Open Access Article
Ricardo J. C.
Fernandes
abcd,
Beatriz D.
Cardoso
e,
Ana Rita
Silva
bc,
Luciana
Pereira
*bc and
Paulo J. G.
Coutinho
ad
aPhysiscs Centre of Minho and Porto Universities (CF-UM-UP), Universitiy of Minho, Campus de Gualtar, 4710-057 Braga, Portugal
bCEB-Centre of Biological Engineering, University of Minho, Campus de Gualtar, 4710-057 Braga, Portugal. E-mail: lucianapereira@ceb.uminho.pt
cLABBELS-Associate Laboratory, Braga/Guimarães, Portugal
dLaPMET-Associate Laboratory, 4169-007 Porto, Portugal
eMechanical Engineering and Resource Sustainability Center (MEtRICs), Mechanical Engineering Department, University of Minho, Campus de Azurém, 4800-058 Guimarães, Portugal
First published on 3rd June 2025
Increased environmental concerns about water pollution and scarcity have driven the development of innovative processes and technologies to address these issues. Photocatalysis using nanomaterials is auspicious, with proven efficiency in degrading different pollutants. Ferrite nanoparticles (MFe2O4) stand out for their dual functionality, photocatalytic activity and potential superparamagnetic behavior, that make them highly appealing for environmental applications, as they facilitate both pollutant degradation and safe recollection, contributing to nanosafety. This work compared two synthesis methods, sol–gel and solvothermal, for synthesizing mixed zinc and copper ferrites (Zn0.5Cu0.5Fe2O4). Characterization results revealed a bandgap suitable for visible light absorption and superparamagnetic properties reaching 37 emu g−1 of saturation magnetization. However, the photocatalytic efficiency of ferrites was minimal due to the recombination of electron–hole pairs (e−/h+). Despite that, the ferrites exhibited a high adsorption capacity for the cationic dye Malachite Green (MG), but photocatalytic degradation was low. To suppress the high e−/h+ recombination, a novel, fast method, was used to incorporate silver onto the ferrite surface. This modification significantly improved the activity of the ferrites, especially in the nanoparticles synthesized by sol–gel, achieving 90% removal. In contrast, ferrites synthesized by the solvothermal method, beside exhibiting a strong capacity to adsorb MG (55.99%), proved to be photocatalytically inefficient. This highlights the distinct functional strengths of the two synthesis methods, with sol–gel favoring photocatalytic activity and solvothermal favoring adsorption. Relevant parameters, including the amount of silver and nanomaterial concentration, were optimized during the study, resulting in a degradation of 61.89%, at a rate of 0.0381 h−1 under visible light. Toxicity tests were also conducted. After irradiation MG samples still exhibited high toxicity towards the bacteria Vibrio fisheri, demonstrating that despite color removal, degradation by-products may still be present causing the toxic effect.
Environmental significanceWater pollution from synthetic dyes poses serious ecological and health risks due to their toxicity, persistence, and resistance to conventional treatment. Addressing this challenge requires efficient, reusable, and sunlight-responsive materials. This study presents magnetic Zn0.5Cu0.5Fe2O4 nanoparticles doped with silver, synthesized via a rapid, low-energy process, as visible-light active photocatalysts. The silver-enhanced sol–gel-derived nanomaterials show strong photocatalytic and adsorption capacities, enabling dye removal even in low-light or dark conditions. Their superparamagnetic behavior ensures easy recovery and reuse, minimizing secondary contamination. These findings support the design of safe, effective, and scalable photocatalysts for the treatment of dye-laden effluents, particularly in the final stages of wastewater treatment, aligning with circular economy principles and advancing sustainable water management strategies. |
The use of nanomaterials as catalysts to accelerate the photocatalytic degradation of dyes is seen as a promising alternative or a complement to the more traditional wastewater treatment approaches. Photocatalysis, is an eco-friendly and efficient method for degrading persistent pollutants like MG in wastewater, using only a catalyst and a light source, without the need for any chemical addition. It allows the breakdown of chemical bonds of organic compounds through the action of the generated reactive species such as photogenerated holes (h+), hydroxyl radicals (OH˙), as well as superoxide anions (O2−).15 It is energy-efficient, selective towards specific contaminants, and easily scalable. Moreover, photocatalysis can be integrated with other advanced treatment methods, enhancing its effectiveness for environmental remediation. However, the efficiency of the photocatalytic process is highly dependent on the light source due to the different bandgaps of the semiconductor materials. For example, nanomaterials based on titanium dioxide (TiO2),16,17 gallium (Ga2O3)18 and indium (In2O3)19,20 