Catalysis Research is an international peer-reviewed Open Access journal published quarterly online by LIDSEN Publishing Inc. This periodical is devoted to publishing high-quality papers that describe the most significant and cutting-edge research in all areas of catalysts and catalyzed reactions. Its aim is to provide timely, authoritative introductions to current thinking, developments and research in carefully selected topics.

Topics contain but are not limited to:

  • Photocatalysis
  • Electrocatalysis
  • Environmental catalysis
  • Biocatalysis, enzymes, enzyme catalysis
  • Catalysis for biomass conversion
  • Organocatalysis, catalysis in organic and polymer chemistry
  • Nanostructured Catalysts
  • Catalytic materials
  • Computational catalysis
  • Kinetics of catalytic reactions

It publishes a variety of article types: Original Research, Review, Communication, Opinion, Comment, Conference Report, Technical Note, Book Review, etc.

There is no restriction on paper length, provided that the text is concise and comprehensive. Authors should present their results in as much detail as possible, as reviewers are encouraged to emphasize scientific rigor and reproducibility.

Publication Speed (median values for papers published in 2023): Submission to First Decision: 4.1 weeks; Submission to Acceptance: 13.0 weeks; Acceptance to Publication: 9 days (1-2 days of FREE language polishing included)

Current Issue: 2024  Archive: 2023 2022 2021
Open Access Original Research

In Water High Yield and Selectivity of CH4 and H2 Production from CO2 Using UVC Light and a SiO2-surface-modified TiO2 Photocatalysts

Oswaldo Núñez 1,2,*, Mohammad Fereidooni 1, Victor Márquez 1, Duangthip Sattayamuk 1, Piyasan Praserthdam 1, Supareak Praserthdam 3

  1. Center of Excellence on Catalysis and Catalytic Reaction Engineering. Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand

  2. Laboratory of Physical Organic Chemistry and Environmental Chemistry, Department of Processes and Systems, Universidad Simón Bolívar (USB), Caracas 1080A, Venezuela

  3. Performance Computing Unit (CECC-HCU), Center of Excellence on Catalysis and Catalytic Reaction Engineering (CECC), Chulalongkorn University, Bangkok 10330, Thailand

Correspondence: Oswaldo Núñez

Academic Editor: Anna Zielińska-Jurek

Special Issue: Green Photocatalysis for Degradation of Emerging Contaminants and Hydrogen Production

Received: August 25, 2023 | Accepted: November 28, 2023 | Published: December 05, 2023

Catalysis Research 2023, Volume 3, Issue 4, doi:10.21926/cr.2304029

Recommended citation: Núñez O, Fereidooni M, Márquez V, Sattayamuk D, Praserthdam P, Praserthdam S. In Water High Yield and Selectivity of CH4 and H2 Production from CO2 Using UVC Light and a SiO2-surface-modified TiO2 Photocatalysts. Catalysis Research 2023; 3(4): 029; doi:10.21926/cr.2304029.

© 2023 by the authors. This is an open access article distributed under the conditions of the Creative Commons by Attribution License, which permits unrestricted use, distribution, and reproduction in any medium or format, provided the original work is correctly cited.

Abstract

To improve CH4 and H2 formation from CO2 photoreduction using non-organic, non-laborious, and inexpensive photocatalysts, we have prepared two surface-silicate-modified TiO2 catalysts: P25-SiO2 and AmTiO2-SiO2 (amorphous TiO2) to be tested in water and using UV light. The last catalyst produces more CH4 and H2 in water than P25 (3:1 TiO2 anatase: rutile) under UV light irradiation of HCO3- and CO2; am-TiO2-SiO2 at pH = 7, produces 8 times more CH4 and H2 than P25 with selectivity at the reactor headspace of 30% and 53%, respectively. Using CO2 (pH = 3), 80 times more CH4 than P25 under the same conditions is obtained with a yield of 71%. This corresponds to a production of 8.9 μmol gcat-1·h-1, one of the highest reported rates of CH4 production from CO2 using carbon-free semiconductors. H2 is also produced by water splitting using Am-TiO2-SiO2 and water at low pH. The enhanced reactivity compared to P25 is attributed to three main factors: a) Low catalyst PZC (4.1) that facilitates CO2 adsorption and proton availability at the active site to catalyze the e transfer from Ti at the TiO2-SiO2-carbonate adduct b) SiO2 acts as electron trap reducing carriers recombination (External intramolecular trapping (EIT) mechanism) and c) am-TiO2-SiO2, light collection efficiency, surface area and irregular atoms distribution. Catalysts were also tested for Methylene blue (MB) photooxidation. P25 is quite a better catalyst in oxidizing MB via OH radicals, probably due to the more positive valence band potentials in the SiO2-modified catalysts that avoid the OH radical formation from water; however, when bicarbonate is added to MB solution, am-TiO2-SiO2 catalysts reactivity increases as a consequence of its valence band down-bending.

Graphical abstract

Click to view original image

Keywords

Amorphous TiO2-SiO2; CO2/HCO3- photoreduction; H2 and CH4 selectivity; electron traps; cyclic voltamnetry; open circuit potential decay; point of zero charge; OH radical formation inhibition

1. Introduction

Production of methane and hydrogen through CO2 photoreduction is of great interest not only from the environmental point of view but also due to the generation of combustible gases that can be converted into value-added products such as Methanol. The last conversion could be achieved via thermo-catalysis with a low performance/price ratio. However, this conversion can be performed via photo-catalytic oxidation of low-concentration methane over decorated graphitic carbon nitride [1,2]. H2 and CH4 are sources of energy storage that can also be produced from CO2. As suggested, co-mixing H2 and CH4 may double the energy storage capacity compared with pure hydrogen or methane storage [3]. Therefore, more energy can be obtained by storing the gases produced by CO2 photo reduction with UV/solar light. CO2 may then be considered as a feedstock. Large surface areas of 2D materials and conjugated microporous polymers (CMPs) allow the storage of combustibles such as H2 and CH4 and their energy storage conversion [4,5].

In water, CO2 photo reduction conversion to CH3OH, HCOOH, and CH3COOH using visible light and 2D/2D zinc vanadium oxide-reduced graphene oxide (ZnV2O6/RGO) nanosheets has been recently reported; up to 5154 μmolgcat-1 of CH3OH is produced [6]. This efficiency is attributed to the trapping and transporting of electrons by reduced graphene oxide (RGO). Also, 3488 μmolgcat-1 CH3OH production has been obtained using ZnV2O6/RGO/g-C3N4 composite [7].

Reports on CH4 and CO production of 225 μmolgcat-1h-1 from CO2 UV-photo reduction using MXene/Clay (2D/2D/2D O-C3N4/Bt/Ti3C2Tx) composite and up to 145 and 133 μmolgcat-1 h-1 with WO3/g-C3N4 and visible light represent quite high yields results [8,9]. CO2 photoreduction using pg/C3N4/Ti3AlC3/TiO2 composite has been reported to yield 91.9 and 4.97 mmolgcat-1h-1 [10]. MXene based cocatalyst: TiO2A/Ti3C2/TiO2R MXene yield85 and 18 mmolgcat-1h-1 of CO and CH4, respectively [11]. S-scheme has been proposed for producing 2103.5 μmolgcat-1h-1 of CH4 with 96.59% selectivity using g-C3N4/TiO2/TiAlC22D/OD/2D composite [12]. However, results using carbon-based semiconductors as g-C3N4 to generate CH4 from CO2 must be carefully interpreted since some CH4 formed may be produced by the reduction of carbon source in the catalyst. Therefore, control experiments without CO2 must be conducted to verify the non-formation of Methane or Methanol under these conditions [13,14,15]. Other studies reported isotopic labeling (13C and 18O) to evaluate carbon and oxygen sources unequivocally [16,17,18,19].

Non-carbon-based photocatalysts such as ZnFe2O4/Ag/TiO2 nanorods have been successfully used to produce CO, CH4 and CH3OH via UV/visible photo reduction of CO2 [20]; up to 1566 μmolgcat-1 CO yield has been obtained using 3D MAX Ti3AlC2/TiO2 in a monolith photo reactor where efficient formation of CH4 via reforming is also reported [21].

