Solution combustion synthesis of I-type heterojunction photocatalyst based on o-YbFeO3/h-YbFeO3/CeO2 toward efficient photo-Fenton degradation of methyl violet Chimica Techno Acta ARTICLE published by Ural Federal University 2021, vol. 8(4), № 20218407 eISSN 2411-1414; chimicatechnoacta.ru DOI: 10.15826/chimtech.2021.8.4.07 1 of 11 Synthesis, structure, and visible-light-driven activity of o-YbFeO3/h-YbFeO3/CeO2 photocatalysts Sofia M. Tikhanova ab* , Lev A. Lebedev a , Svetlana A. Kirillova c , Maria V. Tomkovich a , Vadim I. Popkov a a: Ioffe Institute, 194021, Saint Petersburg, Russia b: Saint-Petersburg State Institute of Technology, 190013, Saint Petersburg, Russia c: Saint Petersburg Electrotechnical University "LETI", 197376, Saint Petersburg, Russia * Corresponding author: tihanova.sof@gmail.com This article belongs to the regular issue. © 2021, The Authors. This article is published in open access form under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/). Abstract Photo-Fenton-like oxidation of organic substances is one of the key ad- vanced oxidation processes based on the reversible Fe2+↔Fe3+ transition and the generation of a strong oxidant ·OH in the presence of H2O2 and is currently considered as a promising method for the purification of polluted aqueous media. However, the absence of effective and stable photocatalysts of this process, operating under the action of visible light, necessitates the exploratory studies, mainly among iron oxides and ferrites of various compositions and structures. In this work, using the method of solution combustion followed by heat treatment in air the heterojunction nanocomposites based on ytterbium orthoferrite and ce- rium dioxide of the composition o-YbFeO3/h-YbFeO3/CeO2 (0–20 mol.%) with high absorption in the visible region and advanced photo-Fenton- like activity were obtained. The nanocomposites were studied by EDS, SEM, XRD, BET, and DRS methods. The photo-Fenton-like activity of the nanocomposites was investigated during the degradation of methyl vio- let under the action of visible (λmax = 410 nm) radiation. As a result, the formation of I-type heterojunction based on stable rhombic (55.4–79.0 nm) and metastable hexagonal (19.5–24.0 nm) modifications of ytterbium orthoferrite (o-YbFeO3 and h-YbFeO3, respectively) and cu- bic cerium dioxide CeO2 (13.2–19.2 nm) nanocrystals was established. It was shown that the obtained nanocomposites had foamy morphology and were characterized by a specific surface in the range of 9.1–25.0 m2/g, depending on the CeO2 content. It was found that nano- crystalline components were chemically and phase-pure, uniformly spa- tially distributed over the nanocomposite, and had multiple contacts with each other. Based on this fact and the established electronic struc- ture of the nanocomposite components, the formation of I-type hetero- junction with the participation of o-YbFeO3 (Eg = 2.15 eV), h-YbFeO3 (Eg = 2.08 eV), and CeO2 (Eg = 2.38 eV) was shown, the presence of which increased photocatalytic activity of the resulting nanocomposite. The optimal content of CeO2 in the nanocomposite was 5%, and the o-YbFeO3/h-YbFeO3/CeO2–5% sample was characterized by the highest rate constant of photo-Fenton-like degradation of methyl violet under the action of visible light equal to k = 0.138 min–1, which was 2.5 to 5 times higher than for nanocomposites based on ytterbium orthoferrite. The obtained results obtained indicate that the developed nanocompo- sites can be considered as a promising basis for the advanced oxidation processes for the purification of aqueous media from organic pollutants. Keywords ytterbium orthoferrite cerium dioxide solution combustion synthesis heterojunction photocatalysts Fenton-like process Received: 29.09.2021 Revised: 27.11.2021 Accepted: 30.11.2021 Available online: 02.12.2021 http://chimicatechnoacta.ru/ https://doi.org/10.15826/chimtech.2021.8.4.07 http://creativecommons.org/licenses/by/4.0/ https://orcid.org/0000-0002-9239-5470 https://orcid.org/0000-0001-9449-9487 https://orcid.org/0000-0001-7912-6033 https://orcid.org/0000-0002-1537-4107 https://orcid.org/0000-0002-8450-4278 Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 2 of 11 1. Introduction Currently, the problem of water purification from organic pollutants is attracting more and more attention. The ad- vanced oxidation processes or AOPs [1] have been actively studied as the basis for promising and effective methods for eliminating pollution from wastewater. These methods are based on the catalytic production and action of hy- droxyl radicals (OH·) having the strong oxidizing ability, due to which complex organic molecules can be oxidized partially to simpler molecules or completely to CO2 and H2O. One of the most promising AOPs is the Fenton process, in which Fe2+ is used as a catalyst for the formation of hydroxyl radicals, and hydrogen peroxide H2O2 is used as their main source [2]. The mechanism of this process can be represented by two main stages following the reaction equations: Fe2+ + H2O2 → Fe3+ + OH + OH− (1) Fe3+ + H2O2 → Fe2+ +·O2H + H+ (2) The main advantages of this process are its relatively high efficiency and ease of