Next Article in Journal
Support Effect of Ga-Based Catalysts in the CO2-Assisted Oxidative Dehydrogenation of Propane
Previous Article in Journal
NO Oxidation on Lanthanum-Doped Ceria Nanoparticles with Controlled Morphology
Previous Article in Special Issue
Application of TiO2-Based Photocatalysts to Antibiotics Degradation: Cases of Sulfamethoxazole, Trimethoprim and Ciprofloxacin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ordered Mesoporous TiO2: The Effect of Structure, Residual Template and Metal Doping on Photocatalytic Activity

1
Department of Heterogeneous Photocatalysis, Leibniz Institute for Catalysis, Albert-Einstein-Str. 29a, 18059 Rostock, Germany
2
Department of Chemistry, College of Education for Girls, University of Mosul, Mosul 41002, Iraq
3
Max Planck Institute for Chemical Energy Conversion, Stiftstraße 34-36, 45470 Mülheim an der Ruhr, Germany
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(5), 895; https://doi.org/10.3390/catal13050895
Submission received: 21 March 2023 / Revised: 19 April 2023 / Accepted: 5 May 2023 / Published: 16 May 2023

Abstract

:
Using a series of ordered mesoporous TiO2 (om-TiO2) with and without Ce or Cu doping, the effects of structure, metal doping and residual template species in the structure are systematically evaluated in terms of products formed during a CO2 photoreduction process. It is found that the ordered mesoporous structure contributes significantly in the hydrogen evolution reaction from the splitting of gaseous water. No cocatalyst was needed to achieve high hydrogen yields. While carbon-containing products are also observed, the presence of remainders of the organic template used in the synthesis process does not allow an unambiguous identification of the source of products. Small amounts of metal doping do not majorly influence the hydrogen evolution, thus the mesoporous structure can eventually be identified as the main cause for the improved performance.

1. Introduction

Photocatalytic water splitting and the reduction of CO2 with water into hydrocarbons by semiconductors are considered as promising technologies for a sustainable synthesis of fuels and platform chemicals [1]. A photocatalyst widely applied in the field of energy and fuel generation is titania (TiO2) [2]. Titania is a low cost, non-toxic, and chemically inert material with a high photostability [3,4]. Three-dimensional titania nanomaterial can be synthesized by sol-gel approach, solvothermal/hydrothermal growth, or by using templates [5].
Recently, mesoporous photocatalysts in general, [6] and titania photocatalysts with mesoporous structures in particular [7,8], have attracted much interest because of their high specific surface area and porous network, leading to an improved diffusion of reactants, better separation of light-generated charge carriers, and increased photocatalytic activity. The synthesis methods of mesoporous titania can be classified in different categories: template free, soft template, hard template, and multi-template routes where each route has its own advantages and disadvantages [8]. For example, the application of soft templates allows the synthesis of mesoporous titania structures with different morphologies and pore sizes ranging from 1-D nanofibers, to 2-D nanosheets, up to 3-D nanostructures. However, the crystallinity of the formed titania structures is relatively low [8]. Zhang et al. synthesized mesoporous TiO2 comprised of small anatase nanoparticles [9]. The materials that had undergone hydrolysis prior to calcination performed better in photocatalytic CO2 reduction than the solely calcined materials and as the benchmark P25. The hydrolysis prevents excessive particle growth, retaining a high surface area.
A subclass of mesoporous titania is ordered mesoporous titania (om-TiO2), in which an array of channels is formed rather than random mesoporosity, which has been proposed as beneficial for mass transfer [10] and metal dispersion. [11] Om-TiO2 has been successful applied in solar cells, lithium-ion batteries, and sensors, and as catalyst support, and photocatalyst [7]. It can be synthesized by sol-gel route using ionic or non-ionic surfactants as a template [12,13] and the size of the ordered mesopores is strongly affected by the used template [7]. In this soft templating pathway, the titanium precursor and the surfactant were co-assembled inside the synthesis mixture. Zimny et al. [14,15] combined the evaporation induced self-assembly (EISA) method with the liquid crystal templating (LCT) pathway. Their synthesized material showed after titania precipitation a hexagonal ordered structure with uniform pore size. A further calcination step led to the formation of semicrystalline walls and the ordered structure was still intact up to a temperature of 450 °C. Moreover, the calcined material showed good photocatalytic activity in the degradation of methyl orange [15]. In contrast to silica-based structures such as SBA-15 the degree of mesoporous ordering of the titania framework decreased with increasing calcination temperature due to the formation of small nanocrystals and their sintering inside the relatively thin framework walls [16]. Liu et al. [17] synthesized om-TiO2 by the EISA technique using cetyltrimethylammoniumbromide (CTAB) as a liquid crystal template showing a 2D hexagonal mesostructure with small particle size and high specific surface area.
Om-TiO2 with mixed anatase-rutile crystal structure has been found to be a beneficial hydrogen evolution photocatalyst, even under visible light irradiation [18]. The higher efficiency compared with P25 at comparable phase ratio has mainly been attributed to the higher surface area and the better interfacial contact of the two phases [18]. Jinhua Ye et al. synthesized om-TiO2 using EISA method [19]. CH4 was the observed main product in photocatalytic CO2 reduction, with an improved efficiency of a factor of 71 compared to the reference P25, and still a factor of 53 times compared to non-ordered mesoporous TiO2. In a comparison of various ordered mesoporous semiconductors prepared by a hard templating method, om-TiO2 showed considerably improved CH4 and CO yields compared to the commercial P25 counterpart, but mesoporous ZnS outperformed this material in terms of CH4 formation [20]. Doping of om-TiO2 with Co was beneficial to improve the formation of CO and CH4, with an optimal Co/Ti ratio of 0.025 [21]. Interestingly, an addition of Co beyond this amount lead to the formation of separate cobalt oxide domains and the almost complete suppression of CO formation [21]. A modification of an om-TiO2 material with gold nanoparticles allowed even the formation of CO and CH4 under visible light [22].
In this work, the photocatalysts have been synthesized employing coprecipitation using EISA combined with the LCT pathway [16] based on the method developed by Zimny [15]. The effect of a doping with Cu or Ce was tested. Hydrogen evolution was significantly enhanced in the mesoporous structure, but in spite of various attempts to remove the organic template, the formation of carbon-based products could not uniquely be attributed to proper photocatalytic CO2 reduction. Our results of significant product formation from carbon sources other than CO2 are particularly relevant for approaches in which organic linkers or CO2 adsorbers are added to mesoporous TiO2, [23] highlighting the need for proper blank experiments in such cases. The influence of the dopant was only moderate.