SiO2-TiO2 materials are used in very different fields [22]. As a photocatalyst, CO2 photo reduction enhancements have been reported. However, in the cases said TiO2 is either embedded in SiO2 matrices or anchored on glass [23,24]. In fact, in these cases, photoredox reactions are faster than TiO2 since the band gap is relatively larger (up to 4.11 eV) than the TiO2 one (3.2 eV); therefore, the CB (Conduction band) reduction and VB (Valence Band) oxidation potentials increase. Nevertheless, SiO2 coating TiO2 P25 has also been used to reduce CO2 to CO. Only CO and H2 were detected, and CO selectivity increased compared to unmodified P25 [25]. Toluene degradation using novel mesoporous SiO2-doped TiO2 nanofibers has been reported [26]. Zirconium-doped TiO2/SiO2 and visible light have been used to degrade VOC, including Toluene [27]. Simulated-sunlight-activated photocatalysis of Methylene Blue (MB) using cerium-doped SiO2/TiO2 nanostructured fibers has also been reported [28].

Crystalline TiO2 modified with tailored nanostructures has been extensively studied [29]; however, many properties associated with semiconductors’ atom arrangements influence their photocatalytic activity. Therefore, using amorphous TiO2 (am-TiO2) may bring essential advantages due to the porosity (high specific area), its high light-harvesting capacity, and the atom disorder that might induce charge separation. C-doped am-TiO2 has been synthesized and used in the photo reduction of CO2, obtaining CH4 production of 4.1 μmolgcat-1h-1 [30]; this catalyst showed a high specific surface area and narrowed band gap of 2.1 eV. Additionally, it has been pointed out that introducing oxygen vacancies on the C-doped am-TiO2 can efficiently accelerate charge separation and transportation of photo-generated hole-electron pairs, enhancing photocatalytic performance [31]. However, am-TiO2 has been investigated less than its crystalline forms, applications that do not require high temperatures have been reported. For instance, its antibacterial capacity has been studied [32]. Density-functional theory (DFT) calculations of structural and electronic properties of am-TiO2 as a photocatalyst for hydrogen production have been published [33]. Mesostructured SiO2-TiO2 mixed oxides with enhanced stability, textural properties, and surface area in the 290-346 m2/g range have been synthesized [34].

We have identified some relevant aspects that may help to increase the CH4 formation rate from HCO3-/CO2 UV light photo reduction [35]. For instance, CO2/HCO3- adsorption and the corresponding adduct concentration increases at pH equal and above P25 point of zero charge (PZC = 6.2) [36]; however, electron transfer from Ti to form the radical anion Ti+4-OCO2- requires protonation, that is also needed to produce H2 via water splitting. Therefore, a TiO2 surface modification must be oriented to increase the negative charge on the surface, but at the same time, more protonation capacity than TiO2 at the catalyst surface is required. CO formation contribution from the oxygen vacancy (Vo) sites must also be minimized to increase CH4 formation during the photocatalytic reaction.

To fulfill the three previous aspects, we have modified am-TiO2 by attaching SiO2 to its surface at room temperature and pressure. Since SiOH is a more vital Bronsted acid than TiOH, PZC of the SiOH at the surface-modified TiO2 must be lower than unmodified TiO2; therefore, at neutral pH, the catalyst surface must be negatively charged, which may improve CO2 adsorption and its photoreduction rate; furthermore, to orient the photoreduction to more CH4 and H2 formation, the proton must also be available (SiOH) at the catalyst active site; therefore, using this surface-SiO2 modified TiO2 catalyst, CH4 formation rate is expected to improve as compared to unmodified TiO2 at neutral pH. Further rate improvement is likely when pH changes to the new motivation PZC value. The amorphous condition of the synthesized catalyst may also improve the CO2 photoreduction rate due to increased surface area, light harvesting, and decreased e/h recombination.

In this work, amorphous TiO2 was synthesized via the sol-gel method, and its surface was modified by its addition to a tetraethyl orthosilicate acid solution. Complete characterization of the solid product obtained is reported. It includes SEM-EDX, TEM, DRS, FT-IR, BET, XPS, XRD, PZC measurement, and Raman spectroscopy. Additionally, catalyst (supported on ITO) current density, de-trapping capacity, surface potential, and electron transfer processes to the electrolyte are also reported using cyclic voltammetry (CV) and open circuit potential decay (OCPD) measurements.

Our approach is to introduce modifications of catalysts without using organic compound additives and as simple as possible to increase CH4 and H2 production and reduce the cost of catalyst preparation. Herein, we report essential CH4 and H2 production improvement from CO2/HCO3- using silicate surface modified amorphous TiO2 and UVC (200-280 nm) light. A necessary amount of H2 is also produced from water splitting. Some insights on the reaction mechanism obtained using electrochemistry and photooxidation measurements are also reported, pointing to the role of the (d-p) π-bonding at the Si-O bond and the non-formation of hydroxyl radical at the catalyst valence band.

2. Materials and Methods

2.1 Materials

The following chemicals and solvents were used: Degussa (P25, Aeoxide); sodium hydroxide (NaOH, 99.0%, Emsure); hydrochloric acid (HCl, 37%, Ensure); sodium phosphate monobasic dehydrate (NaH2PO4 2H2O, 99.0%, Acros Organics); sodium hydrogen carbonate (NaHCO3, 99.0%, Fluka Chemika); acetic acid (CH3COOH, 100%, Merck); titanium (IV) tetraisopropoxide (Ti[OCH(CH3)2]4, 98%, Aldrich); tetraethyl orthosilicate (Si(OC2H5)4, 98%, Aldrich); methylene blue (Sigma-Aldrich); Ethyl Alcohol (C2H6O, 99.9% QRëC) and water (H2O, 17.7 MΩ/cm, Nanopure). The following gases were used: methane (CH4, 99.99%, Linde); hydrogen (H2); carbon monoxide (CO, 99.99%, Boc Gases); carbon dioxide (CO2, 99.99%, Linde); helium (He, >99.99%, Linde); argon (Ar, >99.99%, Linde); nitrogen (N2, >99.99%, Linde); air (>99.99%, Bangkok Industrial Gas).

2.2 Characterization, Photo Reduction, and Gasses Concentration Measurements

Semiconductor morphologies and crystalline structures were obtained by Scanning electron microscopy and energy-dispersive X-ray Spectroscopy (SEM-EDX). An Amicus photoelectron spectrometer with an Mg Kα X-ray source (20 mA and 10 keV) controlled by Kratos Vision2 software. C 1 s peak at 284.6 eV was used for calibration. Using a Lamda 650 UV/Vis instrument. A JEOL JEM-2100 Electron Microscope was used to obtain the Transmission Electron Microscope (TEM) images. Diffuse Reflectance UV-Vis Spectra (DRS) was collected using UV Win Lab software from PerkinElmer. Using Tauc’s equation, (αhν)1/n = A(hν-Eg), the semiconductors band gap (Eg) was obtained. In the last equation, α is the absorption coefficient, h is Planck’s constant, ν and λ, light frequency and wavelength, A is an optical constant, and n = 0.5 for direct transitions. FT-IR Reflectance Spectroscopy (FT-IR) spectra were obtained in a Nicolet 6700 FT-IR instrument. A Horiba 4P-Fluoromax spectrofluorometer with a Xenon lamp (200-800 nm) was used to obtain the corresponding Photoluminescence (PL) spectra. The specific surface area of the catalysts was estimated by the Brunauer-Emmett-Teller (BET) method using a model Asap 2020, Micrometrics instrument. The point of zero charge (PZC) was obtained by plotting a solid catalyst pHi-pHf Vs in water. pHi where pHi and pHf are initial and final-equilibrium pHs; a Metter Toledo Seven multi pH-meter was used. Surface quantitative analysis was obtained using X-ray Photoelectron Spectroscopy (XPS) using an Amicus spectrometer at a voltage of 15 kV and current of 12 mA with Mg Kα X-ray radiation (1253.6 eV) and Al Kα X-ray radiation (1486.6 eV). Crystal structures were determined by X-ray Diffraction (XRD) using a Bruker D8 Advance instrument using a data collection (0.05°s-1) at a 20°-80° angle range. The crystallite size was calculated using the Debye-Scherrer equation. Raman spectra were obtained using a Horiba Scientific LabSpec 6 instrument.

