVA light in a photochemical cell. The photocatalytic activity of TiO2 has been widely investigated ranging from controlled wettability of surfaces to selective conversion of CO2 into fuels (CH4, CH3OH, HCHO and HCOOH) [5] TiO2 nanostructures can also be used to generate hydrogen from water [6] and decomposition of organic molecules (pollutants) [7]. Additionally, the photocatalytic activity of TiO2 can be modified by doping TiO2 nanostructures with metals and non-metals [8,9,10].
There exists also a large variety of TiO2 based photocatalysts such as TiO2/SiO2, Al2O3-TiO2 and TiO2-zeolite. Introduction of additional elements in TiO2 system can enhance the photocatalytic activity. For example, 2% Al2O3-TiO2 nanocomposite [11] for selective degradation of imazapyr showed a two- to three-fold improvement of photocatalytic activity compared to mesoporous TiO2, whereas TiO2-zeolite photocatalysts from metakaolin and rice husk ash [12] showed an enhanced photoactivity compared to powder TiO2. Alternatively, TiO2/SiO2/graphene oxide (GO) photocatalyst [13] demonstrated enhanced visible light photoactivity for self-cleaning surfaces resulting from synergetic action of TiO2 and GO in which TiO2 produces electron-hole pairs and GO quickly transfer them to the surface for redox reactions. Many reports have been devoted to doping TiO2 doping with SiO2 such as TiO2/SiO2 nanocomposite [14,15] for efficient removal of organic pollutants. However, a reduced photocatalytic response of TiO2/SiO2 system [16] has also been observed since SiO2 can form blocking layer on top of TiO2 nanoparticles preventing electron and hole transport onto the nanoparticle surface.
In this paper we use a liquid flame spray (LFS) nanoparticle deposition that is versatile tool for depositing a large range of metal and metal oxide nanoparticles on various substrates [17,18]. The LFS process contains a high velocity and high temperature hydrogen-oxygen flame in which an organometallic precursor is fed. The precursor evaporates, nucleates and forms solid nanoparticles that can be collected on various substrates. The LFS process parameters such as gas flows, precursor feed rate and burner to substrate distance can easily be modified for a deposition of nanoparticles with highly controlled size and amount. Typically, the LFS deposition produces a highly porous network of interconnected nanoparticles with porosity up to 95% [19].
LFS deposited TiO2 nanoparticles are well suitable for photocatalysis. We have previously utilized LFS nanoparticle deposition for photocatalytic wettability conversion of TiO2 nanoparticle coated paperboard [20]. A detailed surface chemical analysis was carried out using both X-ray photoelectron spectroscopy [21] and time-of-flight secondary ion mass spectroscopy [22] that confirmed the role of hydrocarbons and hydroxyl groups generated on the surface via photocatalytic activation. The LFS deposited nanoparticles adhere onto the surface via rather weak van der Waals forces and, therefore, the LFS deposited nanoparticles can easily be removed by wear as has been shown earlier [23]. Multicompound nanomaterials can also be generated using LFS and Teisala et al. [24] showed that LFS generated binary particles sinter better when mixing some amount of SiO2 into TiO2. They showed that even a small fraction of SiO2 will make the LFS deposited TiO2 nanoparticles more stable, i.e., having more wear resistance. The reason for this is that the SiO2 helps particle sintering (neck-forming) or, alternatively, the SiO2 can coat or cover the TiO2 nanoparticles and, thus, support the deposited structure.
Traditionally photocatalytical activity has been characterized using an indirect method based on a color transformation of an optical dye, such as methylene blue (MB) [25]. However, such an indirect method should not be used for photocatalytic activity measurements; this is especially true at visible wavelengths as the incident light itself can bleach the used dye and, therefore, induce an erroneous result [26]. Moreover, during the sample immersion in water solution nanoparticles may detach and migrate from substrate into the solution. Therefore, a much better alternative to measure photocatalytic activity is based on a gas-phase oxidation of small organic compounds such (e.g., methane, acetylene) due to their relatively simple structures which can be easily broken. The mineralization of carbon compounds into CO2 can be directly and simply measured by monitoring the rate of CO2 increase in the reaction chamber. Moreover, the gas-phase detection removes all mechanical stresses induced by liquid phase methods on the nanostructures.
In this work, we report a simple fabrication method for deposition of TiO2/SiO2 multicompound nanoparticles in a single flame. The photocatalytic activity was characterized using an in-house built photoreactor based on decomposition of acetylene into CO2 [27,28]. The multicompound nanoparticles showed a small increase in the adhesion but increase of SiO2 content even up to 5% completely reduced the photocatalytic activity. The observed reduction of photocatalytic activity may be associated to a formation of a thin passivating layer on top of photocatalytically active TiO2 that prevents the diffusion of the generated electrons and holes onto the surface. A thin passivating layer is also supported by the observed water contact angles that are significantly increased in the presence of SiO2.

