Next Article in Journal
The Effect of Repeated Pressing on the Flexural Strength, Color Stability, Vickers Hardness, and Surface Topography of Heat-Pressed Lithium Disilicate
Next Article in Special Issue
Sol-Gel Synthesis and Characterization of Yttrium-Doped MgFe2O4 Spinel
Previous Article in Journal
Eddy Current Testing of Artificial Defects in 316L Stainless Steel Samples Made by Additive Manufacturing Technology
Previous Article in Special Issue
Versatile Zirconium Oxide (ZrO2) Sol-Gel Development for the Micro-Structuring of Various Substrates (Nature and Shape) by Optical and Nano-Imprint Lithography
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reaching Visible Light Photocatalysts with Pt Nanoparticles Supported in TiO2-CeO2

by
Ixchel Alejandra Mejia-Estrella
1,
Alejandro Pérez Larios
2,
Belkis Sulbarán-Rangel
1,* and
Carlos Alberto Guzmán González
3,*
1
Department of Water and Energy, University of Guadalajara, Campus Tonalá, Tonalá 45425, Mexico
2
Department of Engineering, University of Guadalajara, Campus Altos, Tepatitlán de Morelos 47635, Mexico
3
Department of Applied Basic Sciences, University of Guadalajara, Campus Tonalá, Tonalá 45425, Mexico
*
Authors to whom correspondence should be addressed.
Submission received: 28 July 2022 / Revised: 16 September 2022 / Accepted: 23 September 2022 / Published: 30 September 2022

Abstract

:
Nanostructured catalysts of platinum (Pt) supported on commercial TiO2, as well as TiO2-CeO2 (1, 5 and 10 wt% CeO2), were synthesized through the Sol-Gel and impregnation method doped to 1 wt% of Platinum, in order to obtain a viable photocatalytic material able to oxidate organic pollutants under the visible light spectrum. The materials were characterized by different spectroscopy and surface techniques such as Specific surface area (BET), X-ray photoelectron spectroscopy (XPS), XRD, and TEM. The results showed an increase in the diameter of the pore as well as the superficial area of the supports as a function of the CeO2 content. TEM images showed Pt nanoparticles ranking from 2–7 nm, a decrease in the particle size due to the increase of CeO2. The XPS showed oxidized Pt2+ and reduced Pt0 species; also, the relative abundance of the elements Ce3+/Ce4− and Ti4+ on the catalysts. Additionally, a shift in the Eg band gap energy (3.02–2.82 eV) was observed by UV–vis, proving the facticity of applying these materials in a photocatalytic reaction using visible light. Finally, all the synthesized materials were tested on their photocatalytic oxidation activity on a herbicide used worldwide; 2,4-Dichlorophenoxyacetic acid, frequently use in the agriculture in the state of Jalisco. The kinetics activity of each material was measured during 6 h of reaction at UV–Vis 190–400 nm, reaching a removal efficiency of 98% of the initial concentration of the pollutant in 6 h, compared to 32% using unmodified TiO2 in 6 h.

1. Introduction

As the population continues to grow, pollution has increased in all water resources, producing an urgent need to create solutions to remediate it. Studying the literature regarding a reliable technology for water treatment able to oxidate persistent organic molecules heterogeneous catalysts, it has been proven to be able to remove a wide range of contaminants [1,2,3]. By doping the reaction with metallic and nanometric catalysts helps to increase the surface energy of the individual particles, increasing the probability of aggregation, which can reduce the specific surface area of the catalyst and its efficiency since they are widely used materials due to their relatively low cost, and they can be reused [4,5]. To avoid aggregation, it is important to immobilize the active metal nanoparticles on mesoporous solid support, as well as to transform the metal into its oxide form [6].
The morphological structure of the catalyst support determines the dispersion of the nanoparticles and the surface area of the catalytic active sites. Various semiconductor materials including titanium dioxide (TiO2), zinc oxide (ZnO), vanadium pentoxide (V2O5), cerium oxide (CeO2) and tungsten trioxide (WO3) have been extensively studied by photo catalysis reaction [1,7,8]. One of the most researched compositions for a support system is TiO2 because it is an effective, inexpensive, and stable photocatalyst used for the decomposition of organics [1,4]. However, TiO2 can only absorb the ultraviolet portion and can only take advantage of about 4% of the intensity of the sunlight spectrum due to its high bandgap (3.0–3.2 eV) [1,9]. This represents a major limitation since its photocatalytic properties are not fully used; even so, an alternative that has been explored to extend its photo response range to the region is to dope the surface with metallic nanoparticles or combine it with another support [4]. Selected methods like doping and composites have been attempted to achieve photo-initiation into the visible spectrum, therefore decreasing cost and increasing efficiency [6]. Another support that has gained importance recently and that can be used as a photocatalyst is CeO2. This is due to its unique redox properties that consist of reversibly creating and eliminating oxygen vacancies on the surface [10].
The relation between the photocatalytic activity of TiO2 and CeO2 composite under UV and visible light has been studied [8,11,12,13]. Liu et al. in 2005 found that TiO2-CeO2 under visible illumination exhibits more photocatalytic activity than pure TiO2 and CeO2 films. Another study reported by Tian et al. in 2013, who prepared heterostructures of CeO2/TiO2 nanobelts using a hydrothermal method. These authors found that both UV and visible photocatalytic activities of CeO2/TiO2 nanobelt heterostructures were enhanced compared to TiO2 nanobelts and CeO2 nanoparticles. More recently, Henych et al. (2021) found the strong interaction of Ti with Ce within the composites led to the formation of Ce3+ and Ti<4+ states, reduction of titania crystallite size, change of acid-base and surface properties, and synergetic effects that are all responsible for highly improved degradation efficiency of organophosphorus compounds. In addition, the thermocatalytic, photocatalytic, and photothermocatalytic oxidation of some volatile organic compounds, 2-propanol, ethanol, and toluene, were investigated over brookite TiO2-CeO2 composites [11]. Other studies have focused on improving catalyst synthesis methods using green methods [14] or adding doping of TiO2-CeO2 supports to improve photocatalytic properties [15,16].
In order to improve the combination of TiO2 and CeO2 supports, the incorporation of platinum nanoparticles was studied in this research. Platinum nanoparticles are advantageous in biological, biosensor, electro-analytical, analytical, and catalytic applications [17]. They are unique because of their large surface area and their numerous catalytic applications, such as their use as automotive catalytic converters and as petrochemical cracking catalysts [18]. As mentioned above, the present work takes advantage of the combined photoactivity properties of two semiconductors TiO2-CeO2 supports and the platinum nanoparticles forming Pt/TiO2-CeO2 photocatalyst. The TiO2-CeO2 supports were prepared by sol-gel method at different contents of the cerium oxide (2–10 wt%) and the platinum nanoparticles (1.0 wt%) were prepared by the impregnation method to obtain Pt/TiO2-CeO2 photocatalysts. This project is innovative because it will use the semiconductor, TiO2 mixed with CeO2, mechanically alloyed, to shift photo-initiation into the visible range.