Photoreduction was performed in a cylindrical quartz reactor surrounded by six 16-watt Philips Germicidal Ozone UVC (200-280 nm) bulbs. A Shimadzu GC-14A instrument was used to quantify CH4 reactor headspace with a FID detector provided with a Shincarbon ST packed column at 70°C in 10 min of analysis and using He as carrier gas and air as make-up gas with injector and detector temperatures of 150°C. H2 and CO were measured using a Shimadzu GC-8A instrument with a TCD detector and molecular sieve 5A column at 70°C for 18 min. Ar was used as carrier gas with injector and detector temperatures of 150°C.

2.3 Synthesis

2.3.1 Amorphous TiO2 (Am-TiO2)

Following the sol-gel method, 5 mL of acetic acid in 50 mL of Ethanol (95%) solution previously stirred for 30 min and 6.3 mL (0.021 mol) of titanium tetraisopropoxide were added dropwise while stirring for 15 min. Stirring was continued for one hour [31]. 10 ml of the above mixture was left to dry in the air at room temperature. White powder of TiO2 was collected and used directly to obtain am-TiO2-SiO2.

2.3.2 Silicate-Surface Modified TiO2 Catalysts (am-TiO2-SiO2)

While stirring to a solution of 35 ml of Ethanol and 35 ml of distilled water, microliters of HCl solution (35%) were added until pH = 3 was reached. Then, (5 mmol) of tetraethyl ortho silicate was added to the solution. Stirring was continued for 6 h. 2 g of catalyst (P25 or am-TiO2) was added to the final mixture; stirring continued for 1 h. The solution was dried at room temperature for 24 h. Catalyst-SiO2 was air-dried at 110°C for 15 min.

2.3.3 Reuse am-TiO2-SiO2

Catalysts am-TiO2-SiO2 used for CO2 photo reduction (pH = 3) were removed from the reactor, and the water-catalyst mix evaporated at 90°C until the dry catalyst was obtained (0.2 g). The 0.3 g of catalyst needed for the reaction was completed with the catalyst used at pH = 7. The obtained powder was used directly as a reused catalyst for CO2 photoreduction.

2.3.4 Catalysts Supported on ITO

The ITO (Indium Tin Oxide) electrodes were treated with a KOH 0.05% solution and rotated at 1000 RPM to remove the excess of the critical two times before the catalysts' coating. A suspension of P25-SiO2 catalyst in ethanol was prepared using 20 g of catalysts and 20 mL of ethanol and sonicated for 3 minutes (6 rounds, 30 s each time). am-TiO2-SiO2 gel was directly added to ITO. The ITO Electrodes were entirely covered with the suspension/gel and rotated at 1000 RPM to remove the excess material. The covered electrodes were air-dried and then treated at 150°C for 10 minutes. The Catalysts supported on ITO were stored in a closed flask for later use.

2.4 CO2/HCO3- Photo Reduction

A 250 mL three-way cylindrical quartz reactor provided with a sand sparger for bubbling gas into the solution was used. Into this flask, a solution of [CO2]w = 34 mmol/100 mL by bubbling CO2 on 100 mL of water or [HCO3-] = 2.3 mmol/100 mL and 0.3 g silicate surface-modified TiO2 catalyst was added. pH (3-8) was adjusted using HCl, NaOH or 100 mg/L of NaH2PO4 buffer solutions. The reaction was started by turning on the light while stirring. Irradiation was continued for 4 hours. The reactor temperature was maintained at 30°C. The resultant reactor headspace gas samples were taken every 0.5 hours using a 2.5 ml gas syringe (injecting 0.5 mL) and analyzed using GC-FID and GC-TCD. CH4, CO, and H2 calibration curves were used to obtain their concentration during the kinetic runs [36]. Figure S1 shows the reactor setup for CO2 photo reduction, and Figure S2 is a [CH4] calculation example.

2.5 Electrochemical Experiments

All experiments were performed using an Autolab® potentiostat/galvanostat M240, a Metrohm® Ag/AgCl(sat) reference electrode, a Platinized Titanium mesh as the counter electrode, and the catalysts supported on ITO as working electrodes. The solutions were prepared using Deionized Water (DI) (18 MΩ/cm2) and pH adjusted to 7 using a 0.1 M Na2HPO4/NaH2PO4 Buffer solution. Experiments were performed in a conventional three-electrode single-chamber cell. The solution was purged with high-purity nitrogen for 30 min before the measures. The concentration of NaHCO3 used was 0.1 M. To measure Semiconductors´ Open Circuit Potential Decay (OCPD), a 0.8 V potential was applied for 20 s until the current stabilized. After that, the circuit was opened and the working electrode's potential was followed for 10 minutes [37]. To obtain Cyclic Voltammetry (CV) graphs, a scan rate of 10 mV/s was used [38].

2.6 Methylene Blue Photooxidation

Methylene blue (MB) Absorption Vs. a concentration calibration curve (Abs = 0.184 [MB] + 0.0047; R2 = 0.998, at 664 nm) of 4, 2, 1, and 0.5 mg/L were used to follow MB photooxidation using am-TiO2-SiO2 and P25 catalysts.

In each batch, 5 mg of the catalyst was dispersed in 100 ml of 8 ppm MB solution. Before photo-irradiation, the suspension was stirred for 30 min in the dark to reach adsorption/desorption equilibrium. Samples were taken at different time intervals during the photocatalytic reaction, and the catalyst was immediately separated from the solution by a syringe filter. UV-vis spectra was taken, and MB degradation followed at 664 nm. Photodegradation pseudo-first-order rate constants were obtained from plots of ln [MB] Vs. Time. Exponential [MB]/[MB]o Vs. t plots and % photodegradation = ([MB]o - [MB]t)/[MB]o x 100 Vs. t was also obtained. pH was measured at the beginning and the end of each kinetic run.

3. Results and Discussion

3.1 Catalysts Characterization

Figures 1a and 1b show absorption spectra (DRS) and band gaps (Eg) of the synthesized silicate-surface modified catalysts obtained according to Tauc’s equation. These band gaps are equal to or lower than the corresponding P25 one (3.20 eV). For instance, the P25-SiO2 band gap does not change compared to P25. However, amTiO2-SiO2 moves down to 3.13 eV (396 nm). This value, together with the Cu-TiO2-SiO2 one (3.06 eV, 405 nm), is in the violet visible region. Therefore, visible light-photo reduction activity is expected for these catalysts. Cu-P25-SiO2 and N-P25-SiO2 have been previously synthesized in our laboratory, and their band gaps are included just for comparison.

Click to view original image

Figure 1 a Absorption spectra and b Plots of (K.hv)2 vs. hv according to Tauc’s equation: (αhν)1/n = A(hν-Eg), n = 0.5. Obtained band gaps of P25-SiO2 and am-TiO2-SiO2, are shown. Band gaps are also shown for comparison previously synthesized in our laboratory N-P25-SiO2 and Cu-P25-SiO2.

In Table 1, characterization parameters of SiO2-surface modified TiO2 catalysts are shown. am-TiO2-SiO2 and P25-SiO2 catalysts show larger surfaces than P25. These differences may contribute to improving these catalysts´ photo-reduction performances. Table S1 shows BET results between am-TiO2 and am-TiO2-SiO2. The last catalyst shows a larger surface area, pore size, and pore volume than the unmodified am-TiO2 one. The following characteristic IR signals were observed for the two silicate-surface-modified catalysts: 670 cm-1, Ti-O-Ti asymmetric stretching; 1060 cm-1, Si-O-Si symmetric stretching; 960 cm-1 Si-O-Ti stretching; 3750 cm-1, SiO-H extending and a broadband 3700-3500 cm-1 corresponding to water adsorbed on catalysts. These signals have also been identified in SiO2-TiO2 hybrid spheres and sol-gel TiO2-SiO2 films [39,40]. FT IR spectrum of am-P25, am-P25-SiO2, am-TiO2, and am-TiO2-SiO2 are shown in Figure S3. O-Si-O stretching signal at 1022-1080 cm-1 is present only in the -SiO2-modified catalysts.