2. Results and Discussion

Five photocatalysts were produced by the LFS method with different TiO2 and SiO2 content (pure TiO2 and TiO2/SiO2 with ratios of 99.5/0.5, 99.0/1.0, 97.0/3.0 and 95.0/5.0, respectively). Surface morphology of all LFS nanoparticle deposited samples were analyzed by using scanning electron microscopy (SEM). Figure 1 displays the distribution of nanoparticles that is rather uniform for all depositions. Average nanoparticle diameter ranged from 40 nm to 80 nm with majority of particles within 50–70 nm range independent of the used precursor formulation that is in good agreement with our previous studies with similar LFS deposition parameters [20,21].
It can be concluded from Figure 1 that is difficult to see any significant effects in morphology caused by variation in the TiO2/SiO2 content. A small increase in the nanoparticle diameter (to 70–80 nm) with an increased aggregation tendency can be seen with increased SiO2 content. Figure 1f shows UV-Vis absorption spectrum of the samples. Increased SiO2 content is followed by a reduced TiO2 absorption in wavelength range of 200–300 nm. This is expected as SiO2 has a significantly lower absorption in the UV region compared to TiO2 and, thus, increased SiO2 content will reduce UV absorption of the multicompound nanostructures.
To study the actual particle formation photocatalytic activity is a great tool as it is highly surface sensitive, i.e., even a thin SiO2 layer on top of TiO2 can prevent the generated electrons and holes to diffuse onto sample surface. Hence, the expectation would be that the samples with an increasing SiO2 content will be photocatalytically less active in a linear manner than the pure TiO2. Therefore, it is crucial to determine the photocatalytic activity as a function of SiO2/TiO2 ratio but keeping the average particle size almost constant.
Figure 2 shows the measured photocatalytic activity of the multicompound LFS nanoparticles deposited on a glass substrate. Two glasses were used in the reactor chamber to increase the active surface area for improved response from the CO2 detector. Pure TiO2 nanoparticle resulted in the highest photocatalytic activity with CO2 increase of 26.2 ppm in 30 min. All multicompound TiO2/SiO2 samples showed a lower photocatalytic activity compared to the pure TiO2 sample. Additionally, increase of SiO2 content was followed by a lower CO2 increase. With 5.0% SiO2 content, no photocatalytic activity was observed. This is in contrary to our previous work [29], with multicompound nanoparticles used for controlled wettability on a paperboard in which up to 50% SiO2 content was required to convert the initially superhydrophobic surface into a hydrophilic one.
The rapid decrease of the photocatalytic activity by the increase of SiO2 content may be associated with the LFS nanoparticle formation process in the flame during deposition. Both precursor solutions were fed into the same flame simultaneously and the nanoparticle formation process with two precursors is not well understood [30,31]. It is possible that TiO2 nanoparticles nucleate first while SiO2 nanoparticles nucleate later [32]. Hence, the first formed TiO2 nanoparticles can provide a nucleation center for rapid covering layer of SiO2. As a result, multicompound LFS deposition may produce a nanocomposite consisting of a TiO2/SiO2 core-shell structure. Furthermore, SiO2 may also act as bridging agent between multicompound nanoparticles and between the substrate and the multicompound nanoparticles. The bandgap energy of SiO2 is 8.9 eV (corresponding to 139 nm in the deep UVC range) that requires significantly more energetic photons for photocatalytic excitation than anatase TiO2 with a bandgap of 3.2 eV (387 nm) in the UVA range. The observed photocatalytic activity results support the hypothesis of a core-shell particle formation with even a thin SiO2 layer on top of TiO2 nanoparticles that can prevent the diffusion of the excited electron-hole pairs onto the surface.
For a more detailed characterization of the multicompound nanoparticles scanning transmission electron microscopy (STEM) was used with energy-dispersive X-ray (EDS) mapping. Figure 3 shows the STEM images with the corresponding EDS profiles of multicompound TiO2/SiO2 nanoparticles. Figure 3a shows STEM images of pure TiO2 nanoparticles on microscope glass with average diameter of nanoparticles in range of 40–70 nm. Majority of nanoparticles have a spherical shape. The corresponding EDS profile verifies that nanoparticles consist of TiO2 (characteristic peaks at 4.5 and 5.0 keV). Signals of copper and aluminum originate from the sample holder grid (characteristic peaks at 0.9 and 1.4 keV). No SiO2 was observed as expected. Figure 3b,c display STEM images of TiO2/SiO2 99.5/1.5 and 99.0/1.0 samples, respectively. No significant changes were observed in the nanoparticle morphology but the EDS profile indicated presence of SiO2 characteristic peak (1.7 keV). Figure 3d shows the STEM image of TiO2/SiO2 97.0/3.0. The size of nanoparticles was in the same range (40-70 nm) but the EDS profile showed an increase of Si content. Finally, Figure 3e presents the STEM image of TiO2/SiO2 95.0/5.0. A small increase in the average diameter of nanoparticles (50–80 nm) was observed and the EDS profile displayed the highest amount of Si in all samples.
