Next Article in Journal
Enzymatic Hydrolysis of Softwood Derived Paper Sludge by an In Vitro Recombinant Cellulase Cocktail for the Production of Fermentable Sugars
Next Article in Special Issue
Photocatalytic Degradation of Quinoline Yellow over Ag3PO4
Previous Article in Journal
Recent Progress with Pincer Transition Metal Catalysts for Sustainability
Previous Article in Special Issue
Characterization and Comparative Performance of TiO2 Photocatalysts on 6-Mercaptopurine Degradation by Solar Heterogeneous Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photodegradation of Synthetic Dyes Mediated by Ag3PO4-Based Semiconductors under Visible Light Irradiation

Instituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas, Avda. de los Naranjos s/n, E-46022 Valencia, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(7), 774; https://doi.org/10.3390/catal10070774
Submission received: 30 May 2020 / Revised: 6 July 2020 / Accepted: 9 July 2020 / Published: 11 July 2020
(This article belongs to the Special Issue Understanding the Molecular Mechanisms of Photocatalysis)

Abstract

:
Four silver phosphate-based materials were successfully synthesized, characterized, and evaluated, together with TiO2, in the photodegradation of synthetic dyes (tartrazine, Orange II, rhodamine, and Brilliant Blue FCF) under two irradiation sources centered at 420 and 450 nm. Scanning Electron Microscopy (SEM) images showed different topologies of the synthesized materials, whereas diffuse reflectance spectra demonstrated that they display absorption up to 500 nm. Degradation experiments were performed in parallel with the silver materials and TiO2. Upon irradiation centered at 420 nm, the abatement of the dyes was slightly more efficient in the case of TiO2—except for Orange II. Nevertheless, upon irradiation centered at 450 nm, TiO2 demonstrated complete inefficiency and silver phosphates accomplished the complete abatement of the dyes—except for Brilliant Blue FCF. A careful analysis of the achieved degradation of dyes revealed that the main reaction mechanism involves electron transfer to the photogenerated holes in the valence band of silver photocatalysts, together with the direct excitation of dyes and the subsequent formation of reactive species. The performance of TiO2 was only comparable at the shorter wavelength when hydroxyl radicals could be formed; however, it could not compete under irradiation at 450 nm since the formed superoxide anion is not as reactive as hydroxyl radicals.

Graphical Abstract

1. Introduction

The availability of drinking water of good quality is essential to human development. However, there is a stress on natural water resources associated with the increasing population trend that could be mitigated by adopting more sustainable policies to minimize water pollution, such as managing waste effectively. In parallel, more efficient treatments for wastewater are still desirable and constitute a current demand [1,2,3,4]. Synthetic dyes are among the contaminants that are recalcitrant to conventional wastewater treatments [5], and although most of them are not intrinsically toxic, they pose problems to water-based ecosystems because they remain fairly stable under visible light, acting as light filters that would otherwise be used by algae and other aquatic plants [6]. Moreover, they exhibit a huge variety of chemical structures that make their degradation challenging [7].
Advanced oxidation processes such as photo-Fenton and the use of the heterogeneous semiconductor TiO2 are among the photocatalytic approaches that have been tested to produce the abatement of these contaminants [8,9]. In both cases, under solar light irradiation, the highly oxidizing hydroxyl radical is produced, which subsequently degrades most organic compounds, regardless of their chemical structure [10]. However, several disadvantages limit the practical applications of TiO2. Among them are its large intrinsic band-gap (Eg of approximately 3.2 eV), which limits the use of solar light to the UV region (only approximately 4% of solar radiation), and the energy waste associated with fast recombination of the photogenerated free e in the conduction band (CB) and the hole (h+) in the valence band (VB) [11]. Thus, different attempts have been made to enlarge the capability of TiO2 to absorb visible light, such as preparing hybrids with other metals [8,12,13,14,15]. Moreover, the synthesis and use of other semiconductors has also been reviewed [16,17,18].
On the other hand, the potential of a semiconductor to produce a hydroxyl radical from the h+ photogenerated in the VB depends, among other things, on the potential value. When such a value is below the redox potential of the hydroxyl radical (E° OH/H2O = +2.27 V vs. normal hydrogen electrode (NHE)) [9], the generation of this intermediate is not thermodynamically feasible; however, h+ (VB) could be oxidizing enough to produce the abatement of the pollutants through direct electron transfer.
Moreover, if the semiconductor is able to absorb visible light, it could be even more efficient than TiO2 at producing the solar photodegradation of dyes in wastewaters. In this sense, the recently synthesized Ag3PO4 appears to be a suitable alternative, since it can absorb solar light of wavelengths shorter than 530 nm as a result of its indirect band-gap of 2.36 eV and direct transition of 2.45 eV [19,20]. Interestingly, its highly dispersive conduction and valence bands result in an unlikely recombination of electron-hole pairs. The band edge potential values determined for the CB and VB of Ag3PO4 are 0.45 and 2.9 eV, respectively [19,21]. Thus, the formation of OH· radicals from the holes photogenerated in its VB is thermodynamically feasible, whereas the formation of superoxide anions is thermodynamically disfavored because E° (O2/O2· = −0.33 V vs. NHE) [22] is more negative than the CB of Ag3PO4. Furthermore, the CB e of Ag3PO4 could reduce O2 to H2O2 through a two-electron reduction E° (O2/H2O2 = +0.87 V vs. NHE) [22]. Nonetheless, being thermodynamically feasible does not guarantee that a reaction will proceed because kinetics also play a crucial role [23].
In the literature, controversial results have been reported on the generation of hydroxyl radicals and superoxide radical anions upon the illumination of Ag3PO4 in the photocatalytic degradation of a variety of pollutants [24,25,26,27,28,29,30,31,32,33,34,35]. Nevertheless, excited Ag3PO4 could still act as an oxidant via h+ (VB).
Furthermore, new composites such as Ag3PO4/TiO2 (3/97, wt.% ratio) have been demonstrated to be efficient in the photodegradation of p-nitrophenol [36] and in the decomposition of isopropanol by visible light photosensitization of Ag3PO4, probably as a result of the hole transfer from Ag3PO4 to TiO2 and subsequent hydroxyl radical generation, although a rigorous study to evaluate the contribution of the different reactive species to the degradation of isopropanol has not been performed [37].
With this background, in the present paper, we intended to compare the performance of OH· vs. h+ (VB) in the photodegradation of a variety of synthetic dyes. With this purpose, we prepared Ag3PO4 with different morphologies and used them together with commercial TiO2 and a hybrid Ag3PO4–TiO2 to compare their efficiency in the photodegradation of a variety of dyes using artificial light. The selection of dyes included tartrazine, Orange II [38], rhodamine B [39] and Brilliant Blue FCF (Table 1; Figure 1 and Figure S1). Tartrazine and Orange II constitute two examples of azo dyes that have been mainly used as color additives in food and beverages; rhodamine B is a synthetic xanthene-like dye with wide applications in biology as a tracer, and Brilliant Blue FCF belongs to the family of triphenylmethane dyes and is commonly used to color beverages, dairy products, powders, and syrups.