implementation. Since the final products of the decomposition of hydrogen peroxide are oxygen (O2) and water (H2O), this process is safe and en- vironmentally friendly, which is important for the purifi- cation of aqueous media from organic pollutants [3]. However, the classical homogeneous Fenton process has significant disadvantages, such as high specific cost, limited operating pH range, the formation of a large vol- ume of iron sludge (which has a detrimental effect on the environment and leads to the loss of a large amount of catalytic metals), as well as difficulties in catalyst regen- eration (Fe2+). To overcome these disadvantages the im- proved approaches based on the Fenton process are cur- rently being developed and, in particular, a photoinduced Fenton-like heterogeneous process, which eliminates the known limitations. Thus, the development and production of new highly efficient heterogeneous photocatalysts for photoinduced Fenton-like oxidative processes seem to be an urgent task. Nanostructured ferrites, in particular REE orthofer- rites (RFeO3, where R = Sc, Y, Ln), can be used as the basis for such materials. These compounds have been actively studied due to their practically significant structural [4], magnetic [5–8], electrical [8–10], conducting [11] and pho- tocatalytic properties [12–14]. In particular, ytterbium orthoferrite YbFeO3 has important for practical applica- tions electrical and magnetic properties that change de- pending on its crystal structure [15]. Although YbFeO3 has a thermodynamically stable orthorhombic perovskite structure (Pbnm/Pnma), some studies have shown that nanostructured YbFeO3 (mainly in the form of thin films) can have more two more metastable hexagonal phases: nonpolar P63/mmc and polar P63cm, depending on the partial pressure of oxygen [4, 16]. In addition, YbFeO3 na- noparticles, being in a magnetically ordered state, can be used in catalytic processes [13] and then removed from the reaction solution by applying of an external field, which makes it promising to use them as the basis for ef- fective magnetically controlled photocatalysts for the puri- fication of polluted waters, including using photo-Fenton- like processes. However, the use of REE orthoferrites in Fenton-like processes is limited by a high tendency to reverse recom- bination of electron-hole pairs formed during light absorp- tion, as well as by a limited region of radiation absorption (visible region from 500 nm and above). In the previous work, it was shown that the nanocomposite based on h-YbFeO3/o-YbFeO3 exhibits greater photocatalytic activity compared to single-phase o-YbFeO3 at similar specific sur- face areas. The formation of a heterojunction prevents the recombination of electron-hole pairs and thereby increas- es the overall photocatalytic efficiency of the nanocompo- site material [15, 17], but does not significantly affect the characteristic region of radiation absorption. To solve this problem, this work proposes the devel- opment of nanocomposite photocatalysts based on ytterbi- um orthoferrite and a co-catalyst, cubic cerium oxide (CeO2). The use of CeO2 as a co-catalyst shifts the absorp- tion edge to the ultraviolet region of the spectrum [16], which should lead to an increase in the efficiency of the photocatalytic system in the photo-Fenton-like process due to the more efficient absorption of photons with dif- ferent energies. The main methods for obtaining single-phase REE or- thoferrites are hydrothermal [4, 18], sol-gel [19] and mi- crowave [20]. However, the preparation of nanocompo- sites of the above composition presents a certain difficul- ty, since the chemical properties of the corresponding components differ significantly and such systems cannot always be obtained by known approaches and synthetic methods. In this work, we propose the preparation of the o-YbFeO3/h-YbFeO3/CeO2 nanocomposite by the method of solution combustion, carried out in a soft glow mode, fol- lowed by a heat treatment of amorphous products at mod- erate temperatures in air. 2. Experimental Nanocomposites based on ytterbium orthoferrite and cerium dioxide were obtained by solution combustion followed by heat treatment. A schematic representation of the synthesis procedure is shown in Fig. 1. Glycine (C2H5NO2, 98.5%) was added to a solution of oxidizing salts Yb(NO3)3·6H2O (99.5%), Fe(NO3)3·9H2O (98.0%) and Ce(NO3)3·6H2O (99.5%), and ratio G/N = 0.2, pro- vided a glowing combustion mode and the formation of amorphous products [21]. Ytterbium, iron, and cerium nitrates were taken in the ratios required to obtain o-YbFeO3/h-YbFeO3/CeO2 samples with a molar fraction Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 3 of 11 of cerium dioxide of 0, 2.5, 5, 7.5, 10, and 20%. The rea- gents were dissolved in 10 ml of distilled water with con- stant stirring for 15–20 minutes, after which the reaction mixture was heated until almost complete removal of water, transformation into a gel-like mass, and autoigni- tion. A dark brown foamy solid was formed, which was then heat-treated in the air for 24 hours at 800 °C to completely remove organic residues and form crystalline products. The elemental composition of