Next Article in Journal
A Comparison of Shell Theories for Vibration Analysis of Single-Walled Carbon Nanotubes Based on an Anisotropic Elastic Shell Model
Next Article in Special Issue
The Effect of Lithium Ion Leaching from Calcined Li–Al Hydrotalcite on the Rapid Removal of Ni2+/Cu2+ from Contaminated Aqueous Solutions
Previous Article in Journal
Editorial for the Special Issue “Applications of Optical Thin Films”
Previous Article in Special Issue
S-Doped NiFe2O4 Nanosheets Regulated Microbial Community of Suspension for Constructing High Electroactive Consortia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Visible-Light-Enhanced Heterogeneous Photo Degradation of Tetracycline by a Nano-LaFeO3 Catalyst with the Assistance of Persulfate

1
College of Ocean Science and Engineering, Shanghai Maritime University, Shanghai 201306, China
2
Key Laboratory of Original Agro-Environmental Pollution Prevention and Control, Agro-Environmental Protection Institute, Ministry of Agriculture and Rural Affairs, Tianjin 300191, China
3
College of Resources and Environment, Northeast Agricultural University, Harbin 150030, China
4
College of Information Engineering, Shanghai Maritime University, Shanghai 201306, China
5
Key Laboratory of Environmental Biotechnology, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(8), 1388; https://doi.org/10.3390/nano13081388
Submission received: 28 February 2023 / Revised: 6 April 2023 / Accepted: 14 April 2023 / Published: 17 April 2023

Abstract

:
Perovskites with nano-flexible texture structures and excellent catalytic properties have attracted considerable attention for persulfate activation in addressing the organic pollutants in water. In this study, highly crystalline nano-sized LaFeO3 was synthesized by a non-aqueous benzyl alcohol (BA) route. Under optimal conditions, an 83.9% tetracycline (TC) degradation and 54.3% mineralization were achieved at 120 min by using a coupled persulfate/photocatalytic process. Especially compared to LaFeO3-CA (synthesized by a citric acid complexation route), the pseudo-first-order reaction rate constant increased by 1.8 times. We attribute this good degradation performance to the highly specific surface area and small crystallite size of the obtained materials. In this study, we also investigated the effects of some key reaction parameters. Then, the catalyst stability and toxicity tests were also discussed. The surface sulfate radicals were identified as the major reactive species during the oxidation process. This study provided a new insight into nano-constructing a novel perovskite catalyst for the removal of tetracycline in water.

1. Introduction

Water contamination by pharmaceuticals has become an environmental issue of concern in recent decades. Most pharmaceuticals were directly excreted into domestic wastewater without effective treatment by conventional water treatment sites. Up until now, various kinds of drugs, such as antibiotics, hormones, and preservatives, have been identified within water environmental systems [1]. Because of their antibacterial natures, the antibiotic residues within contaminated waters cannot be effectively removed by traditional biological methods. The chemical advanced oxidation processes (AOPs) for wastewater treatment appear then as reliable alternatives to biodegradation processes. AOPs involve the generation of hydroxyl radicals (OH) [2,3,4,5]. In addition to hydroxyl-radicals-based reactions, some other AOPs, including peroxydisulfate-activated reactions, ultrasounds, UV-vis irradiation, and ozonation, have received great attention over the last decades [6,7,8,9,10]. In a “hybrid” AOP, consisting of visible light and the activation of peroxysulfate (PS), the sulfate radical was produced from the decomposition of persulfate ions (S2O82−) by using heat or even transition metal (Equations (1) and (2)) [7]:
S2O82− + heat → 2SO4•−
S2O82− + Men+ →SO4•− + SO42− + Me(n+1)
As shown in Equations (3)–(5), in sulfate-radicals-mediated photo-Fenton-like reactions, PS is an oxidant activated by the catalyst to produce reactive radicals (SO4•−, OH, and others), since it acts as an electron acceptor that can decrease the recombination between photogenerated electrons (e) and holes (h+) [8]. Consequently, h+ reacts with PS to generate SO4•− (Equation (3)). The formed e is able to decompose the PS to generate SO4•− (Equation (4)). In addition, the PS can be activated in the presence of visible light, generating SO4•− (Equation (5)). If a homogeneous ferrous catalyst is used, Fe2+ can be regenerated by the photogenerated e. However, the development of more efficient and stable photoactivation catalysts for sulfate-radical photo-Fenton processes is required.
S2O82− + hvb+ → 2SO4•−
S2O82− + e → 2SO4•−
S2O82− + hv → 2SO4•−
Perovskite-based oxides, of the general formula ABO3, are important heterogeneous catalysts due to the low cost and flexibility of their composition, given their adjustable redox properties [11]. Therefore, they are especially studied within gas-phase catalytic depollution processes, including oxidation or reduction (CO, methane, VOCs, soot, and NOx, etc.) [12,13,14]. Perovskites have been proposed as potential alternatives to the commercially used supported noble metals (mostly based on Pt, Pd formulations) [14]. Among the developed perovskites, iron-based compositions are found to present advantages for catalytic processes, due to their excellent high-temperature thermal and hydrothermal stability, redox activity, and the low cost of their constituting elements [15,16,17,18]. More interesting, in addition to their efficiency for gas-phase effluent treatments, perovskites can also serve as catalysts for wastewater treatment, due to the presence of transition metals with multiple valence states. It is reported, for example, that BiFeO3-type perovskites (BFO) are active catalysts for oxidative acetylation [19], photocatalysis [20,21,22,23,24,25,26], hydrogen generation [27], and the catalytic activation of H2O2 [28]. BiFeO3/ZnFe2O4 nanocomposites have also been employed as catalysts for the removal of azo dyes under visible light irradiation [28].
Owing to its stable crystal structure, excellent optical properties, and narrow bandgap (~2.1 eV), LaFeO3 has been considered as a promising photoactive candidate. However, because LaFeO3 is a p-type semiconductor, the rapid recombination of e and h+ have been reported [29]. In addition, LaFeO3 displays a low activity due to its high charge recombination rates. Therefore, introducing peroxides into the photocatalysis system may be an efficient strategy with the consideration of the capture effect of peroxides for photo-induced electrons on the conduction band of a photocatalyst. In this frame, it could be anticipated that the synergistic effects of the photocatalytic process and sulfate-radical-based performance could exhibit a higher degradation rate for organic pollutants than the Fenton process. As reported, peroxymonosulfate (PMS) was evidenced to be a good candidate for avoiding the recombination of e and h+ and for improving the degradation efficiency of organic compounds [30,31,32,33,34,35]. Compared to H2O2, the longer O–O bond length and asymmetric structure of HSO5 are supposed to guarantee the easier dissociation of PMS over organic compounds. However, some problems still remain, such as the limited catalytic activity of perovskite considering the low surface developed by the perovskites [11], the underutilization of PS [36], and the low stability of the catalyst with iron ions leaching occurring during the reaction [37].
From a fundamental point of view, textural properties, structural properties, and redox properties are interrelated parameters that have a significant effect on the catalytic performance of perovskites. However, using conventional methods for perovskite synthesis, such as calcination milling [17] and even hydroxyacid complexation routes [38,39], lead to a large grain size (D), a low specific surface area (SSA), and a poor activity [24,25]. Indeed, when using a classical complexation approach [38,39], precursor complexation with hydroxyacid allows for the cations to be kept homogeneously dispersed in a hybrid matrix, ensuring crystallization at a lower temperature than that for solid–solid approach, i.e., 600–650 °C. Better catalytic performances are achieved in this last case due to improved textural properties, but an accessible surface area is not the only parameter affecting this catalysis, since the B-cation surface concentration, reducibility, and oxygen mobility of the catalysts are also considered to be key parameters [11,12,26]. Some improved synthetic approaches have then been proposed to obtain higher surface areas (30–100 m2 g−1) and smaller crystal sizes (<20 nm). Typical examples of these include the sol-gel-derived method [13], reactive grinding [40], nanocasting [41], and flame-spray pyrolysis [42], etc. Reactive grinding is an excellent example that allows for the production of high-surface-area materials (reported to reach 100 m2 g−1 [43]), but its popularization is restricted due to the requirement of special equipment, and this approach is better adapted to large-scale synthesis than laboratory-scale synthesis [44,45]. However, it is still a real challenge to directly synthesize size-limited and pure nanocrystalline perovskites with particular porosities. A benzyl alcohol (BA) route was proposed as a valuable approach for oxide and mixed oxide nanoparticle preparations. This route not only offers an excellent control over the particle size, shape, crystallinity, order, purity, and assembly behavior, but also avoids the use of any additional surfactants or templates [46,47,48,49]. The BA route is defined as a non-aqueous synthesis, since during the different competitive reactions, no water is added, which affords a better control over the crystal assembly.
In this work, nano-LaFeO3 was prepared by a citric acid complexation route and benzyl alcohol route. Using an adapted synthesis procedure [49], the benzyl-alcohol-derived material presented quasi-spherical nanosized particles with a much higher specific surface area, as well as a lower crystal size domain size. The reaction parameters for the sulfate-radical–photo-Fenton tetracycline degradation process were optimized (operating pH, oxidant concentration, and catalyst loading) with this catalyst. The catalyst stability was also evaluated upon the reaction cycling, which was followed by metal ion leaching.