Table 1 Catalysts band gap (Eg), EDX-SEM weight percentage (Wt %), surface area, pore size/volume, and IR signals.

Figure 2a and Figure 2b show photoluminescence spectra of am-TiO2-SiO2 and P25-SiO2 catalysts using exciting light at 350 and 250 nm, respectively. More light collection efficiency is shown for am-TiO2-SiO2, especially when exciting light of more energy (UV at 250 nm) is used. It can be ascribed to the removal of electron capture centers on the surface of am-TiO2 or the removal of nonradiative decay channels due to the SiO2 shell around am-TiO2 [41]. This light harvesting efficiency may improve this catalyst photo activity. Figure 2c shows SEM-EDX results of am-TiO2-SiO2. Agglomerations of irregular particle sizes and homogeneous Si and O distribution (EDX) at the catalyst surface spot are observed. TEM images of am-TiO2-SiO2 and P25-SiO2 are shown in Figure S4. Agglomerations and amorphous morphology are detected for amTiO2-SiO2. Meanwhile, crystalline morphology for P25-SiO2 is revealed. According to EDX, prepared am-TiO2-SiO2 contains an average of 53% Ti, 5% Si, and 41% O wt% at the surface. We will use the active site of this surface structure to depict the catalytic aspects covered in this work. Figure 2d shows a PZC value of 4.1 for am-TiO2-SiO2 obtained from pHi-pHf Vs—pHi plot.

Click to view original image

Figure 2 a 320 nm excitation photoluminescence spectra of am-TiO2-SiO2 and P25-TiO2 catalysts. b 250 nm excitation. c SEM-EDX of am-TiO2-SiO2. d am-TiO2-SiO2 PZC measurement.

The oxidation state of the constituent elements in the synthesized catalysts was determined using X-ray photoelectron spectroscopy (XPS), and the results are presented in Figures 3a, 3b, and 3c. The XPS spectrum of Ti 2p in Figure 3a was deconvoluted into two peaks with binding energies of 458 eV and 463.7 eV, attributed to Ti 2p3/2 and Ti 2p1/2, respectively. This confirms an oxidation state of +4 for Ti. Additionally, the Si 2p spectrum (Figure 3b) peaked at a binding energy of 102.2 eV, which was attributed to the Si-O-Ti bond. Furthermore, in the O 1 s XPS spectrum (Figure 3c), three deconvoluted peaks located at 529.8 eV, 530.9 eV, and 532 eV were attributed to Ti-O, Si-O-Ti, and surface hydroxyl groups, respectively.

Click to view original image

Figure 3 Ti (a), Si (b), and O (c) XPS spectra of am-TiO2-SiO2. (d): XRD spectrum. (e) Raman spectrum.

In Figure 3d, the XRD spectrum of am-SiO2-TiO2 is shown. Only one broad signal at ca. 27.5 2θ is observed. This signal is typical of amorphous TiO2, where no crystal TiO2 patterns are observed. Due to the small amounts of Si (5%) present, the corresponding signal at ca. 58 2θ is barely detected in Figure 3d [40].

Figure 3e shows a sharp signal at ca. 3000 cm-1 Raman shift corresponding to O-H stretching. This band is hidden in the corresponding FTIR due to the presence of water. Broad movements centered at 1400, 700, and 170 cm-1 Raman shifts can be identified as titanium species' framework [42]. A signal at 80 cm-1 Raman shift observed in Figure 3e could be assigned to the framework of Si since it is absent in the am-TiO2 Raman spectrum (Figure S5).

3.2 Blanks and Alternative Carbon Sources

As shown in Figure 4, CH4 is formed even without adding any CO2 source (CO2 or NaHCO3). This alternative source could come from the catalyst synthesis or reactor walls' remaining organic materials. Special care is taken to eliminate any solvent from the catalyst by heating the solid catalyst while stirring at 80°C for 24 h before use. Likewise, the glass reactor is cleaned after use using successive 30 min washing of ultrasound (H2O), KOH (EtOH), H2SO4 (H2O2), and HCl (30%). The maximum CH4 observed after 1 h is ca. 0.5 nmol/mL H2O gcat. As observed in Figure 4, the contribution of this source is relatively low compared to the amount of CH4 produced from CO2 reduction. Using a new reactor, blanks with catalyst, CO2 source, and lamp off and without catalysis, CO2 source and lamp on were run. Under these conditions, [CH4] < 0.05 nmol/mL H2O gcat was detected.

Click to view original image

Figure 4 CH4 photoreduction formation without and with CO2 source at two different pHs.

3.3 HCO3-/CO2 Photoreduction

P25 CO2/HCO3- photo reduction proposed mechanism to form CH4, CO, and H2 in the reactor headspace consists of a fast formation of monodentate adduct followed by a rate-limiting adduct radical anion formation [35]. Since the Point of zero charge (PZC) of the silicate surface modified TiO2 catalysts a ca. 2.9 [43] and 4.1 (Figure 2d) for the am-SiO2-TiO2 synthesized in this work, relatively lower values than the PZC (6.2) of P25, negative charge at the silicate surface must be present even at acid pH. Since a negative charge on one of the oxygen (-Si-O-) remains, adduct formation is favored when carbonate or CO2 is present (Scheme 1, center-top). Nevertheless, one -Si-OH is still prorogated at pH ca. 5; therefore, the proton required in the e-transfer from Ti to produce the corresponding radical anion (Scheme 1, left-bottom) is available at the active site. Since this transfer has been pointed out as the photo reduction rate-determining step (rds), faster product formation may be observed when these silicate-surface-modified TiO2 catalysts are used [35].

Click to view original image

Scheme 1 TiO2-SiO2 carbonate adduct formation and electron transfer to form the corresponding adduct radical anion followed by fast product formation.

Figure 5a and Figure 5b show CH4 and H2 production in the photo reduction of HCO3- using the silicate-surface modified TiO2 catalysts synthesized in this work. As expected, am-TiO2-SiO2 is the best catalyst, producing ca. 8 times more methane and H2 than P25 in a non-buffered solution. Given this result in H2 production, we performed water splitting using am-TiO2-SiO2 (Figure 5c). As shown in the figure, H2 is formed with a rate of 2.3 nmol/mLH2O·gcat·h, that is 0.23 μmol/gcat·h. This means that am-TiO2-SiO2 can also be used by adding water at pH = 3 to obtain H2 at a relatively reasonable rate. Scheme 2 shows the proposed mechanism for water splitting and H2 formation when the carbonate adduct is present at the catalyst surface.

Click to view original image

Figure 5 a reactor headspace [CH4] produced in water UV photo reduction of 100 mg of NaHCO3 using 0.3 g of catalyst under non-buffered conditions. b [H2] production. See Table 2 for the reaction's initial pH values. c [H2] formation rate via water splitting, without any carbonate source added at pH = 3. d Reactor headspace [CH4] and [H2] produced in water UV photo reduction of 3.4 mmole/100 mL of CO2 using 0.3 g of am-TiO2-SiO2 catalyst at pH = 3. Re-used catalysts results ([CH4] and [H2]) at pH = 4.6 are also shown.

Click to view original image

Scheme 2 H2 production proposed mechanism for am-TiO2-SiO2. With CO2/HCO3- added (X = O - CO2-) and without carbonate source added (X = O) (water splitting).

Table 2 Production and yield of CH4 and H2 in the HCO3-/CO2 photo reduction using am- TiO2-SiO2. P25 is included as a reference.