exhibit higher efficiency under ultraviolet light. Other nanomaterials, such as ferrites (MFe2O4, M = Ca, Mg, Zn, Cu), are active under less energetic wavelengths, including the visible spectrum, due to their narrow bandgaps.21 Various ferrite-based materials have also been explored for photocatalytic degradation of dyes. For instance, Mg1−xNixFe2O4 mixed ferrites have demonstrated efficient degradation of the dye methylene blue, under visible light, exhibiting adsorption of visible light and significant degradation capacity.22 Similarly, zinc ferrites also demonstrated excellent photocatalytic activity towards the same dye, with a removal efficiency of 96% under visible light.23 These nanomaterials are also chemically inert and stable, exhibit superparamagnetic behaviour, and are highly efficient at degrading a range of pollutants. Their low-cost and versatility make them particularly suitable for industrial applications.24–26 Among the studied ferrites, copper ferrites are especially promising due to their capacity to absorb visible light, a property linked to their narrower bandgap of approximately 1.65 eV.27 The synthesis of mixed ferrites, such as the incorporation of zinc, can enhance the photocatalytic capacity of certain ferrite nanoparticles.28 Structural rearrangements in nanomaterials lead to increase optical, structural, electric and magnetic properties.29 Optimizing the magnetic properties is essential, as it allows for the efficient recovery of nanomaterials after the treatment process, preventing possible additional contamination in the final effluent. Furthermore, magnetic nanomaterials can be easily removed from reactors and reused, reducing operational costs and improving the sustainability of the treatment process.
Regarding photocatalytic activity, ferrites such as copper- and zinc-based nanoparticles face certain limitations, particularly the high recombination rate of e−/h+ pairs.30 Coupling these materials with noble metals, such as silver is an excellent strategy to improve the charge separation efficiency and reducing recombination, thereby improving the photocatalytic activity.31,32 Tsvektov et al.33 demonstrated the preponderance of coupling silver nanoparticles in magnesium ferrites for enhancing photocatalytic degradation of MG. However, no similar positive effect was observed in other ferrites, such as zinc and cobalt ferrites, for the degradation of the same dye.33 Similarly, Palanisamy et al.34 highlighted the critical role of silver in the nanocomposites structure for visible light absorption, which significantly enhanced the photocatalytic degradation of rhodamine B and methylene blue. Recent advances in organic–inorganic composite photocatalysts have demonstrated significant improvements in pollutant degradation efficiency. For example, Zhang et al. (2022)35 prepared a composite photocatalyst, PDSA/BiO2−x, by self-assembling perylene diimide sulfonic acid (PDI) onto BiO2−x. This material showed a remarkable increase in the degradation rates of various pollutants, especially for cationic dyes such as rhodamine B and methylene blue, achieving 21-fold and 18-fold increases, respectively, compared to the single-phase BiO2−x. These findings highlight the potential of tailored composite systems to enhance photocatalytic performance.
In this study, zinc and copper mixed ferrites, specifically ZnFe2O4, CuFe2O4 and Zn0.5Cu0.5Fe2O4, were synthesized by sol–gel (Sol_ZnCu) and solvothermal (SV_ZnCu) methods, with silver subsequently incorporated through a novel reduction methodology at various concentrations. A new and fast synthesis methodology was employed for the coupling of silver onto Zn0.5Cu0.5Fe2O4 nanoparticles (AgZn0.5Cu0.5Fe2O4 nanoparticles), resulting in enhanced photocatalytic performance while preserving key physicochemical properties such as superparamagnetic behavior. The capability of the resulting material to operate efficiently under low energy irradiation, including exclusive visible light, represents also a significant advantage for reducing operational energy costs in practical applications. The potential of these nanomaterials was evaluated via small-scale irradiation assays, using MG dye as a model pollutant. The catalytic efficiency of the nanomaterials synthesized by the two different methods, as well as varying amounts of silver addition, were compared. Additionally, parameters of the photocatalytic process, like the photocatalyst concentration, different irradiation time and wavelength (under UV or visible light) were evaluated.
Zn0.5Cu0.5Fe2O4 prepared by the two methods were calcined under the same conditions (400 °C for 30 minutes). After calcination, the nanomaterials were cleaned with a 1 M hydrochloric acid solution, washed with dimethylsulfoxide (DMSO), and then with ethanol for several cycles. The synthesized Zn0.5Cu0.5Fe2O4 samples were designated as Sol_ZnCu – and SV_ZnCu corresponding to the sol–gel and solvothermal methods, respectively.