2. Materials and Methods

2.1. Materials and Methods

The synthesis was performed by a benzyl alcohol route.
The sample LaFeO3−BA: La(NO3)3·6H2O (2 mmol) was dissolved in anhydrous benzyl alcohol (26.7 mL) at room temperature. After that, iron (III) acetylacetonate (2 mmol) was added dropwise into the benzyl alcohol (40 mL) at room temperature. The reaction mixture was thereafter stirred for 30 min and transferred into a Teflon-lined autoclave. The autoclave was heated to 200 °C for 24 h. The resulting brown suspension was recovered by centrifugation (4 cycles, with washing with ethanol between each cycle). The centrifugation was performed at 8000 rpm for 10 min. The washed material was dried at 80 °C for 12 h, and the dry solid was calcined at 500 °C for 6 h. The prepared sample was labeled as the BA sample.
The synthesis was then performed by a citric acid complexation route.
The sample LaFeO3−CA: aqueous solutions of lanthanum(III) nitrate hexahydrate (20 mM) (LaNO3∙6H2O ≥ 99.0 %, Sigma-Aldrich, Houston, USA) and iron(III) nitrate hexahydrate (20 m mol L−1) (97.7%, Alfa Aesar, Shanghai, China) were prepared with distilled water. Thereafter, an aqueous solution of citric acid (98 %, Prolabo, Paris, France), with a molar equivalence to La + Fe (40 mmol L−1), was added dropwise to the precursor solution under gentle stirring at 80 °C. Ageing was performed until a gel was obtained. The gel was thereafter dried in an oven at 110 °C for 12 h. Subsequently, the obtained foam solid was crushed and calcined at 500 °C for 6 h. The sample was labeled as the CA sample.

2.2. Experimental

The reaction configuration: the experiment was carried out in a photo-reactor consisting of visible LED lamps, as shown in Figure 1. The LED lamp emitted white light at a wavelength of 455 nm and further broadband stokes-shifts were emitted at roughly 500–600 nm.
For the activity measurement, a solution of tetracycline (TC, purity, and source) was prepared with deionized water (in the two hours preceding the reaction). The initial concentration (C0) was fixed at 100 mg L−1 and the C0 exact value was verified by an HPLC analysis. Then, a selected weight of the LaFeO3 was added into 200 mL of the TC solution, with constant stirring that was maintained using a mechanical stirrer (RW 20, IKA, Staufen, Germany) for 15 min, in order to achieve adsorption–desorption equilibrium on the solid catalyst surface. At this time, the [TC]0 value—the concentration of the TC after the adsorption process—was obtained. Then, a persulfate (PS) solution, prepared by the dissolution of 1 mol L−1 (purity and source) in water to obtain a PS concentration of 50 mmol L−1, was added to start the reaction. All the experiments were conducted at ambient temperature (22 ± 2 °C). The reaction solution was sampled (1 mL) over fixed intervals and mixed with 1 mL of methanol (0.1 mol L−1) for at least 15 min to terminate the reaction. Then, the reaction solution was filtered through a 0.22 μm membrane (Millipore Co., Molsheim, France) and mixed with the same volume of methanol to quench the reaction. The TC conversion (XTC) and the total organic carbon (TOC) abatement (XTOC) were calculated by Equations (6) and (7), respectively.
X T C = C 0 C C 0 × 100 %
X T O C = T O C 0 T O C T O C 0 × 100 %
where C is the TC concentration at time t (min), TOC is the total organic carbon time t (min), and TOC0 is the initial total organic carbon after stabilization due to the adsorption on the catalyst surface.