the synthesized samples was studied using energy-dispersive X-ray spectroscopy (EDS) on a Tescan Vega 3 SBH scanning electron micro- scope equipped with Oxford INCA X-act X-ray microanal- ysis. X-ray phase analysis was performed on Rigaku SmartLab 3 powder X-ray diffractometer using Cu Kα (λ = 1.540598 Å). Powder X-ray diffraction data were obtained with a step of 0.01° and a rate of 4.5°/s in the 2θ range from 20° to 60°. Qualitative X-ray phase analy- sis was performed using the ICSD powder diffraction database. The average crystallite size was calculated from the broadening of X-ray diffraction lines using the method of fundamental parameters, implemented in the SmartLab Studio II software package. Quantitative X-ray phase analysis was carried out using the Rietveld method and structural data of crystalline phases. Diffuse reflec- tance spectra in the UV-visible region (DRS) were meas- ured using an Avaspec-ULS2048 spectrometer equipped with an AvaSphere-30-Refl integrating sphere. The photocatalytic activity of the obtained catalysts was studied using the example of the photodegradation of methyl violet (MV) in the presence of hydrogen peroxide H2O2. The process was carried out in a transparent glass beaker, under a light source λmax = 410 nm with constant stirring. In the course of the experiment, 1.5 mg of the catalyst were suspended in 1 ml of distilled water and added to 1 ml of MV solution (3 mmol/L), after which 10 ml of an H2O2 solution with a concentration of 20 mmol/L were added to start the reaction. In photocata- lytic experiments, the concentration of the dye was deter- mined spectrophotometrically. AvaLight-XE light source and Avaspec-ULS2048 spectrometer were used to measure the absorption spectra. Decolorization MV measurements were taken every 10 minutes. Fig. 1 Procedure for the preparation of o-YbFeO3/h-YbFeO3/CeO2 nanocomposite Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 4 of 11 3. Results and discussion According to the results of the EDS analysis, in the sam- ples obtained by the method described above, in the gen- eral case, the presence of four main elements is observed – ytterbium (Yb), cerium (Ce), iron (Fe), and oxygen (O) (Fig. 2a). No impurity elements were found, which indi- cates a high chemical purity of the obtained compositions. The result of the quantitative analysis of a series of sam- ples is presented in Fig. 2b. The error of the determination method for heavy elements is about 0.5 wt.%, it was stat- ed that the compositions of the composites for the main elements (Yb, Ce, Fe) were in good agreement with the composition specified during their synthesis. The stoichi- ometric relationship is also confirmed by X-ray diffraction data since no reflections of ytterbium oxide or any of the iron oxides are observed. These facts confirm the success- ful synthesis of samples with a molar content of cerium of 0, 2.5, 5, 7.5, 10, and 20%. To determine the uniformity of distribution of key chemical elements over the volume of o-YbFeO3/h-YbFeO3/CeO2–5% composition, SEM study, combined with EDS mapping, was carried out; its results are presented in Fig. 3. Fig. 2 Typical EDS spectrum of the o-YbFeO3/h-YbFeO3/CeO2–5% nanocomposite (a) and the composition of the obtained samples con- cerning the main elements (in wt.%) (b) Fig. 3 SEM images (a, b), multi-element (c) and single-element (d–f) EDS mappings of the o-YbFeO3/h-YbFeO3/CeO2–5% sample Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 5 of 11 The area selected based on the SEM results for map- ping (Fig. 3a, b) reflects the typical foam-like morphology of powders – products of solution combustion. At the same time, a comparison of the results of multi-element and single-element mapping indicates a high uniformity of the distribution of all the main elements – Yb, Ce, and Fe. No areas with enrichment in one of the elements were found in the results of the EDS mapping. This also indirectly con- firms the presence of a large number of contacts between Yb, Ce, and Fe-containing phases in the composition of the obtained composites. Fig. 4a shows the diffraction patterns of the samples with different CeO2 contents subjected to heat treatment at 800 °С for 24 hours. According to the qualitative X-ray phase analysis, the YbFeO3 sample, which does not contain cerium oxide, is a single-phase orthorhombic o-YbFeO3. The samples composed of CeO2 contain two phases of yt- terbium orthoferrite – metastable hexagonal h-YbFeO3 and stable orthorhombic o-YbFeO3. As the mole fraction of CeO2 increases, a broadening and an increase in the inten- sity of the peaks corresponding to cerium oxide are ob- served, while the relative intensity of the peaks corre- sponding to o-YbFeO3 does not change. It is known that cerium is capable of forming perovskite-like compounds of the o-CeFeO3 type [22]; however, due to the relatively large ionic radius of Ce3+ and its low stability in the air, such compounds are oxidized at elevated temperatures with decomposition to more stable CeO2 and iron oxides. For this reason, the cerium orthoferrite phase was not detected in the composition of the samples, and the pres- ence of cerium was observed exclusively in the form