A tribology test with a cotton cloth square tip was performed to test adhesion of multicompound nanoparticles on glass substrate. The surfaces after tribological wear are presented on Figure 4 and show a more broken and smeared outlook with disordered nanoparticles. Agglomerated nanoparticles after tribological wear had a diameter of approximately 70–90 nm. Figure 4f shows the photographs of pure TiO2 nanoparticle coated surface before and after tribological wear. The tribologically worn area is clearly seen in the center of image with photocatalytically active coating reduced approximately by 50%. Hence, the photocatalytic activity is expected to be reduced by a factor of two.
Figure 5 shows photocatalytic activity results of TiO2/SiO2 multicompound nanoparticles after tribological wear. As shown on Figure 4f, approximately 50% of the active area was tribologically worn. The removal of the active area roughly by half reduces the photoactivity in average by two as shown on Figure 5 with remaining activity ranging from 52.6% (TiO2 sample) to 46.5% (TiO2/SiO2 97.0/3.0). The highest rate of photo oxidation of C2H2 was observed with pure TiO2 photocatalyst. As the content of SiO2 was increased, the photocatalytic activity was decreased to 8.9 ppm, 6.6 ppm, 3.7 ppm and 0 ppm SiO2 content of 0.5%, 1%, 3% and 5%, respectively.
The photocatalytic activation of TiO2 is facilitated by absorption of a photon with energy equal to, or greater than, the band gap of TiO2 (3.2 eV). This results in a generation of an electron–hole pair that can easily migrate onto the surface of the TiO2 nanoparticle (Equation (1)). The surface water molecules react with oxygen in Ti-O-Ti bond and form a hydroxyl ion and a proton (See Equation (3)). Electrons formed upon absorption of light can react with absorbed oxygen generating highly reactive oxygen species (ROS, Equation (4)) whereas holes can interact with hydroxyl ion producing hydroxyl ion radical (Equation (5)). Hydroxyl ion radicals react with absorbed acetylene that is followed by photodegradation into CO2 and H2O (Equation (6)). The overall reactions (1)–(6) are given below:
TiO2 + hν → TiO2 (ecb + h+vb),
TiO2 (ecb + h+vb) → TiO2* + heat
TiO2* + H2Oads → TiO2 + OH + H+
ecb + O2ads → O2
h+vb + OH → OH•
2OH• + C2H2 + 2O2 → 2CO2 + 2H2O
The surface chemistry of the deposited multicompound nanoparticles was also characterized by water contact angle (WCA) measurements. Figure 6 shows the measured WCA values of the TiO2/SiO2 multicompound nanoparticles deposited on glass. For multicompound nanoparticles, the lowest value of 11.8° was observed with 0.5 silica content that was increased up to 34.1° with 5% SiO2 content. These observations are in agreement with literature [33] of sol-gel deposited TiO2/SiO2 films that displayed a lower contact angle and increased photocatalytic degradation of MB with increased content of TiO2 phase. The observed increase in the WCA supports the formation of a thin passivating SiO2 layer on top of photocatalytically active TiO2 particles. Capillary water absorption coefficients [34] can provide additional information about the wettability of multicompound nanoparticle coated surfaces and we plan to return to this issue in a future communication.
The observed results are in contrast to our previous work with wettability of multicompound TiO2/SiO2 nanoparticles on paperboard [29] that showed wettability conversion approximately at 50% SiO2 increase, whereas here on glass even a 3% SiO2 content resulted in a significant increase of the WCA. It is worth emphasizing here that our previous study was carried out on a paperboard in which the deposited nanoparticles collect volatile hydrocarbons evaporated from the deposited paperboard substrate. On paperboard the wettability formed an S-shaped curve as a function of the TiO2 fraction (in fact, the ratio of Ti/(Si + Ti) content in the used precursor for LFS) when scanned through all the SiO2/TiO2 compositions. This meant that 0% and 10% of TiO2 deposits collected the least amount of hydrocarbons, whereas 90% and 100% of TiO2 deposits collect the most hydrocarbons. This was followed by the observed S-curve behavior from highly wetting state to highly non-wetting state around 50% TiO2 content. In this study, no volatile organic compounds were present as glass was used as a substrate.