2. Results and Discussion

2.1. Characterization

Four different photocatalysts were prepared following slight modifications of previously reported protocols. The morphology of the synthesized materials was determined by means of SEM (Figure 2). Figure 2a shows cubic Ag3PO4 in a variation in magnification (2600×, top left, to 28,200×, bottom right). Regular cubes are seen, and these have equidistant edges and flat faces with an average diameter of 1360 nm [20]. Figure 2b shows the rhombic dodecahedron Ag3PO4 in a range of magnifications (6000×, top left, to 34,700×, bottom right). Though irregular, a common rhombic dodecahedron shape can be seen throughout, which accords with previously reported composites. The images show a smaller average diameter than those previously described (a mean diameter of 340 nm vs. reported 600 nm) and a higher polydispersity [20]. Figure 2c illustrates synthesized spherical Ag3PO4 catalysts with irregular sizes in a variation in magnification from 11,300× (top left) to 90,400× (bottom right). Finally, Figure 2d shows the hybrid Ag3PO4–TiO2 catalyst in a variation in magnification from 3600× (top right) to 56,700× (top left). The bottom image illustrates large Ag3PO4 particles with flat surfaces and smaller TiO2 particles deposited on the surface [48].
The potential of the synthesized catalysts to act as visible-light photocatalysts was initially investigated by recording their diffuse reflectance spectra. The absorbance (F(R)) was obtained by a transformation based on the Kubelka–Munk function (Figure 3a). All Ag3PO4 photocatalysts significantly absorbed in the visible region, with the rhombic dodecahedrons, spheres, and hybrid heterostructures absorbing up to 500 nm and the cubes extending the absorption by up to 525 nm. They showed a clear improvement in the absorbance of the solar spectrum when compared to TiO2, which has a measured upper absorbance limit of 385 nm. The band-gap values were determined from the intercepts of the Tauc plots obtained by plotting (F(R) * h)1/2 versus photon energy (h) in eV (Figure 3b). The determined indirect band-gap values of Ag3PO4 were slightly dependent on the morphologies: 2.32, 2.35, and 2.26 eV for the cubes, rhombic dodecahedrons, and spheres, respectively. The hybrid Ag3PO4–TiO2 catalyst gave two values: 3.09 and 2.28 eV, in agreement with the present two substances; finally, the determined band-gap value for the commercial TiO2 was 3.23 eV. All the experimental values were similar to those reported in the literature for analogous materials [9,19,37].

2.2. Photocatalytic Activity

The photocatalytic performance of the different photocatalysts was evaluated by assessing the degradation of the selected synthetic dyes under artificial irradiation. To compare the efficiency of the h+ vs. the OH· or the injection of an e from the excited states of the dyes into the CB of the photocatalysts, two different irradiation sources were tested: (i) fluorescent lamps with emission centered at 420 nm where all the semiconductors and the azo dyes absorbed and (ii) LEDs centered at 450 nm where only Ag3PO4, Ag3PO4–TiO2 and the azo dyes absorbed, although a small contribution of rhodamine B (RhB) and Brilliant Blue FCF (BB) in the absorption cannot be disregarded. The irradiation times were fixed to ensure that bleaching was not competitive, and controls in the dark were also performed (see Figures S4–S6).
Figure 4 shows the photodegradations achieved upon irradiation centered at 420 nm, and Table 2 contains the pseudo-first order rate constants determined for all the combinations of dye/photocatalyst (see Figure S8 for the linear fittings to a pseudo-first order model). A degradation percentage higher than 90% was achieved in the case of tartrazine (TZ), regardless of the photocatalyst used. The Ag3PO4 cubes showed less efficiency than other Ag3PO4 materials, while TiO2 appeared to be slightly more successful than the rest of the Ag3PO4 catalysts. In the case of Orange II (OII), the observed photodegradation followed a completely different trend, as in this case, Ag3PO4 photocatalysts were much more efficient than TiO2. Finally, in the case of RhB and BB, their photodegradations in the presence of Ag3PO4 were generally much slower than those observed for TZ and OII, while when TiO2 was used, similar values to those found for the azo dyes were determined. Thus, the filter action of the dyes did not appear to have an influence on the performance of TiO2, regardless of the absorbance at 420 nm for OII being approximately one third of that of TZ or the fact that RhB and BB could not absorb/filter light at 420 nm (Table 1 and Figure S1). Analogously, the potential contribution of the direct excitation of the dyes followed by the injection of an e into the CB of TiO2 looked negligible; thus, results pointed to the hydroxyl radical as the oxidant responsible for the degradation of the studied dyes in the photodegradations catalyzed by TiO2 [44,49]. On the contrary, in the case of the Ag3PO4 materials, the values of the pseudo-first order rate constants for the photodegradations of these dyes could be related to their chemical structure, as RhB and BB are more reluctant to oxidation than the TZ and OII azo dyes.
Figure 5 shows the photodegradations achieved upon irradiation centered at 450 nm, and Table 3 contains the pseudo-first order rate constants determined for all the combinations of dye/photocatalyst (see Figure S9 for the linear fittings to a pseudo-first order model). The abatement of the two azo dyes, TZ and OII, catalyzed by the silver materials was very efficient, while TiO2 demonstrated a complete inefficiency at this wavelength for those pollutants. For the case of RhB, TiO2 was only able of produce the abatement of RhB in a 40% yield, while Ag3PO4 was very efficient [21,50]. Lastly, the photodegradation of BB at 450 nm only arrived at 60% abatement when the hybrid was used. In fact, the photodegradation of BB was only successful at percentages higher than 90% when TiO2 was employed at 420 nm. Overall, the performance of the hybrid material was analogous to the rest of the silver-based ones [51,52]. The injection of an e from the excited RhB into the CB of TiO2 appeared as the more likely photodegradation pathway for the case of RhB, while this was not operating in the case of TZ, OII, or BB.

2.3. Formation of Reactive Oxygen Species

Different reactive species including OH·, O2·, and h+ can be involved in heterogeneous photocatalytic oxidation processes. To assess their role in the reaction mechanisms, a variety of scavengers can be used. Sometimes, this may be misleading due to the incomplete species selectivity of the scavengers, but it is the method of choice when the direct detection of the reaction intermediates using time-resolved spectroscopy is not possible. Hence, the effect of 2-propanol (IPA), p-benzoquinone (BQ), and ammonium oxalate (AO) as scavengers of OH·, O2·, and h+, respectively, was investigated in the photodegradation of TZ, AOII, RhB, and BB in an attempt to elucidate the operating reaction mechanisms [21,53,54]. The obtained results for the irradiations performed at 420 and 450 nm are shown in Figure 6 and Figure 7, respectively. As illustrated in Figure 6a, the photodegradation of TZ was remarkably suppressed by the addition of BQ in all cases, while AO only decreased the efficiency of Ag3PO4 and Ag3PO4–TiO2; finally, the effect of IPA was higher for Ag3PO4–TiO2 than for TiO2. The results obtained for OII (Figure 6b) showed that in the case of Ag3PO4 and Ag3PO4–TiO2, the effect of the scavengers was BQ > AO > IPA, while for TiO2, the order was IPA > BQ > AO. Interestingly, in the photodegradation of RhB (Figure 6c) by Ag3PO4, only AO was effective, while BQ >>> AO was observed in the presence of Ag3PO4–TiO2; finally, BQ >> IPA when TiO2 was employed. For BB (Figure 6d), the obtained results indicated that the main scavengers were BQ >>> AO or IPA for Ag3PO4 and Ag3PO4–TiO2, while the order was AO > BQ in the case of TiO2.
The photo-oxidation of TZ at 450 nm (Figure 7a) was mainly inhibited by AO when Ag3PO4 was employed and by BQ >> IPA for Ag3PO4–TiO2. In the abatement of OII (Figure 7b), the most effective scavenger was BQ in all cases, although marginal effects of IPA for Ag3PO4 and Ag3PO4–TiO2 cannot be disregarded, as well as AO for Ag3PO4. In the case of RhB (Figure 7c), the main scavenger was again BQ for all the photocatalysts. Lastly, the abatement of BB by Ag3PO4 and Ag3PO4–TiO2 was affected by both BQ and AO.
Furthermore, the formation of OH· from the photocatalysts upon illumination was evaluated by fluorescence using terephthalic acid, which is known to react directly with OH· to produce 2-hydroxyterephthalic acid, a highly fluorescent product with an emission maximum at 425 nm upon excitation at 315 nm [55,56]. The results are shown in Figure 8a for the Ag3PO4 cubes, in Figure 8b for TiO2, and in Figure S7 for all the catalysts.
Sampling was performed for up to 120 min, and the results presented in Figure 6 clearly show that OH· was not formed in the case of the silver-based photocatalysts, while it was clearly formed in the case of TiO2.