of cerium (IV) dioxide. In addition, for the samples contain- ing CeO2, the formation of the h-YbFeO3 phase is observed, which was not previously seen in this system under simi- lar conditions without the addition of cerium [23]. The formation of the hexagonal phase of ytterbium orthofer- rite in this case can be caused by the influence of the formed cerium (IV) dioxide, which creates spatial re- strictions during crystallization and restricts mass trans- fer. The combination of these factors creates the condi- tions for the formation of metastable h-YbFeO3, which, like the hexagonal orthoferrites of other REEs, are stable only for small nanocrystals (usually less than 15 nm) [24]. The concentration dependencies in Fig. 4b were ob- tained with the Rietveld method by using Rigaku SmartLab Studio II software. The parameters of refinement and cri- teria of obtained refinement are presented in Fig. 5 and Table 1. The Rwp, Rp, and Re parameters show that the calculated XRD pattern is in good correspondence with measured data. Fig. 4 X-ray diffraction results of o-YbFeO3/h-YbFeO3/CeO2 (a), molar fraction (mol.%) (b) and average crystallite sizes (c) versus CeO2 content (mol.%) Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 6 of 11 Fig. 5 Rietveld refinement result for the X-ray diffraction of o-YbFeO3/h-YbFeO3/CeO2 Table 1 The parameters of refinement by the Rietveld method and criteria of obtained refinement The change in the degree of conversion for crystalline phases, calculated from the data of X-ray diffraction, is presented in Fig. 4b. With an increase in the cerium con- tent in the composite, the molar fraction of o-YbFeO3 steadily decreases, while the molar fraction of CeO2 in- creases. In addition, in the presence of cerium in the com- position, the appearance of the h-YbFeO3 phase is ob- served, the fraction of which remains practically the same with a change in the cerium content. In addition to the change in the molar fraction of crystalline phases the gradual broadening of the diffraction peaks also occurs, based on which it is possible to estimate the change in the average size of the corresponding crystallites depending on the cerium content (Fig. 4c). It was found that the crys- tallite sizes varied in the range 55.4–79.0 nm for o- YbFeO3, 19.5–24.0 nm, and 13.2–19.2 nm for h-YbFeO3 and CeO2, respectively. The absence of a noticeable change in the average crystallite size of the h-YbFeO3 and CeO2 phases is explained by their relatively small fractions in the composition of the composites, which does not allow them to increase their size due to recrystallization pro- cesses – the dominant o-YbFeO3 phase prevents such mass transfer. Earlier, using a similar system as an example, it has been demonstrated that for the structural transition hexagonal → orthorhombic REE orthoferrite nanocrystals of the hexagonal phase should reach the critical size [25]. Therefore, due to the limited mass transfer, h-YbFeO3 nanocrystals do not acquire the critical size and remain in a metastable state. Such mutual influence of the compo- nents in the o-YbFeO3/h-YbFeO3/CeO2 system suggests the presence of multiple contacts of individual phases and confirms the heterojunction structure of the resulting composite. The results of scanning electron microscopy (Fig. 6) in- dicate that the obtained samples have a highly porous, foamy structure, which is a consequence of the used syn- thetic method used, characterized by the violent release of gaseous reaction products and foaming of solid-phase prod- ucts [23]. In this case, even after heat treatment, the foamy morphology is retained at the microlevel, while at the level of individual nanoparticles there are visible morphological changes. Thus, in the case of pure YbFeO3 without added cerium (Fig. 6a), the distinguishable nanoparticles are weakly agglomerated, have a spherical morphology, and are aggregated into foam-like structures of micron sizes. As shown earlier, the products of solution combustion have foam-like morphology and orthoferrite nanocrystals are formed upon heat treatment of the initial amorphous prod- uct from the pore walls while maintaining the foam micro- structure [24]. Alternatively, in the case of samples ob- tained with the addition of cerium (Fig. 6b–f), there is a more monolithic aggregation of individual particles, which are less distinguishable. This may point to the presence of more finely dispersed h-YbFeO3 and CeO2 phases in the space between the main coarse-crystalline o-YbFeO3 phase. It is worth noting that the size of o-YbFeO3 nanoparticles regularly decreases with the addition of cerium, which is also confirmed by the XRD results (Fig. 4c). The results of determining the specific surface area of the samples by the BET method are presented in Fig. 7. According to these data, when cerium is added to the sys- tem based on YbFeO3 up to a content of 5.0 mol.%, a 2.5-fold increase in the specific surface area is observed. However, with a further increase in the cerium content, this escalation is replaced by a reduction, and at 20 mol.