:
1, 1
:
2, 1
:
3.5, 1
:
5, or 1
:
10) dispersed in 1 mL of ultrapure water were added drop-by-drop. After that, 500 μL of 1 mol L−1D-glucose in ultra-pure water and 300 μL of 0.1 mol L−1 NaOH were added drop-by-drop. The solution changed from a brownish color, characteristic of ferrites, to black due to the silver reduction at the surface of the nanoparticles. To finalize the process, the nanoparticles were washed repeatedly with water and ethanol for several cycles to remove reaction remains and were placed in an oven for 6 hours at 100 °C. The following samples were synthesized: Sol_AgZn0.5Cu0.5Fe2O4 synthesized by sol–gel and functionalized with silver (S1); SV_AgZn0.5Cu0.5Fe2O4 synthesized by solvothermal and functionalized with silver (S1); and samples with different amounts of silver, i.e., Sol_AgZnCu (1
:
1); Sol_AgZnCu (1
:
2); Sol_AgZnCu (1
:
3.5); Sol_AgZnCu (1
:
5); Sol_AgZnCu (1
:
10).
| (αhv)n = A(hv − Eg) | (1) |
X-ray diffraction (XRD) measurements were made using a PAN'Alytical X'Pert Pro diffractometer (Malvern Panalytical Ltd, Malvern, UK) in a Bragg–Brentano configuration operating with Cu Kα radiation (λ = 0.154060 nm), at the Electron Microscopy Unit of the University of Trás-os-Montes and Alto Douro (UTAD), Vila Real, Portugal. The magnetization measurements were evaluated on an MPMMS3 Superconducting Quantum Interference Device (SQUID) magnetometer Quantum Design MPMS5XL (Quantum Design Inc., San Diego, CA, USA) at IFIMUO (University of Porto, Portugal). The zeta(ζ)-potential values were obtained on a DLS equipment Litesizer™ 500 from Anton-Paar (Anton-Paar GmbH, Graz, Austria) equipped with a laser diode of λ 658 nm. TEM images were obtained using JEOL JEM-1011 high-contrast microscope operating with Cu Kα operating at 100 kV (Centro de Apoio Científico-Technolóxico à investigación (CACTI), Vigo, Spain). The samples were placed under ultrasonication treatment and then deposited on copper grids with carbon and Formvar. Image J was used to analyze the obtained images.
000 rpm for 10 minutes to remove any solid content. A Shimadzu UV-3600 plus UV-vis-NIR Spectrophotometer (Shimadzu Corporation, Kyoto, Japan) was used to assess the color removal. The difference between the initial dye absorption and the remaining value after stabilization in the dark was used to evaluate the adsorption (Fadsorption) onto nanomaterials. The removal under irradiation was calculated by the difference between absorbance at initial irradiation time (t = 0) and at the end of the assay. For the final removal (Fremoval) assessment, the correspondent removal under irradiation was calculated and added to the initial adsorbed dye. The dye degradation rate was determined by applying a first-order kinetics model (eqn (2)–(4).![]() | (2) |
![]() | (3) |
| Ct/C0 = F∞ + (1 − F∞)e−kt | (4) |
Before toxicity analysis, samples were centrifuged (10 min at 10
000 rpm), filtered (Whatman SPARTAN syringe filters, regenerated cellulose, 0.2 μm pore size), and the pH set to values between 6 and 9 with HCl or sodium hydroxide. Salinity and oxygen concentration were tuned to 2% NaCl and 3 mg L−1, respectively. Negative control was prepared with a solution of 2% NaCl, while the positive control was a potassium dichromate solution (105.8 mg L−1).40,41
Luminescence was recorded in relative light units (RLU Sec−1), in kinetic mode using a microplate reader (Biotek® Cytation3, Fisher Scientific, Korea), and luminescence inhibition percentage was calculated after 30 min of contact, as stated in ref. 40–42.
m:1) was adapted to a stoichiometric distribution of Cu and Zn cations across both the tetrahedral and octahedral sites over a spinel structure with an arbitrary inversion degree. The sample obtained from solvothermal method exhibited only the mixed ferrite crystalline phase, while the sol–gel method additionally resulted in the formation of a hematite (CIF file nr 9000139; space group R
c:H) phase, which constitutes 12.0% weight fraction. This new phase is responsible for the additional peaks observed in Fig. 2B. The main results are shown in Table 1.