2. Materials and Methods

2.1. Materials

Cerium (IV) oxide reagent grade, 97%, was used as a support for the catalyst in the sol-gel method. The TiO2 P25 reagent grade 99.5%, commercial salt of hexachloroplatinic for a precursor of nanoparticles of Pt at 37.5% purity (H2PtCL6 * 6H2O), the reactive used to reach pH in the preparation methods was nitric acid (HNO3) reagent at 65%, ethanol (C2H5OH) at 99.8% hydrochloric acid (HCl) at 99.9%. All reagents were obtained from Sigma-Aldrich (Toluca, México).

2.2. Support Preparation

2.2.1. Impregnation Synthesis

The TiO2 supports was prepared using TiO2 Degussa P25 (Aldrich, 99.5%) it was first placed in a thermal treatment at 500 °C for 2 h with an airflow of 50 mL/min. The CeO2 was incorporated into the TiO2 with different contents (1, 5, and 10% by weight of CeO2) with an aqueous solution of Ce (NO3)3 * 6H2O. This solution was added to the TiO2 that was placed in a ball flask. The mixture was stirred for 3 h on a rotary evaporator. Afterwards, the samples were dried under vacuum in a 60 °C water bath. Subsequently they were dried in an oven at 120 °C for 12 h and calcined at 500 °C for 4 h with an air flow (50 mL/min) and a heating ramp of 2 °C/min. This process was performed in duplicate.

2.2.2. Sol-Gel Synthesis

To prepare the TiO2 support material by sol-gel, titanium IV butoxide (Aldrich, 97%) was used as a TiO2 precursor with water, ethanol, and a few drops of HNO3 to fix the pH in the solution to 3. The preparation of the supports was made in a three-necked flask, mixing dropwise the n-Butoxide into the mix in the water/alkoxide solution (8:1 molar ratio). Then the mixture was placed to reflux and stirred vigorously for 24 h. The temperature of the preparation was maintained in a range between 75–80 °C. The samples were dried under vacuum on a rotary evaporator with a 75 °C water bath. Finally, the supports were calcined at 500 °C for 4 h with an airflow (50 mL/min) and a heating ramp of 2 °C/min [19].
The process to add the CeO2 into the support web was the same as described previous for the TiO2 sol-gel, with the difference of adding the reagent Ce (NO3)3 * 6H2O by previously preparing a solution in order to obtain the desired percentages in the support web. This process was performed in duplicate and in parallel.

2.3. Pt Catalysts Preparation

The catalysts were prepared by wet impregnation, using the support material previously synthesized with TiO2 and CeO2. Prior to catalyst preparation, both TiO2 and TiO2-CeO2 supports were previously air-dried at 100 °C for 24 h. Subsequently, the supports were added to a ball flask to which a 0.001 M hydrochloric acid solution was used to adjust the pH. Then the mixed solution with the support was left stirring until it became homogenized, after which the Pt solution was added, and commercial salt of H2PtCl6 * 6H2O at 37.5% purity was employed as a precursor for Pt in order to obtain a semiconductor material with a 1% weight metal content. The suspended solution was heated to 60 °C with vigorous stirring for 4 h. The leftover solids were dried in the oven at 120 °C for 24 h and calcined at 500 °C with an airflow of 1 cm3/s and heating rate of 2 °C/min. Finally, the samples were placed in a vacuum desiccator in amber glass vials wrapped in foil paper to mitigate the exposure to light [20].