% cerium, the specific surface area becomes almost the same as that of the pure YbFeO3 sample. The observed pattern is in good agreement with the results of scanning electron microscopy (Fig. 6) and the increase in the specific surface area in the first range is explained by the appearance of h- YbFeO3 and CeO2 nanocrystals with a larger specific sur- face area than o-YbFeO3 due to smaller crystallite sizes. Parame- ters o-YbFeO3 h-YbFeO3 CeO2 Initial CIF ICSD#189735 isostructural to ICSD#73361 ICSD#20194 Crystal system Orthorhombic Hexagonal Cubic Space group 62: Pbnm 194: P63/mmc 225: 𝐹𝑚3𝑚 Fraction, mol.% 43.58 13.67 42.75 Average crystal size, nm 47.2 27.8 25.0 a, Å 5.2395 3.50703 5.39149 b, Å 5.5628 3.50703 5.39149 c, Å 7.5788 11.65325 5.39149 α, ° 90 90 90 β, ° 90 90 90 γ, ° 90 120 90 Rwp, % 4.26 4.25 4.25 Rp, % 3.35 3.35 3.35 Re, % 4.31 4.31 4.31 S 0.9867 0.9853 0.9854 Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 7 of 11 Fig. 6 SEM images of o-YbFeO3/h-YbFeO3/CeO2 samples with different cerium content: 0% (a), 2.5% (b), 5% (c), 7.5% (d), 10% (e) and 20% (f) The decrease in the specific surface area in the second composition range is associated with the intensification of aggregation processes due to the increased cerium con- tent, as a result of which the part of the nanocrystal sur- face is inaccessible for contact. Thus, the o-YbFeO3/h-YbFeO3/CeO2 sample has the largest specific surface area and it is 25.0 m2/g, which is noticeably higher than for samples based on pure o-YbFeO3 and o-YbFeO3/h-YbFeO3 composite [23]. The diffuse reflectance spectra of the obtained samples (Fig. 8a) reveal that they intensively absorb the radiation of the visible spectrum. Several inflections in these de- pendences confirm the presence of several phases with different optical absorption edges in the samples. To de- termine the band gap of these phases, the experimental data were recalculated into the Tauc coordinates (Fig. 8b), and the calculation results for o-YbFeO3, h-YbFeO3, and CeO2 are shown in Fig. 8c. According to the calculations, the band gaps for all phases do not depend on the cerium content in the samples, as evidenced by their non- systematic variation within the error of the determination method. The average values of the band gaps were 2.15 eV for o-YbFeO3, 2.08 eV for h-YbFeO3, and 2.38 eV for CeO2, which is in good agreement with the literature data [4, 16, 18–20]. Thus, the synthesis of a nanocomposite based on the separate o-YbFeO3, h-YbFeO3, and CeO2 phas- es with the absence of incorporation of cerium into the structure of ytterbium orthoferrite and ytterbium into the structure of cerium (IV) dioxide is confirmed. Based on the data on the average values of the band gap, the boundaries of the valence and conduction bands for all phases were determined using the empirical formu- las [26]. The transition energies of orthorhombic and hex- agonal ytterbium ferrites are close to each other, but their difference affects the electronic structure. Fig. 7 BET surface area of the o-YbFeO3/h-YbFeO3/CeO2 samples versus CeO2 content Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 8 of 11 Fig. 8 DRS spectra (a), Tauc plots (b) and band gap energies (c) of the o-YbFeO3/h-YbFeO3/CeO2 samples The calculation results are presented in Table 2 and their comparison allows us to conclude that a type I heterojunction structure is formed within the obtained composites, which is suitable for suppressing the processes of reverse recombina- tion of electron-hole pairs in photocatalytic processes. Table 2 Parameters of the electronic structure of o-YbFeO3/h-YbFeO3/CeO2 composite Eg, eV χ, eV ECB, eV EVB, eV o-YbFeO3 2.15 5.574 0.00 2.15 h-YbFeO3 2.08 5.574 0.03 2.11 CeO2 2.38 5.56 –0.10 2.28 The photocatalytic activity of the obtained samples was studied in the process of photo-Fenton-like decomposition of methyl violet in the presence of hydrogen peroxide un- der the action of visible light. Fig. 9a shows the MV spec- tra of the reaction solution obtained at different times after the start of the photocatalytic experiment. From these data, it follows that in the presence of o-YbFeO3/h-YbFeO3/CeO2–5% composite, almost complete decolorization of the solution occurs within 90 minutes, which characterizes the composite as an effective catalyst for Fenton-like processes [27]. To assess the effectiveness of the nanocomposites ob- tained in this work thekinetic studies were carried out, the results of which are presented in Fig. 9b. Before the start of photocatalytic experiments, the reaction solution with the particles of the composite distributed in it was continuously mixed and kept in the dark until the adsorption-desorption equilibrium was established. At the same time, the drop in the concentration of the dye in the solution for the samples differed and correlated with the values of their specific sur- face area – the higher the specific surface area of the sam- ple, the more MV was adsorbed on its surface (see the ini- tial interval in Fig. 9b). After the equilibrium was estab- lished the series of photocatalytic experiments were carried out and the obtained kinetic dependences indicated a differ- ent methyl violet photodegradation rate in the presence of the obtained nanocomposites and pure ytterbium orthofer- rite. By the end of the