2.4. Photocatalytic Mechanism

The redox potentials of the ground and excited states of the selected dyes are shown in Scheme 1 to illustrate that, in principle, they can be oxidized by the h+ generated in the VB of the photocatalysts and they can also inject an e in the CBs of the semiconductors upon the absorption of visible light of the appropriate wavelength, since both processes are thermodynamically feasible for all these dyes.
The photodegradation achieved upon the irradiation of the dyes in the presence of TiO2 could have been the result of Equation (1) (Scheme 2) followed by the generation of hydroxyl radicals (Equation (4)) that were eventually responsible for the oxidation of dyes (Equation (8)). The direct oxidation of dyes by h+ (VB), as described in Equation (7), would not have competed in this case with the favorable formation of OH·, according to the different concentrations of H2O and dye molecules on the surface of TiO2 and the known efficiency of the formation of OH· by TiO2 [57]. In addition, the absorption of light by the dyes followed by the injection of an electron from their excited states into the CB, and the subsequent oxidation (Equations (9)–(11)) could, in principle, have been operating for the azo dyes upon the illumination centered at 420 or 450 nm or in the case of RhB and BB at 450 nm. However, a careful analysis of the results obtained for the case of TZ and OII at the two wavelengths (Figure 4 and Figure 5) and the effect of the addition of scavengers (Figure 6 and Figure 7) suggested the formation of O2· as a result of Equations (1) and (3) or Equations (9) and (10), together with the formation of OH·(Equation (4)), which was eventually more efficient in producing the abatement of the azo dyes and thus explained that they remained stable at 450 nm. The same was true for the abatement of RhB at 420 nm, while the reactivity observed for this dye at 450 nm cannot be explained by direct irradiation of TiO2; it could only be explained as a result of the processes described in Equations (9)–(11).
Similar values for the pseudo-first order rate constants were determined for all dyes upon irradiation at 420 nm in the presence of TiO2, thus indicating a similar reactivity. This could be mainly attributed to the generation and subsequent reactivity of the hydroxyl radical. Analogously, upon irradiation at 450 nm, the determined values for the pseudo-first order rate constants for all dyes were in the same order of magnitude except for RhB. Thus, the injection of an electron from the excited state of TZ, OII, and BB into the conduction band of TiO2 and subsequent reactivity of the generated radical cations constitute a very inefficient degradation pathway.
The observed photodegradation of dyes in the presence of Ag3PO4 can be explained according to Equations (2) or (9) (Scheme 2), followed by either Equation (7) (since no formation of OH was observed; see Figure 8 and Figure S7) or Equations (10) and (11), although the contribution of the processes described in Equations (5) and (6) could not be disregarded (see the effect of IPA in the reactivity in Figure 7 and Figure 8 for OII and BB at 420 nm, and for OII when irradiation was performed at 450 nm). Nevertheless, the main participation of Equation (7) over the other alternatives in all cases could be demonstrated upon careful analyses of the photodegradation achieved in the cases of RhB and BB. These two dyes were efficiently photodegraded in the presence of Ag3PO4 semiconductors at 420 nm, which is a piece of evidence of the oxidative potential of h+ (VB)Ag3PO4. Though the formation of O2· from the e injected in the (VB)Ag3PO4 was not thermodynamically feasible, experiments in the presence of BQ (Figure 7 and Figure 8) pointed to the potential of the excited dyes to directly generate O2·, as has been already described for dissolved organic matter or drugs under irradiation [58,59,60,61,62].
Though the results obtained at the different wavelengths were not comparable, the participation of the process described in Equations (10) and (11) could not be disregarded for RhB and BB when the photodegradations were performed at 450 nm, where both dyes absorbed.
The performance of the hybrid material was comparable to the rest of the silver-based catalysts, pointing to an efficient e transfer from the dyes to the photogenerated holes in the VBAg3PO4. Nevertheless, the results from the experiments in the presence of scavengers suggested the formation of the less reactive O2·, as well as the marginal generation of hydroxyl radicals.
Overall, the obtained results pointed to a high oxidative potential of h+ (VB)Ag3PO4, comparable to or even higher than that of OH· (photogenerated only in the case of TiO2 upon irradiation centered at 420 nm), although a quantitative analysis with different illumination sources was difficult to make. In other words, the photodegradation of dyes at 450 nm in the presence of Ag3PO4 materials, even though hydroxyl radicals were not generated, appears as an efficient alternative. In this sense, efforts in trying to increase the stability of Ag3PO4 or to investigate of the potential of other semiconductors with the ability to absorb visible light and with high oxidant h+ (VB)—in addition to other efforts focused on doping—could be alternative options to explore TiO2 to enlarge its absorption into the visible region.

3. Materials and Methods

3.1. Chemicals

Tartrazine, Orange II, rhodamine B, Brilliant Blue FCF, AgNO3, CH3CO2Ag, Na2HPO4, Na3PO4.12H2O, TiO2 P25 Degussa, 2-propanol, p-benzoquinone, ammonium oxalate, terephthalic acid, and NH3 were from Sigma Aldrich. The water used in all the experiments was Milli-Q grade. All the reagents of this work were of analytical grade and used without further purifications.