2. Results

2.1. Structural Characteristics of the TiO2 Samples

The small angle X-ray scattering (SAXS) patterns of the mesoporous materials (Figure 1A) show a d100 Bragg reflection which indicates the presence of ordered domains inside the titania framework. The presence of the doping metals (Cu or Ce) had a relatively low impact on the shape of its reflection indicating that low doping levels affect the formation of the ordered mesoporous structure only to a minor degree. The X-ray diffraction (XRD) powder patterns of pristine calcined titania and materials synthesized with 0.5 mol% metal doping are displayed in Figure 1B. All patterns exhibited intense reflections of the anatase phase (International Center of Diffraction Data (ICDD) 03-065-5714) together with low intense reflections of β-TiO2 (ICDD 00-035-0088). Bragg reflections from phases containing the doping metal or an expansion of anatase cell volume due to presence of the doping metal were not observed. The average crystallite sizes which were estimated using the Scherrer equation are listed in Table 1. The crystallite size of the pristine om-TiO2 and of the materials synthesized with 0.5 mol% doping metal were similar and ranged between 4–5 nm.

2.2. Textural Properties

N2 adsorption–desorption was applied to explore the pore structure, the Brunauer–Emmett–Teller (BET) surface area, and pore size distribution of the prepared om-materials. The corresponding isotherms of the calcined pristine and doping metal containing titania samples are presented in Figure 2. All samples exhibited a type IV isotherm, which is characteristic for mesoporous materials with hysteresis loops between H2 and H3 [24].
The BET surface areas (s), pore volumes (Vp), and maximum of the pore diameter distribution (dp) are summarized in Table 1. The BET surface area of pristine om-titania and materials synthesized with 0.5 mol% metal doping were relatively high compared to P25. Repeated synthesis (six times) of pristine titania shows an average BET surface area of 176 ± 16 m2·g−1. The differences in BET surface area between different batches might be attributed to the solvent evaporation step because the geometrical structure of the titania network before precipitation with gaseous ammonia seems to strongly depend on the amount of evaporated solvent. With increasing solvent evaporation time the viscosity of the formed gel increases. Since the viscosity change can only be observed visually, it is difficult to always achieve the same point even if the same time and temperature are used.

