2.1. Characterization of the Photocatalysts
The as-synthesized photocatalysts were characterized and their photocatalytic activity was studied by the degradation of LFX and CFX in aqueous solution. The most efficient catalyst in the degradation of LFX, 5%Au@ZnONPs-3%MoS2-1%rGO, was selected to be fully characterized.
The BET surface area of the catalysts is shown in
Table S1. As can be seen, the bare ZnONPs showed a surface area of 67 m
2g
-1, increasing with the incorporation of Au nanoparticles, and MoS
2 and rGO nanosheets. This increase in surface area by the addition of other cocatalysts has been previously described [
14,
15,
16,
17]. The trend of the results shows that the higher percentages of Au and MoS
2, the higher the surface area of the catalyst. The 5%Au@ZnONPs-5%MoS
2-1%rGO composite showed the highest surface area, 151 m
2g
-1, which represents a difference of 84 m
2g
-1 compared to bare ZnONPs.
The precursors of the catalysts were characterized by field emission scanning electron microscopy, FESEM, (see
Figure 1). ZnONPs at different magnifications (
Figure 1a,b) show inhomogeneous particles with diameters ranging from ca. 15 to 20 nm. rGO (
Figure 1c) consists of inhomogeneous particles formed by sheets that pack very close together, with different diameters (ca. 1 μm to 5 μm), similar to what has been observed in other works [
29]. The MoS
2 was previously delaminated by ultrasound treatment in aqueous solution (
Figure 1d). As expected, delaminated MoS
2 exhibits a layered structure with sizes ranging from ca. 1 µm to 6 µm [
30].
Figure 2 shows the high-resolution transmission electron microscopy (HRTEM) images of the precursors.
Figure 2a,b show the ZnONPs after the incorporation of AuNPs. As can be seen there, the ZnONPs are highly crystalline, showing distinct lattice fringes with an interplanar spacing of ca. 0.28 nm (inset of 2b), corresponding to the d-spacing of the [001] planes [
14]. This is consistent with the results shown by the selected area electron diffraction (SAED) in the inset of
Figure 2b, as well as by XRD analysis to be discussed later. The AuNPs also presented high crystallinity, with diameters of ca. 10 nm (
Figure 2a,b ).
Figure 2c corresponds to the rGO micrograph, clearly showing the lattice fringes with an interplanar distance of ca. 0.24 nm, which represent the d-spacing of the [002] planes [
31]. The atomic structure of a highly exfoliated MoS
2 is shown in
Figure 2d. The SAED of MoS
2 (inset of 2d) indicates a high crystalline material. It is possible that due to the ultrasound exfoliation process some structural defects may have been generated. The presence of these defects has not yet been verified, but it could influence the activity of the catalysts.
Figure 3 shows the Raman spectra of rGO, ZnONPs, MoS
2, and the 5%Au@ZnONPs-3%MoS
2-1%rGO catalyst. rGO (
Figure 3a) shows two peaks at 1350 cm
-1 and 1586 cm
-1, corresponding at the D and G bands, respectively, and represent the presence of carbon atom lattice defects and in-plane stretching vibration from sp
2 hybridization of carbon [
32]. The ZnONPs (
Figure 3b) shows different peaks at ca. 327 cm
-1, 437 cm
-1, 550 cm
-1, and 1125 cm
-1. The 327 cm
-1 band is attributed to the second-order Raman spectrum [
14], whereas the 437 cm
-1 band was assigned to the
E2 modes of Zn motion, corresponding to the band characteristic of the wurtzite phase [
14]. The 550 cm
-1 band is assigned to the
E1 mode and usually originates from second-order Raman scattering [
33]. The band at 1125 cm
-1 was assigned to overtones and/or combination bands [
14]. The Raman spectrum of MoS
2 (see
Figure 3c) shows two characteristic bands at ca. 383 cm
-1 and 407 cm
-1, which have been assigned to the E
12g and A
1g modes, respectively [
30], being attributed to the exfoliation process and the formation MoS
2 flakes with few layers [
34,
35]. The 5%Au@ZnONPs-3%MoS
2-1%rGO catalyst (
Figure 3d) only showed the G-band of rGO, possibly due to the low concentration of rGO (1% by weight) in the sample. The catalyst (
Figure 3d) showed the four characteristic bands of ZnONPs, but with lower intensity and some small changes. These differences are attributed to the interaction with the other additives. The presence of the two main MoS
2 bands was also evident in the catalyst. The presence of the most significant peaks of all the catalyst components confirmed the heterostructured nature of the composite.