Click to view original image

Figure 6 CH4, H2, and CO produced in the photocatalyzed UV reduction of CO2 at pH = 3.

Since the reactor [CO2]w = 3.4 mmol/100 mL when using 1 atm of CO2 pressure, faster CH4 and H2 productions than when NaHCO3 is used (2.3 mmol/100 mL) are expected. However, as shown in Figure 5d, when am-TiO2-SiO2 is used, CH4 is produced 80 times faster (8.9 μmol gcat-1h-1) than P25. This is quite a remarkable result for a solid-liquid condition since this rate is even more significant than when Pt-TiO2 is used; it is reported at 1.4 and 5.2 μmolgcat-1h-1 using solid-liquid and solid-gas conditions [44]. For hollow anatase, single crystal and mesocrystal values of 0.2, 1, and 1.7 μmolgcat-1h-1 have been reported for solid-gas needs [45]. Using carbon-doped amorphous TiO2, 4.1 μmolgcat-1h-1 has been reported [32]. A similar rate of 8.9 μmolgcat-1h-1 has been obtained using 2%MgO-Pt-TiO2 under solid-gas conditions [44]. This value is one of the best yields reported for CO2 photo reduction for non-carbon-based photocatalysts. In the kinetics shown in Figure 5d, pH = 3, CH4 is produced more significantly than H2. This is quite an important result since at pH = 8 (Figure 5a and 5b), H2 is favored instead. This means that acid pH favors CH4 formation.

In Figure 5d, reuse am-TiO2-SiO2 results are also shown. As depicted, the CH4 rate (3 μmolgcat-1h-1) decreases, and the H2 rate increases (0.6 μmolgcat-1h-1) compared to catalysts used for the first time at pH = 3. This difference may be due to the difference in pH: 4.6 Vs. 3. As reported in Table 2 at pH = 5, CH4 rate = 5 μmolgcat-1h-1 and H2 rate = 0.1 μmolgcat-1h-1. However, catalyst deactivation cannot be discarded since the reused catalyst was obtained via evaporation of water at 90°C for several hours.

To study the H2/CH4 selectivity change with pH in more detail, we performed additional experiments in the 5-7 pH range using NaHCO3 and NaH2PO4 as buffers. Two critical remarks can be pointed out from Figure S6a's and Figure S6b's results. As observed for pH = 5, without buffer there is more selectivity towards CH4. It is the maximum rate of CH4 produced using NaHCO3. A 5 μmolgcat-1h-1 is obtained. A relatively more significant rate than the reported values for other catalysts in solid-liquid or solid-gas conditions [32,44,45]. However, this rate is reduced if NaH2PO4 buffer is used at the same pH = 5. This means that an inhibition induced by the phosphate buffer occurs. Probably due to the additional consumption of CO2 via the equilibrium: CO2 + H2O + Na2HPO4 = Na2CO3 + H3PO4. Therefore, adding phosphate buffer decreases CH4 and H2 formation rates, as experimentally observed. However, since H2 is also formed without CO2 (Scheme 1), CH4 formation is more affected than H2 (Figure S6c and Figure S6d) when buffer is added. These results can be pointed out as experimental evidence for the Scheme 2 mechanism in which H2 and CH4 are formed in two paths via partitioning competition of the radical anion formed in the rds. For practical purposes, selectivity CH4/H2 can be reversed by adding buffer to the solution and changing pH. Table 2 resumes the H2 and CH4 produced (μmolgcat-1h-1) of am-SiO2-TiO2 catalysts under different photo reduction conditions. Water splitting is also shown.

As further evidence for the mechanisms proposed in Scheme 1 and Scheme 3, we measured [CO] and compared its formation rate with [CH4] and [H2] ones at pH = 3 (Figure 6). We have also measured product ratios ([CH4]/[H2]) with time at two different pHs to obtain the kinetic (t = 0) and thermodynamic main products using ratios extrapolation to t = 0 and the slope of the ratios Vs. Time plots (Figure S7 left and right).

Click to view original image

Scheme 3 CH4 Vs. H2 formation in the CO2/HCO3- reduction using am-TiO2-SiO2. CH4 rate and selectivity increase by lowering the pH since more protons (4) are required to produce CH4 than the one needed (2) to yield H2.

As shown in Figure 6, mainly methane is produced using CO2 and at pH ca 3. As depicted in the figure inset, some H2 and CO are also formed in comparable ca. 8 nmol/mL H2O gcat concentrations after 240 min. From the [CH4]/[H2] plot shown in Figure S7 (left), it can be interpreted that both products are formed from the same intermediate that cleaves preferentially to form CH4 and H2 (Scheme 3) with some cleavage through the CO path. Although, CH4 and H2 are produced from the cleavage of the same intermediate, CH4 is thermodynamically (t = 240 min, [CH4]/[H2] = 40) and kinetically (t = 0, [CH4]/[H2] = 28) more favorable at pH = 3. However, at pH = 8, H2 is favorably formed. As shown in Figure S7 (right), at t = 0, H2 and CH4 are formed in equal concentration, although H2/CH4 ratio increases up to 3 after 90 min, probably due to the limited amount of H+ at this pH. In fact, after 90 min, the ratio slope changes, indicating more than one mechanism for H2 formation: water splitting (Figure 5c) and Scheme 3. Since in the water splitting mechanism CH4 is not produced, this mechanism contributes more at the initial [H2]/[CH4] reaction times (positive slope, Figure S7 (right)): meanwhile, Scheme 3 one contributes at final reaction times (slightly negative slope, Figure S7 (right)).

3.4 Electrochemical Measurements

Figure 7a shows CVs of P25 and am-TiO2-SiO2 coated on ITO. The current intensity is larger for P25 than am-TiO2-SiO2, but its de-trapping (positive current) is slower [46]. Less current and slow de-trapping in am-TiO2-SiO2 is associated with the catalyst surface capacity to retain a charge. SiO2 at the surface efficiently includes e; therefore, de-trapping to the solution becomes relatively slow. In the plot of E Vs. t embedded in Figure 7a, an OCPD with smoother decay is observed for am-TiO2-SiO2. Figure 7a (embedded) and reported values in Table 3 show that am-TiO2-SiO2 halftime is 122 s. Meanwhile, in P25, two slopes are observed with halftimes of 19 s and 103 s. The fast process in P25 corresponds to P25 electron transfer to the electrolyte from its CB and the slow one to electron transfer from oxygen vacancy (Vo) traps. We have reported these two mechanisms, and this OCPD measurement confirms our previous proposal [35]. Another aspect worth mentioning from Figure 7a is the equilibrium potential (surface potential [47]) reached after decay by the two catalysts; as observed in the figure and reported in Table 3, am-TiO2-SiO2 reaches a less negative potential than P25. Although the former catalyst´s PZC value (4.1) is relatively lower than the P25 one (6.2), there is more surface negative charge at pH = 7 for P25. This counterintuitive result can be explained in terms of better water solvation in am-TiO2-SiO2 compared to P25.

Click to view original image

Figure 7 a CV of ITO-P25 and ITO-am-TiO2-SiO2. Inset: corresponding open circuit potential decays, after applying a potential of -0.8 V. b ln (E-E(t)) Vs. time of ITO-P25 and ITO-am-TiO2-SiO2. Rates (straight-line slopes) are shown. Values are summarized in Table 3. c CV of ITO-am-TiO2-SiO2 and ITO-am-TiO2-SiO2 adding NaHCO3- to the solution. Inset: corresponding open circuit potential decay after applying a potential of -0.8 V.

Table 3 Rate and half-life time values of potential decay ITO-P25, ITO-am-TiO2-SiO2, and ITO-am-TiO2-SiO2 with NaHCO3 added.