2.3. Characterization and Analysis

X-ray diffraction (XRD) was performed. The X-ray diffraction patterns were obtained on a D5005 diffractometer (Bruker, Karlsruhe, Germany) equipped with Cu Kα radiation (λ = 1.5406 Å). The signal was recorded for 2θ between 20° and 80°, with a recording step of 0.05° (3 s per step). A phase identification was performed by a comparison with the JCPDS database. The crystal size was determined using the Scherrer equation after applying Warren’s correction for the instrumental broadening.
For the N2-physisorption, N2 adsorption measurements were performed on a Micromeritics TRISTAR 3000 instrument. A desorption isotherm was used to determine the pore size distribution using the Barret Joyner Halenda (BJH) method. The specific surface areas (SSA) of the catalysts were obtained by applying the Brunauer–Emmett–Teller (BET) equation (0.10 < P/P0 < 0.35). Prior to the analysis, the samples (0.5 g) were heated at 250 °C under a vacuum for 6 h.
The specific surface area (ssa) can then be derived:
s s a = V m N A a m v m m s
where NA is the Avogadro’s number (6.022 × 1023 mol−1), am is the effective cross-section area of one adsorbed molecule, vm is the molar volume of one adsorbed molecule (22,400 mL of the volume occupied by 1 mol of the adsorbate gas under standard conditions), and ms is the mass of the substrate/adsorbent.
Then, transmission electron microscopy (TEM) was used. The catalyst morphology was determined by a TEM analysis coupled with energy-dispersive X-ray spectroscopy (EDXS). The micrographs were collected on a JEM 2100 instrument (JEOL, Tokyo, Japan), operated at 200 kV with a LaB6 source and equipped with a Gatan Ultra scan camera. The EDX spectroscopy was carried out with a Hypernine (Premium, JEOL, Japan) detector (active area: 30 mm2), using the SM-JED 2300T software for the data acquisition and treatment.
All the catalysts were systematically characterized using X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), N2-physisorption, and ICP-OES (see the Supplementary Data for details).
The tetracycline concentration was determined by using a high-performance liquid chromatography (HPLC) instrument that was equipped with an Aminex HPX-87 (BioRad, Hercules, CA, USA) column. The TOC content was determined using a TOC meter. The concentration of the PS was analyzed with an iodometric titration method, according to the procedure described in ref [6]. The amount of leached iron was determined by an analysis of the Fe ion concentration in the aqueous phase. Spectrophotometry at λ = 510 nm, after adding 1,10-phenanthroline to form an iron–phenanthroline complex, was used. Finally, the stability and reusability of the catalyst were evaluated.

3. Results and Discussion

3.1. Physicochemical and Structure Properties

The samples that were prepared by non-aqueous benzyl alcohol routes (BA) and citrate complexation routes (CA) showed reflections that were characteristic of the perovskite phase (Figure 2), being in perfect agreement with the orthorhombic structure of LaFeO3 (JCPDS card 74-2203). A small, broad, and poorly defined reflection of 28.220° (101) was visible on the baseline of the BA sample pattern. This reflection was identified in the La(OH)3 external phase, according to the JCPDS reference JCPDS No. 36-1481. The larger FWHM of the XRD peaks observed for the BA sample show that this material presented a lower crystal size than the CA sample at the iso-crystallization temperature (18.4 nm for BA vs. 22.3 nm for CA): an 18% lower crystal size was measured for the BA. These results confirm that the BA route was adapted for the production of crystalline-mixed oxide nanoparticles, starting from an organic–inorganic hybrid precursor [50]. A larger specific surface area (32.5 m2 g−1) was also obtained for the BA sample, which was much higher than the value for the LaFeO3 sample that was obtained by the CA route (15.3 m2 g−1), with a 112% higher SSA.
Starting from the value of the average crystal domain size obtained from the XRD, it is possible to calculate a theoretical surface area, assuming a cubic-shaped crystal and no contact between the adjacent elementary particles. Then, a ratio R between the experimental specific surface area and the theoretical surface area (R = SBET/Sth, Table 1) gives an indication of the agglomeration degree of the elementary crystals in the different materials.
Then, when R = 1, the developed surface area is equal to the theoretical surface area, which characterizes a poorly agglomerated material. When R approaches 0, meaning that the theoretical surface area is larger than the experimental specific surface area, dense and poorly porous agglomerates are occured in the material [51,52]. A comparison between the R ratios obtained for the two samples showed that the BA sample presented a significantly higher ratio compared to that of the CA sample. This result indicates that the BA route prevented the potential agglomeration of the perovskite, leading to a greater accessible surface area.
The morphologies of the BA and CA LaFeO3 samples were investigated by TEM (Figure 3).
Based on the images in Figure 3A,B, the LaFeO3 sample that was synthesized via the BA route was composed of almost spherical LaFeO3 nanoparticles, with diameters ranging from 10 to 30 nm. These observed sizes were consistent with the average crystal domain sizes determined by the XRD (18.4 nm). The elementary particles formed small agglomerates, in which the presence of contrasted zones evidenced the formation of internal porosity. Slight differences were observed for the CA sample. While the low-magnification images evidenced the formation of agglomerates of more comparable sizes than those for the BA sample (Figure 3C), the high-magnification observation evidenced the agglomeration of elementary particles of sizes between 10 and 80 nm (Figure 3D). The particles showed inhomogeneity in their sizes and shapes. This morphology, mostly that of the larger average particle size, was at the origin of the lowest SSA (15.3 m2 g−1).