2.3. Optical Properties and Surface Composition

The UV/Vis diffuse reflectance spectra (DRS) of the om-materials are presented in Figure 3. DRS spectra of P25, pure CeO2, or CuO phase are presented in Figure S1 (Supplementary Materials). The om-TiO2 sample showed only absorbance in the wavelength region <400 nm [25,26] similar to P25 due to band gap excitation, which can be understood as the ligand-metal charge transfer between titanium Ti4+ (3 d) and oxygen ligand O2- (2p) [27]. The samples synthesized with 0.5 mol% of doping metal showed some absorption at wavelengths higher than 400 nm. The extent of the red shift was stronger for Cu than for Ce.
Figure 3b shows modified Tauc plots used for estimation of the band gap [28,29,30]. The band gap of all three samples was around 3.2 eV (see Table 1) indicating that the low dopant concentration does not affect the TiO2 band gap energy but it might create additional levels adjacent to the absorption edge.
When Cu was added, absorption shifting occurred at just above 400 nm together with a second absorption band at higher wavelengths. The absorption between 400 and 500 nm was previously attributed to interfacial charge transfer from the O 2p valence band to the Cu2+ state attached to TiO2 [31] or more generally to the formation of a heterojunction between CuxO and TiO2 [32]. The Cu2+ state might be present either as Cu(II) ion, Cu(II) nanoclusters, or in the form of an amorphous CuO oxide phase [31]. Absorption at wavelengths higher than 550 nm (second band) is mainly due to the d-d transition of Cu2+ in the crystalline environment of TiO2 [31,33], to an interband transition in Cu(II) clusters deposited over TiO2 [32], or to the presence of crystalline bulk CuO [34]. The EPR spectrum of the Cu-doped samples is shown in Figure S2. The spectra showed a hyperfine structure of Cu2+ ions due to interaction of the electron spin (S = ½) with its nucleus spin (I = 3/2). The value for the parallel component of the g-tensor was g = 2.347 and the hyperfine splitting amounted to A = 110 G. The values for the normal components could not be resolved. The spectrum indicates the location of Cu2+ ions in an axially distorted octahedral environment [31] occupying sites vacated by lattice Ti4+ ions [35]. The signal of isolated Cu2+ ions seems to be superimposed by a broader peak which results from Cu2+ interaction in a CuO cluster. The appearance of a broad Cu2+ signal was recently explained by Cu segregation to the titania surface during the crystallization process [36].
TiO2 samples synthesized with Ce also showed a red shift in the absorption toward longer wavelengths. Absorption until 500 nm was previously observed for CeO2 phase [37] and connected with the presence of Ce3+ ions in mesoporous CeO2 or TiO2/CeO2 samples [38]. Therefore, it might be assumed that some parts of the Ce4+ ions have formed metal oxide clusters, probably located on the surface of titania. Otherwise, a red shift in light absorption was also explained by incorporation of Ce4+ ions into the titania lattice where they substitute Ti4+ ions [39]. The amount of Ce integrated into the TiO2 crystallites is expected to be rather low because of the big difference in the ionic radii between Ce4+ and Ti4+. This assumption is supported by XRD results (Figure 1B) which show no lattice distortion in the presence of Ce. Figure S3 shows HAADF-STEM images of Ce-doped mesoporous titania. The images confirm the mesoporous structure und support the assumption that Ce is finely distributed inside or on the surface of the titania framework.
The Ti 2p core level XP spectra of the mesoporous titania samples are presented in Figure S4. The Ti 2p binding energy was in the range of 458.4–458.8 which indicates that the titanium ions are mostly present as Ti(IV), whether the TiO2 was doped or not. Core level O 1 s and C 1 s spectra are shown in Figure 4. The O 1 s spectrum of om-TiO2 shows a signal at 529.4 eV which is attributed to the lattice oxygen in TiO2. The slight shift to higher binding energy in the case of the doped samples might occur from the additional O-Cu or O-Ce bonds. A shoulder peak around 531.7 eV was assigned previously to the presence of physisorbed water. The C 1 s XP spectrum for the three samples is nearly identical, showing the presence of carbon containing C-C, C=O, and C=OOH bonds which originate mainly from the surfactant.
X-ray photoelectron spectroscopy (XPS) was also used to evaluate the surface composition of the materials. In addition to the elements listed in Table 2, the materials also contained small amounts of Si and F, which may result from sealing grease of the used glassware. The low nitrogen content likely originated from the precipitation of the TiO2 with gaseous NH3. The presence of nitrogen indicated that the adsorbed ammonia species were not completely removed from the titania surface during the following surfactant extraction and the thermal annealing. The relatively high surface carbon content indicates that despite repeating the surfactant removal step four times in boiling ethanol, a larger portion of the surfactant was still present on the catalyst surface. The results demonstrate that even after a thermal treatment at 450 °C in pure oxygen the organic template cannot be removed completely. As will become evident in the following paragraphs, this high carbon loading can act as a carbon pool, leading to formation of carbonaceous species that do not originate from the reaction of CO2 with the tested photocatalysts.

2.4. Effect of Mesoporous Structure and Metal Loading in Product Evolution

The materials were tested in CO2 photoreduction in the high purity photoreactor presented in detail below. The measurements were conducted in such a way that a CO2-free experiment under continuous flow (6000 ppm H2O in 6 mL min−1 He for 45 min, followed by a He flow of 6 mL min−1) preceded a batch experiment, in which CO2 was also offered to the photocatalyst, to determine simultaneously the reactivity of the samples in CO2 photoreduction and H2 evolution. The preceding flow experiment was employed as a method to photocatalytically clean the material from the residual organic template that had been used for the synthesis of the mesoporous structure and from impurities adsorbed on the material from the atmosphere. This process has been described and applied previously [40]. The main product formed in the batch experiments from all mesoporous and metal-modified samples in the presence of CO2 and light was hydrogen (H2). The H2 yields for all samples are presented in Figure 5.
It is evident that the mesoporous structure significantly affects the production of H2 when compared to P25 which was used as the benchmark material. This fully agrees with an earlier work of ours, [41] in which the low activity of bare P25 for H2 evolution has been described. The ordered mesoporous TiO2 (om-TiO2) sample produces more than an order of magnitude more H2 than P25-TiO2 in half the reaction time, with a final yield of 2520 ppm gcat−1 h−1 (Figure 4). It must be clearly noted that no cocatalyst (such as Pt) was used in this work. Although the surface area increases only ~threefold in the om-TiO2 material compared to P25, the hydrogen evolution increases by a factor of ~25. Thus, the improvement must be attributed to more than just the larger surface area. Our current study does not allow us to unambiguously trace a specific property of the om-TiO2 as the cause for the improved performance. The possibilities may be the presence of an increased amount of defect sites in the ordered mesoporous structure, an improved separation and migration of charge carriers, and/or a better diffusion of reactants in the structure, in agreement with Refs. [7,8]. Lastly, also the interface formation between anatase and β-TiO2 in om-TiO2, instead of anatase and rutile as in P25, may have a beneficial effect. Further studies are needed to elucidate the specific causes.
The addition of 0.5%mol of Cu in the mesoporous structure leads to a 5.8% percent increase in H2 production, from 2520 to 2666 ppm gcat−1 h−1. The addition of Ce though seems to be detrimental to H2 production as the 0.5%mol Ce/om-TiO2 sample exhibited 8.9% less H2 than the om-TiO2 sample (2295 ppm gcat−1 h−1). The higher SSA and actual metal loading of the Cu-modified sample could explain the better performance of the om-0.5% mol Cu/TiO2 when compared to the 0.5%mol Ce one. Similarly, the slightly lower surface area caused by Ce may have an influence, which is in line with our assumption that any direct influence of the dopant on the reaction is small. Another possibility is that Cu causes a slightly improved absorption of visible light (Figure 3). Cu is also more likely to be reduced to a metallic state under irradiation, which may then provide active sites for H2 formation. As is evident from Figure 5, the presence of the mesoporous structure is the main reason for the much higher H2 evolution compared to P25. While the addition of Cu leads to an increase in H2 production the extra cost associated with the preparation of the Cu-doped om-TiO2 material is not justified by that marginal increase in productivity.
Apart from H2, C-containing products were also identified when reacting the samples with CO2. In the case of P25-TiO2, 66.7 ppm of CO and 17.9 ppm of CH4 were formed after 4.5 h (Figure S1). In previous studies of ours [40,41] we have reproducibly seen that the main product of CO2 photoreduction using P25 is CH4, with a concentration of 80 ppm when using 1.5% CO2 (in He) and 0.6% H2O. In the experiments of the current work, a more diluted concentration was selected (0.5% CO2, 0.6% H2O), and a much-decreased CH4 concentration was observed. The product distribution may also be influenced by residual impurities (see below).
For the mesoporous materials, CH4 and CO and traces of ethane (C2H6) were observed (Figure S5). While these identified C-containing molecules might be considered as products of the reaction of CO2 with the respective samples, this seems not to be the case in these series of measurements (for P25 as well as the mesoporous samples). As was evident from the XPS surface composition analysis of the tested photocatalysts (Table 2), all the mesoporous samples have a carbon-rich surface due to the presence of leftover species from the incomplete removal of the template. These species form a carbon pool which (because of the high thermodynamic stability of CO2) are likely to react faster than CO2 under light irradiation. In this way, an overestimation of the efficiency of the samples toward C-products from CO2 photoreduction is to be expected. It is clear, though, that even if all C-containing species were considered as products of CO2 reduction (which they should not), the observed concentration of H2 formed would be many times higher. For the sake of completeness, the concentrations of the C-containing species formed from the decomposition of the template can be seen in Figure S5. The two metal loaded samples seem to be somewhat more affected by the residual template: much higher amounts of CH4 were formed in the batch experiments compared to the om-TiO2 sample (Figure S5). This may indicate that the metal dopant catalyzes the degradation of the residual template. Although the amounts of CH4 and CO are rather similar regardless of the dopant, C2H6 was also observed for the Ce-doped sample (Figure S5).
It may be assumed that the reaction of the leftover species could also be considered the origin of the H2 formed. However, in an experiment conducted by our group, 100 μmL of isopropanol (iPrOH) were added to the surface of a P25 pellet (total mass 1 g) [42]. Then 1.5% CO2 (in He) was added to the reaction mix and an Hg/Xe lamp (200 W, light intensity 200 mW cm−2) was used to irradiate the reaction chamber. While various C-containing molecules were formed, including CH4, C2H6, and C2H4, no H2 product could be identified, i.e., the concentration was below ~20 ppm. Based on these results, the formation of major amounts of H2 from hydrocarbons leftover in om-TiO2, especially in the amounts presented in Figure 4, is not likely to occur.