3. Materials and Methods

3.1. Liquid Flame Spray (LFS) Multicompound Nanoparticle Deposition

Functional multicompound nanoparticles were generated using a liquid flame spray (LFS) nanoparticle deposition that allows a cost-efficient deposition of various metal and metal oxide nanoparticles in atmospheric conditions for large areas [17,18] as shown schematically on Figure 7a. LFS contains a high temperature and a high velocity flame in which an organometallic precursor evaporates, nucleates and forms solid nanoparticles of the final material. LFS multicompound nanoparticle deposition was carried out using single nozzle type burner and the microscope glass samples were attached in a rotating carousel for the nanoparticle deposition. The nozzle was placed 6 cm away from the sample surface. Hydrogen and oxygen were used for the combustion gases with gas flow rates of 50 L/min and 15 L/min, respectively. The precursor feed rate was set to 12 mL/min. As the combined total concentration of titanium and silicon was fixed to 50 mg/mL, this resulted in production rate of 600 mg of TiO2/SiO2 nanoparticles in minute. Ratio of Si was varied between 0 and 5% and a total of 5 different TiO2/SiO2 ratios were used.

3.2. Photocatalytic Activity Characterization by Gas-Phase Oxidation of Acetylene

An in-house built gas-phase reactor for photocatalytic activity characterization was constructed from stainless steel with a shell diameter of 145 mm and volume of 15.4 cm3 as shown on Figure 7b. A removable lid consists of UV transparent glass and flange connection with reactor shell. The reactor was equipped with CO2 detector (Vaisala GMP343, Vantaa, FI), temperature and humidity sensor (Thorlabs TSP01, Newton, NJ, US) and pressure meter (Wika PGT10, USB mode, Klingenberg, DE). All detectors were USB connected to the PC for a real-time measurement [27,28].
The photocatalytic activity measurement was carried out using acetylene (C2H2) and technical air that were supplied to the mixing chamber. The technical air played a role as a carrier of acetylene to the reaction chamber and proportion of technical air to acetylene was set to 3:1. Only gas-phase constituents were utilized in photocatalytic activity detection and there is no need for a buffer solution in gas-phase detection in contrast to cyclic voltammetry. The gases flow inside the reactor was continuous until the CO2 concentration becomes constant. Once the concentration of CO2 was stabilized, the valves were shut down and the concentration increase of photogenerated CO2 was monitored (a batch type reactor). The ultraviolet A (UVA) lamp (UVP Black-Ray® B-100AP High Intensity, Cambridge, UK, 100 W) with peak emission at 365 nm and intensity of approximately 20 mW/cm2 was located 17 cm above the reactor lid.
The detection method is based on oxidation of acetylene into CO2 and H2O. The main parameters of the photocatalytical reaction such as pressure and humidity during photocatalytical reaction did not change significantly and were assumed to be constant (40–50% and atmospheric pressure). An external ventilation was applied to remove excess heat from the UVA lamp and the overall temperature was kept in the range of 27–30 °C. After completion of the measurement, the valves can be opened and the reactor can be filled with technical air to remove all reaction components before the subsequent measurements. All photocatalytic activity results are an average of three, almost overlapping, parallel measurements that verifies the stability and reproducibility of both the used measurement system and the multicompound photocatalysts with no mechanical stresses induced on the nanostructured surface during the gas-phase detection.