Figure 4 shows the diffraction pattern of the catalyst 5%Au@ZnONPs-3%MoS
2-1%rGO, along with that of ZnONPs, 5%Au@ZnONPs, MoS
2 and rGO for comparison purposes. The diffraction peaks of ZnONPs (
Figure 4a) can be unambiguously indexed to the ZnO phase of hexagonal wurtzite [
36], whose reflections are dominant in the 5%Au@ZnONPs-3%MoS
2-1%rGO catalyst, as observed in
Figure 4e. Incorporation of 5%AuNPs (
Figure 4b) does not reveal the appearance of a new peak at ca. 38.1°, corresponding to Au (111), possibly due to the high dispersion of the metal [
37]. MoS
2 (
Figure 4c) shows several diffraction peaks at ca. 32°, 36°, 39°, 49°, and 58° that have been ascribed to (100), (102), (103), (105), and (110) crystalline planes of 2H-type MoS
2 hexagonal phase (JCPDS # 75–1539), respectively [
30,
38,
39]. rGO (
Figure 4d) shows a peak at ca. 23.8°, assigned to the (002) crystal plane, indicating that in this reduced material most of the functional groups with oxygen, which are characteristic of graphene oxide, have been removed [
40,
41]. rGO shows a second peak close to 40°, at a slightly lower angle than expected (43°), which has been assigned to the (100) plane of the hexagonal carbon structure.
Figure 4e shows the different diffraction peaks corresponding to the most active catalyst. To facilitate the identification of components, the same color code has been used. As can be seen, the most prominent components are shown in the catalyst although, as observed in 4b, no peak corresponding to gold is observed.
The representative elements of the most efficient catalyst (5%Au@ZnONPs-3%MoS
2-1%rGO) were characterized by X-ray photoelectron spectroscopy (XPS). Zn2p (see
Figure 5a) shows two components at 1044.2 eV and 1020.6 eV, with a characteristic spin-orbit splitting of 23.6 eV, that were ascribed to the Zn2p
1/2 and Zn2p
3/2 transitions of Zn
2+, respectively [
42,
43]. Both transitions are very symmetrical, and the fitting to other possible states of Zn did not give results, so any additional contribution was ruled out.
Figure 5b shows the transition corresponding to O1s. The transition is clearly asymmetric, and it has been possible to deconvolute into three components at ca. 530.1, 532.2 eV and 535.1 eV. The peak at ca. 530.1 eV, was assigned to O
2- species in the ZnO network, and the one observed at 532.2 eV was assigned to O
2- in oxygen-deficient regions, respectively [
44]. Graphene oxide (rGO) should show a component below 530 eV, although this contribution should be masked by the peak at 530.1 eV. The component observed at the highest binding energy (535.1 eV) must correspond to species generated by the interaction of the ZnO nanoparticles with rGO. In fact, O1s components have been observed in rGO at BE above 535 eV, although their origin is not entirely clear [
45].
Figure 5c shows the Au4f transition, with peaks at 83.9 eV and 87.6 eV that have been ascribed to Au4f
7/2 and Au4f
5/2, respectively. Both peaks, together with a characteristic spin-orbit splitting of 3.7 eV, evidence the presence of metallic gold [
46]. The C1s spectrum (
Figure 5d) is quite asymmetric and has been deconvolved into three components at 284.6, 287.2 and 289.2 eV, respectively. The most important contribution is observed at 284.6 eV, and has been ascribed to the sp
2 carbon of rGO. The other two components observed at 287.2 and 289.2 eV can be attributed to C-OH and O=C-OH, respectively, and are possibly due to the presence of structural defects in rGO, produced during the graphite exfoliation process, and subsequent reduction of graphene [
47,
48].