2.4. Characterization Techniques

The TiO2-CeO2 supports and the platinum nanoparticles that form the Pt/TiO2-CeO2 photocatalyst synthesized by sol-gel and impregnation at different concentrations were characterized to determine which method improved their photocatalytic properties. The determination of the specific surface area was carried out using the standard Brunauer, Emmett and Teller (BET) method using nitrogen physisorption in Micromeritics ASAP 2020. The X-ray diffraction (XRD) was used to determine their phases and crystallinity. This was carried out using an Empyrean by Malvern Panalytical, Almelo, equipped with Cu-Kα radiation (λ = 0.154 nm). The phase content of anatase and rutile were calculated with the XRD intensity of the characteristic peaks of the phases [21,22], as shown in Equation (1).
W A = K A I A ( K A I A + I R )
where WA is the mole fractions of anatase, IA and IR are the X-ray integrated intensities of the anatase and the rutile, respectively, and KA = 0.886.
The presence of elements in the catalyst and the percentage of each element were determined using X-ray photoelectron spectrometry (XPS) Phoipos 150 (ESCALAB 210, VG Scientific Ltd., East Grinstead, UK) and Raman spectroscopy (Cora 5500 Anton Paar, Anton Paar, Germany). Transmission Electron Microscopy (TEM) has been used to explicate the innermost structure, morphology, and exact particle size of the composite system (FEI TITAN G2 80–300, Hillsborough, OR, USA) operated at 300 keV. The UV–Vis (UV–Visible) spectrophotometer (Shimadzu UV-2600, Kyoto, Japan) was used to determine the energy level of the band gap for all composites of Pt/TiO2-CeO2 photocatalyst synthesized.

2.5. Photocatalytic Reaction

The photodegradation experiments of 2,4-D was carried out at room temperature, using a slurry reactor, a glass beaker of 200 mL capacity with 150 mL of a mother solution with 200 ppm of 2,4-Dimethylamine salt, and 200 mg of catalyst mixed with air (BOYUS air pump 4000 B, with pressure of 0.012 MPa and an output of 3.2 L per minute). The reaction was kept in agitation for 30 min in complete darkness until the adsorption desorption equilibrium had reached light striking the reactor with a UV lamp (4 watts). The run time for adsorption tests was 6 h under darkness, at the natural pH of the slurry. An example was obtained every 30 min, using a filtrating syringe to extract 4 mL of the slurry and using a membrane to separate the suspension material. Every example was measured in a UV–Vis (UV-2600 Shimadzu) at 190–400 nm. To obtain the kinetics activity of each material. The concentration of the reaction was calculated from the absorption band at 282 nm, applying the equation of Beer–Lambert. The conversion percentage was calculated using the Equation (2).
X 2 , 4 D = 2 , 4 D 0 2 , 4 D f 2 , 4 D 0 × 100 %
where X2,4D is the percentage of the 2,4-D conversion, 2,4D0 is the concentration of the pollutant at the beginning of the reaction, and 2,4Df is the concentration of the pollutant at the end of the reaction.
The heterogenous photocatalysts were carried out by employing the following catalyst: Aeroxide P-25® Commercial TiO2, TiO2-CeO2 (5, 10 wt%), Pt-TiO2-CeO2 (5, 10 wt%) synthesis by impregnation and TiO2-CeO2 (1, 5, 10 wt%), Pt-TiO2-CeO2 (1, 5, 10 wt%) synthesis by sol-gel.

3. Results and Discussion

3.1. Specific Area by the BET Method

In order to investigate the effect that was created in the surface of the support with the addition of CeO2 in different concentrations, several material characterization techniques were made. Starting with the specific area determined by the BET method, as well as the average pore diameter of the TiO2 and TiO2-CeO2 supports, we can observe that the effect of adding CeO2 to TiO2 increased the pore diameter and the specific area decreased. The impregnation method generated a higher diameter of pores and low specific surface area than the sol-gel method, probably because the sol-gel method had better dispersion of CeO2 in the TiO2. Similar effects have been reported by other authors [23]. Table 1 shows the results of the specific surface area, and the pore diameter of the catalyst support.

3.2. X-ray Diffraction (XRD)

An X-ray diffraction characterization test was performed to evaluate the content of the anatase and rutile phases in the supports for both methods. The crystalline phases of TiO2 and TiO2-CeO2 can be seen in the diffraction patterns found in Figure 1. For the Pt-TiO2 catalyst and the mixed oxides TiO2-CeO2 synthesized by sol-gel and impregnation, the presence of the peaks 2ϴ = 25.19, 37.60, 53.95, 54.36, 62.68, 75.04, and 82.7 are attributed to the anatase phase, corresponding to the plane (JCPDS no. 21-1272). Sol-gel presented the rutile phase whose peaks associated with the phase are 2ϴ = 70.16 and impregnation in 2ϴ = 35.81, 41.04, and 70, according to JCPDS with reference number 23-0278. The anatase phase is dominant in both methods due to the heat treatment to which the material was subjected. This is good since the anatase phase is a better photocatalyst than rutile because the exciton diffusion is twice as long [6]. In addition, in the sol-gel method it is possible to observe a decrease in the peaks of the anatase phase due to the increase of CeO2 in the network of the support. For CeO2 the peaks 2ϴ = 47.80, are associated with the cerenite phase that corresponds to a cubic packing of CeO2. With the impregnation method, CeO2 was observed at 2ϴ = 27.29, 47.80, and 56.54. In Pt/TiO2-CeO2 catalysts, the cerenite phase was observed very little, probably because it is very dispersed within the TiO2 structure, this can be observed in the shift to the right of the characteristic peaks of the TiO2 (JCPDS, no. 04-0802).