experiment, the highest photodegra- dation efficiency was observed for samples with a cerium dioxide molar concentration of 2.5 and 5.0% and amounted to 94.0% and 96.4%, respectively. Thus, the nanocompo- sites exhibit almost 1.5 times higher efficiency of photodeg- radation than a sample of pure ytterbium orthoferrite, which is explained by the heterojunction structure of nano- composites increasing the generation of electron-hole pairs and reducing their recombination. The activity of pure ceri- um dioxide in this process is very low, which is explained by the absence of photons in the visible radiation spectrum with energy, sufficient for the transition of an electron from the valence band to the conduction band. For a quantitative comparison of the photocatalytic performance of nanocomposites and pure ytterbium ortho- ferrite, the obtained kinetic dependences were rearranged into logarithmic coordinates, –ln(C/C0) = f(t). The lineari- zation of all kinetic dependences in these coordinates con- firms the occurrence of photodegradation of methyl violet accompanied by nanocomposites in the pseudo-first-order characteristic of Fenton-like processes [3]. From the slope of the obtained dependences, the rate constants of the photodegradation reaction were determined. These and the other characteristics of the photocatalysts are summa- rized in Table 3 and a visual comparison of the obtained reaction rate values is shown in Fig 9d. According to the results obtained, with an increase in the CeO2 content, the rate constant values first increase up to 5% CeO2 and then decrease. The initial trend is associ- ated with promoting charge separation/transfer and re- ducing overpotential for oxidation, and its change into a downward trend is explained by covering active sites on photocatalyst, shielding the light absorption, decreasing the surface area and activity, increasing the charge re- combination. Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 9 of 11 Fig. 9 Photo-Fenton-like degradation of MV over o-YbFeO3/h-YbFeO3/CeO2–5% catalyst under visible light (a) kinetic curves (b) and logarithmic kinetic curves of the pseudo-first-order process (c) and photodecomposition rate constant (d) versus cerium dioxide content Table 3 Kinetic parameters of Fenton-like degradation of methyl violet in the presence of photocatalysts based on YbFeO3 under visible light irradiation Thus, the content of cerium dioxide in the nanocomposite equal to 5% is optimal for the given system and the method of its preparation. The corresponding value of the constant k = 0.138 min–1 is 2.5–5 times higher than that for the similar systems based on ytterbium orthoferrite (Table 3). Based on the results of this work, the information ob- tained on the electronic structure of nanocomposites and their photocatalytic properties, as well as the literature data, a schematic representation of the mechanism of pho- to-Fenton-like degradation of methyl violet in the pres- X(CeO2), mol.% Organic compound Half-life time, min Degradation efficiency, % The reaction rate constant, min–1 R Ref. 0 MV 35 68.90 0.0067 0.929 This work 2.50 MV 20 96.40 0.0127 0.957 5 MV 30 94.00 0.0138 0.990 7.5 MV 60 69.43 0.0054 0.985 10 MV 45 73.80 0.0064 0.986 20 MV 40 74.70 0.0069 0.986 100 MV 65 57.60 0.0045 0.985 YbFO–700 MV – 66.20 0.0048 0.991 [23] YbFO–800 MV – 47.40 0.0031 0.988 [23] LaFeO3 Acid Orange 7 – 40.70 0.0097 0.952 [28] LaFeO3/CeO2 Acid Orange 7 – 28.40 0.0070 0.971 [28] BiFeO3 BR46 – 95.00 0.0292 0.970 [29] CoFe2O4 BR46 – 44.00 0.0054 0.9663 [30] Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 10 of 11 ence of YbFeO3/CeO2 nanocomposite under the action of visible light was proposed (Fig. 10). Under the action of visible radiation the electron-hole pairs are formed in the components of the composite based on ytterbium orthoferrite, which absorbs in the vis- ible region. Since the energy of the conduction band for CeO2 is higher than that for o-YbFeO3 and h-YbFeO3, and the energy of the valence band, on the contrary, is lower, the resulting nanocomposite has a combination of type I heterojunctions cascaded into each other. This arrange- ment allows the holes to migrate from CeO2 to o-YbFeO3 and from there to h-YbFeO3, oxidizing OH– to ·OH, which is the strongest oxidizing agent. In addition, electrons can also cascade from CeO2 to o-YbFeO3, and then to h-YbFeO3, initiating the Fe3+ → Fe2+ reduction. The presence of H2O2 makes it possible for the Fenton-like process to arise. As shown earlier in [23], Fe2+ is formed on the surface of h-YbFeO3 during the reduction of Fe3+ by electrons trans- ferred from the o-YbFeO3 conduction band. During the reaction, Fe2+ is oxidized by H2O2 to Fe3+ and ·OH radicals are formed. Due to the addition of CeO2 to the system, a similar process can occur on the o-YbFeO3 surface through electrons supplied from the conduction band of cerium oxide. The interaction of ·OH with methyl violet leads to its oxidation to the most stable products – CO2 and H2O. Since the Fenton-like process of the formation of ·OH radi- cals is activated both on the surface of h-YbFeO3 and o-YbFeO3, the heterojunction h-YbFeO3/o-YbFeO3/CeO2 nanocomposite exhibits greater photocatalytic activity than the previously obtained h-YbFeO3/o-YbFeO3, pure o-YbFeO3, and pure CeO2. 