3. Materials and Methods

Triblock copolymer P123 (EO20PO70EO20, Sigma-Aldrich, Darmstadt, Germany) was used as a structure–directing agent (surfactant) and titanium isopropoxide (Ti(OiPr)4, 97%, Sigma-Aldrich, Germany) was used as the titania precursor. Cerium chloride heptahydrate (CeCl3·7H2O, 99.9%) was also purchased from Sigma-Aldrich (Germany). Copper chloride dihydrate (CuCl2·2H2O, 99%) and HCl (37%) were obtained from VWR International (Darmstadt, Germany). Absolute ethanol was obtained from J. T. Baker (Darmstadt, Germany). All chemicals were used as received without any further purification.

3.1. Synthesis of Un-Doped and Metal Doped Ordered Mesoporous Titania

An amount of 1 g of P123 was dissolved in 20 g of ethanol in a two-necked round bottom flask (100 mL) at room temperature (RT) under stirring. After 30 min of stirring at RT, 2.1 g of HCl (37%), 3 g of Ti(OiPr)4, × g of metal source (for more details see Table S1), and 2 g of distilled water were added to the P123-containing solution. Immediately after water addition, the flask was flushed with argon to avoid the formation of a titania precipitate. The two-necked round bottom flask containing the clear reaction mixture was immersed in a water bath which was heated to a temperature of 40 °C. After the reaction mixture had achieved the desired temperature, the argon flow was stopped and the solvent removal was started under reduced pressure (250 mbar). With an increasing amount of evaporated solvent, the pressure was reduced finally to a value of 55 mbar. The gels obtained after solvent evaporation were dried in an oven at 40 °C in air atmosphere for 15 to 18 h. The aged materials were deposited as a thin layer on Petri dishes by using a spatula and placed in a desiccator. The air in the desiccator was exchanged for ammonia and the solid material was treated at RT with ammonia for 18–22 h to allow the TiO2 precipitation. The P123 surfactant was removed from the titania framework by treatment of the solid material with boiling ethanol for 1 h. This procedure was repeated 4 times. Finally, the obtained powder was heated under argon flow (100 mL/min) from 22–450 °C for 6.5 h and then 2 h under oxygen flow (100 mL/min) at 450 °C. Pristine mesoporous TiO2 was also prepared using the same procedure without the addition of metal source.