3.3. Raman Spectroscopy

On the other hand, as seen in Figure 2, in the Raman spectra of the TiO2 supports and the mixed oxides TiO2-CeO2, which were prepared by the sol-gel and impregnation method, peaks corresponding to the anatase phase 398–400, 518–520, and 640 cm−1 are observed. Observing a slight Raman shift, which means that CeO2 has been integrated into the structure of the TiO2 support. It is assumed that the ≡ Ti-O-Ti ≡ bonds of the corresponding TiO2 network of the anatase phase are disturbed by the presence of cerium oxide, which suggests some substitutions of the Ti4+ by Ce4+ that form ≡ Ti-O-Ce ≡ bonds in the structure of titanium oxide.

3.4. Transmission Electron Microscope (TEM)

In the transmission electron microscope (TEM) information obtained is the particle size by analyzing the images in software capable of measuring the diameter of the particle on a nanometric scale. Their respective alpha images of each of the elements and how it is dispersed inside and outside the support is shown in Figure 3. The particle size dispersion histograms were obtained by analyzing the series of data obtained in the TEM images, as shown in Figure 3.
Figure 4 shows how the particle size decreases, with the addition of CeO2 to the network of the support, that can suggest an increment of the specific surface area of the support. The micrographs reveal that the sizes of the metallic particles for the photocatalysts ranged from 2 to 6 nm. The smallest Pt particle size was observed in Pt/TiO2 CeO2 10 wt%.

3.5. X-ray Photoelectron Spectrometry (XPS)

The binding energy and the atomic ratios of Pt, Ti and Ce, for the Pt/TiO2-CeO2 catalysts (5 and 10 wt%) prepared by impregnation and sol-gel are reported in Table 2. The relative abundance of the Pt0− Pt2+ and Ce3+ Ce4+ species were calculated from the area under the curve of the respective peaks of the XPS spectra for the different catalysts (Figure 5). In Table 2, the corresponding binding energies for Pt 4f(7/2) are shown; the values of the binding energies for the Pt/TiO2 and Pt/TiO2-CeO2 catalysts are around 73.0 to 75.9 eV corresponding to Pt0 and Pt2+. A shift in the binding energy towards higher energies can be observed with an increase the amount of cerium oxide in the Pt/TiO2-CeO2 catalysts (5% by weight and 10%). This is due mainly to the fact that CeO2 is considered as an oxygen supplier which makes the platinum species in the reduced state Pt0 transform to the oxidized species of Pt2+ [24,25,26].
Table 2 reports the binding energies for TiO2, which can have values ranging between 457.7–458.1 eV [27], as shown in Figure 6, which indicates that there was no modification due to the doping effect with the CeO2 content, nor with the preparation method of the Pt supports. The binding energy was also determined for the Ce 3d5/2 level (Table 2 and Figure 6); it was found in the region of 870–920 eV [28]. Relative abundance calculated from these XPS spectra showed that Ce4+ (oxidized) species increased with increasing CeO2 content relative to Ce3+ (reduced). Coinciding with Rocha et al. (2015), it is possible to observe that at a lower concentration of CeO2, a greater number of atoms in the Ce3+ oxidation state will be obtained.

3.6. UV–Vis for Band Gap

The photophysical properties UV–Vis absorption spectra of the catalysts were evaluated to investigate the effect of CeO2 on the support network. Figure 6 shows the spectra of UV–Vis materials by diffuse reflectance for sol-gel and impregnation methods. All samples have a shift between these wavelengths, which can be attributed to the transitions of the Ti-O electrons of the TiO2 and TiO2-CeO2 nanocrystals.
Table 3 shows the results where a change in activation energy (3.02–2.8 eV) was observed for the TiO2-CeO2 samples from 1% to 5% by weight of CeO2, compared to the reference TiO2 in anatase phase (3.4 eV). The band gap energies were calculated by a linear fit of the slope to the abscissa and are reported in Table 3. It diminished from 3.45 eV, for the bare TiO2, to 2.82 eV, for the TiO2-CeO2 (at 5 wt%) sample. It is evident that cerium oxide modifies the bulk semiconductor properties of TiO2. The shift of the Eg band gap to a lower energy can be attributed to the incorporation of Ce4+ cations, which substitute some Ti4+ cations.