3.3. Tribological Wear of the Nanoparticle Coated Microscope Glasses

Tribological wear test was performed using a tribometer (CSM+ Instruments Tribometer TRN S/N 18-347, Needham, MA, USA). The tribometer with a square tip (area of 1 × 1 cm2) covered with a cotton layer by double-side tape was used with two cycles with a 3 N force during 10 s. The tribologically worn area was approximately 50% of the deposited nanoparticle area as displayed on Figure 4. The friction data was collected with a ModelIX software (CSM+ Instruments, Needham, MA, USA).

3.4. Scanning and Scanning Transmission Electron Microscope (SEM/STEM) Imaging

The morphology and average structural sizes of TiO2 nanoparticles were observed with Hitachi S-4800 FE-SEM (field emission scanning electron microscope, Tokyo, Japan). Scanning electron microscope (SEM) images were acquired after cutting microscope slide to small squares (2 × 2 cm2).

3.5. Absorption Spectrum

A PerkinElmer Lambda 900 UV−vis/NIR spectrometer (Waltham, MA, USA) with a 150 mm integrating sphere was used to measure the absorption spectra of the multicompound nanoparticles.

3.6. Wettability Characterization

Water contact angle (WCA) measurements were carried out using a KSV Cam 200 (KSV Instruments Ltd., Helsinki, Finland) contact angle goniometer with 4 µL deionized water droplet at a room temperature. A Young–Laplacian fitting protocol was utilized to calculate the WCA value from the measured droplet profile.

4. Conclusions

We investigated LFS deposited TiO2/SiO2 multicompound nanoparticles for photocatalytic activity. The addition of SiO2 content did not improve nanoparticle adhesion but significantly reduced the photocatalytic activity, with the TiO2/SiO2 95.0/5.0 sample showing no photocatalytic activity. During a simultaneous deposition in the LFS flame, SiO2 was probably deposited on top of TiO2 nanoparticles, followed by significantly reduced photocatalytic activity. The deposition of SiO2 on top of TiO2 was also supported by the changes in the surface wettability.
Our future studies will involve multicompound metal (Ag or Au) combined with photocatalytic metal oxide (TiO2 or ZnO) nanoparticles in which even a visible light can activate the photocatalysis via plasmonic activation of the metallic nanomaterials. For improved adhesion of TiO2 nanoparticles on glass surface, a precursor layer will be studied. We believe that multicompound nanomaterials will find many applications in solar-driven photocatalysis and chemistry in future.