3.2. Characterization

SAXS measurements were carried out using a Kratky-type instrument (SAXSess, Anton Paar, Graz, Austria) operated at 40 kV and 50 mA in slit collimation using a two-dimensional CCD detector cooled to –40 °C or image plates digitalized with a Cyclone Plus (Perkin Elmer, Downers Grove, IL, USA). A Göbel mirror was used to convert a divergent polychromatic X-ray beam into a collimated line-shaped beam of CuKα radiation (λ = 0.154 nm). Slit collimation of the primary beam was aimed to increase the flux and to improve the signal quality. The sample cell consisted of a metal body with two windows for the beam. The samples were sealed between two layers of adhesive tape. Scattering profiles of the mesoporous materials were obtained by subtraction of the detector current background and the scattering pattern of adhesive tape from the experimental scattering patterns. The 2D scattering patterns were converted with SAXS Quant software (Anton Paar) into one-dimensional scattering curves as a function of the magnitude of the scattering vector q = (4/λ) sin (θ/2). Correction of the instrumental broadening effects (smearing) was carried out with SAXS Quant software using the slit length profile determined in a separate experiment. All SAXS measurements were carried out at 25 °C.
XRD powder patterns were recorded either on a Panalytical X’Pert diffractometer (Almelo, the Netherlands) equipped with an Xcelerator detector or on a Panalytical Empyrean diffractometer equipped with a PIXcel3D detector system, both used with automatic divergence slits and Cu Kα radiation (40 kV, 40 mA; λ = 0.015406 nm). The measurements were performed in 0.0167° steps and 25 s of data collecting time per step. Phase analysis was performed with the Panalytical HighScore Plus software package using the PDF-2 2015 database of the ICDD. Crystallite sizes were calculated by applying the Scherrer equation and the usage of the integral breadth under the assumption of spherically shaped crystallites. K is set to 1.0747. The average value of the calculated sizes of diffraction peaks between 20 ≤ 2θ ≤ 50° is presented here.
Nitrogen adsorption–desorption isotherms were collected at –196 °C on BELSORP-mini II (BEL Japan, Inc., Osaka, Japan). The specific surface area and pore size distribution were calculated from the adsorption and desorption branches of the isotherm, respectively, applying the BET equation for the N2 relative pressure range of 0.05 < P/P0 < 0.3 and the Barrett–Joyner–Halenda (BJH) method for the pressure range of 0.3 < P/P0 < 0.99.
UV/Vis spectra were measured with the AvaSpec-2048 fiber optic spectrometer (Avantes B.V., Apeldoorn, the Netherlands) using the FCR7UV-400-400-2-Me-HT Fiber optic and pure barium sulfate BaSO4 powder as the white reflection standard of referencing.
The oxidation states and the surface compositions were determined by XPS. The measurements were performed with an ESCALAB 220iXL (Thermo Fischer Scientific, Waltham, MA, USA) with monochromatic Al Kα radiation (E = 1486.6 eV). The samples were fixed on a stainless steel sample holder with double adhesive carbon tape. For charge compensation a flood gun was used. The spectra were referenced to the C1s peak at 284.8 eV. After background subtraction the peaks were fitted with Gaussian-Lorentzian curves to determine the positions and areas of the peaks. The surface composition was calculated from the peak areas divided by the element-specific Scofield factor and the transmission function of the spectrometer.

3.3. Photocatalytic Experiments

Photocatalytic experiments were conducted in a high purity gas-phase photoreactor system as reported previously [40,41]. The system is designed specifically to prevent carbonaceous impurities resulting from reactor construction. A 200 W Hg/Xe lamp with water filter to remove the IR radiation (UV+vis irradiation with an intensity of approximately 200 mW cm−1) was used for all experiments. The measurement procedure comprised a pure gas-phase water splitting experiment under conditions of continuous flow. High purity helium (purity 99.9999%) at a flow rate of 6 mL min−1 was mixed with 0.6% of water using a stainless steel saturator cooled to 5 °C. Water dosing was switched off after 45 min, and from that time onwards, only adsorbed water on the samples was present. Apart from checking the activity in pure water splitting, this measurement also serves to photocatalytically degrade organic impurities still left on the mesoporous titania surface. Gas analysis was carried out with a calibrated online gas chromatograph (Shimadzu TRACERA, Duisburg, Germany) with a barrier discharge ionization detector (BID). The gas analytics can detect CH4, CO, CO2, H2O, H2, C2H4, C2H6, O2, and N2. For CH4, the detection limit is ~0.1 ppm; for all other gases ~1 ppm.
The flow experiment was followed by a batch experiment in which the reactor was filled with 0.5% CO2 (purity 99.9995%) in helium (purity 99.9999%) to an initial pressure of 1500 mbar. Gas analysis was similar to the system described above. Due to gas sampling from the batch reactor, the pressure in the reactor drops upon each measurement. This pressure drop was corrected for by the calibration curves used for the GC analytics.

4. Conclusions

An ordered mesoporous TiO2 was presented with very high activity in photocatalytic hydrogen evolution in absence of any cocatalyst. The yields were more than an order of magnitude larger than for commercial P25 and were not just related to the larger specific surface area, which increased only ~threefold. The ordered mesoporous structure, rather than the doping with Cu or Ce, was the main factor leading to the high activity. The dopants had only a moderate effect, which was positive for Cu and detrimental for Ce. Although carbon-containing products were also observed, the inability to completely remove the template from the ordered mesoporous structure makes it impossible to unambiguously assign these species to products from CO2 reduction. This again highlights the need for proper reference experiments, whenever the synthesis procedure involves considerable amounts of organic species.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal13050895/s1, Figure S1: UV/Vis DRS spectra of pure P25, CeO2, and CuO. Figure S2: Room temperature EPR spectrum of 0.5 mol% Cu/TiO2. Figure S3: HAADF-STEM images of om-0.5 mol% Ce/TiO2 using two different magnifications. Figure S4: Ti 2p XP spectra of (A) om-TiO2, (B) 0.5 mol% Cu/TiO2, and (C) 0.5 mol% Cu/TiO2. Figure S5: Overview of formed carbon-containing species on various samples.