![]() | ||
| Fig. 2 XRD diffractograms and corresponding Rietveld analysis of obtained Zn/Cu mixed ferrites by solvothermal (A) and sol–gel (B) methods. | ||
| Sample | Ox,y,z1 | i | Phase size (nm) | R P | χ 2 | |
|---|---|---|---|---|---|---|
| Lattice constant (nm) | ||||||
| Zn, Cu ferrites | Hematite | |||||
| a 1, value of Ox,y,z in CIF file 2300615 is 0.2535; i, inversion degree; (*), 3a = 0.50362 and c = 1.3757; Rp, Rietveld refinement pattern fits; χ2, chi-squared value. | ||||||
| Sol_ZnCu | 0.3813 | 0.34 | 26.4 | 49 (*) | 7.6 | 1.43 |
| 0.84112 | ||||||
| SV_ZnCu | 0.3799 | 0.51 | 37.9 | — | 8.4 | 1.23 |
| 0.83891 | ||||||
Reasonable fits were obtained, with RP values of 7.6 and 8.4 for, respectively, the nanoparticles Sol_ZnCu and SV_ZnCu. The included lattice constant for of zinc ferrite in the used CIF file is 0.83910 nm. It correspond to ZnFe2O4 nanoparticles obtained from microwave-assisted growth.46 Using a solvothermal process the synthesis resulted in nanoparticles with a higher lattice parameter of 0.8420 nm. The lattice parameters of the Zn, Cu mixed ferrites follow similar variations with the preparation method, with the sol–gel route originating a higher lattice value. Also, the similar size of Cu2+ and Zn2+ (73 pm vs. 74 pm) makes us expect a very small impact on ferrite crystal structure upon substitution of Zn2+ with Cu2+. Size estimation of the Zn, Cu mixed ferrite was obtained using the size broadening effect implemented by BGMN and resulted in 37.9 nm and 26.4 nm for respectively the SV_ZnCu and Sol_ZnCu samples. The size of the hematite phase in the latter sample was estimated to be 49 nm.
The absorption spectrum of Zn0.5Cu0.5Fe2O4 synthesized by sol–gel and solvothermal methods is represented in Fig. 3. Both synthesized nanomaterials show an absorption that steadily increases with energy in measured spectral region. The results demonstrate that silver alters the spectrum, especially in higher wavelengths. The non-appearance of a plasmonic band of silver could be explained by a thick layer of silver, reflecting light, leading to an apparent constant absorption. By applying a Tauc plot, two slightly different bandgaps were estimated. A direct bandgap of 1.97 eV was calculated for sol–gel synthesized nanomaterials, while the solvothermal method determined a direct bandgap of 1.82 eV. Both results fell within the range of possible values reported in the literature.47 The calculated bandgaps allow efficient visible light absorption.
The nanoparticles' crystal structure, morphology, and size distribution were determined using TEM. The images of sol–gel, solvothermal and silver functionalized nanoparticles elucidate a cubic structure with slightly more rounded nanoparticles in Sol_ZnCu (Fig. 4a–c). The cubic structure seems to be much more well-defined for nanoparticles obtained by solvothermal methodology (Fig. 4d–f). Silver deposition over the nanoparticles is more visible in image g than in image h of Fig. 4, due to lower agglomeration of nanoparticles. Silver is deposited heterogeneously into the surface of the nanoparticles. This can promote the internal flow of electrons and suppress the undesired recombination usually observed in ferrite nanoparticles.
The two methods of synthesis resulted in distinct morphological features. Sol_ZnCu nanoparticles exhibited a more uniform dispersion pattern, although predominantly anchored onto an apparent large structure. In contrast, a large number of SV_ZnCu nanoparticles tend to form large agglomerates, with noticeable clustering among smaller particles and other components. This tendency to aggregate, may negatively impact their performance. Despite thorough post-synthesis washing, residual structures are still evident, as observed in Fig. 4, where larger aggregates appear to serve as deposition sites for the nanoparticles.
For size estimation, the nanoparticles were manually outlined, considering the diameter of a circle with an equivalent area. A size of (24.6 ± 11.22) nm was calculated for Sol_ZnCu, while SV_ZnCu exhibited a slightly higher size (37.2 ± 10.74) nm. These values are compatible with the ones estimate from XRD measurements. The aspect ratio of the nanoparticles was also analysed using a fit to a rectangle for each individually outlined nanoparticle. The aspect ratio is defined as the ratio of the rectangle's length to its width, serving as an indicator of particle elongation: values close to 1 correspond to approximately spherical particles, whereas higher values indicate more elongated or irregular shapes.