Author Contributions

F.T. carried out the photocatalytic measurements, data analysis and wrote the first draft of the manuscript. J.H. performed the LFS nanoparticle depositions. J.M.M. supervised the LFS nanoparticle depositions. J.J.S. coordinated the joint research and finalized the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Finnish Cultural Foundation for a research grant to F.T. J.J.S. acknowledges the Faculty of Science and Forestry at the University of Eastern Finland for the financial support (grant no. 579/2017) and the Academy of Finland Flagship for Photonics Research and Innovation (PREIN, decision no. 320166).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article here. The data presented in this study are also available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nickheslat, A.; Amin, M.M.; Izanloo, H.; Fatehizadeh, A.; Mousavi, S.M. Phenol Photocatalytic Degradation by Advanced Oxidation Process under Ultraviolet Radiation Using Titanium Dioxide. J. Environ. Res. Public Health 2013, 2013, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zaleska-Medynska, A.; Gołąbiewska, A.; Kobylański, M.P. Metal Oxide-Based Photocatalysis. Fundamentals and Prospects for Application. In Metal Oxide-Based Photocatalysis; Korotcenkov, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 3–50. [Google Scholar]
  3. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  4. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37. [Google Scholar] [CrossRef]
  5. Ran, J.; Jaroniec, M.; Qiao, S.-Z. Cocatalysts in semiconductor-based photocatalytic CO2 reduction: Achievements, challenges, and opportunities. Adv. Mater. 2018, 30, 1704649. [Google Scholar] [CrossRef]
  6. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  7. Karapati, S.; Giannakopoulou, T.; Todorova, N.; Boukos, N.; Dimotikali, D.; Trapalis, C. Eco-efficient TiO2 modification for air pollutants oxidation. Appl. Catal. B Environ. 2015, 176, 578–585. [Google Scholar] [CrossRef]
  8. Khairy, M.; Zakaria, W. Effect of metal-doping of TiO2 nanoparticles on their photocatalytic activities toward removal of organic dyes. Egypt. J. Pet. 2014, 23, 419–426. [Google Scholar] [CrossRef] [Green Version]
  9. Yalç, Y.; Murat, K.; Zekiye, Ç. The Role of Non-Metal Doping in TiO2 Photocatalysis. J. Adv. Oxid. Technol 2010, 13, 281–296. [Google Scholar]
  10. Quesada, J.; Arreola-Sánchez, R.; Faba, L.; Díaz, E.; Rentería-Tapia, V.M.; Ordóñez, S. General Effect of Au nanoparticles on the activity of TiO2 for ethanol upgrading reactions. Appl. Catal. A 2018, 551, 23–33. [Google Scholar] [CrossRef]
  11. Ismail, A.A.; Abdelfattah, I.; Faycal Atitar, M.; Robben, L.; Bouzid, H.; Al-Sayari, S.A.; Bahnemann, D.W. Photocatalytic degradation of imazapyr using mesoporous Al2O3–TiO2 nanocomposites. Separ. Purif. Technol. 2015, 145, 147–153. [Google Scholar] [CrossRef]
  12. Setthaya, N.; Chindaprasirt, P.; Yin, S.; Pimraksa, K. TiO2-zeolite photocatalysts made of metakaolin and rice husk ash for removal of methylene blue dye. Powder Technol. 2017, 313, 417–426. [Google Scholar] [CrossRef]
  13. Gao, J.; Li, W.; Zhao, X.; Wang, L.; Pan, N. Durable visible light self-cleaning surfaces imparted by TiO2/SiO2/GO photocatalyst. Textile Res. J. 2019, 89, 517–527. [Google Scholar] [CrossRef]
  14. Nabih, S.; Esmail Shalan, A.; Samy Abu Serea, E.; Goda, M.A.; Fathi Sanad, M. Photocatalytic performance of TiO2@SiO2 nanocomposites for the treatment of different organic dyes. J. Mat. Sci. Mater. Electron. 2019, 30, 9623–9633. [Google Scholar] [CrossRef]
  15. Bellardita, M.; Addamo, M.; Di Paola, A.; Marcì, G.; Palmisano, L.; Cassar, L.; Borsa, M. Photocatalytic activity of TiO2/SiO2 systems. J. Hazard. Mater. 2010, 174, 707–713. [Google Scholar] [CrossRef]
  16. Xu, L.; Shen, Y.; Ding, Y.; Wang, L. Superhydrophobic and ultraviolet-blocking cotton fabrics based on TiO2/SiO2 composite. Nanopart. J. Nanosc. Nanotechnol. 2018, 18, 6879–6886. [Google Scholar] [CrossRef] [PubMed]