3.2. Process Configuration Effect on the Degradation of the TC Degradation

Different configurations of the process were tested to evaluate their respective efficiencies in the TC degradation reaction (Figure 4):
  • The PS alone (single process);
  • The catalyst alone (single process);
  • A combination of visible light/PS (dual process);
  • A combination of PS/BA (dual process);
  • A combination of visible light/BA (dual process);
For the combination of visible light/PS/BA (ternary process), the quantum yield was estimated to examine the photodegradation efficiency [53], which was found to be 2.3 × 10−2.
The photo stability of the tetracycline was also assessed. In order to explore the individual contribution of the photo-catalysis to the removal of the TC, the TC solution was irradiated by visible light alone until the equilibrium concentration of the TC, [TC]0, was obtained. A 15.5% decrease in the TC concentration was obtained by photolysis, given a [TC]0 value of 98.9 mg L−1 (corrected concentration). The retaining of more than 80% of the initial concentration evidenced the resistance of the TC toward to photo-catalytic degradation under visible irradiation alone.
In the single processes (PS or catalyst alone), the TC chemical degradation rate remained very slow under the selected operating conditions (22 °C, pH = 3.7). At the end of the reaction time, i.e., t = 120 min, the remaining concentration of the TC stayed above 92% of the [TC]0 value (Figure 4). The experiments realized with only the BA catalyst added to the solution also demonstrated a very limited efficiency, with the value of the TC concentration’s decrease remaining below 20%. Such an abatement (<10%) then demonstrates the limited ability of the catalyst surface for the simple adsorption of the organic pollutant.
Among the three configurations of the dual processes, the least efficient was the visible light/BA configuration, with an abatement that was limited to 14%, which is not surprising considering that, in this configuration, no efficient oxidant was added. Then, a slightly better efficiency was reached with the dual visible light/PS configuration, with a TC concentration abatement of 30% that was obtained after 120 min. However, this result shows that the process was poorly efficient when considering the stoichiometry of the reaction, being in excess of the PS (molar ratio PS/TC = 208.3) to afford the complete degradation of the pollutant. The best results were obtained during the last dual process, the coupling of PS and BA without light irradiation: a TC concentration abatement of 39% was obtained after 120 min. These results clearly indicate that the PS could be activated by the LaFeO3 to generate active radical oxidants.
When perovskites are irradiated with visible light energy that is equivalent to their band gap energy, the generation of photo-generated electrons and holes takes place. These photo-generated electrons and holes then participate in series of reactions and lead to the generation of active radical species (OH), as shown below in Equations (9)–(13).
LaFeO3 + hv → LaFeO3 (e + h+)
e + O2 → O2•−
O2•− + H+ → HO2
2e + HO2 + H+OH + OH
h+ + OHOH
h+ + H2O → H+ + OH
Additionally, these photo-generated electrons and holes (h+) can undergo reactions with PS to enhance the production of SO4•−, as shown below in Equations (14) and (15):
e + S2O82− → SO4•− + HO
h+ + HOOH
Besides the role of photocatalysis, the activation of PS on the iron oxide surface induces the formation of radicals, including SO4•− (Equations (17)–(19)) via a reaction on the catalyst surface [34], which will thereafter react with the TC. From a mechanistic point of view, ref. [35] demonstrated that the formation of these radicals involves an Fe3+/Fe2+ redox cycle, according to the following equations:
Fe2+ + HSO5 → Fe3+ +SO4•− + HO
Fe2+ + HSO5 → Fe3+ + SO42− + HO
Fe3+ + HSO5 → Fe2+ + SO5•− + H+
Thereafter, the formed radicals (sulfate and/or hydroxyl) can attack the TC molecules and the TC molecules can be directly oxidized by the OH and SO4•−, thus generating by-products. These detected intermediates are then, to some extent, further mineralized to CO2 (Equation (20)).
OH/SO4•− + TC ↔intermediates ↔ CO2 + H2O + SO42−
For the ternary process, when compared to the dual conditions, the degradation efficiency largely increased upon the visible irradiation being added to the coupled PS/BA. Indeed, the apparent constant rate value, kapp (calculated assuming a pseudo-first-order reaction, list the different hypotheses), that was measured for the visible light/PS/BA was 0.0218 min−1, which was more than 6.8 times than the sum of the kapp obtained for the visible light/PS (0.0032 min−1) and PS/LaFeO3 (0.0050 min−1). Especially compared to LaFeO3-CA (LaFeO3 syntheses by a citric acid complexation route), the pseudo-first-order reaction rate constant increased by 1.8 times. This indicated the synergistic effect of coupling the three conditions on the TC degradation rate. Furthermore, we evaluated the decomposition of the added PS that took place during the reaction. The utilization degree of the PS was 63% in the ternary visible light/PS/BA process, while it was only 21% for the dark dual PS/BA process. TC can absorb visible light energy into the excited state, facilitating an electron transfer to the Fe3+ surface site and resulting in PS decomposition into SO4•− radicals, with an associated reduction in the iron surface site that is available on the catalyst surface, as proposed the above in Equations (16)–(18).
In order to confirm this mechanism (surface activated/dependent or not), the CA and BA catalysts were tested under comparable conditions (Figure 5). While a kapp of 0.0218 min−1 was obtained with the BA catalyst, a value of 0.0122 min−1 was obtained with the CA catalyst.
This result demonstrated the beneficial effect of using a BA catalyst instead of a CA catalyst. Considering the surface-activated mechanism for oxidant radical formation, both the surface area (being higher for the BA, Table 1) and the Fe3+ surface site’s reducibility (also being higher for the BA) were expected to be at the origin of the highest performances obtained with the BA catalyst. TPR experiments indeed demonstrated an activation of the Fe3+ surface site with the BA at a temperature of approx. 400 °C below the temperature that was obtained for the CA, which was already obtained by our previous results on the oxygen mobility/reducibility of iron-based perovskites produced by a benzyl alcohol route [44].

3.3. Operating Parameter Impact on the Process Global Efficiency

3.3.1. Catalyst Dosage and Oxidant Concentration Effect

The effect of the concentration of the BA on the TC degradation efficiency was also investigated (Figure 6). It was found that the TC removal efficiency and kapp increased with an increase in the catalyst loading when it increased from 0.1 g L−1 to 0.5 g L−1. At a BA loading of 0.5 g L−1, an 83.9% abatement of the [TC]0 was obtained after 120 min, corresponding to a kapp of 0.0218 min−1. A further increase of catalysts loading (BA) did not improve the TC abatement (84%) after 120 min.
The effects of the oxidants and PS concentrations on the degradation of the TC are presented in Figure 7 and also show that an increase in the oxidant concentration is beneficial to the TC abatement.
In this last case, the kapp value was linearly dependent on the PS concentration. Then, in the interval of the optimization selected, the catalyst loading and initial PS concentration values that led to the highest TC abatement, at a pH of 3.7, were 0.5 g L−1 and 50 mmol L−1, respectively.

3.3.2. pH Value Effect

To investigate the pH effect on the degradation of the tetracycline, three different values of the initial pH (3.7, 7.0, and 9.0) were tested under acidic, neutral, and basic conditions. The results presented in Figure 8 show that the calculated TC removal efficiency and kapp, assuming a first-order reaction rate, decreased accordingly with an increase in the pH value.
When compared to the value obtained at pH = 3.5 (kapp = 0.012 min−1), the value of the kapp decreased by 20% at pH = 7.0 to 0.0096 min−1. While the kapp decreased significantly, the final TC abatement value was 82.3% after 120 min, which was very close to the value obtained at pH = 3.5, i.e., 83.9%. From a mechanism point of view, carboxyl acids (such as acetic acid) are produced during the reaction. Therefore, when the TC is degraded, the production of acidic intermediates induces a pH value decrease in the reaction medium when the reaction is performed without a buffer solution (such as in our case). This explains why, with a reaction starting at pH = 3.7, a final pH value of 2.6 was obtained after 120 min of the reaction. In the case of an initial pH of 7.0, the final value obtained was 2.7. Then, starting from 3.7 or 7.0, the final pH values of the reaction were comparable, explaining why comparable TC abatements were obtained in the two cases. The lowest value of the kapp, 0.0073 min−1, was obtained when the initial pH value was increased to 9.0 and the TC abatement reached only 53%, a value far below the 82.3–83.9% obtained at the pHs of 3.7 and 7.0. In the case of an initial pH of 9.0, the pH value obtained at the end of the reaction was 4.2. These results demonstrate that the reaction was far more efficient in acidic conditions. However, the choice to perform the reaction without any buffer (such as NaHCO3/KHSO4 or Na2HPO4/KH2PO4) originated from the fact that buffer ions may act as scavengers of radicals such as SO4•−, as demonstrated in Equations (21) and (22) [44]:
SO4•− + HCO3− → SO42− + HCO3
HCO3 + H+ ⇋ CO32−