The histograms are also shown in Fig. 4(i)–(l). Compared to previous works, ferrite nanoparticle sizes seem similar despite synthesis method and chemical composition alterations. For instance, Zn0.5Ca0.5Fe2O4, synthesised using a similar sol–gel methodology, reaches close average particle size values (15 ± 2) nm,31 while the size was reduced to (10 ± 3 nm) using a coprecipitation method.48 In another study, solvothermal methodology, resulted in smaller size particles when substituting Cu with Mg (17.4 ± 8.0 nm).32 The sol–gel method shows limited variation in particle size with precursor variation.
Fig. 5 shows the magnetic measurements for Zn0.5Cu0.5Fe2O4 and silver functionalized Zn0.5Cu0.5Fe2O4. Incorporating Cu2+ ions into the zinc structure enhanced the magnetic moment, increasing the saturation magnetization of zinc ferrites due to substituting non-magnetic Zn2+ with magnetic Cu2+.49 There is a considerable difference between the Ms values of sol–gel and solvothermal synthesised nanomaterials (Fig. 5 and Table 1). The first one presents a Ms of 32.83 emu g−1, which follows the results of other studies.50,51 On the other hand, the synthesis by solvothermal method leads to a low Ms of the nanomaterials, 2.80 emu g−1. This may be attributed to the inefficiency of the cleaning process (Fig. 4 TEM), the weight of the non-magnetic mass, and/or even the formation of non-magnetic phases, such as zinc and copper oxides, all of them playing a significant role.
The process of silver functionalization in solvothermal nanomaterials, using the described reduction method, increases the Ms. SV_AgZnCu to 18.76 emu g−1, despite the increased mass addition caused by silver, which, theoretically, should decrease the Ms. Due to the reduction reaction at the surface of the nanomaterials, some of the of the reaction remained by-products, mainly unreacted precursors, residual surfactants, and non-magnetic oxide phases, are washed out during the post-synthesis treatment. Specifically, during the silver functionalization step via the described reduction method, Ag+ ions are reduced to metallic Ag0 on the nanoparticle surface. This redox reaction not only leads to the deposition of Ag0 but also promotes the removal of loosely bound or unreacted by-products formed during synthesis, such as residual organics or metal hydroxides, which are non-magnetic in nature. This removal leads to a relative decrease in the non-magnetic mass fraction and consequently leaving the magnetic phase relatively more concentrated. This, in turn, results in an apparent increase in specific magnetization (Ms, emu g−1), as the magnetic response becomes more prominent relative to the reduced overall mass. Silver functionalization under reducing conditions may also promote partial crystallization or reorganization of surface magnetic domains, further contributing to the increased Ms. Sol_ZnCu, contrarily to SV_ZnCu, shows a typical behaviour for silver doping, i.e., a decrease in the Ms attributed to the increase in total mass resulting from silver deposition without significant residue elimination or magnetic phases reordering.
All the nanomaterials presented superparamagnetic behaviour, as evidenced by the ratio between remnant magnetization (Mr) and Ms, which was less than 0.1.51 The small size of the nanoparticles plays a significant role in promoting this behavior.52
The values of ζ-potential obtained for the different synthesized nanomaterials in ultra-pure water (solvent used in the irradiation assays) are presented in Table 2 and in Fig. 6. Zinc and copper ferrites present a typical negative ζ-potential despite variances in ion substitution and synthesis methods.39,53,54 This was confirmed by the negativity ζ-potential for Sol_ZnCu (−24.97 eV) and SV_ZnCu (−17.18 eV). ζ-Potential indicates affinity for adsorbing cationic molecules such as the MG dye. The reasonably high absolute values of the measured ζ-potential indicate a moderate colloidal stability of the prepared nanoparticles in water. Functionalization with silver increased the ζ-potential to (−17.01 ± 0.98) eV for Sol_AgZnCu and (−8.02 ± 0.31) eV for SV_AgZnCu, although both values remained negative (Fig. 6) so that MG should still be adsorbed.