  17. Mäkelä, J.M.; Aromaa, M.; Teisala, H.; Tuominen, M.; Stepien, M.; Saarinen, J.J.; Toivakka, M.; Kuusipalo, J. Nanoparticle Deposition from Liquid Flame Spray onto Moving Roll-to-Roll Paperboard Material. J. Aerosol Sci. 2011, 68, 827–837. [Google Scholar] [CrossRef] [Green Version]
  18. Mäkelä, J.M.; Haapanen, J.; Harra, J.; Juuti, P.; Kujanpää, S. Liquid Flame Spray—A Hydrogen-Oxygen Flame Based Method for Nanoparticle Synthesis and Functional Nanocoatings. KONA Powder Part. J. 2017, 34, 141–154. [Google Scholar] [CrossRef] [Green Version]
  19. Mädler, L.; Roessler, A.; Pratsinis, S.E. Direct formation of highly porous gas-sensing films by in situ thermophoretic deposition of flame-made Pt/SnO2 nanoparticles. Sens. Actuators B Chem. 2006, 114, 283–295. [Google Scholar] [CrossRef]
  20. Stepien, M.; Saarinen, J.J.; Teisala, H.; Tuominen, M.; Aromaa, M.; Kuusipalo, J.; Mäkelä, J.M.; Toivakka, M. Adjustable wettability of paperboard by liquid flame spray nanoparticle deposition. Appl. Surf. Sci. 2011, 257, 1911–1917. [Google Scholar] [CrossRef]
  21. Stepien, M.; Saarinen, J.J.; Teisala, H.; Tuominen, M.; Aromaa, M.; Kuusipalo, J.; Mäkelä, J.M.; Toivakka, M. Surface chemical analysis of photocatalytic wettability conversion of TiO2 nanoparticle coating. Surf. Coat. Technol. 2012, 208, 73–79. [Google Scholar] [CrossRef]
  22. Stepien, M.; Saarinen, J.J.; Teisala, H.; Tuominen, M.; Aromaa, M.; Haapanen, J.; Kuusipalo, J.; Mäkelä, J.M.; Toivakka, M. ToF-SIMS analysis of UV-switchable TiO2-nanoparticle coated paper surface. Langmuir 2013, 29, 3780–3790. [Google Scholar] [CrossRef]
  23. Stepien, M.; Saarinen, J.J.; Teisala, H.; Tuominen, M.; Haapanen, J.; Kuusipalo, J.; Mäkelä, J.M.; Toivakka, M. Compressibility of porous TiO2 nanoparticle coating on paperboard. Nanoscale Res. Lett. 2013, 8, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Teisala, H.; Geyer, F.; Haapanen, J.; Juuti, P.; Mäkelä, J.M.; Vollmer, D.; Butt, H.J. Ultrafast processing of hierarchical nanotexture for a transparent superamphiphobic coating with extremely low roll-off angle and high impalement pressure. Adv. Mater. 2018, 30, 1706529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Xu, C.; Rangaiah, G.P.; Zhao, X.S. Photocatalytic degradation of methylene blue by titanium dioxide: Experimental and modeling study. Ind. Eng. Chem. Res. 2014, 53, 14641–14649. [Google Scholar] [CrossRef]
  26. Yan, X.; Ohno, T.; Nishijima, K.; Abe, R.; Ohtani, B. Is methylene blue an appropriate substrate for a photocatalytic activity test? A study with visible-light responsive titania. Chem. Phys. Lett. 2006, 429, 606–610. [Google Scholar] [CrossRef] [Green Version]
  27. Temerov, F.; Ankudze, B.; Saarinen, J.J. TiO2 inverse opal structures with facile decoration of precious metal nanoparticles for enhanced photocatalytic activity. Mater. Chem. Phys. 2020, 242, 122471. [Google Scholar] [CrossRef]
  28. Pham, K.; Temerov, F.; Saarinen, J.J. Multicompound inverse opal structures with gold nanoparticles for visible light photocatalytic activity. Mater. Des. 2020, 194, 108886. [Google Scholar] [CrossRef]
  29. Haapanen, J.; Aromaa, M.; Teisala, H.; Tuominen, M.; Stepien, M.; Saarinen, J.J.; Heikkilä, M.; Toivakka, M.; Kuusipalo, J.; Mäkelä, J.M. Binary TiO2/SiO2 nanoparticle coating for controlling the wetting properties of paperboard. Mater. Chem. Phys. 2015, 150, 230–237. [Google Scholar] [CrossRef]
  30. Keskinen, H.; Mäkelä, J.M.; Aromaa, M.; Ristimäki, J.; Kanerva, T.; Levänen, E. Effect of silver addition on the formation and deposition of titania nanoparticles produced by liquid flame spray. J. Nanopart. Res. 2007, 9, 569–588. [Google Scholar] [CrossRef]
  31. Ulrich, G.D. Theory of Particle Formation and Growth in Oxide Synthesis Flames. Combust. Sci. Technol. 1971, 4, 47–57. [Google Scholar] [CrossRef]