3.4. Catalyst Stability Test

The BA catalyst stability was investigated by the means of recycling experiments: the BA catalyst was repeatedly used in five successive batch experiments, as shown in Figure 9. After each run, the BA catalyst in the reaction solution was centrifuged, recovered, and dried at 100 °C for 12 h. The dried powder was directly used for the next reaction cycle, applying exactly similar conditions to those for the first run.
As can be observed in Figure 9, the BA was capable of being reutilized for at least five cycles, and the reused catalyst retained an acceptable catalytic activity. After the third cycle, the total removal efficiency of the TC was 83.9% at the first cycle. At the fifth cycle, the TC abatement always reached 79.3%. At this fifth cycle, the TC removal efficiency and the leached iron concentration were also evaluated. In addition, we also tested the TOC value. The result found that the TOC removal efficiency was 54.3%, which was an excellent level of mineralization considering the difficulty of degrading TC molecules. Furthermore, the leaching of Fe and La from LaFeO3 was also evaluated, quantifying the dissolved Fe and La in the reaction medium by ICP-OES: [La] = 0.03 mg L−1 and [Fe] = 0.02 mg L−1. This corresponded to a very limited leaching of the elements from the catalyst, in accordance with the maintenance of an excellent abatement during the five successive cycles at pH 3.7. To validate the heterogeneous character of the reaction, the reaction was performed with iron and lanthanum, which were added by the dissolution of nitrate precursors at the same concentrations as those obtained in the final reaction medium. A TC abatement of 3.81% was obtained after 120 min, which can be considered negligible in comparison to the 83.9% obtained in the presence of the catalyst. Then, this result confirmed the absence of a contribution of the dissolved species (either Fe or La) to the efficiency measured. It also allowed for the conclusion of a heterogeneous, mediated reaction, in which the nano-LaFeO3 behaved as an efficient and stable system when coupled with persulfate and visible irradiation.

3.5. Toxicity Evolution

Even though 83.9% of the TC was converted in the visible light/LaFeO3-BA/PS process after 120 min, the TOC conversion of the TC was only 54.3%. In this case, the variation in the acute toxicity was evaluated by a 48 h immobilization assay with D. magna. A 60% mortality rate of the D. magna was reached with a raw TC solution. After a 120 min reaction in the visible light/LaFeO3-BA/PS system, the toxicity reached a maximum value and 100% of the D. magna was immobilized, indicating that the formation of primary degradation by-products had a considerably higher toxicity than the TC. When the reaction time was extended to 240 min, the inhabitation percentage decayed, i.e., the mortality rate was only 45%. Therefore, prolonging the reaction time to reduce its acute toxicity is necessary for the TC treatment.

4. Conclusions

Tetracycline in an aqueous solution was efficiently degraded by a visible light/heterogeneous LaFeO3/persulfate process. The use of nano-LaFeO3 perovskite and visible LED lamps provided sensible advantages, including: (i) a low cost for the heterogeneous catalyst, (ii) convenient operation conditions (ambient temperature and no pressure), and (iii) a low power consumption (LED lamps irradiation). The properties of the nano-LaFeO3, especially its surface area, impact the efficiency of the global process. The other parameters influencing the final tetracycline degradation degree are: the catalyst loading, the persulfate loading, and the reaction medium pH. Satisfying the stability of the catalyst is obtained upon the reaction cycling, with a very limited leaching of the La and Fe. Nano-LaFeO3 can be considered to be an efficient heterogeneous catalyst, allowing for the use of persulfate as an oxidant to replace the more classically used H2O2, in view of the pollutant degradation in water.

Author Contributions

Writing—original draft preparation, conceptualization, methodology, L.H. and Y.W.; software, F.Z.; validation, S.L. and L.F.; formal analysis, L.W. and W.X.; project administration, funding acquisition, L.W. and C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Science and Technology Commission of Shanghai Municipality (No. 22692107400), the National Natural Science Foundation of China (No. 21976197), the Foreign Experts Program of the Ministry of Science and Technology of China (No. G2022051018L).

Data Availability Statement

Not applicable.

Conflicts of Interest

The inefficient Fe(III)/Fe(II) cycle and high persulfate dosage severely restricted the practical application prospects of the persulfate activation. The development of a novel heterogeneous catalyst with a high surface area, high Fe(III)/Fe(II) cycle, and in situ persulfate generation is of significance in the wastewater advance oxidation process. In this study, we designed and fabricated a nano-sized LeFeO3 as an efficient persulfate activator. Tetracycline in an aqueous solution was efficiently degraded by a visible light/heterogeneous LaFeO3/persulfate process. The low power consumption (LED lamps irradiation) is adaptable for industry utilization. The surface oxygen vacancies of LaFeO3 as photo-generated electron transfer mediators enhance the redox cycle efficiency of Fe(II)/Fe(III). The interfacial reaction mechanism experiment indicates that the surface sulfate radicals groups participate in the persulfate activation as the main active sites. We believe this study provides a facile route for efficient persulfate activation and strengthens the inconvenient role of in situ persulfate generation on pollutants degradation.