3.7. Photocatalysts Degradation of 2,4-Dichlorophenoxyacetic Acid

The photocatalytic degradation reactions of 2,4-D acid were carried out at room temperature at 298 K for 6 h, with a concentration of 200 ppm of the reagent, followed by the UV absorption band of 283 that corresponds mainly to the transition electron n → π*, which is mainly attributed to the C-Cl bond [29,30]. The percentage conversion as a function of time for both supports and catalysts impregnated and prepared by the sol-gel method at 360 min of reaction are shown in Figure 7 and Table 4.
The photocatalytic degradation of 2,4-D in the absence of support or catalyst had a conversion of 32% while the maximum conversion reached was 95% and 97% for the Pt/TiO2-CeO2 1% and Pt/TiO2-CeO2 catalysts 5% prepared by the sol-gel method (Figure 7B), and the catalysts prepared by the impregnation method reached a maximum of 62% conversion. On the other hand, the supports prepared by impregnation reached a maximum of 49% TiO2 while those prepared by the sol-gel method reached up to 61% (Figure 7A). The highest yield achieved in the catalysts prepared by sol-gel Pt/TiO2-CeO2 1% and Pt/TiO2-CeO2 5% could be attributed to an optimal concentration of CeO2, which allows the insertion within the CeO2 of the TiO2 and leads to the deformation of the lattice, modifying the mobility of the oxygen atoms and favoring the oxidation-reduction process [26]. In contrast, the results of XPS in the Pt/TiO2-CeO2 1% and Pt/TiO2-CeO2 5% catalysts showed that the proportion of oxidized species of Pt 2+ and Ce 4+ are essential to function as oxygen scavengers, which are important in oxidation-reduction processes. Additionally, the smallest particle size in the catalysts prepared with the supports by sol-gel was in a range of 2 to 6 nm. This is due to a greater specific area due to a good integration of CeO2, which favors a better dispersion of the metallic nanoparticles on the surface of the supports, favoring the catalytic activity in the degradation of 2,4-D.

4. Conclusions

In the present work, the TiO2 and TiO2-CeO2 supports, prepared by the sol-gel method and increasing the CeO2 concentration in a 1–10 ratio in the TiO2 support network, significantly increased the pore diameter, affecting the specific surface area for the catalyst. On the other hand, in the supports prepared by impregnation, no important modification was observed, either in the area or in the pore diameter due to the addition of CeO2, since these remained constant. However, when comparing the results of both materials we can conclude that sol-gel supports can obtain pore diameters four times smaller than those obtained with impregnation. By having less exposed area, the Pt catalyst particles will be larger because they tend to agglomerate, as they do not have enough space to disperse efficiently. Affecting the catalytic activity of the material, the Pt particles, being well dispersed, favored the catalytic activity of the material. Another important fact is that it was possible to obtain Pt nanoparticles on the sol-gel supports in the order of 2 and 6 nm, dependent of the CeO2 content in the support. A cerium oxide shift in the energy band gap was observed in the Pt/TiO2-CeO2 photocatalysts. It is proposed that the high activity showed by the Pt/TiO2-CeO2 photo-catalysts can be due to a synergetic effect between the cerium oxide and the platinum of oxidizing agent.