Figure 5e shows the Mo3d and S2s transitions. Mo3d shows two well-defined and symmetrical peaks at 232.1 and 229.0 eV, which have been attributed to the Mo3d
3/2 and Mo3d
5/2 doublet, respectively, and assigned to the Mo
4+ state in MoS
2 [
47,
48]. The observed peak at ca. 226.4 eV corresponds to the contribution of S2s [
48], characteristic of MoS
2.
The different precursors and the most efficient catalyst (5%Au@ZnONPs-3%MoS
2-1%rGO) were characterized by diffuse reflectance spectroscopy. From the reflectance in Kubelka-Munk units, the Tauc plots were obtained, which have allowed establishing the bandgaps of the different precursors and the most efficient catalyst (see
Figure 6). The bandgap energy of the wurtzite crystalline phase of ZnO has been reported to be ca. 3.37 eV [
49], while the synthesized ZnONPs (
Figure 6a) showed a value of 3.24 eV. This slight difference could be due to the morphology and particle size of the semiconductor [
49]. After the incorporation of 5 wt.% of AuNPs onto the bare ZnONPs, the bandgap energy decreased to 3.19 eV (
Figure 6b). This reduction was expected and has been previously reported [
13,
14,
15]. However, even when the bandgap is decreased, it is still in the ultraviolet region of the electromagnetic spectrum [
15]. The MoS
2 bandgap energy showed a value of 2.45 eV (
Figure 6c). Depending on the degree of delamination, and the number of layers of the material, this value can vary widely from ca. 0.9 eV to values above 2.50 eV [
50]. This effect has also been observed to depend on the degree of quantum confinement of the material along the c axis [
50]. The rGO bandgap energy was 2.00 eV (
Figure 6e), although, as previously described in other investigations, the level of reduction can greatly affect this value, with bandgaps ranging from ca. 0.20 eV to 2.00 eV [
51]. Both MoS
2 and rGO can absorb in the visible region, which undoubtedly improves the catalytic properties of the composite, promoting energy absorption in the visible range. The bandgap energy of the most efficient catalyst (5%Au@ZnONPs-3%MoS
2-1%rGO) was 2.15 eV (
Figure 6d), indicating that the catalyst can efficiently use visible light for catalytic processes, as confirmed in the degradation studies that will be discussed later.
2.2. Photodegradation of Levofloxacin and Ciprofloxacin
The ideals conditions in terms of antibiotic concentration, catalyst loading, and pH were determined for LFX, using the 5%Au@ZnONPs-1%MoS
2-1%rGO, 5%Au@ZnONPs-3%MoS
2-1%rGO, and 5%Au@ ZnONPs-5%MoS
2-1%rGO catalysts, and the results can be found in the supplementary information section. In the case of CFX, these conditions have been previously determined in other studies by our research group [
17,
52].
It is reported [
18] that the concentration of the antibiotic must be in a suitable range to improve the interaction with the active sites of the catalyst, avoiding the interaction between the LFX molecules, which could decrease the degradation rate. For that reason, the first parameter studied was the evaluation of the initial concentration of LFX (see
Figure S1). To carry out the experiments, the catalyst loading, and pH of the solution was set to be 1.0 g/L and 7, respectively. As observed, the concentration of LFX varied between 2 μM – 50 μM, being 10 μM the concentration at which the three evaluated catalysts obtained the highest percentage of degradation. From 2 μM – 10 μM the photocatalytic activity increased, suggesting a good interaction between the catalyst and LFX. After 10 μM, the concentration began to decrease, and this was associated to the byproducts formed during the photodegradation. These byproducts can compete for the catalyst's active sites, reducing its efficiency. Also, at higher concentrations, the surface of the catalysts can become saturated leading to a decrease in the degradation rate [
18]. The catalyst with the highest percentage of degradation at 10 μM was 5%Au@ZnONPs-3%MoS
2-1%rGO (99.8%), followed by 5%Au@ZnONPs-5%MoS
2-1%rGO (98%), and 5%Au@ ZnONPs-1%MoS
2-1%rGO (95%), respectively.