References

  1. Ogunlaja, A.; Ogunlaja, O.O.; Olukannia, O.D.; Taylor, G.O.; Olorunnisola, C.G.; Dougnon, V.T.; Mousse, W.; Fatta-Kassinos, D.; Msagati TA, M.; Unuabonah, E.I. Antibiotic resistomes and their chemical residues in aquatic environments in Africa. Environ. Pollut. 2022, 312, 119783. [Google Scholar] [CrossRef]
  2. Do, S.H.; Jo, J.H.; Jo, Y.H.; Lee, H.K.; Kong, S.H. Application of a peroxymonosulfate Cobalt (PMS/Co(II)) system to treat diesel-contaminated soil. Chemosphere 2009, 77, 1127–1131. [Google Scholar] [CrossRef] [PubMed]
  3. Rastogi, A.; Al-Abed, S.R.; Dionysiou, D.D. Sulfate radical-based ferrous–peroxymonosulfate oxidative system for PCBs degradation in aqueous and sediment systems. Appl. Catal. B Enivron. 2009, 85, 171–175. [Google Scholar] [CrossRef]
  4. Kolthoff, I.M.; Miller, I.K. The chemistry of persulfate. I. The kinetics and mechanism of the decomposition of the persulfate ion in aqueous medium. J. Phys. Chem. B 1951, 73, 3055–3059. [Google Scholar]
  5. Anipsitakis, P.G.; Stathatos, E.; Dionysiou, D.D. Heterogeneous Activation of Oxone Using Co3O4. J. Phys. Chem. B 2005, 109, 13052–13055. [Google Scholar] [CrossRef]
  6. Cai, C.; Zhang, H.; Zhong, X.; Hou, L. Ultrasound enhanced heterogeneous activation of peroxymonosulfate by a bimetallic Fe–Co/SBA-15 catalyst for the degradation of Orange II in water. J. Hazard. Mater. 2015, 283, 70–79. [Google Scholar] [CrossRef]
  7. House, D.A.; House, D.A. Kinetics and mechanism of oxidations by peroxydisulfate. Chem. Rev. 1962, 62, 185–203. [Google Scholar] [CrossRef]
  8. Wang, Y.; Liu, C.; Zhang, Y.; Meng, W.; Yu, B.; Pu, S.; Yuan, D.; Qi, F.; Xu, B.; Chu, W. Sulfate radical-based photo-Fenton reaction derived by CuBi2O4 and its composites with α-Bi2O3 under visible light irradiation: Catalyst fabrication, performance and reaction mechanism. Appl. Catal. B Enivron. 2018, 235, 264–273. [Google Scholar] [CrossRef]
  9. Chudzińska, J.; Woźniak, B.; Sprynskyy, M.; Nowak, I.; Feliczak-Guzik, A. Photoremoval of bisphenol A using hierarchical zeolites and diatom biosilica. Int. J. Mol. Sci. 2023, 24, 2878. [Google Scholar] [CrossRef] [PubMed]
  10. Sprynskyy, M.; Szczyglewska, P.; Wojtczak, I.; Nowak, I.; Witkowski, A.; Buszewski, B.; Feliczak-Guzik, A. Diatom biosilica doped with palladium(II) chloride nanoparticles as new efficient photocatalysts for methyl orange degradation. Int. J. Mol. Sci. 2021, 22, 6734. [Google Scholar] [CrossRef]
  11. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as substitutes of noble metals for heterogeneous catalysis: Dream or reality. Chem. Rev. 2014, 114, 10292–10368. [Google Scholar] [CrossRef] [PubMed]
  12. Royer, S.; Bérubé, F.; Kaliaguine, S. Effect of the synthesis conditions on the redox and catalytic properties in oxidation reactions of LaCo1−xFexO3. Appl. Catal. A 2005, 282, 273–284. [Google Scholar] [CrossRef]
  13. Kakihana, M.; Yashima, M.; Yoshimura, M.; Ikeda, N.; Sugitani, Y. Synthesis of high surface area LaMnO3+d by a polymerizable complex method. J. Alloys Compd. 1999, 283, 102–105. [Google Scholar] [CrossRef]
  14. Kim, C.H.; Qi, G.; Dahlberg, K.; Li, W. Strontium-doped perovskites rival platinum catalysts for treating NOx in simulated diesel exhaust. Science 2010, 327, 1624–1627. [Google Scholar] [CrossRef] [PubMed]
  15. Russo, N.; Fufori, S.; Fino, D.; Saracco, G.; Specchia, V. Lanthanum cobaltite catalysts for diesel soot combustion. Appl. Catal. B Enivron. 2008, 83, 85–95. [Google Scholar] [CrossRef]
  16. Peňa, M.A.; Fierro, J.L. Chemical structures and performance of perovskite oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef]
  17. Szabo, V.; Bassir, M.; Neste, A.V.; Kaliaguine, S. Perovskite-type oxides synthesized by reactive grinding: Part IV. Catalytic properties of LaCo1xFexO3 in methane oxidation. Appl. Catal. B Enivron. 2003, 43, 81–92. [Google Scholar] [CrossRef]
  18. Royer, S.; Neste, A.V.; Davidson, R.; Mclntyre, S.; Kaliaguine, S. Methane Oxidation over Nanocrystalline LaCo1-xFexO3: Resistance to SO2 Poisoning. Ind. Eng. Chem. Res. 2004, 43, 5670–5680. [Google Scholar] [CrossRef]
  19. Meríno, N.A.; Barbero, B.P.; Grange, P.; Cadús, L.E. La1xCaxCoO3 perovskite-type oxides: Preparation, characterisation, stability, and catalytic potentiality for the total oxidation of propane. J. Catal. 2005, 231, 232–244. [Google Scholar] [CrossRef]
  20. Nie, Y.; Zhang, L.; Li, Y.; Hu, C. Enhanced Fenton-like degradation of refractory organic compounds by surface complex formation of LaFeO3 and H2O2. J. Hazard. Mater. 2015, 294, 195–200. [Google Scholar] [CrossRef] [PubMed]
  21. Li, S.; Jing, L.; Fu, W.; Yang, L.; Xin, B.; Fu, H. Photoinduced charge property of nanosized perovskite-type LaFeO3 and its relationships with photocatalytic activity under visible irradiation. Mater. Res. Bull. 2007, 42, 203–212. [Google Scholar] [CrossRef]
  22. Parida, K.M.; Reddy, K.H.; Martha, S.; Das, D.P.; Biswal, N. Fabrication of nanocrystalline LaFeO3: An efficient sol–gel auto-combustion assisted visible light responsive photocatalyst for water decomposition. Int. J. Hydro. Energy 2010, 35, 12161–12168. [Google Scholar] [CrossRef]
  23. Tang, P.; Tong, Y.; Chen, H.; Cao, F.; Pan, G. Microwave-assisted synthesis of nanoparticulate perovskite LaFeO3 as a high active visible-light photocatalyst. Curr. Appl. Phys. 2013, 13, 340–343. [Google Scholar] [CrossRef]
  24. Su, H.; Jing, L.; Shi, K.; Yao, C.; Fu, H. Synthesis of large surface area LaFeO3 nanoparticles by SBA-16 template method as high active visible photocatalysts. J. Nanoparticle. Res. 2010, 12, 967–974. [Google Scholar] [CrossRef]