Author Contributions

Conceptualization, N.S. and J.S.; formal analysis, A.M.M. and A.E.B.; investigation, A.M.M., A.E.B. and N.S.; resources, N.S. and J.S.; supervision, N.S. and J.S.; validation, A.M.M., A.E.B., S.R. and N.S.; visualization, A.M.M. and N.S.; writing—original draft, A.M.M., N.S. and J.S.; writing—review and editing, S.R., N.S. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The APC was funded by the Leibniz Institute for Catalysis.

Data Availability Statement

Data is available from the authors upon reasonable request.

Acknowledgments

We would like to thank Stephan Bartling and Henrik Lund, Leibniz Institute for Catalysis, for the XPS and Powder XRD measurements, respectively. Jabor Rabeah, Leibniz Institute for Catalysis, is gratefully acknowledged for EPR measurements. We are thankful to Nikolaos Moustakas, Leibniz Institute for Catalysis, for assistance in data processing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moustakas, N.G.; 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. [Google Scholar] [CrossRef] [PubMed]
  2. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and perspectives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic processes. Energy Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef]
  3. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for environmental photocatalytic applications: A review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  4. Xiong, Z.; Zhao, Y.; Zhang, J.; Zheng, C. Efficient photocatalytic reduction of CO2 into liquid products over cerium doped titania nanoparticles synthesized by a sol–gel auto-ignited method. Fuel Process. Technol. 2015, 135, 6–13. [Google Scholar] [CrossRef]
  5. Fattakhova-Rohlfing, D.; Zaleska, A.; Bein, T. Three-dimensional titanium dioxide nanomaterials. Chem. Rev. 2014, 114, 9487–9558. [Google Scholar] [CrossRef]
  6. Cherevan, A.S.; Deilmann, L.; Weller, T.; Eder, D.; Marschall, R. Mesoporous Semiconductors: A New Model to Assess Accessible Surface Area and Increased Photocatalytic Activity? ACS Appl. Energy Mater. 2018, 1, 5787–5799. [Google Scholar] [CrossRef]
  7. Zhang, R.; Elzatahry, A.A.; Al-Deyab, S.S.; Zhao, D. Mesoporous titania: From synthesis to application. Nano Today 2012, 7, 344–366. [Google Scholar] [CrossRef]
  8. Zhang, W.; Tian, Y.; He, H.; Xu, L.; Li, W.; Zhao, D. Recent advances in the synthesis of hierarchically mesoporous TiO2 materials for energy and environmental applications. Natl. Sci. Rev. 2020, 7, 1702–1725. [Google Scholar] [CrossRef]
  9. Zhang, T.; Low, J.; Koh, K.; Yu, J.; Asefa, T. Mesoporous TiO2 Comprising Small, Highly Crystalline Nanoparticles for Efficient CO2 Reduction by H2O. ACS Sustain. Chem. Eng. 2018, 6, 531–540. [Google Scholar] [CrossRef]
  10. Lo, A.-Y.; Taghipour, F. Ordered mesoporous photocatalysts for CO2 photoreduction. J. Mater. Chem. A 2021, 9, 26430–26453. [Google Scholar] [CrossRef]
  11. Parlett, C.M.; Wilson, K.; Lee, A.F. Hierarchical porous materials: Catalytic applications. Chem. Soc. Rev. 2013, 42, 3876–3893. [Google Scholar] [CrossRef]
  12. Ismail, A.A.; Bahnemann, D.W. Mesoporous titania photocatalysts: Preparation, characterization and reaction mechanisms. J. Mater. Chem. 2011, 21, 11686–11707. [Google Scholar] [CrossRef]
  13. Bonelli, B.; Esposito, S.; Freyria, F.S. Mesoporous Titania: Synthesis, Properties and Comparison with Non-Porous Titania. In Titanium Dioxide; InTechOpen: Rijeka, Croatia, 2017. [Google Scholar]
  14. Zimny, K.; Ghanbaja, J.; Carteret, C.; Stébé, M.-J.; Blin, J.-L. Highly ordered mesoporous titania with semi crystalline framework templated by large or small nonionic surfactants. New J. Chem. 2010, 34, 2113–2117. [Google Scholar] [CrossRef]
  15. Zimny, K.; Roques-Carmes, T.; Carteret, C.; Stébé, M.; Blin, J. Synthesis and photoactivity of ordered mesoporous titania with a semicrystalline framework. J. Phys. Chem. C 2012, 116, 6585–6594. [Google Scholar] [CrossRef]
  16. Mohammed, A.M.; Sebek, M.; Kreyenschulte, C.; Lund, H.; Rabeah, J.; Langer, P.; Strunk, J.; Steinfeldt, N. Effect of metal ion addition on structural characteristics and photocatalytic activity of ordered mesoporous titania. J. Sol-Gel Sci. Technol. 2019, 91, 539–551. [Google Scholar] [CrossRef]
  17. Chen, L.; Li, Y.-J.; Peng, X.; Li, Z.-S.; Zeng, M.-X. Preparation and improved photocatalytic activity of ordered mesoporous TiO2 by evaporation induced self-assembly technique using liquid crystal as template. T. Nonferr. Metal. Soc. 2014, 24, 1072–1078. [Google Scholar]