4. Conclusions As a result of the work, new I-type heterojunction nanocomposites were successfully developed based on various structural forms of ytterbium orthoferrite and cerium dioxide, which were obtained using the method of solution combustion and heat treatment. A detailed analysis of the composition and structure of o-YbFeO3/h-YbFeO3/CeO2 nanocomposites made it pos- sible to determine the features of the formation of this polycrystalline system and the mutual influence of the components on the growth and transformation of the corresponding nanocrystals. The study of the morpho- logical features of nanocomposites and their specific surface area established a positive effect of the foamy microstructure and developed surface on the photocata- lytic activity of the samples. The formation of I-type heterojunction had a positive effect on the resulting efficiency of photocatalysts in the Fenton-like process of methyl violet degradation, and the presence of an optimum cerium dioxide content of 5% indicated a complex mutual influence of the components associated with covering active sites on the photocatalyst, shield- ing the light absorption, decreasing the surface area and activity, increasing the charge recombination. The nanocomposite photocatalysts developed as a result of this work may be of interest for potential use in photo - Fenton-like processes of oxidation of organic pollutants and other advanced oxidation processes. Fig. 10 Schematic mechanism of Fenton-like photodegradation of methyl violet in the presence of heterojunction nanocomposite o-YbFeO3/h-YbFeO3/CeO2 Chimica Techno Acta 2021, vol. 8(4), № 20218407 ARTICLE 11 of 11 Acknowledgments This work is supported by the Grant of President of the Rus- sian Federation МК-795.2021.1.3. The XRD, SEM, and EDS studies were performed on the equipment of the Engineer- ing Center of Saint Petersburg State Institute of Technology. References 1. Matafonova G, Batoev V. Recent advances in application of UV light-emitting diodes for degrading organic pollutants in water through advanced oxidation processes: A review. Wa- ter Res. 2018;132:177–189. doi:10.1016/j.watres.2017.12.079 2. Skvortsova LN, Bolgaru KA, Sherstoboeva MV, Dychko KA. Degradation of diclofenac in aqueous solutions under condi- tions of combined homogeneous and heterogeneous photoca- talysis. Russ J Phys Chem A. 2020;94:1248–1253. doi:10.1134/S0036024420060242 3. Makhotkina OA, Preis SV, Parkhomchuk EV. Water delignifi- cation by advanced oxidation processes: Homogeneous and heterogeneous Fenton and H2O2 photo-assisted reactions. Appl Catal B Environ. 2008;84:821–826. doi:10.1016/j.apcatb.2008.06.015 4. Hosokawa S, Jeon HJ, Inoue M. Thermal stabilities of hexagonal and orthorhombic YbFeO3 synthesized by solvothermal method and their catalytic activities for methane combustion. Res Chem Intermed. 2011;37:291–296. doi:10.1007/s11164-011-0251-9 5. Cao S, Sinha K, Zhang X, Zhang X, Wang X, Yin Y. Electronic structure and direct observation of ferrimagnetism in mul- tiferroic hexagonal YbFeO3. Phys Rev B. 2017;95. doi:10.1103/PhysRevB.95.224428 6. Albadi Y, Sirotkin AA, Semenov VG, Abiev RS, Popkov VI. Syn- thesis of superparamagnetic GdFeO3 nanoparticles using a free impinging-jets microreactor. Russ Chem Bull. 2020;69:1290–1295. doi:10.1007/s11172-020-2900-x 7. Sklyarova A, Popkov VI, Pleshakov IV, Matveev VV, Štěpánková H, Chlan V. Peculiarities of 57Fe NMR Spectrum in micro- and nanocrystalline europium orthoferrites. Appl Magn Reson. 2020. doi:10.1007/s00723-020-01224-y 8. Shen T, Hu C, Yang WL, Liu HC, Wei XL. Theoretical investi- gation of magnetic, electronic and optical properties of or- thorhombic YFeO3: A first-principle study. Mater Sci Semicond Process. 2015;34:114–120. doi:10.1016/j.mssp.2015.02.015 9. Makoed II, Liedienov NA, Pashchenko AV, Levchenko GG, Tatarchuk DD, Didenko YV. Influence of rare-earth doping on the structural and dielectric properties of orthoferrite La0.5R0.5FeO3 ceramics synthesized under high pressure. J Al- loys Compd. 2020;842:155859. doi:10.1016/j.jallcom.2020.155859 10. Downie LJ, Goff RJ, Kockelmann W, Forder SD, Parker JE, Morrison FD. Structural, magnetic and electrical properties of the hexagonal ferrites MFeO3 (M=Y, Yb, In). J Solid State Chem. 2012;190:52–60. doi:10.1016/j.jssc.2012.02.004 11. Polat O, Coskun M, Coskun FM, Zlamal J, Kurt BZ, Durmus Z. Co doped YbFeO3: exploring the electrical properties via tun- ing the doping level. Ionics.2019;25:4013–4029. doi:10.1007/s11581-019-02934-5 12. Dhinesh Kumar R, Thangappan R, Jayavel R. Synthesis and characterization of LaFeO3/TiO2 nanocomposites for visible light photocatalytic activity. J Phys Chem Solids. 2017;101:25–33. doi:10.1016/j.jpcs.2016.10.005 13. Martinson KD, Kondrashkova IS, Omarov SO, Sladkovskiy DA, Kiselev AS, Kiseleva TY. Magnetically recoverable catalyst based on porous nanocrystalline HoFeO3 for processes of n- hexane conversion. Adv Powder Technol. 2020;31:402–408. doi:10.1016/j.apt.2019.10.033 14. Baeissa ES. Environmental remediation of aqueous methyl orange dye solution via photocatalytic oxidation using Ag- GdFeO3 nanoparticles. J Alloys Compd. 