  25. Tijare, S.N.; Joshi, M.V.; Padole, P.S.; Mangrulkar, P.A.; Rayalu, S.S.; Labhsetwar, N.K. Photocatalytic hydrogen generation through water splitting on nano-crystalline LaFeO3 perovskite. Inter. J. Hydrogen Engery 2012, 37, 10451–10456. [Google Scholar] [CrossRef]
  26. Libby, W.F. Promising catalyst for auto exhaust. Science 1971, 171, 499–500. [Google Scholar] [CrossRef]
  27. Safizade, B.; Masoudpanah, S.M.; Hasheminiasari, M.; Ghasemi, A. Photocatalytic activity of BiFeO3/ZnFe2O4 nanocomposites under visible light irradiation. RSC Adv. 2018, 8, 6988–6995. [Google Scholar] [CrossRef]
  28. Shi, J.; Guo, L. ABO3-based photocatalysts for water splitting. Prog. Nat. Sci. Mat. Int. 2012, 22, 592–615. [Google Scholar] [CrossRef]
  29. Xu, K.; Feng, J. Superior photocatalytic performance of LaFeO3/gC3N4 heterojunction nanocomposites under visible light irradiation. RSC Adv. 2017, 7, 45369–45376. [Google Scholar] [CrossRef]
  30. Lu, S.; Wang, G.L.; Chen, S.; Yu, H.T.; Quan, X. Heterogeneous activation of peroxymonosulfate by LaCo1-xCuxO3 perovskites for degradation of organic pollutants. J. Hazard. Mater. 2018, 353, 401–409. [Google Scholar] [CrossRef] [PubMed]
  31. Gao, P.P.; Tian, X.K.; Nie, Y.L.; Yang, C.; Zhou, Z.X.; Wang, Y.X. Promoted peroxymonosulfate activation into singlet oxygen over perovskite for ofloxacin degradation by controlling the oxygen defect concentration. Chem. Eng. J. 2019, 359, 828–839. [Google Scholar] [CrossRef]
  32. Rao, Y.F.; Zhang, Y.Y.; Fan, J.H.; Wei, G.L.; Wang, D.; Han, F.M.; Huang, Y.; Croue, J.P. Enhanced peroxymonosulfate activation by Cu-doped LaFeO3 with rich oxygen vacancies: Compound-specific mechanisms. Chem. Eng. J. 2022, 435, 134882. [Google Scholar] [CrossRef]
  33. Zhu, Y.F.; Liu, Y.F.; Miller, K.A.; Zhu, H.Y.; Egap, E. Lead halide perovskite nanocrystals as photocatalysts for PET-RAFT polymerization under visible and near-infrared irradiation. ACS Macro. Lett. 2020, 9, 725–730. [Google Scholar] [CrossRef] [PubMed]
  34. Jin, X.Y.; Ma, K.L.; Gao, H.F. Tunable luminescence and enhanced polar solvent resistance of perovskite nanocrystals achieved by surface-initiated photopolymerization. J. Am. Chem. Soc. 2022, 144, 20411–20420. [Google Scholar] [CrossRef]
  35. Zhu, Y.F.; Liu, Y.F.; Ai, Q.; Gao, G.H.; Yuan, L.; Fang, Q.Y.; Tian, X.Y.; Zhang, X.; Egap, E.; Ajayan, P.M.; et al. In situ synthesis of lead-free halide perovskite–COF nanocomposites as photocatalysts for photoinduced polymerization in both organic and aqueous phases. ACS Materials Lett. 2022, 4, 464–471. [Google Scholar] [CrossRef]
  36. Koo, P.L.; Jaafar, N.F.; Yap, P.S.; Oh, W.D. A review on the application of perovskite as peroxymonosulfate activator for organic pollutants removal. J. Environ. Chem. Eng. 2022, 10, 107093. [Google Scholar] [CrossRef]
  37. Courty, P.; Ajot, H.; Marcilly, C.; Delmon, B. Oxydes mixtes ou en solution solide sous forme très divisée obtenus par décomposition thermique de précurseurs amorphes. Powder Technol. 1973, 7, 21–38. [Google Scholar] [CrossRef]
  38. Baythoun, M.S.G.; Sale, F.R. Production of strontium-substituted lanthanum manganite perovskite powder by the amorphous citrate process. J. Mater. Sci. 1982, 17, 2757–2769. [Google Scholar] [CrossRef]
  39. Dacquin, J.P.; Dujardin, C.; Granger, P. Surface reconstruction of supported Pd on LaCoO3: Consequences on the catalytic properties in the decomposition of N2O. J. Catal. 2008, 253, 37–49. [Google Scholar] [CrossRef]
  40. Lisi, L.; Bagnasco, G.; Ciambelli, P.; Rossi, S.D.; Porta, P.; Russo, G.; Turco, M. Perovskite-type oxides: II. Redox properties of LaMn1− xCuxO3 and LaCo1−xCuxO3 and methane catalytic combustion. J. Solid State Chem. 1999, 146, 176–183. [Google Scholar] [CrossRef]
  41. Zhang, R.; Villanueva, A.; Alamdari, H.; Kaliaguine, S. SCR of NO by propene over nanoscale LaMn1−xCuxO3 perovskites. Appl. Catal. A 2006, 307, 85–97. [Google Scholar] [CrossRef]
  42. Kaliaguine, S.; Neste, A.V.; Szabo, V.; Gallot, J.E.; Bassir, M.; Muzychuk, R. Perovskite-type oxides synthesized by reactive grinding: Part I. Preparation and characterization. Appl. Catal. A 2001, 209, 345–358. [Google Scholar] [CrossRef]
  43. Levasseur, B.; Kaliaguine, S. Methanol oxidation on LaBO3 (B = Co, Mn, Fe) perovskite-type catalysts prepared by reactive grinding. Appl. Catal. A 2008, 343, 29–38. [Google Scholar] [CrossRef]
  44. Alamdari, H.; Royer, S. Handbook of Perovskites and Related Mixed Oxides, Chapter 2: Mechanochemistry; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2016; pp. 25–46. [Google Scholar]
  45. Properties and Applications of Perovskite-Type Oxides; Tejuca, L.G.; Fierro, J.L.G. (Eds.) Marcel Dekker: New York, NY, USA, 1992; p. 271. [Google Scholar]
  46. Campagnoli, E.; Tavares, A.; Fabbrini, L.; Rossetti, I.; Dubitsky, Y.A.; Zaopo, A.L. Forni, Effect of preparation method on activity and stability of LaMnO3 and LaCoO3 catalysts for the flameless combustion of methane. Appl. Catal. B Enivron. 2005, 55, 133–139. [Google Scholar] [CrossRef]
  47. Yang, W.; Zhang, R.; Chen, B.; Bion, N.; Duprez, D.; Hou, L.; Zhang, H.; Royer, S. Design of nanocrystalline mixed oxides with improved oxygen mobility: A simple non-aqueous route to nano-LaFeO3 and the consequences on the catalytic oxidation performances. Chem. Comm. 2013, 49, 4923–4925. [Google Scholar] [CrossRef] [PubMed]
  48. Chiarello, G.L.; Rossetti, I.; Forni, L. Flame-spray pyrolysis preparation of perovskites for methane catalytic combustion. J. Catal. 2005, 236, 251–261. [Google Scholar] [CrossRef]
  49. Kleitz, F.; Choi, S.H.; Ryoo, R. Cubic Ia 3 d large mesoporous silica: Synthesis and replication to platinum nanowires, carbon nanorods and carbon nanotubes. Chem. Comm. 2003, 17, 2136–2137. [Google Scholar] [CrossRef] [PubMed]