These experiments were performed at pH 7 using NaH2PO4 buffer with reduced dissolved O2 using previously blowing N2 into the solution; therefore, electron from de-trapping are used to reduce water in the solution. If an additional electron acceptor is added, we expect to observe changes in the corresponding OCPD. As observed in Figure 7c, when NaHCO3 is added to the solution, the capacity of am-TiO2-SiO2 to retain charge increases. Also, as shown in Figure 7c, the embedded plot of E Vs. t, the surface charge value obtained is more damaging than the am-TiO2-SiO2 and P25 ones (Table 3). This can be interpreted as an additional trap present when NaHCO3 is added. The Adduct shown in Scheme 1 (top right) then acts as a trap that makes more electrons available for photo reduction at the CB. As the figure shows, the destabilization of the two negative charges is diminished since the distance between them in the adduct decreases. This charge accumulation capability is one of the factors contributing to the efficient methane and hydrogen production when am-TiO2-SiO2 is used.

3.5 Photo Reduction Vs. Photo Oxidation

We have used the bleaching induced by OH radical on methylene blue (MB) to test its production by photooxidation of water at the TiO2 VB [49,50]. Figure S8a and Table 4 show that quenching effectively occurs when using P25 but does not proceed when am-TiO2-SiO2 and P25-TiO2-SiO2 are used. We conclude that OH. is not an intermediate in the photooxidation of water. O2 is then preferentially produced directly or produced from peroxide. The reduction potential for one-electron oxidation to produce hydroxyl radical from water is 2.81 V [51], a higher value than the VB potential of am-TiO2-SiO2 (1.93 V) and P25-SiO2 (2.48 V) [52]. Therefore, one electron transfer is thermodynamically unfavorable; It is only the driven force of the 4 e oxidation to yield O2 (0.81 V at pH = 7) [52] that allows the first electron transfer to occur, but the product (OH-) does not accumulate at all. However, the 2 e and 4 e oxidation to produce H2O2 and O2 are feasible since the corresponding potentials at pH = 7 are 1.34 V and 0.81 V, respectively [53]. Conditions for water splitting (H2O = H2 + 1/2O2) without a cocatalyst or sacrificial reagent are then provided for am-TiO2-SiO2 since the conduction band reduction potential is -1.0 V [53]. Meanwhile, H+/H2 at pH 3 and 7 are 0.17 V and -0.41 V, respectively. The non formation of hydroxyl radicals is an excellent advantage for water splitting and CO2 photoreduction since a sacrificial reagent is unnecessary. Some MB degradation is observed when HCO3- is added to the MB solution (Figure S8b and Table 4). Adding HCO3- increases the negative charge on its surface, as shown in the surface potential values of Table 3. Increasing the negative charge on the surface induces down-bending on the valence band [54]. Therefore, the VB potential becomes more positive, allowing the one-electron transfer from water to produce a hydroxyl radical that MB quenches. Figure S9 describes the RedOx processes at the am-TiO2-SiO2 surface.

Table 4 Methylene Blue (MB) ([MB] = 8 ppm) photooxidation pseudo-first-order rate constants, half times and degradation efficiency after 30 min UV light irradiation. Adding [HCO3-] = 4 ppm, last column) significantly improves the activity of less active silicate surface-modified catalysts.

The more negative surface charge also favors positively charged organic adsorptions. Efficient MB removal and E. coli bacteria inactivation have been reported using P25-SiO2. In both cases, an effective negative (P25-SiO2)-positive (MB and protein) interaction is claimed as the leading cause for the improvement in photooxidation rate [55]. The adsorption process is then crucial in these photocatalytic reactions. Likewise, it may play a role in MB's binding inhibition of ACE2 (angiotensin-converting enzyme 2) to SARS-COV2-RBD [56]. The more negative protein the ACE2 active site is, the more efficient MB antiviral activity is predicted.

4. Conclusions

We have prepared silicate-surface modified amorphous TiO2 (am-TiO2-SiO2) and P25 (P25-SiO2) catalysts. These catalysts yield more CH4 and H2 than P25 using UV light. In fact, in 4 h, using 3.4 mmol of CO2 (1 atm)/100 mL H2O and 0.3 g of am-TiO2-SiO2, 80 times more CH4 (8.9 μmolgcat-1h-1) than P25 at pH= 3 is produced and 8 times more H2 (0.84 μmolgcat-1h-1) at pH = 7 when 2.3 mmol of NaCO3/100 mL is used with selectivity’s of 71% and 53%, respectively. To our knowledge, 8.9 μmolgcat-1h-1 of CH4 production in water from CO2 photo reduction is one of the higher rates reported when using non-carbon-based photocatalysts. H2 is also formed just by adding water to the am-TiO2-SiO2 catalyst. H2/CH4 selectivities can be modulated by changing pH and/or using a buffer. H2 production is favored at neutral pH, and CH4 at acid pH. Buffer NaH2PO4 addition diminishes CH4 production and barely affects H2 formation. Due to the low PZC value (4.1) of am-TiO2-SiO2 as compared to P25, at pH < 7, faster photo reduction rates are observed since there is a negative charge (-SiO-) on the surface that promotes CO2 adsorption, but there is also a proton available (-SiOH) at the active site. Therefore, once the carbonate-catalyst adduct is formed, the electron transfer from Ti to produce the radical anion can be catalyzed by -SiOH. Other factors may also be favoring am-TiO2-SiO2 performance as photo-reduction catalyst. Cyclic voltammetry and open circuit potential decay experiments show less intense current and less de-traping for am-TiO2-SiO2 than P25, and there is only one smooth decay slope for the former and two slopes for the last catalyst. Therefore, the attached Si at the surface also acts as an external intramolecular trap (EIT) that promotes electron transfer from Ti. When HCO3- is added to the solution and the carbonate-catalysts adduct is formed at the surface, a more negative surface potential is measured, meaning that the carbonate at the surface also facilitates the electron transfer from Ti. Additionally, due to the disordered distribution of atoms at the amorphous TiO2, less e/h recombination occurs. For instance, this catalyst also shows a high light harvesting efficiency with P25-SiO2. A larger surface area than P25 may also contribute. Meanwhile, CH4 and H2 further oxidations at the catalyst interface are avoided in the photo reduction reactions due to the relatively large hydrogen and methane gas-water coefficients, which allow them to accumulate in the reactor headspace.

P25 is a better catalyst than the silicate-modified TiO2 catalysts for MB photooxidation. The less positive VB potential in the last catalyst makes the OH radical formation from water oxidation unfavorable. Therefore, there is no hydroxyl radical to be quenched by MB. However, when NaHCO3 is added to the solution, these catalyst´s rates increase more than 10 times at neutral pH. The valence band down-bending induced by the more negative surface charge allows some OH radical formation and its corresponding quenching by MB.

The results and mechanism shown in this work have contributed to developing new procedures to increase CO2 photo-reduction as is the case of coated films at the reactor headspace.

Acknowledgments

This research project is supported by the Second Century Fund (C2F), Chulalongkorn University.

Author Contributions

Oswaldo Núñez: Designing, directing, interpreting and writing manuscript. Mohammad Fereidooni: Synthesis, characterization, photo RedOx experiments and revising manuscript. Victor Márquez: Electrochemical experiments. Duangthip Sattayamuk: Synthesis and characterization. Piyasan Praserthdam: Directing and financing. Supareak Praserthdam: Preliminary calculations and revising manuscript.

Competing Interests

The authors declare no competing financial interest.

Additional Materials

1. Figure S1: Photo reactor setup and CO2 photo reduction procedure.

2. Figure S2: GC-FID and concentration calculation example.

3. Figure S3: FT IR spectrum, top to bottom of am-P25, am-P25-SiO2, am-TiO2 and am-TiO2-SiO2. O-Si-O stretching signal at 1022-1080 cm-1 is present only in the -SiO2 catalysts.

4. Figure S4: 80000 x and 200000 x TEM images. Left: am-TiO2-SiO2. Right: P25-TiO2-SiO2.

5. Figure S5: am-TiO2 Raman spectrum.

6. Figure S6: a and c Reactor headspace [CH4] and [H2] produced in water UV photo reduction of 100 mg of NaHCO3 using 0.3 g of am-TiO2-SiO2 at different initial pH (without buffer); b and d Reactor headspace [CH4] and [H2] produced in water UV photo reduction of 100 mg of NaHCO3 using 0.3 g of am-TiO2-SiO2 using NaH2PO4 buffer.