  18. Zhang, W.; He, H.; Tian, Y.; Lan, K.; Liu, Q.; Wang, C.; Liu, Y.; Elzatahry, A.; Che, R.; Li, W.; et al. Synthesis of uniform ordered mesoporous TiO2 microspheres with controllable phase junctions for efficient solar water splitting. Chem. Sci. 2019, 10, 1664–1670. [Google Scholar] [CrossRef]
  19. Wang, T.; Meng, X.G.; Li, P.; Ouyang, S.X.; Chang, K.; Liu, G.G.; Mei, Z.W.; Ye, J.H. Photoreduction of CO2 over the well-crystallized ordered mesoporous TiO2 with the confined space effect. Nano Energy 2014, 9, 50–60. [Google Scholar] [CrossRef]
  20. Lee, Y.Y.; Jung, H.S.; Kim, J.M.; Kang, Y.T. Photocatalytic CO2 conversion on highly ordered mesoporous materials: Comparisons of metal oxides and compound semiconductors. Appl. Catal. B-Environ. 2018, 224, 594–601. [Google Scholar] [CrossRef]
  21. Wang, T.; Meng, X.; Liu, G.; Chang, K.; Li, P.; Kang, Q.; Liu, L.; Li, M.; Ouyang, S.; Ye, J. In situ synthesis of ordered mesoporous Co-doped TiO2 and its enhanced photocatalytic activity and selectivity for the reduction of CO2. J. Mater. Chem. A 2015, 3, 9491–9501. [Google Scholar] [CrossRef]
  22. Xue, H.; Wang, T.; Gong, H.; Guo, H.; Fan, X.; Gao, B.; Feng, Y.; Meng, X.; Huang, X.; He, J. Constructing Ordered Three-Dimensional TiO2 Channels for Enhanced Visible-Light Photocatalytic Performance in CO2 Conversion Induced by Au Nanoparticles. Chem. Asian J. 2018, 13, 577–583. [Google Scholar] [CrossRef] [PubMed]
  23. Xiong, J.; Zhu, X.; Chen, Z.; Ren, Y.; Wang, W.; Han, C.; Li, W.; Li, S.; Cheng, G. Oxygen Vacancy and Metallic Silver Site Coinjection Associates Photocatalytic CO2 Reduction upon Mesoporous NH2–TiO2 Nanoparticles Assembly. Solar RRL 2022, 6, 2200657. [Google Scholar] [CrossRef]
  24. Sing, K.S.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  25. Zuas, O.; Budiman, H. Synthesis of nanostructured copper-doped titania and its properties. Nano-Micro Lett. 2013, 5, 26–33. [Google Scholar] [CrossRef]
  26. Wang, Q.; Jin, R.; Zhang, M.; Gao, S. Solvothermal preparation of Fe-doped TiO2 nanotube arrays for enhancement in visible light induced photoelectrochemical performance. J. Alloys Compd. 2017, 690, 139–144. [Google Scholar] [CrossRef]
  27. Zhang, D. Chemical synthesis of Ni/TiO2 nanophotocatalyst for UV/visible light assisted degradation of organic dye in aqueous solution. J. Sol-Gel Sci. Technol. 2011, 58, 312–318. [Google Scholar] [CrossRef]
  28. Brik, M.; Srivastava, A.; Popov, A. A few common misconceptions in the interpretation of experimental spectroscopic data. Opt. Mater. 2022, 127, 112276. [Google Scholar] [CrossRef]
  29. Jubu, P.; Yam, F.; Igba, V.; Beh, K. Tauc-plot scale and extrapolation effect on bandgap estimation from UV–vis–NIR data–a case study of β-Ga2O3. J. Solid State Chem. 2020, 290, 121576. [Google Scholar] [CrossRef]
  30. Makuła, P.; Pacia, M.; Macyk, W. How to correctly determine the band gap energy of modified semiconductor photocatalysts based on UV–Vis spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef]
  31. Choudhury, B.; Dey, M.; Choudhury, A. Defect generation, d-d transition, and band gap reduction in Cu-doped TiO2 nanoparticles. Int. Nano Lett. 2013, 3, 25. [Google Scholar] [CrossRef]
  32. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO–TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef]
  33. Anitha, B.; Khadar, M.A. Dopant concentration dependent magnetism of Cu-doped TiO2 nanocrystals. J. Nanopart. Res. 2016, 18, 149. [Google Scholar] [CrossRef]
  34. Colon, G.; Maicu, M.; Hidalgo, M.S.; Navio, J. Cu-doped TiO2 systems with improved photocatalytic activity. Appl. Catal. B 2006, 67, 41–51. [Google Scholar] [CrossRef]
  35. Li, G.; Dimitrijevic, N.M.; Chen, L.; Rajh, T.; Gray, K.A. Role of surface/interfacial Cu2+ sites in the photocatalytic activity of coupled CuO−TiO2 nanocomposites. J. Phys. Chem. C 2008, 112, 19040–19044. [Google Scholar] [CrossRef]
  36. Baltazar, P.; Lara, V.; Cordoba, G.; Arroyo, R. Kinetics of the amorphous—Anatase phase transformation in copper doped titanium oxide. J. Sol-Gel Sci. Technol. 2006, 37, 129–133. [Google Scholar] [CrossRef]
  37. Yao, X.; Tang, C.; Ji, Z.; Dai, Y.; Cao, Y.; Gao, F.; Dong, L.; Chen, Y. Investigation of the physicochemical properties and catalytic activities of Ce 0.67 M 0.33 O2 (M = Zr4+, Ti4+, Sn4+) solid solutions for NO removal by CO. Catal. Sci. Technol. 2013, 3, 688–698. [Google Scholar] [CrossRef]
  38. Zeng, M.; Li, Y.; Mao, M.; Bai, J.; Ren, L.; Zhao, X. Synergetic effect between photocatalysis on TiO2 and thermocatalysis on CeO2 for gas-phase oxidation of benzene on TiO2/CeO2 nanocomposites. ACS Catalysis 2015, 5, 3278–3286. [Google Scholar] [CrossRef]