Author Contributions

Methodology, I.A.M.E. and C.A.G.G.; formal analysis, A.P.L., I.A.M.E. and C.A.G.G.; writing—original draft preparation, I.A.M.E. and B.S.-R.; writing—review and editing, B.S-R.; supervision, C.A.G.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A graduate scholarship for Ixchel Alejandra Mejía-Estrella was provided by the National Council for Science and Technology (CONACyT) of México.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gopinath, K.; Madhav, N.; Krishnan, A.; Malolan, R.; Rangarajan, G. Present applications of titanium dioxide for the photocatalytic removal of pollutants from water: A review. J. Environ. Manag. 2020, 270, 110906. [Google Scholar] [CrossRef] [PubMed]
  2. Borges, M.E.; Sierra, M.; Cuevas, E.; García, R.D.; Esparza, P. Photocatalysis with solar energy: Sunlight-responsive photocatalyst based on TiO2 loaded on a natural material for wastewater treatment. Sol. Energy 2016, 135, 527–535. [Google Scholar] [CrossRef]
  3. Van Deelen, T.W.; Hernández Mejía, C.; de Jong, K.P. Control of metal-support interactions in heterogeneous catalysts to enhance activity and selectivity. Nat. Catal. 2019, 2, 955–970. [Google Scholar] [CrossRef]
  4. Yunarti, R.T.; Isa, I.D.; Dimonti, L.C.C.; Dwiatmoko, A.A.; Ridwan, M.; Ha, J.-M. Study of Ag2O/TiO2 nanowires synthesis and characterization for heterogeneous reduction reaction catalysis of 4-nitrophenol. Nano-Struct. Nano-Objects 2021, 26, 100719. [Google Scholar] [CrossRef]
  5. Ayesha, B.; Jabeen, U.; Naeem, A.; Kasi, P.; Malghani, M.N.K.; Khan, S.U.; Akhtar, J.; Aamir, M. Synthesis of zinc stannate nanoparticles by sol-gel method for photocatalysis of commercial dyes. Results Chem. 2020, 2, 100023. [Google Scholar] [CrossRef]
  6. Fonseca-Cervantes, O.; Pérez-Larios, A.; Romero Arellano, V.; Sulbaran-Rangel, B.; Guzmán González, C. Effects in Band Gap for Photocatalysis in TiO2 Support by Adding Gold and Ruthenium. Processes 2020, 8, 1032. [Google Scholar] [CrossRef]
  7. Pragathiswaran, C.; Smitha, C.; Mahin Abbubakkar, B.; Govindhan, P.; Anantha Krishnan, N. Synthesis and characterization of TiO2/ZnO–Ag nanocomposite for photocatalytic degradation of dyes and anti-microbial activity. Mater. Today Proc. 2021, 45, 3357–3364. [Google Scholar] [CrossRef]
  8. Tian, J.; Sang, Y.; Zhao, Z.; Zhou, W.; Wang, D.; Kang, X.; Liu, H.; Wang, J.; Chen, S.; Cai, H.; et al. Enhanced Photocatalytic Performances of CeO2/TiO2 Nanobelt Heterostructures. Small 2013, 9, 3864–3872. [Google Scholar] [CrossRef]
  9. Chatterjee, D.; Dasgupta, S. Visible light indu ced photocatalytic degradation of organic pollutants. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 186–205. [Google Scholar] [CrossRef]
  10. Matte, L.P.; Kilian, A.S.; Luza, L.; Alves, M.C.M.; Morais, J.; Baptista, D.L.; Dupont, J.; Bernardi, F. Influence of the CeO2 Support on the Reduction Properties of Cu/CeO2 and Ni/CeO2 Nanoparticles. J. Phys. Chem. C 2015, 119, 26459–26470. [Google Scholar] [CrossRef]
  11. Bellardita, M.; Fiorenza, R.; Urso, L.; Spitaleri, L.; Gulino, A.; Compagnini, G.; Scirè, S.; Palmisano, L. Exploring the Photothermo-Catalytic Performance of Brookite TiO2-CeO2 Composites. Catalysts 2020, 10, 765. [Google Scholar] [CrossRef]
  12. Henych, J.; Šťastný, M.; Němečková, Z.; Mazanec, K.; Tolasz, J.; Kormunda, M.; Ederer, J.; Janoš, P. Bifunctional TiO2/CeO2 reactive adsorbent/photocatalyst for degradation of bis-p-nitrophenyl phosphate and CWAs. Chem. Eng. J. 2021, 414, 128822. [Google Scholar] [CrossRef]
  13. Liu, B.; Zhao, X.; Zhang, N.; Zhao, Q.; He, X.; Feng, J. Photocatalytic mechanism of TiO2-CeO2 films prepared by magnetron sputtering under UV and visible light. Surf. Sci. 2005, 595, 203–211. [Google Scholar] [CrossRef]
  14. Gnanasekaran, L.; Rajendran, S.; Priya, A.K.; Durgalakshmi, D.; Vo, D.-V.N.; Cornejo-Ponce, L.; Gracia, F.; Soto-Moscoso, M. Photocatalytic degradation of 2,4-dichlorophenol using bio-green assisted TiO2-CeO2 nanocomposite system. Environ. Res. 2021, 195, 110852. [Google Scholar] [CrossRef]
  15. García-Domínguez, Á.E.; Torres-Torres, G.; Arévalo-Pérez, J.C.; Silahua-Pavón, A.; Sánchez-Trinidad, C.; Godavarthi, S.; Ojeda-López, R.; Sierra-Gómez, U.A.; Cervantes-Uribe, A. Urea assisted synthesis of TiO2-CeO2 composites for photocatalytic acetaminophen degradation via simplex-centroid mixture design. Results Eng. 2022, 14, 100443. [Google Scholar] [CrossRef]