3.2. Preparation of the Photocatalysts

Several Ag3PO4 photocatalysts were prepared with different morphologies following slight modifications of the protocol developed by Bi et al. [20,63]. Moreover, a hybrid Ag3PO4–TiO2 with a molar ratio 3:10 was synthesized using the procedure described by Yao et al. [48]. More specific typical procedures follow:
For the Ag3PO4 rhombic dodecahedrons, 0.8 g of CH3CO2Ag was dissolved in distilled water at 60 °C whilst stirring until a saturated solution was obtained. Then, 32 mL of aqueous Na2HPO4 (0.45 M) were added dropwise at 60 °C. Immediately, a yellow precipitate formed. The mixture was stirred in the dark for one hour at 60 °C to ensure that the reaction was complete, after which it was cooled naturally to room temperature. The product was centrifuged at 6000 rpm for 30 min and then cleaned with distilled water before being re-centrifuged. This cleaning process was repeated three times before leaving the Ag3PO4 rhombic dodecahedrons in a vacuum desiccator until dry.
For the Ag3PO4 cubes, 0.4 g of AgNO3 was dissolved in distilled water at 60 °C whilst stirring to make a saturated solution. Then, aqueous NH3 (0.1 M) was added dropwise until the solution turned transparent. Next, 31.4 mL of aqueous Na2HPO4 (0.45 M) were added dropwise at 60 °C with continuous stirring until a fine yellow precipitate was formed. The mixture was left to stir for an hour to ensure that the reaction was complete, and then it was cooled naturally to room temperature. The crude was centrifuged at 6000 rpm for 30 min and washed twice with distilled water. The wet Ag3PO4 cubes were left in a vacuum desiccator until dry.
For the Ag3PO4 spheres, 0.34 g of AgNO3 was dissolved in distilled water at 60 °C whilst stirring to make a saturated solution. Then, 30 mL of aqueous Na3PO4 (0.07 M) were added dropwise at 60 °C whilst continuously stirring until a yellow precipitate was formed. The solution was left to stir for an hour and then cooled naturally to room temperature. The product was centrifuged at 6000 rpm for 30 min and then cleaned with distilled water before being re-centrifuged. This cleaning process was repeated three times, and the wet Ag3PO4 spheres were left in a vacuum desiccator until dry.
For the Ag3PO4–TiO2 hybrid heterostructures, in a typical procedure, the heterostructures were prepared by dispersing 0.8 g of commercial TiO2 P25 Degussa in 25 mL of distilled water and sonicating for 5 min. Immediately, 1.5 g of AgNO3 was added to the aqueous TiO2 suspension and stirred for 10 min. Next, 25 mL of aqueous Na3PO4 (0.13 M) were added dropwise to the reaction mixture until the suspension turned from white to yellow. The mixture was left to stir for 3 h and centrifuged at 6000 rpm for 30 min. The solid was cleaned with distilled water and re-centrifuged. This cleaning process was repeated three times, and the wet Ag3PO4–TiO2 hybrid was left in a vacuum desiccator until dry.

3.3. Scanning Electron Microscopy Analysis

A Leica EM MED020 high vacuum coater (Leica, Wetzlar, Germany) was used to deposit a fine layer of gold nanoparticles over the sample to avoid charging and modifying the Ag3PO4 topography. Pictures were taken with a Zeiss Ultra 55 field emission scanning electron microscope (FESEM) (Zeiss, Oberkochen, Germany), which has a 1.5 nm resolution when working at 3 kV. Samples were ground using a pestle and mortar to ensure as little agglomeration of nanoparticles as possible. These finely ground particles were connected to a sample holder using double-sided conductive tape to avoid charging of the sample by the electron beam. Examination took place in a vacuum. To avoid bias sampling, several images were taken randomly. The average size of particles was determined by measuring the diameter of a minimum of 50 particles from the random images.

3.4. Photophysical Experiments

The diffuse reflectance of the Ag3PO4 photocatalysts and the hybrid Ag3PO4–TiO2 were recorded using a Cary 5000 from Agilent Technologies (Agilent Technologies, Santa Clara, USA) equipped with an integrating sphere. Barium sulphate (BaSO4) was used as a white standard. Then, the absorption spectra were obtained from the reflectance spectra by means of Kubelka–Munk transformations. A Shimadzu UV-2101PC spectrophotometer (Shimadzu Corporation, Kyoto, Japan) was employed to record the UV/Vis absorption spectra of the food dyes. All the spectra were recorded at room temperature using quartz cells of 1 cm optical path length.
The formation of hydroxyl radicals was evaluated using a Photon Technology International (PTI) LPS-220B spectrometer (Horiba Ltd., Kyoto, Japan). Emissions from solutions were recorded between 350 and 600 nm at different times (λexc = 315 nm). As a matter of fact, the emissions centered at approximately 426 nm corresponding to 2-hydroxyterephthalic acid acted as evidence for the formation of hydroxyl radicals.

3.5. Photocatalytic Degradation of Dyes

Experiments were carried out in parallel to permit direct comparison. Typically, 2.5 mg of photocatalysts were added to 5 mL of an aqueous solution of each dye (5 × 10−5 M). Pyrex test-tubes were stirred in the dark for 30 min to ensure that the adsorption/desorption equilibrium of dye on photocatalyst had been reached prior to irradiation. Irradiation at 420 nm was performed using a Luzchem photoreactor (model LZC-4 V) (Luzchem Research Inc., Ottawa, Canada) equipped with 8 fluorescent lamps emitting at max = 420 nm, while irradiation at 450 nm was performed using a home-made circular photoreactor made with a spiral set-up of 2.5 m strip green LEDs (λem centered at 450 nm), Samsung SMD5630IP20 of 40 W from LEDBOX.
Monitoring the photodegradation of the dyes was based on the absorbance of the aliquots (taken at different times and centrifuged in a Hettich EBA 21 centrifuge) recorded on an Agilent Cary 50 UV–visible spectrophotometer (Agilent Technologies, Santa Clara, USA).
A series of experiments were performed in the presence of IPA (1 × 10−1 M), BQ (1 × 10−4 M), and AO (1 × 10−4 M) as scavengers for OH·, O2·, and h+, respectively, under the experimental conditions described above.
To evaluate the formation of hydroxyl radicals, individual mixtures of each photocatalyst (0.5 g/L) in aqueous solutions containing NaOH (2 × 103 M) and terephthalic acid (5 × 104 M) were irradiated at 450 nm for the Ag3PO4-based photocatalysts and in UV for the TiO2.

4. Conclusions

Silver phosphate photocatalysts have demonstrated that the hydroxyl radical is not the only reactive oxygen species for semiconductor photocatalysts to produce the abatement of synthetic dyes. In fact, the synthesized Ag3PO4 materials have proven successful, even if they are not to be able to generate hydroxyl radicals. This is mainly due to the oxidative potential of the photogenerated holes in the VB, although the formation of superoxide anion directly from the excited dyes has been evidenced in the presence of scavengers, and the direct electron transfer from the excited dyes to the CB of the photocatalysts cannot be disregarded. By contrast, the performance of TiO2 is only comparable at shorter wavelengths due to the generation of hydroxyl radicals; however, it loses efficiency under visible light irradiation since the formed superoxide radical anion is not as efficient as hydroxyl radicals in producing the abatement of dyes. Therefore, visible-light absorbing photocatalysts with an unlikely recombination of electron-hole pairs and an appropriate potential of their CB constitute good alternatives that can be efficiently applied for wastewater remediation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/7/774/s1, Figure S1: Normalized absorbance of tartrazine (pink), Orange II (red), rhodamine B (black) and Brilliant Blue FCF (blue); Figure S2: Emission spectrum of Luzchem photoreactor (model LZC-4 V) with 14 bulbs emitting at 420 nm (data kindly provided by Luzchem Research, Inc.); Figure S3: Emission spectrum of homemade photoreactor emitting at 450 nm; Figure S4: Photolysis of TZ (A), OII (B), RhB (C), and BB (D) under 420 nm light; Figure S5. Photolysis of TZ (A), OII (B), RhB (C), and BB (D) under 450 nm LED light; Figure S6. Dark experiments of TZ (A), OII (B), RhB (C), and BB (D) in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic dodecahedrons (), Ag3PO4 spheres (), hybrid TiO2/Ag3PO4 (), or TiO2 (■); Figure S7: Emission spectra of a solution of terephthalic acid (0.5 mM in 2 mM aqueous NaOH) recorded at different irradiation times in the presence of Ag3PO4 cubes (A), Ag3PO4 rhombic dodecahedrons (B), Ag3PO4 spheres (C), hybrid Ag3PO4–TiO2 (D), or TiO2 (E), (λexc = 315 nm); Figure S8. Plot of the ln of the relative concentration of TZ (a); OII (b), RhB (c) and BB (d) vs time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic (), Ag3PO4 spheres (), hybrid Ag3PO4-TiO2 () or TiO2 (■), under 420 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media; Figure S9. Plot of the ln of the relative concentration of TZ (a); OII (b), RhB (c) and BB (d) vs time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic (), Ag3PO4 spheres (), hybrid Ag3PO4-TiO2 () or TiO2 (■), under 450 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.