  50. Xu, J.; Li, J.; Wu, F.; Zhang, Y. Rapid photooxidation of As(III) through surface complexation with nascent colloidal ferric hydroxide. Environ. Sci. Technol. 2014, 48, 272–278. [Google Scholar] [CrossRef]
  51. Zhou, D.; Chen, L.; Zhang, C.; Yu, Y.; Zhang, L.; Wu, F. A novel photochemical system of ferrous sulfite complex: Kinetics and mechanisms of rapid decolorization of Acid Orange 7 in aqueous solutions. Water Res. 2014, 57, 87–95. [Google Scholar] [CrossRef]
  52. Gao, P.; Tian, X.K.; Fu, W.; Wang, Y.X.; Nie, Y.L.; Yang, C.; Deng, Y. Copper in LaMnO3 to promote peroxymonosulfate activation by regulating the reactive oxygen species in sulfamethoxazole degradation. J. Hazard. Mater. 2021, 411, 125163. [Google Scholar] [CrossRef]
  53. Liu, D.D.; Tang, Y.B.; Hao, Z.K.; Chen, D.Q.; Li, T.Q.; Jiang, L.P.; Tian, B.; Yan, C.P.; Luo, Y.; Jia, B.Y. Comparative study of H2O2/PDS-based advanced oxidation process using nanopaticles for Rhodamine B degradation: Mechanism, stability and applicability. J. Water. Pro. Eng. 2022, 47, 102757. [Google Scholar] [CrossRef]
Figure 1. Scheme of the experimental reaction configuration with a LED strip as light source.
Figure 1. Scheme of the experimental reaction configuration with a LED strip as light source.
Nanomaterials 13 01388 g001
Figure 2. XRD results for LaFeO3 samples prepared by the BA and CA routes. Bottom of the figure: vertical lines for reference LaFeO3 (JCPDS card 74-2203).
Figure 2. XRD results for LaFeO3 samples prepared by the BA and CA routes. Bottom of the figure: vertical lines for reference LaFeO3 (JCPDS card 74-2203).
Nanomaterials 13 01388 g002
Figure 3. TEM images of BA catalyst (A,B), and CA catalyst (C,D) samples, after thermal stabilization at 500 °C.
Figure 3. TEM images of BA catalyst (A,B), and CA catalyst (C,D) samples, after thermal stabilization at 500 °C.
Nanomaterials 13 01388 g003
Figure 4. Process combination effect on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, [PS] = 50 mM, and pH0 = 3.7.
Figure 4. Process combination effect on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, [PS] = 50 mM, and pH0 = 3.7.
Nanomaterials 13 01388 g004
Figure 5. Impact of the preparation method on the catalyst efficiency in the visible light/PS/catalyst process for the tetracycline degradation. Conditions: C0 = 100 mg L−1, catalyst (BA or CA) = 0.5 g L−1, [PS] = 50 mmol L−1, and pH0 = 3.7.
Figure 5. Impact of the preparation method on the catalyst efficiency in the visible light/PS/catalyst process for the tetracycline degradation. Conditions: C0 = 100 mg L−1, catalyst (BA or CA) = 0.5 g L−1, [PS] = 50 mmol L−1, and pH0 = 3.7.
Nanomaterials 13 01388 g005
Figure 6. Effect of the BA catalyst dosage on the TC degradation. Conditions: C0 = 100 mg L−1, [PS] = 50 mM, and pH0 = 3.7.
Figure 6. Effect of the BA catalyst dosage on the TC degradation. Conditions: C0 = 100 mg L−1, [PS] = 50 mM, and pH0 = 3.7.
Nanomaterials 13 01388 g006
Figure 7. Effect of the sodium persulfate concentration on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, and pH0 = 3.7.
Figure 7. Effect of the sodium persulfate concentration on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, and pH0 = 3.7.
Nanomaterials 13 01388 g007
Figure 8. Effect of reaction medium initial pH on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, and [PS] = 50 mmol L−1.
Figure 8. Effect of reaction medium initial pH on the tetracycline degradation. Conditions: C0 = 100 mg L−1, [BA] = 0.5 g L−1, and [PS] = 50 mmol L−1.
Nanomaterials 13 01388 g008
Figure 9. Evolution of the BA catalyst activity with the reaction cycles of TC degradation under visible light/PS/catalyst coupled process. Conditions: C0 = 100 mg L−1, [PS] = 50 mmol L−1, [BA] = 0.5 g L−1, and pH0 = 3.7.
Figure 9. Evolution of the BA catalyst activity with the reaction cycles of TC degradation under visible light/PS/catalyst coupled process. Conditions: C0 = 100 mg L−1, [PS] = 50 mmol L−1, [BA] = 0.5 g L−1, and pH0 = 3.7.
Nanomaterials 13 01388 g009
Table 1. Selected physico-chemical characteristics obtained for BA and CA catalysts.
Table 1. Selected physico-chemical characteristics obtained for BA and CA catalysts.
SamplesBA SampleCA Sample
Chemical compositionLaxFeyO3±δLaxFeyO3±δ
DXRD/a nm18.422.3
Sth/b m2 g−149.236.9
SBET/c m2 g−132.515.3
R/d -0.660.41
a, DXRD is the crystal domain size determined using the Scherrer equation after correction of the FWHM for instrumental broadening; b, Sth is the theoretical surface area calculated from the surface area developed by a cubic particle volume of separated single crystal grains of DXRD size, assuming no contact between adjacent particles (no agglomeration); c, SBET is the surface area determined using the B.E.T. equation; and d, R = SBET/Sth, characterize the degree of agglomeration in the material (R = 1, no agglomeration).
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

Hou, L.; Wang, Y.; Zhou, F.; Liu, S.; Fu, L.; Wang, L.; Zhang, C.; Xue, W. A Visible-Light-Enhanced Heterogeneous Photo Degradation of Tetracycline by a Nano-LaFeO3 Catalyst with the Assistance of Persulfate. Nanomaterials 2023, 13, 1388. https://doi.org/10.3390/nano13081388

AMA Style

Hou L, Wang Y, Zhou F, Liu S, Fu L, Wang L, Zhang C, Xue W. A Visible-Light-Enhanced Heterogeneous Photo Degradation of Tetracycline by a Nano-LaFeO3 Catalyst with the Assistance of Persulfate. Nanomaterials. 2023; 13(8):1388. https://doi.org/10.3390/nano13081388

Chicago/Turabian Style

Hou, Liwei, Yanan Wang, Fan Zhou, Shuangyue Liu, Lin Fu, Lei Wang, Changbo Zhang, and Weijie Xue. 2023. "A Visible-Light-Enhanced Heterogeneous Photo Degradation of Tetracycline by a Nano-LaFeO3 Catalyst with the Assistance of Persulfate" Nanomaterials 13, no. 8: 1388. https://doi.org/10.3390/nano13081388

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