  16. Petrović, S.; Stanković, M.; Pavlović, S.; Mojović, Z.; Radić, N.; Mojović, M.; Rožić, L. Nickel oxide on mechanochemically synthesized TiO2-CeO2: Photocatalytic and electrochemical activity. React. Kinet. Mech. Catal. 2021, 133, 1097–1110. [Google Scholar] [CrossRef]
  17. Khan, M.A.R.; Mamun, M.S.A.; Ara, M.H. Review on platinum nanoparticles: Synthesis, characterization, and applications. Microchem. J. 2021, 171, 106840. [Google Scholar] [CrossRef]
  18. Jeyaraj, M.; Gurunathan, S.; Qasim, M.; Kang, M.-H.; Kim, J.-H. A Comprehensive Review on the Synthesis, Characterization, and Biomedical Application of Platinum Nanoparticles. Nanomaterials 2019, 9, 1719. [Google Scholar] [CrossRef]
  19. Guzmán, C.; del Ángel, G.; Gómez, R.; Galindo-Hernández, F.; Ángeles-Chavez, C. Degradation of the herbicide 2,4-dichlorophenoxyacetic acid over Au/TiO2-CeO2 photocatalysts: Effect of the CeO2 content on the photoactivity. Catal. Today 2011, 166, 146–151. [Google Scholar] [CrossRef]
  20. Guzmán, C.; Del Angel, G.; Fierro, J.L.G.; Bertin, V. Role of Pt Oxidation State on the Activity and Selectivity for Crotonaldehyde Hydrogenation Over Pt–Sn/Al2O3–La and Pt–Pb/Al2O3–La Catalysts. Top. Catal. 2010, 53, 1142–1144. [Google Scholar] [CrossRef]
  21. Zhang, H.; Banfield, J.F. Understanding Polymorphic Phase Transformation Behavior during Growth of Nanocrystalline Aggregates:  Insights from TiO2. J. Phys. Chem. B 2000, 104, 3481–3487. [Google Scholar] [CrossRef]
  22. He, Z.; Cai, Q.; Fang, H.; Situ, G.; Qiu, J.; Song, S.; Chen, J. Photocatalytic activity of TiO2 containing anatase nanoparticles and rutile nanoflower structure consisting of nanorods. J. Environ. Sci. 2013, 25, 2460–2468. [Google Scholar] [CrossRef]
  23. Rocha-Ortiz, G.; Tessensohn, M.E.; Salas-Reyes, M.; Flores-Moreno, R.; Webster, R.D.; Astudillo-Sánchez, P.D. Homogeneous electron-transfer reaction between anionic species of anthraquinone derivatives and molecular oxygen in acetonitrile solutions: Electrochemical properties of disperse red 60. Electrochim. Acta 2020, 354, 136601. [Google Scholar] [CrossRef]
  24. Yang, W.-D.; Hsu, Y.-C.; Lin, W.-C.; Huang, I.-L. Characterization and photocatalytic activity of N and Pt doped titania prepared by microemulsion technique. Adv. Mater. Sci. 2018, 3, 1–5. [Google Scholar] [CrossRef]
  25. Dauscher, A.; Hilaire, L.; Le Normand, F.; Müller, W.; Maire, G.; Vasquez, A. Characterization by XPS and XAS of supported Pt/TiO2—CeO2 catalysts. Surf. Interface Anal. 1990, 16, 341–346. [Google Scholar] [CrossRef]
  26. Rocha, M.A.L.; Del Ángel, G.; Torres-Torres, G.; Cervantes, A.; Vázquez, A.; Arrieta, A.; Beltramini, J.N. Effect of the Pt oxidation state and Ce3+/Ce4+ ratio on the Pt/TiO2-CeO2 catalysts in the phenol degradation by catalytic wet air oxidation (CWAO). Catal. Today 2015, 250, 145–154. [Google Scholar] [CrossRef]
  27. Thermo Fisher Scientific. Titanium Transition Metal Primary XPS Region: Ti2p. Available online: https://xpssimplified.com/elements/titanium.php (accessed on 24 July 2022).
  28. Thermo Fisher Scientific. Platinum Transition Metal. Primary XPS Region: Pt4f. Available online: https://xpssimplified.com/elements/platinum.php#appnotes (accessed on 24 July 2022).
  29. Ramos-Ramírez, E.; Gutiérrez-Ortega, N.L.; Tzompantzi-Morales, F.; Barrera-Rodríguez, A.; Castillo-Rodríguez, J.C.; Tzompantzi-Flores, C.; Santolalla-Vargas, C.E.; Guevara-Hornedo, M.d.P. Photocatalytic Degradation of 2,4-Dichlorophenol on NiAl-Mixed Oxides Derivatives of Activated Layered Double Hydroxides. Top. Catal. 2020, 63, 546–563. [Google Scholar] [CrossRef]
  30. Ba-Abbad, M.M.; Kadhum, A.A.H.; Mohamad, A.B.; Takriff, M.S.; Sopian, K. Photocatalytic degradation of chlorophenols under direct solar radiation in the presence of ZnO catalyst. Res. Chem. Intermed. 2013, 39, 1981–1996. [Google Scholar] [CrossRef]
Figure 1. XRD of the catalysts Pt/TiO2 and Pt/TiO2-CeO2: (a) Sol-gel and (b) Impregnation.
Figure 1. XRD of the catalysts Pt/TiO2 and Pt/TiO2-CeO2: (a) Sol-gel and (b) Impregnation.
Materials 15 06784 g001
Figure 2. Raman spectrum for the support system TiO2, TiO2-CeO2 (1, 5, 10 wt%): (a) Sol-gel and (b) Impregnation.
Figure 2. Raman spectrum for the support system TiO2, TiO2-CeO2 (1, 5, 10 wt%): (a) Sol-gel and (b) Impregnation.
Materials 15 06784 g002