Author Contributions

Conceptualization, M.A.M. and M.L.M.; data curation, A.P.; formal analysis, A.P.; funding acquisition, M.A.M. and M.L.M.; investigation, A.P., A.B., and P.F.J.; methodology, A.P., A.B., and P.F.J.; project administration, M.A.M. and M.L.M.; Supervision, M.L.M.; validation, A.P.; writing—original draft, A.P., A.B., and P.F.J.; writing—review and editing, M.A.M. and M.L.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Spanish Government (Grant SEV-2016-0683), Generalitat Valenciana (Prometeo Program) and H2020/Marie Skłodowska-Curie Actions under the AQUAlity project (Reference: 765860).

Acknowledgments

The authors would like to acknowledge H2020/Marie Skłodowska-Curie Actions under the AQUAlity project (Reference: 765860). Consellería d’Educació, Investigació, Cultura i Esport (PROMETEO/2017/075 and GRISOLÍAP/2017/005) is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boczkaj, G.; Fernandes, A. Wastewater treatment by means of advanced oxidation processes at basic pH conditions: A review. Chem. Eng. J. 2017, 320, 608–633. [Google Scholar] [CrossRef]
  2. Miklos, D.B.; Remy, C.; Jekel, M.; Linden, K.G.; Drewes, J.E.; Hubner, U. Evaluation of advanced oxidation processes for water and wastewater treatment—A critical review. Water Res. 2018, 139, 118–131. [Google Scholar] [CrossRef] [PubMed]
  3. Gagol, M.; Przyjazny, A.; Boczkaj, G. Wastewater treatment by means of advanced oxidation processes based on cavitation-A review. Chem. Eng. J. 2018, 338, 599–627. [Google Scholar] [CrossRef]
  4. Rizzo, L. Bioassays as a tool for evaluating advanced oxidation processes in water and wastewater treatment. Water Res. 2011, 45, 4311–4340. [Google Scholar] [CrossRef] [PubMed]
  5. Vaiano, V.; Iervolino, G.; Rizzo, L.; Sannino, D. Advanced Oxidation Processes for the Removal of Food Dyes in Wastewater. Curr. Org. Chem. 2017, 21, 1068–1073. [Google Scholar] [CrossRef]
  6. Fernandez, C.; Larrechi, M.S.; Callao, M.P. An analytical overview of processes for removing organic dyes from wastewater effluents. TrAC Trend Anal. Chem. 2010, 29, 1202–1211. [Google Scholar] [CrossRef]
  7. Gregory, P. Azo dyes-Structure carcinogenicity relationships. Dye. Pigment. 1986, 7, 45–56. [Google Scholar] [CrossRef]
  8. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations: A review. Appl. Catal. B 2004, 49, 1–14. [Google Scholar] [CrossRef]
  9. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C 2000, 1, 1–21. [Google Scholar] [CrossRef]
  10. Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (•OH). Chem. Rev. 2015, 115, 13051–13092. [Google Scholar] [CrossRef]
  11. Mills, A.; Le Hunte, S. An overview of semiconductor photocatalysis. J. Photochem. Photobiol. A 1997, 108, 1–35. [Google Scholar] [CrossRef]
  12. Han, F.; Kambala, V.S.R.; Srinivasan, M.; Rajarathnam, D.; Naidu, R. Tailored titanium dioxide photocatalysts for the degradation of organic dyes in wastewater treatment: A review. Appl. Catal. A 2009, 359, 25–40. [Google Scholar] [CrossRef]
  13. Rauf, M.A.; Meetani, M.A.; Hisaindee, S. An overview on the photocatalytic degradation of azo dyes in the presence of TiO2 doped with selective transition metals. Desalination 2011, 276, 13–27. [Google Scholar] [CrossRef]
  14. Ismael, M. Highly effective ruthenium-doped TiO2 nanoparticles photocatalyst for visible-light-driven photocatalytic hydrogen production. New J. Chem. 2019, 43, 9596–9605. [Google Scholar] [CrossRef]
  15. Ismael, M. Enhanced photocatalytic hydrogen production and degradation of organic pollutants from Fe (III) doped TiO2 nanoparticles. J. Environ. Chem. Eng. 2020, 8, 103676. [Google Scholar] [CrossRef]
  16. Rajeshwar, K.; Osugi, M.E.; Chanmanee, W.; Chenthamarakshan, C.R.; Zanoni, M.V.B.; Kajitvichyanukul, P.; Krishnan-Ayer, R. Heterogeneous photocatalytic treatment of organic dyes in air and aqueous media. J. Photochem. Photobiol. C 2008, 9, 171–192. [Google Scholar] [CrossRef]
  17. Ismael, M.; Elhaddad, E.; Taffa, D.H.; Wark, M. Synthesis of Phase Pure Hexagonal YFeO3 Perovskite as Efficient Visible Light Active Photocatalyst. Catalysts 2017, 7, 326. [Google Scholar] [CrossRef] [Green Version]
  18. Ismael, M.; Wark, M. Perovskite-type LaFeO3: Photoelectrochemical Properties and Photocatalytic Degradation of Organic Pollutants Under Visible Light Irradiation. Catalysts 2019, 9, 342. [Google Scholar] [CrossRef] [Green Version]
  19. Yi, Z.; Ye, J.; Kikugawa, N.; Kako, T.; Ouyang, S.; Stuart-Williams, H.; Withers, R.L. An orthophosphate semiconductor with photooxidation properties under visible-light irradiation. Nat. Mater. 2010, 9, 559–564. [Google Scholar] [CrossRef]
  20. Bi, Y.; Ouyang, S.; Umezawa, N.; Cao, J.; Ye, J. Facet Effect of Single-Crystalline Ag3PO4 Sub-microcrystals on Photocatalytic Properties. J. Am. Chem. Soc. 2011, 133, 6490–6492. [Google Scholar] [CrossRef]
  21. Jinfeng, Z.; Tao, Z. Preparation and characterization of highly efficient and stable visible-light-responsive photocatalyst AgBr/Ag3PO4. J. Nanomater. 2013, 2013, 565074. [Google Scholar] [CrossRef] [Green Version]
  22. Wardman, P. Reduction potentials of one-electron couples involving free-radicals in aqueous solution. J. Phys. Chem. Ref. Data 1989, 18, 1637–1755. [Google Scholar] [CrossRef] [Green Version]
  23. Pitre, S.P.; McTiernan, C.D.; Scaiano, J.C. Understanding the Kinetics and Spectroscopy of Photoredox Catalysis and Transition-Metal-Free Alternatives. Acc. Chem. Res. 2016, 49, 1320–1330. [Google Scholar] [CrossRef]
  24. Cao, J.; Luo, B.; Lin, H.; Xu, B.; Chen, S. Visible light photocatalytic activity enhancement and mechanism of AgBr/Ag3PO4 hybrids for degradation of methyl orange. J. Hazard. Mater. 2012, 217–218, 107–115. [Google Scholar] [CrossRef] [PubMed]