7. Figure S7: Left: [CH4]/[H2] ratio Vs. time at pH = 3. Right: [H2]/[CH4] ratio Vs. time at pH = 7.4.

8. Figure S8: a 8 ppm of MB photo oxidation using silicate surface-modified TiO2 catalysts. b ln (C/Co) Vs. time plot of MB UV photo oxidation using P25-SiO2 and am-TiO2-SiO2 catalysts. Pseudo-first order rate constants are found from the corresponding slopes and reported in Table 4.

9. Figure S9: am-TiO2-SiO2 RedOx photo-activity. Si acts as an external intramolecular trap (EIT) promoted by the (d-p) π-bonding at the Si-O bond. e/h recombination is then reduced. CH4 is produced 80 times faster than when P25 is used and H2 is produced 8 times faster. Only O2 is produced at the valence band due to its relatively low potential with a value lower than the OH-/H2O one. However, when adding HCO3- the negative surface charge increases promoting VB down-bending and some OH is produced that is quenched by MB.

10. Table S1: am-TiO2 Vs. am-TiO2-SiO2 BET results.

References

  1. Li Y, Li J, Zhang G, Wang K, Wu X. Selective photocatalytic oxidation of low concentration methane over graphitic carbon nitride-decorated tungsten bronze cesium. ACS Sustain Chem Eng. 2019; 7: 4382-4389. [CrossRef]
  2. Tahir B, Tahir M, Amin NA. Ag-La loaded protonated carbon nitrides nanotubes (pCNNT) with improved charge separation in a monolithic honeycomb photoreactor for enhanced bireforming of methane (BRM) to fuels. Appl Catal B. 2019; 248: 167-183. [CrossRef]
  3. Xue Q, Wu M, Zeng XC, Jena P. Co-mixing hydrogen and methane may double the energy storage capacity. J Mater Chem A. 2018; 6: 8916-8922. [CrossRef]
  4. Hayat A, Sohail M, El Jery A, Al Zaydi KM, Raza S, Ali H, et al. Recent advances, properties, fabrication and opportunities in two-dimensional materials for their potential sustainable applications. Energy Storage Mater. 2023; 59: 102780. [CrossRef]
  5. Hayat A, Sohail M, El Jery A, Al Zaydi KM, Raza S, Ali H, et al. Recent advances in ground-breaking conjugated microporous polymers-based materials, their synthesis, modification and potential applications. Mater Today. 2023; 64: 180-208. [CrossRef]
  6. Bafaqeer A, Tahir M, Amin NA. Synergistic effects of 2D/2D ZnV2O6/RGO nanosheets heterojunction for stable and high performance photo-induced CO2 reduction to solar fuels. Chem Eng J. 2018; 334: 2142-2153. [CrossRef]
  7. Bafaqeer A, Tahir M, Ali Khan A, Saidina Amin NA. Indirect Z-scheme assembly of 2D ZnV2O6/RGO/g-C3N4 nanosheets with RGO/pCN as solid-state electron mediators toward visible-light-enhanced CO2 reduction. Ind Eng Chem Res. 2019; 58: 8612-8624. [CrossRef]
  8. Tahir M, Tahir B. 2D/2D/2D O-C3N4/Bt/Ti3C2Tx heterojunction with novel MXene/clay multi-electron mediator for stimulating photo-induced CO2 reforming to CO and CH4. Chem Eng J. 2020; 400: 125868. [CrossRef]
  9. Tahir B, Tahir M, Nawawi MG. Highly stable 3D/2D WO3/g-C3N4 Z-scheme heterojunction for stimulating photocatalytic CO2 reduction by H2O/H2 to CO and CH4 under visible light. J CO2 Util. 2020; 41: 101270. [CrossRef]
  10. Tahir M. Construction of a stable two-dimensional MAX supported protonated graphitic carbon nitride (pg-C3N4)/Ti3AlC2/TiO2 Z-scheme multiheterojunction system for efficient photocatalytic CO2 reduction through dry reforming of methanol. Energy Fuels. 2020; 34: 3540-3556. [CrossRef]
  11. Khan AA, Tahir M, Zakaria ZY. Synergistic effect of anatase/rutile TiO2 with exfoliated Ti3C2TR MXene multilayers composite for enhanced CO2 photoreduction via dry and bi-reforming of methane under UV-visible light. J Environ Chem Eng. 2021; 9: 105244. [CrossRef]
  12. Tahir M, Tahir B. Constructing S-scheme 2D/0D g-C3N4/TiO2 NPs/MPs heterojunction with 2D-Ti3AlC2 MAX cocatalyst for photocatalytic CO2 reduction to CO/CH4 in fixed-bed and monolith photoreactors. J Mater Sci Technol. 2022; 106: 195-210. [CrossRef]
  13. Yu J, Wang K, Xiao W, Cheng B. Photocatalytic reduction of CO2 into hydrocarbon solar fuels over g-C3N4-Pt nanocomposite photocatalysts. Phys Chem Chem Phys. 2014; 16: 11492-11501. [CrossRef]
  14. Chen X, Wang J, Huang C, Zhang S, Zhang H, Li Z, et al. Barium zirconate: A new photocatalyst for converting CO2 into hydrocarbons under UV irradiation. Catal Sci Technol. 2015; 5: 1758-1763. [CrossRef]
  15. Xie K, Umezawa N, Zhang N, Reunchan P, Zhang Y, Ye J. Self-doped SrTiO3-δ photocatalyst with enhanced activity for artificial photosynthesis under visible light. Energy Environ Sci. 2011; 4: 4211-4219. [CrossRef]
  16. Yang CC, Yu YH, van der Linden B, Wu JC, Mul G. Artificial photosynthesis over crystalline TiO2-based catalysts: Fact or fiction? J Am Chem Soc. 2010; 132: 8398-8406. [CrossRef]
  17. Li P, Zhou Y, Li H, Xu Q, Meng X, Wang X, et al. All-solid-state Z-scheme system arrays of Fe2V4O13/RGO/CdS for visible light-driving photocatalytic CO2 reduction into renewable hydrocarbon fuel. Chem Commun. 2015; 51: 800-803. [CrossRef]
  18. Cybula A, Klein M, Zaleska A. Methane formation over TiO2-based photocatalysts: Reaction pathways. Appl Catal B. 2015; 164: 433-442. [CrossRef]
  19. Moustakas NG, Strunk J. Photocatalytic CO2 reduction on TiO2-based materials under controlled reaction conditions: Systematic insights from a literature study. Chem Eur J. 2018; 24: 12739-12746. [CrossRef]
  20. Tahir M. Well-designed ZnFe2O4/Ag/TiO2 nanorods heterojunction with Ag as electron mediator for photocatalytic CO2 reduction to fuels under UV/visible light. J CO2 Util. 2020; 37: 134-146. [CrossRef]
  21. Tahir M. Enhanced photocatalytic CO2 reduction to fuels through bireforming of methane over structured 3D MAX Ti3AlC2/TiO2 heterojunction in a monolith photoreactor. J CO2 Util. 2020; 38: 99-112. [CrossRef]
  22. Gao X, Wachs IE. Titania-silica as catalysts: Molecular structural characteristics and physico-chemical properties. Catal Today. 1999; 51: 233-254. [CrossRef]
  23. Inoue H, Matsuyama T, Liu BJ, Sakata T, Mori H, Yoneyama H. Photocatalytic activities for carbon dioxide reduction of TiO2 microcrystals prepared in SiO2 matrices using a sol-gel method. Chem Lett. 1994; 23: 653-656. [CrossRef]
  24. Anpo M, Chiba K. Photocatalytic reduction of CO2 on anchored titanium oxide catalysts. J Mol Catal. 1992; 74: 207-212. [CrossRef]
  25. Yuan L, Han C, Pagliaro M, Xu YJ. Origin of enhancing the photocatalytic performance of TiO2 for artificial photoreduction of CO2 through a SiO2 coating strategy. J Phys Chem C. 2016; 120: 265-273. [CrossRef]