The second parameter studied for LFX was the catalyst loading (see
Figure S2). The LFX concentration used to carry out the experiment was 10 μM at pH=7. The amount of the catalysts varied between 0.2 g/L and 1.5 g/L, being 1.1 g/L the catalyst loading with the best results for the three composites tested. It is suggested that between 0.2 g/L and 1.1 g/L the interaction between LFX and the catalyst is enhanced, but at loadings over 1.1 g/L a decrease is observed. This might be associated to the poor interaction between LFX and the catalyst and/or the scattering of the irradiation due to the high amount of catalyst present in the solution [
18]. The catalyst with the higher percentage of degradation was 5%Au@ZnONPs-3%MoS
2-1%rGO (99.8%), followed by 5%Au@ZnONPs-1%MoS
2-1%rGO (97.5%), and 5%Au@ ZnONPs-5%MoS
2-1%rGO (94.8%), respectively.
The third parameter studied was pH (
Figure S3). To carry out this experiment, only the highest efficiency catalyst (5%Au@ZnONPs-3%MoS
2-1%rGO) was used, with a catalyst loading of 1.1 g/L and a concentration of 10 μM of LFX. As can be seen, the pH varied between 4 and 11, with pH=8 being the one with the highest percentage of degradation. At acidic pH (< 7), the photocatalytic process was not favorable and could be attributed to the repulsive forces between the positively charged ZnO surface and the LFX molecules [
53]. At alkaline pH (>7), the photodegradation rate increased until pH=8, and then slowly decreased until reaching pH=9 and 11, suggesting attractive forces between the positively charged ZnO and the
-OH ions present in the solution [
53,
54]. According to the experiments carried out, the ideal conditions for the catalytic photodegradation of LFX consisted of a 10 μM LFX concentration at pH=8 and with a catalyst loading of 1.1 g/L. For the CFX photodegradation reaction, the ideal conditions were 10 µM of CFX at pH 7 with a catalyst loading of 1.0 g/L.
Figure 7 shows the rate of catalytic photodegradation of CFX and LFX as a function of time. The photodegradation behavior of both antibiotics was very different depending on the catalyst. In the case of CFX (
Figure 7a–c), it was observed that after 60 min the degradation varied from 70% to 96%. The three catalysts with the highest degradation percentages were 1%Au@ZnONPs-3%MoS
2-1%rGO (96%), followed by 3%Au@ZnONPs-3%MoS
2-1%rGO (90%), and 1 %Au@ZnONPs-1%MoS
2-1%rGO (86%). The catalyst that produced less degradation (70%) was 5%Au@ZnONPs-5%MoS
2-1%rGO. The observed results indicate that at higher percentages of AuNPs and MoS
2, the degradation decreases. A possible explanation for this behavior could be that at higher loadings, AuNPs and MoS
2 nanosheets can cause a slight scattering of radiation. On the other hand, the agglomeration of the particles could be improving the catalyst-catalyst contact instead of favoring the catalyst-CFX contact, which would imply a decrease in activity. [
18]. In the case of LFX (
Figure 7d–f), experiments were performed for 120 min instead of 60 min as done with CFX. This decision was made when the percentages of degradation of both antibiotics were compared at 60 minutes. In that time range, all catalysts could degrade 70-96% CFX, but only 65-80% LFX. This difference suggests that CFX degrades faster while LFX is more recalcitrant to degrade. Similar results have been reported before [
17], and are attributed to factors such as pH and the chemical structure of the antibiotics. The catalyst with the highest percentage of LFX degradation during a reaction time of 120 min was 5%Au@ZnONPs-3%MoS
2-1%rGO (99.8%), followed by 5%Au@ZnONPs-5%MoS
2- 1%rGO (99%), and 3%Au@ZnONPs-3%MoS
2-1%rGO (98%). The catalyst with the lowest percentage of degradation was 1%Au@ZnONPs-1%MoS
2-1%rGO, (89%). Unlike what was observed in CFX, LFX degradation was favored with high percentages of both AuNPs and MoS
2. It is possible that for LFX the high percentages of AuNPs and MoS
2 improve the catalyst-LFX contact by creating new active sites for photocatalytic activity. Other research groups [
3] have reported similar results with high percentages of other cocatalysts.