  25. Ge, M.; Zhu, N.; Zhao, Y.; Li, J.; Liu, L. Sunlight-Assisted Degradation of Dye Pollutants in Ag3PO4 Suspension. Ind. Eng. Chem. Res. 2012, 51, 5167–5173. [Google Scholar] [CrossRef]
  26. Ge, M. Photodegradation of rhodamine B and methyl orange by Ag3PO4 catalyst under visible light irradiation. Chin. J. Catal. 2014, 35, 1410–1417. [Google Scholar] [CrossRef]
  27. Qamar, M.; Elsayed, R.B.; Alhooshani, K.R.; Ahmed, M.I.; Bahnemann, D.W. Chemoselective and highly efficient conversion of aromatic alcohols into aldehydes photo-catalyzed by Ag3PO4 in aqueous suspension under simulated sunlight. Catal. Commun. 2015, 58, 34–39. [Google Scholar] [CrossRef]
  28. Taheri, M.E.; Petala, A.; Frontistis, Z.; Mantzavinos, D.; Kondarides, D.I. Fast photocatalytic degradation of bisphenol A by Ag3PO4/TiO2 composites under solar radiation. Catal. Today 2017, 280, 99–107. [Google Scholar] [CrossRef]
  29. Li, X.; Xu, P.; Chen, M.; Zeng, G.; Wang, D.; Chen, F.; Tan, X. Application of silver phosphate-based photocatalysts: Barriers and solutions. Chem. Eng. J. 2019, 366, 339–357. [Google Scholar] [CrossRef]
  30. Zwara, J.; Grabowska, E.; Klimczuk, T.; Lisowski, W.; Zaleska-Medynska, A. Shape-dependent enhanced photocatalytic effect under visible light of Ag3PO4 particles. J. Photochem. Photobiol. A 2018, 367, 240–252. [Google Scholar] [CrossRef]
  31. Petala, A.; Spyrou, D.; Frontistis, Z.; Mantzavinos, D.; Kondarides, D.I. Immobilized Ag3PO4 photocatalyst for micro-pollutants removal in a continuous flow annular photoreactor. Catal. Today 2019, 328, 223–229. [Google Scholar] [CrossRef]
  32. Cruz-Filho, J.F.; Costa, T.M.S.; Lima, M.S.; Silva, L.J.; Santos, R.S.; Cavalcante, L.S.; Luz, G.E. Effect of different synthesis methods on the morphology, optical behavior, and superior photocatalytic performances of Ag3PO4 sub-microcrystals using white-light-emitting diodes. J. Photochem. Photobiol. A 2019, 377, 14–25. [Google Scholar] [CrossRef]
  33. Zhu, C.; Li, Y.; Yang, Y.; Chen, Y.; Yang, Z.; Wang, P.; Feng, W. Influence of operational parameters on photocatalytic decolorization of a cationic azo dye under visible-light in aqueous Ag3PO4. Inorg. Chem. Commun. 2020, 115, 107850. [Google Scholar] [CrossRef]
  34. Raza, N.; Raza, W.; Gul, H.; Azam, M.; Lee, J.; Vikrant, K.; Kim, K.-H. Solar-light-active silver phosphate/titanium dioxide/silica heterostructures for photocatalytic removal of organic dye. J. Clean. Prod. 2020, 254, 120031. [Google Scholar] [CrossRef]
  35. Tab, A.; Bellal, B.; Belabed, C.; Dahmane, M.; Trari, M. Visible light assisted photocatalytic degradation and mineralization of Rhodamine B in aqueous solution by Ag3PO4. Optik 2020, 214, 164858. [Google Scholar] [CrossRef]
  36. Hamrouni, A.; Azzouzi, H.; Rayes, A.; Palmisano, L.; Ceccato, R.; Parrino, F. Enhanced Solar Light Photocatalytic Activity of Ag Doped TiO2-Ag3PO4 Composites. Nanomaterials 2020, 10, 795. [Google Scholar] [CrossRef] [Green Version]
  37. Rawal, S.B.; Sung, S.D.; Lee, W.I. Novel Ag3PO4/TiO2 composites for efficient decomposition of gaseous 2-propanol under visible-light irradiation. Catal. Commun. 2012, 17, 131–135. [Google Scholar] [CrossRef]
  38. Ma, J.; Zou, J.; Li, L.; Yao, C.; Zhang, T.; Li, D. Synthesis and characterization of Ag3PO4 immobilized in bentonite for the sunlight-driven degradation of Orange II. Appl. Catal. B 2013, 134–135, 1–6. [Google Scholar] [CrossRef]
  39. Molla, M.; Tateishi, I.; Furukawa, M.; Katsumata, H.; Suzuki, T.; Knaeco, S. Evaluation of Reaction Mechanism for Photocatalytic Degradation of Dye with Self-Sensitized TiO2 under Visible Light Irradiation. Open J. Inorg. Non-Met. Mater. 2017, 7, 1–7. [Google Scholar]
  40. Trasatti, S. The absolute electrode potential: An explanatory note (Recommendations 1986). Pure Appl. Chem. 1986, 58, 955. [Google Scholar] [CrossRef]
  41. Ahmad, I.; Murtaza, S.; Ahmed, S. Electrochemical and photovoltaic study of sunset yellow and tartrazine dyes. Mon. Chem. 2015, 146, 1631–1640. [Google Scholar] [CrossRef]
  42. Ghoreishi, S.M.; Behpour, M.; Golestaneh, M. Simultaneous voltammetric determination of Brilliant Blue and Tartrazine in real samples at the surface of a multi-walled carbon nanotube paste electrode. Anal. Methods 2011, 3, 2842–2847. [Google Scholar] [CrossRef]
  43. Taniguchi, M.; Lindsey, J.S. Database of Absorption and Fluorescence Spectra of >300 Common Compounds for use in PhotochemCAD. Photochem. Photobiol. 2018, 94, 290–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Vinodgopal, K.; Wynkoop, D.E.; Kamat, P.V. Environmental Photochemistry on Semiconductor Surfaces:  Photosensitized Degradation of a Textile Azo Dye, Acid Orange 7, on TiO2 Particles Using Visible Light. Environ. Sci. Technol. 1996, 30, 1660–1666. [Google Scholar] [CrossRef]
  45. Vinodgopal, K.; Kamat, P.V. Photochemistry of textile azo dyes. Spectral characterization of excited state, reduced and oxidized forms of Acid Orange 7. J. Photochem. Photobiol. A 1994, 83, 141–146. [Google Scholar] [CrossRef]
  46. Romero, N.A.; Nicewicz, D.A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef]
  47. Chebotarev, A.N.; Bevziuk, K.V.; Snigur, D.V.; Bazel, Y.R. The brilliant blue FCF ion-molecular forms in solutions according to the spectrophotometry data. Russ. J. Phys. Chem. A 2017, 91, 1907–1912. [Google Scholar] [CrossRef]
  48. Yao, W.; Zhang, B.; Huang, C.; Ma, C.; Song, X.; Xu, Q. Synthesis and characterization of high efficiency and stable Ag3PO4/TiO2 visible light photocatalyst for the degradation of methylene blue and rhodamine B solutions. J. Mater. Chem. 2012, 22, 4050–4055. [Google Scholar] [CrossRef]