Figure 3. Transmission electron microscopy images of the Pt/TiO2-CeO2 10%: (a) Sol-gel to 100 nm scale, (b) Sol-gel take α of the Ti, (c) Sol-Gel take α of the Ce, (d) Sol-Gel take α of the Pt, (e) Impregnation to 100 nm scale, (f) Impregnation take α of the Ti, (g) Impregnation take α of the Ce, and (h) Impregnation take α of the Pt.
Figure 3. Transmission electron microscopy images of the Pt/TiO2-CeO2 10%: (a) Sol-gel to 100 nm scale, (b) Sol-gel take α of the Ti, (c) Sol-Gel take α of the Ce, (d) Sol-Gel take α of the Pt, (e) Impregnation to 100 nm scale, (f) Impregnation take α of the Ti, (g) Impregnation take α of the Ce, and (h) Impregnation take α of the Pt.
Materials 15 06784 g003
Figure 4. Size particle distribution of the Pt determinates by TEM for the photocatalyst Pt/TiO2-CeO2 10 wt% (a) Sol-gel and (b) Impregnation.
Figure 4. Size particle distribution of the Pt determinates by TEM for the photocatalyst Pt/TiO2-CeO2 10 wt% (a) Sol-gel and (b) Impregnation.
Materials 15 06784 g004
Figure 5. XPS spectra (Ce 3d region) for photocatalyst (a) Sol-gel Pt/TiO2-CeO2 5%, (b) Sol-gel Pt/TiO2-CeO2 10%, (c) Impregnation Pt/TiO2-CeO2 5%, and (d) Impregnation Pt/TiO2-CeO2 10%.
Figure 5. XPS spectra (Ce 3d region) for photocatalyst (a) Sol-gel Pt/TiO2-CeO2 5%, (b) Sol-gel Pt/TiO2-CeO2 10%, (c) Impregnation Pt/TiO2-CeO2 5%, and (d) Impregnation Pt/TiO2-CeO2 10%.
Materials 15 06784 g005
Figure 6. UV–vis spectra for the TiO2-CeO2 supports Sol-Gel and Impregnation.
Figure 6. UV–vis spectra for the TiO2-CeO2 supports Sol-Gel and Impregnation.
Materials 15 06784 g006
Figure 7. Photocatalysts degradation of 2,4-Dichlorophenoxyacetic acid. (a) Impregnation and (b) Sol-Gel materials.
Figure 7. Photocatalysts degradation of 2,4-Dichlorophenoxyacetic acid. (a) Impregnation and (b) Sol-Gel materials.
Materials 15 06784 g007
Table 1. Specific surface area, and pore diameter for TiO2-CeO2 catalysts support.
Table 1. Specific surface area, and pore diameter for TiO2-CeO2 catalysts support.
SupportMethodDiameter Pore (Å)Specific Surface Area (m2/g)
TiO2Sol-gel52.54185.59
TiO2-CeO2 1%Sol-gel53.08181.46
TiO2-CeO2 10%Sol-gel77.51104.36
TiO2-CeO2 1%Impregnation295.0643.56
TiO2-CeO2 10%Impregnation304.0543.61
Table 2. Binding energy and relative abundance of the different species obtained by XPS of the catalysts.
Table 2. Binding energy and relative abundance of the different species obtained by XPS of the catalysts.
Support MethodBinding Energy (eV)Relative Abundance (%)
Pt (4f7/2)Ti (2p3/2)Ce (3d5/2)Pt0− Pt2+Ti4+Ce3+ Ce4+
Pt/TiO2Sol-gel73.0
75.9
458.1-80–20100-
Pt/TiO2-CeO2 5%Sol-gel75.4
77.09–78.09
458880–90053–4710056–44
Pt/TiO2-CeO2 10%Sol-gel75.9
77.09–78.09
458.1881–900.147–5310052–47
Pt/TiO2-CeO2 5%Impregnation-457.7880–900-10048.8–51.19
Pt/TiO2-CeO2 10%Impregnation-465880–900-10038–62
Table 3. Band Gap Energy and Wavelengths.
Table 3. Band Gap Energy and Wavelengths.
Catalyst ID NameBand Gap (eV)Wavelengths (nm)
TiO23.45359
Pt/TiO23.39365
Pt/TiO2-CeO2- 1%3.05406
Pt/TiO2-CeO2- 5%2.82439
Table 4. Photocatalysts degradation of 2,4 Dichlorophenoxiacetyc acid.
Table 4. Photocatalysts degradation of 2,4 Dichlorophenoxiacetyc acid.
CatalystsMethodPt (wt%)X%Cf (ppm)
TiO2Commercial-38160
TiO2-CeO2 5%Impregnation-49128
TiO2-CeO2 10%Impregnation-45138
Pt-TiO2-CeO2 5%Impregnation16295
Pt-TiO2-CeO2 10%Impregnation156110
TiO2 Sol-Gel-38155
TiO2-CeO2 1%Sol-Gel-6198
TiO2-CeO2 5%Sol-Gel-6166
TiO2-CeO2 10%Sol-Gel-45138
Pt-TiO2-CeO2 1%Sol-Gel19511
Pt-TiO2-CeO2 5%Sol-Gel1977
Pt-TiO2-CeO2 10%Sol-Gel18927
Photolysis--32169
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mejia-Estrella, I.A.; Pérez Larios, A.; Sulbarán-Rangel, B.; Guzmán González, C.A. Reaching Visible Light Photocatalysts with Pt Nanoparticles Supported in TiO2-CeO2. Materials 2022, 15, 6784. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15196784

AMA Style

Mejia-Estrella IA, Pérez Larios A, Sulbarán-Rangel B, Guzmán González CA. Reaching Visible Light Photocatalysts with Pt Nanoparticles Supported in TiO2-CeO2. Materials. 2022; 15(19):6784. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15196784

Chicago/Turabian Style

Mejia-Estrella, Ixchel Alejandra, Alejandro Pérez Larios, Belkis Sulbarán-Rangel, and Carlos Alberto Guzmán González. 2022. "Reaching Visible Light Photocatalysts with Pt Nanoparticles Supported in TiO2-CeO2" Materials 15, no. 19: 6784. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15196784

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