  32. Teleki, A.; Pratsinis, S.E.; Wegner, K.; Jossen, R. Flame-coating of titania particles with silica. J. Mater. Res. 2005, 20, 1336–1347. [Google Scholar] [CrossRef] [Green Version]
  33. Kwon, C.H.; Kim, J.H.; Jung, I.S.; Shin, H.; Yoon, K.H. Preparation and characterization of TiO2–SiO2 nano-composite thin films. Cer. Int. 2003, 29, 851–856. [Google Scholar] [CrossRef]
  34. Kapridaki, C.; Maravelaki-Kalaitzaki, P. TiO2–SiO2–PDMS nano-composite hydrophobic coating with self-cleaning properties for marble protection. Prog. Org. Coat. 2013, 76, 400–410. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) TiO2, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0, (e) TiO2/SiO2 95.0/5.0 and (f) UV-Vis absorption spectra of the corresponding samples.
Figure 1. SEM images of (a) TiO2, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0, (e) TiO2/SiO2 95.0/5.0 and (f) UV-Vis absorption spectra of the corresponding samples.
Inorganics 09 00021 g001
Figure 2. Measured photocatalytic activity of (a) TiO2 100% and TiO2/SiO2 multicompound nanoparticles with (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0.
Figure 2. Measured photocatalytic activity of (a) TiO2 100% and TiO2/SiO2 multicompound nanoparticles with (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0.
Inorganics 09 00021 g002
Figure 3. STEM images and EDS profiles of (a) TiO2 100%, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0).
Figure 3. STEM images and EDS profiles of (a) TiO2 100%, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0).
Inorganics 09 00021 g003
Figure 4. SEM images of (a) TiO2, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0 after tribological wear. Photographs of the nanoparticle deposited microscope glass before and after the tribological wear are shown in (f).
Figure 4. SEM images of (a) TiO2, (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0 after tribological wear. Photographs of the nanoparticle deposited microscope glass before and after the tribological wear are shown in (f).
Inorganics 09 00021 g004
Figure 5. Measured photocatalytic activity of (a) TiO2 100% and TiO2/SiO2 multicompound nanoparticles with (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0 after tribological wear.
Figure 5. Measured photocatalytic activity of (a) TiO2 100% and TiO2/SiO2 multicompound nanoparticles with (b) TiO2/SiO2 99.5/0.5, (c) TiO2/SiO2 99.0/1.0, (d) TiO2/SiO2 97.0/3.0 and (e) TiO2/SiO2 95.0/5.0 after tribological wear.
Inorganics 09 00021 g005
Figure 6. Water contact angle images of (a) TiO2/SiO2 99.5/0.5, (b) TiO2/SiO2 99.0/1.0, (c) TiO2/SiO2 97.0/3.0 and (d) TiO2/SiO2 95.0/5.0. 100% TiO2 sample completely wetted the surface with WCA close to 0.
Figure 6. Water contact angle images of (a) TiO2/SiO2 99.5/0.5, (b) TiO2/SiO2 99.0/1.0, (c) TiO2/SiO2 97.0/3.0 and (d) TiO2/SiO2 95.0/5.0. 100% TiO2 sample completely wetted the surface with WCA close to 0.
Inorganics 09 00021 g006
Figure 7. Schematic picture of (a) LFS nanoparticle deposition and (b) the used gas-phase detector for photocatalytic activity characterization.
Figure 7. Schematic picture of (a) LFS nanoparticle deposition and (b) the used gas-phase detector for photocatalytic activity characterization.
Inorganics 09 00021 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Temerov, F.; Haapanen, J.; Mäkelä, J.M.; Saarinen, J.J. Photocatalytic Activity of Multicompound TiO2/SiO2 Nanoparticles. Inorganics 2021, 9, 21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040021

AMA Style

Temerov F, Haapanen J, Mäkelä JM, Saarinen JJ. Photocatalytic Activity of Multicompound TiO2/SiO2 Nanoparticles. Inorganics. 2021; 9(4):21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040021

Chicago/Turabian Style

Temerov, Filipp, Janne Haapanen, Jyrki M. Mäkelä, and Jarkko J. Saarinen. 2021. "Photocatalytic Activity of Multicompound TiO2/SiO2 Nanoparticles" Inorganics 9, no. 4: 21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040021

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