  49. Molla, M.A.I.; Tateishi, I.; Furukawa, M.; Katsumata, H.; Suzuki, T.; Kaneco, S. Photocatalytic Decolorization of Dye with Self-Dye-Sensitization under Fluorescent Light Irradiation. ChemEngineering 2017, 1, 8. [Google Scholar] [CrossRef] [Green Version]
  50. Baiocchi, C.; Brussino, M.C.; Pramauro, E.; Prevot, A.B.; Palmisano, L.; Marci, G. Characterization of methyl orange and its photocatalytic degradation products by HPLC/UV–VIS diode array and atmospheric pressure ionization quadrupole ion trap mass spectrometry. Int. J. Mass Spectrom. 2002, 214, 247–256. [Google Scholar] [CrossRef]
  51. Liu, R.; Hu, P.; Chen, S. Photocatalytic activity of Ag3PO4 nanoparticle/TiO2 nanobelt heterostructures. Appl. Surf. Sci. 2012, 258, 9805–9809. [Google Scholar] [CrossRef]
  52. Wang, P.; Li, Y.; Liu, Z.; Chen, J.; Wu, Y.; Guo, M.; Na, P. In-situ deposition of Ag3PO4 on TiO2 nanosheets dominated by (001) facets for enhanced photocatalytic activities and recyclability. Ceram. Int. 2017, 43, 11588–11595. [Google Scholar] [CrossRef]
  53. Kim, W.J.; Pradhan, D.; Min, B.-K.; Sohn, Y. Adsorption/photocatalytic activity and fundamental natures of BiOCl and BiOClxI1−x prepared in water and ethylene glycol environments, and Ag and Au-doping effects. Appl. Catal. B 2014, 147, 711–725. [Google Scholar] [CrossRef]
  54. Cao, J.; Luo, B.; Lin, H.; Chen, S. Synthesis, characterization and photocatalytic activity of AgBr/H2WO4 composite photocatalyst. J. Mol. Catal. A Chem. 2011, 344, 138–144. [Google Scholar] [CrossRef]
  55. Ishibashi, K.-i.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Detection of active oxidative species in TiO2 photocatalysis using the fluorescence technique. Electrochem. Commun. 2000, 2, 207–210. [Google Scholar] [CrossRef]
  56. Xiao, Q.; Si, Z.; Zhang, J.; Xiao, C.; Tan, X. Photoinduced hydroxyl radical and photocatalytic activity of samarium-doped TiO2 nanocrystalline. J. Hazard. Mat. 2008, 150, 62–67. [Google Scholar] [CrossRef]
  57. Hoffmann, M.R.; Martin, S.T.; Choi, W.Y.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  58. Reszka, K.; Chignell, C.F. Spectroscopic studies of cutaneous photosensitizing agents IV. The photolysis of benoxaprofen, an anti-inflammatory drug with phototoxic properties. Photochem. Photobiol. 1983, 38, 281–291. [Google Scholar] [CrossRef]
  59. Burns, J.M. Methods for reactive oxygen species (ROS) detection in aqueous environments. Aquat. Sci. 2012, 74, 683–734. [Google Scholar] [CrossRef]
  60. Zhang, D.; Yan, S.; Song, W. Photochemically Induced Formation of Reactive Oxygen Species (ROS) from Effluent Organic Matter. Environ. Sci. Technol. 2014, 48, 12645–12653. [Google Scholar] [CrossRef]
  61. Hayyan, M.; Hashim, M.A.; AlNashef, I.M. Superoxide Ion: Generation and Chemical Implications. Chem. Rev. 2016, 116, 3029–3085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. D, N.; Kondamareddy, K.K.; Bin, H.; Lu, D.; Kumar, P.; Dwivedi, R.K.; Fu, D. Enhanced visible light photodegradation activity of RhB/MB from aqueous solution using nanosized novel Fe-Cd co-modified ZnO. Sci. Rep. 2018, 8, 10691. [Google Scholar] [CrossRef] [PubMed]
  63. Bi, Y.; Ouyang, S.; Cao, J.; Ye, J. Facile synthesis of rhombic dodecahedral AgX/Ag3PO4 (X = Cl, Br, I) heterocrystals with enhanced photocatalytic properties and stabilities. Phys. Chem. Chem. Phys. 2011, 13, 10071–10075. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of the selected synthetic dyes.
Figure 1. Chemical structures of the selected synthetic dyes.
Catalysts 10 00774 g001
Figure 2. SEM Images of the synthesized photocatalysts: (a) Cubic Ag3PO4, (b) rhombic dodecahedron Ag3PO4, (c) spherical Ag3PO4, and (d) hybrid Ag3PO4–TiO2.
Figure 2. SEM Images of the synthesized photocatalysts: (a) Cubic Ag3PO4, (b) rhombic dodecahedron Ag3PO4, (c) spherical Ag3PO4, and (d) hybrid Ag3PO4–TiO2.
Catalysts 10 00774 g002
Figure 3. Kubelka–Munk diffusion reflectance UV–visible spectra (a) and Tauc plots for indirect band-gap determination (b) of the synthesized Ag3PO4-based catalysts and the commercial TiO2: cubes (red), rhombic dodecahedrons (blue), spheres (pink), hybrid Ag3PO4–TiO2 (green), and TiO2 (black). Linear fittings to estimate the band-gaps are shown.
Figure 3. Kubelka–Munk diffusion reflectance UV–visible spectra (a) and Tauc plots for indirect band-gap determination (b) of the synthesized Ag3PO4-based catalysts and the commercial TiO2: cubes (red), rhombic dodecahedrons (blue), spheres (pink), hybrid Ag3PO4–TiO2 (green), and TiO2 (black). Linear fittings to estimate the band-gaps are shown.
Catalysts 10 00774 g003
Figure 4. Relative concentration of tartrazine (TZ) (a), Orange II (OII) (b), rhodamine blue (RhB) (c), and Brilliant Blue FCF (BB) (d) vs. time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic dodecahedrons (), Ag3PO4 spheres (), hybrid Ag3PO4–TiO2 (), or TiO2 (■) under 420 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Figure 4. Relative concentration of tartrazine (TZ) (a), Orange II (OII) (b), rhodamine blue (RhB) (c), and Brilliant Blue FCF (BB) (d) vs. time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic dodecahedrons (), Ag3PO4 spheres (), hybrid Ag3PO4–TiO2 (), or TiO2 (■) under 420 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Catalysts 10 00774 g004
Figure 5. Relative concentration of TZ (a), OII (b), RhB (c) and BB (d) vs. time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic dodecahedrons (), Ag3PO4 spheres (), hybrid Ag3PO4–TiO2 (), or TiO2 (■), under 450 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Figure 5. Relative concentration of TZ (a), OII (b), RhB (c) and BB (d) vs. time in the presence of Ag3PO4 cubes (), Ag3PO4 rhombic dodecahedrons (), Ag3PO4 spheres (), hybrid Ag3PO4–TiO2 (), or TiO2 (■), under 450 nm irradiation. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Catalysts 10 00774 g005