To study the effect and contribution of parameters such as catalyst, radiation and presence of oxygen, control experiments were carried out for the antibiotic LFX (
Figure S4). For the anoxic experiments, the solution was purged with nitrogen gas (N
2) for 180 min, whereas the photolytic experiments were performed without the presence of the catalyst (5%Au@ZnONPs-3%MoS
2-1%rGO). For catalytic experiments, the radiation source was removed, and the solution was kept in the dark. As can be seen (
Figure S4), LFX degradation is negligible when the oxygen source (anoxic), radiation source (photolysis) and catalyst (catalysis) are removed from the system. Without the oxygen source, radicals are not formed and oxidation and ultimately degradation of LFX does not occur. Without radiation sources, there is no activation of the catalyst, so electron-hole pairs (e
-/h
+) will not be generated, and degradation cannot continue. If the catalyst is removed, the degradation does not continue as the radiation source is not sufficient to degrade the LFX molecules. The stability of this fluoroquinolone in water is consistent with studies where CFX and LFX have been detected in surface waters [
7,
8].
To study the recyclability of the most efficient catalysts for the degradation of LFX (5%Au@ZnONPs-3%MoS
2-1%rGO), and CFX (1%Au@ZnONPs-3%MoS
2-1%rGO), 15 cycles were performed (
Figure S5). The experiments consisted of recovering the catalyst after each cycle by centrifugation (3000 rpm for 20 min), followed by washing with deionized water and ethanol to remove any byproducts, and drying for 5 hours at 60 °C. After drying the catalyst, the same parameters were used for antibiotic degradation, as discussed above. As can be seen (
Figure S5a), the degradation of LFX after 15 cycles experienced a low decrease in efficiency, until reaching a degradation percentage of 92.8 %, which represents a decrease in activity close to 7 %. In the case of CFX, the loss of efficiency is greater, reaching a degradation percentage close to 85.7 % after 15 cycles of use. This difference could eventually be associated with the different loading of gold in both catalysts. The most efficient catalyst for CFX photodegradation has only 1 % AuNPs, so a possible gold leaching could have a much greater effect (as observed in
Figure S5b), compared to what happens with LFX, whose most efficient catalyst has 5%AuNPs (
Figure S5a). The results obtained are certainly promising, taking into account that the catalysts are made up of three components whose synergistic behavior is maintained with few variations after each cycle of use. Furthermore, these results suggest that the catalysts could be used for longer cycles without excessively compromising degradation efficiency.
In semiconductor catalysis, electron-hole pairs (e
-/h
+) are formed when an electron leaves the valence band (VB) of the semiconductor and reaches the conduction band (CB) [
13,
14,
15,
16,
17,
18,
19,
20]. The hole (h
+) that is formed acts as an oxidizing agent and can degrade substances that are prone to oxidation. If the electron that left the VB, for some reason, does not reach the CB and returns, recombination occurs. This recombination is one of the main disadvantages of semiconductor photocatalysis [
14,
15,
16,
17]. To reduce the probability of recombination, hole scavengers are often used [
19]. The idea behind this is to incorporate a substance that is more susceptible to oxidation than the contaminant of interest. The substance will feel attraction to the h
+ formed in the VB of the catalyst and will be oxidized, reducing the probability of recombination. The excited electrons that reach the CB of the catalyst can be gained by oxygen species, forming superoxide radicals (O
2-) that can act as even stronger oxidizing agents than the valence band’s holes. These radicals then can degrade the antibiotics, leading to a higher degradation percentage. In this sense, and to evaluate the photodegradation mechanism, some scavengers were added to the reaction mixture: tert-butanol (t-butanol), 1,4-benzoquinone (1,4-BQ) and disodium salt of ethylenediaminetetraacetic acid, EDTA-Na
2 (
Figure S6). t-Butanol, 1,4-BQ and EDTA-Na
2, were adopted as hydroxyl radical (
•OH), superoxide radical (
•O
2−), and h
+ scavenger, respectively. As can be seen, t-butanol hindered photoactivity noticeably, suggesting the main role of
•OH reactive species in the photodegradation process. 1,4-BQ hindered the reaction, although clearly to a lesser extent, which supports the fact that ·O
2− does not play as prominent a role as the hydroxyl radical in the degradation process. Finally, the presence of EDTA practically did not affect the reaction, thus it is evident that the holes generated during the catalytic process do not intervene in the photodegradation of LFX. Similar effects were observed for CFX (see
Figure S6b), although in this case the effects of all scavengers were certainly greater.