  26. Zhan S, Yang Y, Gao X, Yu H, Yang S, Zhu D, et al. Rapid degradation of toxic toluene using novel mesoporous SiO2 doped TiO2 nanofibers. Catal Today. 2014; 225: 10-17. [CrossRef]
  27. Kim CS, Shin JW, An SH, Jang HD, Kim TO. Photodegradation of volatile organic compounds using zirconium-doped TiO2/SiO2 visible light photocatalysts. Chem Eng J. 2012; 204-206: 40-47. [CrossRef]
  28. Liu Y, Yu H, Lv Z, Zhan S, Yang J, Peng X, et al. Simulated-sunlight-activated photocatalysis of methylene blue using cerium-doped SiO2/TiO2 nanostructured fibers. J Environ Sci. 2012; 24: 1867-1875. [CrossRef]
  29. Li X, Yu J, Jaroniec M, Chen X. Cocatalysts for selective photoreduction of CO2 into solar fuels. Chem Rev. 2019; 119: 3962-4179. [CrossRef]
  30. Sun S, Song P, Cui J, Liang S. Amorphous TiO2 nanostructures: Synthesis, fundamental properties and photocatalytic applications. Catal Sci Technol. 2019; 9: 4198-4215. [CrossRef]
  31. Wang P, Yin G, Bi Q, Huang X, Du X, Zhao W, et al. Efficient photocatalytic reduction of CO2 using carbon-doped amorphous titanium oxide. ChemCatChem. 2018; 10: 3854-3861. [CrossRef]
  32. Vargas MA, Rodríguez Páez JE. Amorphous TiO2 nanoparticles: Synthesis and antibacterial capacity. J Non Cryst Solids. 2017; 459: 192-205. [CrossRef]
  33. Kaur K, Singh CV. Amorphous TiO2 as a photocatalyst for hydrogen production: A DFT study of structural and electronic properties. Energy Procedia. 2012; 29: 291-299. [CrossRef]
  34. Calleja G, Serrano DP, Sanz R, Pizarro P. Mesostructured SiO2-doped TiO2 with enhanced thermal stability prepared by a soft-templating sol-gel route. Microporous Mesoporous Mater. 2008; 111: 429-440. [CrossRef]
  35. Núñez O, Sattayamuk D, Saelee T, Yamashita H, Kuwahara Y, Mori K, et al. A closer look inside TiO2(P25) photocatalytic CO2/HCO3- reduction with water. Methane rate and selectivity enhancements. Chem Eng J. 2021; 409: 128141. [CrossRef]
  36. Zhu M, Wang H, Keller AA, Wang T, Li F. The effect of humic acid on the aggregation of titanium dioxide nanoparticles under different pH and ionic strengths. Sci Total Environ. 2014; 487: 375-380. [CrossRef]
  37. Bisquert J, Zaban A, Greenshtein M, Mora Seró I. Determination of rate constants for charge transfer and the distribution of semiconductor and electrolyte electronic energy levels in dye-sensitized solar cells by open-circuit photovoltage decay method. J Am Chem Soc. 2004; 126: 13550-13559. [CrossRef]
  38. Fabregat Santiago F, Mora Seró I, Garcia Belmonte G, Bisquert J. Cyclic voltammetry studies of nanoporous semiconductors. Capacitive and reactive properties of nanocrystalline TiO2 electrodes in aqueous electrolyte. J Phys Chem B. 2003; 107: 758-768. [CrossRef]
  39. Nowacka M, Ambrożewicz D, Jesionowski T. TiO2-SiO2/Ph-POSS functional hybrids: Preparation and characterisation. J Nanomater. 2013; 2013: 680821. doi: 10.1155/2013/680821. [CrossRef]
  40. Song CF, Lü MK, Yang P, Xu D, Yuan DR. Structure and photoluminescence properties of sol-gel TiO2-SiO2 films. Thin Solid Films. 2002; 413: 155-159. [CrossRef]
  41. Gu F, Li CZ, Wang SF, Lü MK. Solution-phase synthesis of spherical zinc sulfide nanostructures. Langmuir. 2006; 22: 1329-1332. [CrossRef]
  42. Li C, Xiong G, Liu J, Ying P, Xin Q, Feng Z. Identifying framework titanium in TS-1 zeolite by UV resonance Raman spectroscopy. J Phys Chem B. 2001; 105: 2993-2997. [CrossRef]
  43. Salgado BC, Valentini A. Evaluation of the photocatalytic activity of SiO2@TiO2 hybrid spheres in the degradation of methylene blue and hydroxylation of benzene: Kinetic and mechanistic study. Braz J Chem Eng. 2020; 36: 1501-1518. [CrossRef]
  44. Xie S, Wang Y, Zhang Q, Deng W, Wang Y. MgO-and Pt-promoted TiO2 as an efficient photocatalyst for the preferential reduction of carbon dioxide in the presence of water. Acs Catal. 2014; 4: 3644-3653. [CrossRef]
  45. Jiao W, Wang L, Liu G, Lu GQ, Cheng HM. Hollow anatase TiO2 single crystals and mesocrystals with dominant {101} facets for improved photocatalysis activity and tuned reaction preference. Acs Catal. 2012; 2: 1854-1859. [CrossRef]
  46. Bertoluzzi L, Badia Bou L, Fabregat Santiago F, Gimenez S, Bisquert J. Interpretation of cyclic voltammetry measurements of thin semiconductor films for solar fuel applications. J Phys Chem Lett. 2013; 4: 1334-1339. [CrossRef]
  47. Bisquert J, Cendula P, Bertoluzzi L, Gimenez S. Energy diagram of semiconductor/electrolyte junctions. J Phys Chem Lett. 2014; 5: 205-207. [CrossRef]
  48. Bischoff M, Biriukov D, Předota M, Roke S, Marchioro A. Surface potential and interfacial water order at the amorphous TiO2 nanoparticle/aqueous interface. J Phys Chem C. 2020; 124: 10961-10974. [CrossRef]
  49. Satoh AY, Trosko JE, Masten SJ. Methylene blue dye test for rapid qualitative detection of hydroxyl radicals formed in a Fenton's reaction aqueous solution. Environ Sci Technol. 2007; 41: 2881-2887. [CrossRef]
  50. Nosaka Y, Nosaka A. Understanding hydroxyl radical (OH) generation processes in photocatalysis. ACS Energy Lett. 2016; 1: 356-359. [CrossRef]
  51. Armstronga DA, Huie RE, Lymar S, Koppenol WH, Merényi G, Neta P, et al. Standard electrode potentials involving radicals in aqueous solution: Inorganic radicals. BioInorg React Mech. 2013; 9: 59-61. [CrossRef]
  52. Fereidooni M, Núñez O, Márquez V, Paz CV, Villanueva MS, Tun MZ, et al. Effect of substrate conductivity on charge transfer and CO2 photoreduction in water vapor over silica-modified TiO2 films. Appl Surf Sci. 2023; 611: 155595. [CrossRef]
  53. Wood PM. The potential diagram for oxygen at pH 7. Biochem J. 1988; 253: 287-289. [CrossRef]
  54. Zhang Z, Yates Jr JT. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chem Rev. 2012; 112: 5520-5551. [CrossRef]
  55. Martín Sómer M, Benz D, van Ommen JR, Marugán J. Multitarget evaluation of the photocatalytic activity of P25-SiO2 prepared by atomic layer deposition. Catalysts. 2020; 10: 450. doi: 10.3390/catl10040450. [CrossRef]
  56. Bojadzic D, Alcazar O, Buchwald P. Methylene blue inhibits the SARS-CoV-2 spike-ACE2 protein-protein interaction-a mechanism that can contribute to its antiviral activity against COVID-19. Front Pharmacol. 2021; 11. doi: 10.3389/fphar.2020.600372. [CrossRef]
Newsletter
Download PDF Supplementary File Download Citation
0 0

TOP