Figure 6. Effect of different scavengers on the relative degradation efficiency of TZ (a), OII (b), RhB (c), and BB (d) in the presence of Ag3PO4, hybrid Ag3PO4–TiO2, and TiO2, under 420 nm irradiation. No scavenger is shown in black, 2-propanol (1 × 10−1 M) is shown in green, p-benzoquinone (1 × 10−4 M) is shown in yellow, and ammonium oxalate (1 × 10−4 M) is shown in blue. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Figure 6. Effect of different scavengers on the relative degradation efficiency of TZ (a), OII (b), RhB (c), and BB (d) in the presence of Ag3PO4, hybrid Ag3PO4–TiO2, and TiO2, under 420 nm irradiation. No scavenger is shown in black, 2-propanol (1 × 10−1 M) is shown in green, p-benzoquinone (1 × 10−4 M) is shown in yellow, and ammonium oxalate (1 × 10−4 M) is shown in blue. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media.
Catalysts 10 00774 g006
Figure 7. Effect of different scavengers on the relative degradation efficiency of TZ (a), OII (b), RhB (c), and BB (d) in the presence of Ag3PO4, hybrid Ag3PO4–TiO2, and TiO2, under 450 nm irradiation. No scavenger is shown in black, 2-propanol (1 × 10−1 M) is shown in green, p-benzoquinone (1 × 10−4 M) is shown in yellow, and ammonium oxalate (1 × 10−4 M) is shown in blue. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media. Note that ammonium oxalate was not employed for TiO2.
Figure 7. Effect of different scavengers on the relative degradation efficiency of TZ (a), OII (b), RhB (c), and BB (d) in the presence of Ag3PO4, hybrid Ag3PO4–TiO2, and TiO2, under 450 nm irradiation. No scavenger is shown in black, 2-propanol (1 × 10−1 M) is shown in green, p-benzoquinone (1 × 10−4 M) is shown in yellow, and ammonium oxalate (1 × 10−4 M) is shown in blue. Initial concentrations: 5 × 10−5 M (dye) and 0.5 g/L (photocatalyst) in aerated aqueous media. Note that ammonium oxalate was not employed for TiO2.
Catalysts 10 00774 g007
Figure 8. Emission spectra of a solution of terephthalic acid (0.5 mM in 2 mM aqueous NaOH) and 0.5 g/L of Ag3PO4 cubes (a) or TiO2 (b), recorded at different irradiation times: 0 (black), 15 (red), 45 (blue), 60 (green), and 120 min (pink). (λexc = 315 nm).
Figure 8. Emission spectra of a solution of terephthalic acid (0.5 mM in 2 mM aqueous NaOH) and 0.5 g/L of Ag3PO4 cubes (a) or TiO2 (b), recorded at different irradiation times: 0 (black), 15 (red), 45 (blue), 60 (green), and 120 min (pink). (λexc = 315 nm).
Catalysts 10 00774 g008
Scheme 1. Redox potentials (in V vs. NHE) of the semiconductors and synthetic dyes.
Scheme 1. Redox potentials (in V vs. NHE) of the semiconductors and synthetic dyes.
Catalysts 10 00774 sch001
Scheme 2. Potential steps in the degradation of dyes photocatalyzed by TiO2 or Ag3PO4.
Scheme 2. Potential steps in the degradation of dyes photocatalyzed by TiO2 or Ag3PO4.
Catalysts 10 00774 sch002
Table 1. Photophysical properties and electrochemical data of the selected synthetic dyes [40]. NHE: normal hydrogen electrode.
Table 1. Photophysical properties and electrochemical data of the selected synthetic dyes [40]. NHE: normal hydrogen electrode.
DyeMax (nm)(M−1 cm−1)Eox (dye+/dye, V vs. NHE)E (HOMO–LUMO Gap, eV)Eox * (dye+/1dye *, V vs. NHE)
Tartrazine (TZ)429.5 [41]20,810 [41]+1.25 [42]2.33−1.08 [41]
Orange II (OII)480 [43]15,400 [43]+0.76 [44]2.36 [45]−1.60 [45]
Rhodamine B (RhB)545 [43]106,000 [43]+1.15 [46]2.22 [46]−1.07 [46]
Brilliant Blue FCF (BB)625 [47]97,000 [47]+1.06 [42]1.86−0.80
Table 2. Pseudo-first order constants (min−1) determined for the photodegradations performed under 420 nm irradiation.
Table 2. Pseudo-first order constants (min−1) determined for the photodegradations performed under 420 nm irradiation.
PhotocatalystTZOIIRhBBB
Ag3PO4 cubes6.9 × 10−32.0 × 10−18.1 × 10−38.8 × 10−3
Ag3PO4 rhombic dodecahedrons1.8 × 10−38.5 × 10−26.1 × 10−35.4 × 10−4
Ag3PO4 spheres1.6 × 10−21.3 × 10−15.4 × 10−35.1 × 10−3
TiO2–Ag3PO41.8 × 10−22.0 × 10−14.4 × 10−34.0 × 10−3
TiO21.9 × 10−21.8 × 10−21.9 × 10−21.3 × 10−2
Table 3. Pseudo-first order constants (min−1) determined for the photodegradations performed under 450 nm irradiation.
Table 3. Pseudo-first order constants (min−1) determined for the photodegradations performed under 450 nm irradiation.
PhotocatalystTZOIIRhBBB
Ag3PO4 cubes3.5 × 10−24.1 × 10−14.8 × 10−21.2 × 10−2
Ag3PO4 rhombic dodecahedrons4.3 × 10−22.8 × 10−12.5 × 10−21.7 × 10−2
Ag3PO4 spheres4.4 × 10−23.2 × 10−15.6 × 10−21.0 × 10−2
TiO2–Ag3PO43.4 × 10−23.0 × 10−14.4 × 10−21.6 × 10−2
TiO25.7 × 10−46.7 × 10−44.7 × 10−38.1 × 10−4

Share and Cite

MDPI and ACS Style

Pavanello, A.; Blasco, A.; Johnston, P.F.; Miranda, M.A.; Marin, M.L. Enhanced Photodegradation of Synthetic Dyes Mediated by Ag3PO4-Based Semiconductors under Visible Light Irradiation. Catalysts 2020, 10, 774. https://doi.org/10.3390/catal10070774

AMA Style

Pavanello A, Blasco A, Johnston PF, Miranda MA, Marin ML. Enhanced Photodegradation of Synthetic Dyes Mediated by Ag3PO4-Based Semiconductors under Visible Light Irradiation. Catalysts. 2020; 10(7):774. https://doi.org/10.3390/catal10070774

Chicago/Turabian Style

Pavanello, Alice, Alejandro Blasco, Peter F. Johnston, Miguel A. Miranda, and Maria Luisa Marin. 2020. "Enhanced Photodegradation of Synthetic Dyes Mediated by Ag3PO4-Based Semiconductors under Visible Light Irradiation" Catalysts 10, no. 7: 774. https://doi.org/10.3390/catal10070774

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