  39. Deng, C.; Li, B.; Dong, L.; Zhang, F.; Fan, M.; Jin, G.; Gao, J.; Gao, L.; Zhang, F.; Zhou, X. NO reduction by CO over CuO supported on CeO2-doped TiO2: The effect of the amount of a few CeO2. Phys. Chem. Chem. Phys. 2015, 17, 16092–16109. [Google Scholar] [CrossRef]
  40. Dilla, M.; Schlögl, R.; Strunk, J. Photocatalytic CO2 Reduction Under Continuous Flow High-Purity Conditions: Quantitative Evaluation of CH4 Formation in the Steady-State. ChemCatChem 2017, 9, 696–704. [Google Scholar] [CrossRef]
  41. Pougin, A.; Dilla, M.; Strunk, J. Identification and exclusion of intermediates of photocatalytic CO2 reduction on TiO2 under conditions of highest purity. Phys. Chem. Chem. Phys. 2016, 18, 10809–10817. [Google Scholar] [CrossRef]
  42. Moustakas, N.G.; Lorenz, F.; Dilla, M.; Peppel, T.; Strunk, J. Pivotal Role of Holes in Photocatalytic CO2 Reduction on TiO2. Chemistry 2021, 27, 17213–17219. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) SAXS and (B) XRD powder patterns of pristine and doped TiO2 materials prepared with 0.5 mol% doping metal after annealing at 450 °C; a—pristine om-TiO2, b—om-0.5 mol% Cu/TiO2, c—om-0.5 mol% Ce/TiO2 (the single SAXS pattern was shifted by multiplication with a constant factor).
Figure 1. (A) SAXS and (B) XRD powder patterns of pristine and doped TiO2 materials prepared with 0.5 mol% doping metal after annealing at 450 °C; a—pristine om-TiO2, b—om-0.5 mol% Cu/TiO2, c—om-0.5 mol% Ce/TiO2 (the single SAXS pattern was shifted by multiplication with a constant factor).
Catalysts 13 00895 g001
Figure 2. N2 adsorption–desorption isotherms and pore size distribution of pristine and doped mesoporous om-TiO2.
Figure 2. N2 adsorption–desorption isotherms and pore size distribution of pristine and doped mesoporous om-TiO2.
Catalysts 13 00895 g002
Figure 3. UV/Vis diffuse reflectance spectra (DRS) of (A)—pristine om-TiO2 and materials synthesized with 0.5 mol% doping metal and (B) the corresponding Tauc plots for estimation of the band gap.
Figure 3. UV/Vis diffuse reflectance spectra (DRS) of (A)—pristine om-TiO2 and materials synthesized with 0.5 mol% doping metal and (B) the corresponding Tauc plots for estimation of the band gap.
Catalysts 13 00895 g003
Figure 4. Deconvoluted O 1 s (AC) and C 1 s (DF) core level spectra of om-TiO2 (A,D), om-0.5 mol%Cu/TiO2 (B,E), and om-0.5 mol%Ce/TiO2 (C,F).
Figure 4. Deconvoluted O 1 s (AC) and C 1 s (DF) core level spectra of om-TiO2 (A,D), om-0.5 mol%Cu/TiO2 (B,E), and om-0.5 mol%Ce/TiO2 (C,F).
Catalysts 13 00895 g004
Figure 5. H2 production from tested samples under batch conditions in the presence of humidified CO2.
Figure 5. H2 production from tested samples under batch conditions in the presence of humidified CO2.
Catalysts 13 00895 g005
Table 1. Characteristics of calcined om-TiO2 prepared with 0.5 mol% of Cu or Ce; n.d. = not determined.
Table 1. Characteristics of calcined om-TiO2 prepared with 0.5 mol% of Cu or Ce; n.d. = not determined.
SampleCrystal Phases
m2/g
Vp
cm3/g
dp
nm
Cryst. Size *
nm
Band Gap
eV
P25Anatase/Rutile50n.d.n.d.n.d.3.2
om-TiO2Anatase/β-TiO21660.273.953.17 ± 0.04
0.5 mol% CuAnatase/β-TiO21690.233.953.18 ± 0.04
0.5 mol% CeAnatase/β-TiO21350.243.943.22 ± 0.04
* Anatase.
Table 2. Near surface composition of the different materials determined by XPS.
Table 2. Near surface composition of the different materials determined by XPS.
SampleTi
Atom%
O
Atom%
C
Atom%
N
Atom%
Dopant
Atom%
P25
pristine om-TiO223.855.118.80.5n.a.
om-TiO2/0.5 mol% Cu23.856.018.50.60.3
om-TiO2/0.5 mol% Ce25.352.020.10.60.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohammed, A.M.; Becerikli, A.E.; Ristig, S.; Steinfeldt, N.; Strunk, J. Ordered Mesoporous TiO2: The Effect of Structure, Residual Template and Metal Doping on Photocatalytic Activity. Catalysts 2023, 13, 895. https://doi.org/10.3390/catal13050895

AMA Style

Mohammed AM, Becerikli AE, Ristig S, Steinfeldt N, Strunk J. Ordered Mesoporous TiO2: The Effect of Structure, Residual Template and Metal Doping on Photocatalytic Activity. Catalysts. 2023; 13(5):895. https://doi.org/10.3390/catal13050895

Chicago/Turabian Style

Mohammed, Ahmed M., Ahmet E. Becerikli, Simon Ristig, Norbert Steinfeldt, and Jennifer Strunk. 2023. "Ordered Mesoporous TiO2: The Effect of Structure, Residual Template and Metal Doping on Photocatalytic Activity" Catalysts 13, no. 5: 895. https://doi.org/10.3390/catal13050895

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