2016;678:267–272. doi:10.1016/j.jallcom.2016.04.007 15. Liu J, He F, Chen L, Qin X, Zhao N, Huang Y. Novel hexagonal- YFeO3/α-Fe2O3 heterojunction composite nanowires with en- hanced visible light photocatalytic activity. Mater Lett. 2016;165:263–266. doi:10.1016/j.matlet.2015.12.008 16. Yang Z, Wu H, Liao J, Li W, Song Z, Yang Y. Infrared to visible upconversion luminescence in Er3+/Yb3+ co-doped CeO2 in- verse opal. Mater Sci Eng B. 2013;178:977–981. doi:10.1016/j.mseb.2013.06.007 17. Shao Z, Meng X, Lai H, Zhang D, Pu X, Su C. Coralline-like Ni2P decorated novel tetrapod-bundle Cd0.9Zn0.1S ZB/WZ homojunctions for highly efficient visible-light photocatalytic hydrogen evolution. Chinese J Catal. 2021;42:439–449. doi:10.1016/S1872-2067(20)63597-5 18. Hosokawa S, Matsumoto S, Tada R, Shibano T, Tanaka T. Development of Mn-modified hexagonal YbFeO3 catalyst for reducing the use of precious metal resources. J Japan Soc Powder Powder Metall. 2017;64:583–588. doi:10.2497/jjspm.64.583 19. Tang P, Yu L, Min J, Yang J, Chen H. Preparation of nanocrys- talline YbFeO3 by sol-gel method and its visible-light photo- catalytic activities. Ferroelectrics. 2017;521:71–76. doi:10.1080/00150193.2017.1390962 20. Jia LX, Zhu JY, Lin TT, Jiang Z, Tang CW, Tang PS. Prepara- tion YbFeO3 by microwave assisted method and its visible- light photocatalytic activity. Adv Mater Res. 2013;699:708– 711. doi:10.4028/www.scientific.net/AMR.699.708 21. Kondrashkova IS, Martinson KD, Zakharova NV, Popkov VI. Synthesis of Nanocrystalline HoFeO3 Photocatalyst via Heat Treatment of Products of Glycine-Nitrate Combustion. Russ J Gen Chem. 2018;88:2465–2471. doi:10.1134/S1070363218120022 22. Petschnig LL, Fuhrmann G, Schildhammer D, Tribus M, Schottenberger H, Huppertz H. Solution combustion synthe- sis of CeFeO3 under ambient atmosphere. Ceram Int. 2016;42:4262–4267. doi:10.1016/j.ceramint.2015.11.102 23. Tikhanova SM, Lebedev LA, Martinson KD, Chebanenko MI, Buryanenko IV, Semenov VG. The synthesis of novel heterojunc- tion h-YbFeO3/o-YbFeO3 photocatalyst with enhanced Fenton- like activity under visible-light. New J Chem. Royal Society of Chemistry. 2021;45:1541–1550. doi:10.1039/d0nj04895j 24. Popkov VI, Almjasheva OV, Nevedomskiy VN, Sokolov VV, Gusarov VV. Crystallization behavior and morphological fea- tures of YFeO3 nanocrystallites obtained by glycine-nitrate combustion. Nanosyst Physics, Chem Math. 2015;6:866–874. doi:10.17586/2220-8054-2015-6-6-866-874 25. Popkov VI, Almjasheva OV, Nevedomskiy VN, Panchuk VV, Semenov VG, Gusarov VV. Effect of spatial constraints on the phase evolution of YFeO3-based nanopowders under heat treatment of glycine-nitrate combustion products. Ceram Int. 2018;44:20906–20912. doi:10.1016/j.ceramint.2018.08.097 26. Kusmierek E. A CeO2 semiconductor as a photocatalytic and photoelectrocatalytic material for the remediation of pollutants in industrial wastewater: A review. Catalysts. 2020;10:1–54. 27. Fan X, Hao H, Shen X, Chen F, Zhang J. Removal and degrada- tion pathway study of sulfasalazine with Fenton-like reac- tion. J Hazard Mater. 2011;190:493–500. doi:10.1016/j.jhazmat.2011.03.069 28. Wu S, Lin Y, Yang C, Du C, Teng Q, Ma Y. Enhanced activation of peroxymonosulfte by LaFeO3 perovskite supported on Al2O3 for degradation of organic pollutants. Chemosphere. 2019;237:124478. doi:10.1016/j.chemosphere.2019.124478 29. Saeed K, Khan I, Gul T, Sadiq M. Efficient photodegradation of methyl violet dye using TiO2/Pt and TiO2/Pd photocata- lysts. Appl Water Sci. 2017;7:3841–3848. doi:10.1007/s13201-017-0535-3 30. Mazarji M, Esmaili H, Bidhendi GN, Mahmoodi NM, Minkina T, Sushkova S. Green synthesis of reduced graphene oxide-CoFe2O4 nanocomposite as a highly efficient visible-light-driven catalyst in photocatalysis and photo Fenton-like reaction. Mater Sci Eng B. 2021;270:115223. doi:10.1016/j.mseb.2021.115223 https://doi.org/10.1016/j.watres.2017.12.079 https://doi.org/10.1134/S0036024420060242 https://doi.org/10.1016/j.apcatb.2008.06.015 https://doi.org/10.1007/s11164-011-0251-9 https://doi.org/10.1103/PhysRevB.95.224428 https://doi.org/10.1007/s11172-020-2900-x https://doi.org/10.1007/s00723-020-01224-y https://doi.org/10.1016/j.mssp.2015.02.015 https://doi.org/10.1016/j.jallcom.2020.155859 https://doi.org/10.1016/j.jssc.2012.02.004 https://doi.org/10.1007/s11581-019-02934-5 https://doi.org/10.1016/j.jpcs.2016.10.005 https://doi.org/10.1016/j.apt.2019.10.033 https://doi.org/10.1016/j.jallcom.2016.04.007 https://doi.org/10.1016/j.matlet.2015.12.008 https://doi.org/10.1016/j.mseb.2013.06.007 https://doi.org/10.1016/S1872-2067(20)63597-5 https://doi.org/10.2497/jjspm.64.583 https://doi.org/10.1080/00150193.2017.1390962 https://doi.org/10.4028/www.scientific.net/AMR.699.708 https://doi.org/10.1134/S1070363218120022 https://doi.org/10.1016/j.ceramint.2015.11.102 https://doi.org/10.1039/d0nj04895j https://doi.org/10.17586/2220-8054-2015-6-6-866-874 https://doi.org/10.1016/j.ceramint.2018.08.097 https://doi.org/10.1016/j.jhazmat.2011.03.069 https://doi.org/10.1016/j.chemosphere.2019.124478 https://doi.org/10.1007/s13201-017-0535-3 https://doi.org/10.1016/j.mseb.2021.115223