The intermediates of the photodegradation of LFX and CFX by the catalysts that showed higher efficiency (5%Au@ZnONPs-3%MoS
2-1%rGO and 1%Au@ZnONPs-3%MoS
2-1%rGO, respectively), were characterized by GC-MS. Based on these results, a possible degradation pathway has been established (see
Figure 8). For both contaminants, photodegradation was very fast. In the case of LFX (
Figure 8a), three different and simultaneous degradation pathways have been suggested. Pathway 1 involves a first piperazin ring cleavage (m/z 308), followed by the loss of methyl groups (m/z 279) [
56,
57], and then mineralization. Pathway 2 shows a first degradation stage similar to 1 (m/z 279) [
56,
57], followed by decarboxylation and subsequent mineralization. In pathway 3, a depiperazinylation and defluorination is observed, for subsequent mineralization. In the case of CFX, photodegradation was even faster than in LFX, generating different compounds that have allowed establishing four possible pathways (
Figure 8b). The first three pathways are characterized by simultaneous depiperazinylation and defluorination, leading to various intermediates (m/z: 284, 216 and 205) [
58], with subsequent mineralization. Pathway 3 also experiences the loss of the cyclopropane ring. Pathway 4 shows the piperazin ring cleavage (m/z 220), followed by the cleavage of the moiety corresponding to the heterocycle with nitrogen and the cyclopropane ring (m/z 141), to continue with subsequent mineralization.
2.3. Mechanism for the Photodegradation of Levofloxacin and Ciprofloxacin
A possible mechanism of the catalytic photodegradation of LFX and CFX is shown in
Figure 9. The band edge position and the migration direction of the photogenerated charge carriers were determined by the bandgap energies (
Figure 6) and the Mulliken electronegativity theory [
59], using the following equations:
E
CB is the edge potential of the conduction band, X is the absolute electronegativity, E
C has a value of 4.50 eV and corresponds to the energy of free electrons on the hydrogen scale [
60,
61], E
g is the bandgap energy, and E
VB is the edge potential of the valence band. The absolute electronegativity for ZnO and MoS
2 are 5.75 eV and 5.32 eV, respectively, whereas the E
CB and E
VB edge potentials for Au@ZnO were -0.37 eV and 2.87 eV. In the case of MoS
2, the edge potentials were -0.405 eV (E
CB) and 2.045 eV (E
VB). These values are similar to those reported by other research groups [
17,
62].
Under visible light, photons do not have enough energy to remove an electron from the VB of ZnO, so the route of degradation relies, mainly, on the other components of the system. In the case of MoS
2, visible light is capable of removing electrons from the VB to the CB. Those electrons have enough energy to reduce oxygen (O
2) molecules into superoxide radicals (
•O
2-), which in turn react with water to form hydroxyl radicals. Hydroxyl and superoxide radicals can oxidize and degrade CFX and LFX. Since MoS
2 has a more negative E
CB potential edge (-0.405 eV) than Au@ZnONPs (-0.37 eV), photoexcited electrons from the VB of MoS
2 can be injected into the CB of Au@ZnONPs. Once there, O
2 can gain those electrons by forming
•O
2- and hydroxyl radicals, which in turn can oxidize and degrade antibiotics. Multiples studies [
63,
64,
65] have reported that the AuNPs can act as an electron sink, reducing the recombination of e
-/h
+ and providing active sites for the catalytic processes. Another advantage of using AuNPs is that under visible light (~580 nm depending on the AuNPs particle size) the phenomenon of surface plasmon resonance (SPR) occurs [
63,
64,
65]. The SPR provides photoexcited electrons with enough energy to reduce the O
2 molecules into
•O
2- radicals. Reduced graphene oxide (rGO) has a bandgap of 2.0 eV (
Figure 6), allowing the use of visible light to form h
+ and photoexcited electrons capable of producing superoxide and hydroxyl radicals. In addition, rGO has a high surface area which allows to create active sites for the photocatalytic process.