| Nanomaterial | M s (emu g−1) | M r (emu g−1) | C (Oe) | M r/Ms | ζ-Potential (eV) |
|---|---|---|---|---|---|
| Sol_ZnCu | 32.83 | 0.62 | 9.44 | 0.019 | −24.97 ± 0.83 |
| SV_ZnCu | 2.80 | 0.10 | 16.12 | 0.04 | −17.18 ± 0.83 |
| Sol_AgZnCu | 23.10 | 0.08 | 2.18 | 0.003 | −17.01 ± 0.98 |
| SV_Ag_ZnCu | 18.76 | 0.72 | 18.64 | 0.04 | −8.02 ± 0.31 |
| Nanomaterials | Method of silver deposition | Nanomaterial concentration (g L−1) | Light source | Initial adsorption (%) | Removal under irradiation (%) | Irradiation time (h) | Final removal (%) | Final MG concentration (mg L−1) | Rate constant (h−1) |
|---|---|---|---|---|---|---|---|---|---|
| a Removal under these conditions was attributed merely to the adsorption step. | |||||||||
| Photolysis | None | UV | — | 7.62 | 6 | 7.62 | 9.24 | 0.0132 | |
| Sol_ZnCu | n.a. | 0.36 | Visible | 18.51 | 1.03 | 6 | 19.35 | 8.07 | 0.0020 |
| Sol_ZnCu | n.a. | 0.36 | UV | 31.93 | 17.24 | 6 | 43.67 | 5.63 | 0.0299 |
| SV_ZnCu | n.a. | 0.36 | Visible | 43.25 | 4.06 | 6 | 45.55 | 5.44 | 0.0029 |
| SV_ZnCu | n.a. | 0.36 | UV | 35.51 | 4.78 | 6 | 38.59 | 6.14 | 0.0188 |
| Sol_AgZnCu | S1 | 0.36 | Visible | 24.99 | 5.74 | 6 | 29.30 | 7.93 | 0.0074 |
| Sol_AgZnCu | S1 | 0.36 | UV | 36.15 | 85.37 | 6 | 90.66 | 0.93 | 0.1589 |
| SV_AgZnCu | S1 | 0.36 | Visible | 55.99 | 0.00 | 6 | 55.99 | 4.40 | (*)a |
| SV_AgZnCu | S1 | 0.36 | UV | 49.12 | 0.00 | 6 | 49.12 | 5.09 | (*)a |
Sol_AgZnCu (1 : 1) |
S2 | 0.36 | Visible | 16.26 | 0.29 | 6 | 16.50 | 8.35 | 0.0111 |
Sol_AgZnCu (1 : 2) |
S2 | 0.36 | Visible | 31.27 | 7.78 | 6 | 36.61 | 6.34 | 0.0186 |
Sol_AgZnCu (1 : 3.5) |
S2 | 0.36 | Visible | 25.98 | 3.60 | 6 | 28.65 | 7.14 | 0.0282 |
Sol_AgZnCu (1 : 5) |
S2 | 0.36 | Visible | 19.78 | 25.69 | 6 | 40.39 | 5.96 | 0.0665 |
Sol_AgZnCu (1 : 10) |
S2 | 0.36 | Visible | 0 | 3.52 | 6 | 3.52 | 9.65 | 0.0068 |
Sol_AgZnCu (1 : 5) |
S2 | 0.18 | Visible | 0 | 19.55 | 6 | 19.55 | 8.04 | 0.0334 |
Sol_AgZnCu (1 : 5) |
S2 | 0.36 | Visible | 19.79 | 25.69 | 6 | 40.39 | 5.96 | 0.0665 |
Sol_AgZnCu (1 : 5) |
S2 | 0.50 | Visible | 19.49 | 23.39 | 6 | 38.32 | 6.17 | 0.0491 |
Sol_AgZnCu (1 : 5) |
S2 | 1.00 | Visible | 24.26 | 35.40 | 6 | 51.08 | 4.89 | 0.0982 |
Sol_AgZnCu (1 : 5) |
S2 | 0.36 | UV | 21.47 | 63.51 | 12 | 71.35 | 2.87 | 0.1033 |
Sol_AgZnCu (1 : 5) |
S2 | 0.36 | Visible | 18.40 | 30.20 | 18 | 43.05 | 5.69 | 0.0218 |
Sol_AgZnCu (1 : 5) |
S2 | 0.75 | UV | 20.18 | 47.50 | 12 | 58.52 | 4.15 | 0.0730 |
Sol_AgZnCu (1 : 5) |
S2 | 0.75 | Visible | 24.52 | 49.50 | 18 | 61.89 | 3.81 | 0.0381 |
Upon the incorporation of silver (batch S1), an increase in the dispersion and stability of the nanomaterials was observed. An improvement in the photocatalytic activity was obtained with Sol_AgZnCu, with a 10-fold increase in the photodegradation rate (90.66%, 0.1589 h−1) under UV irradiation when compared with Sol_ZnCu. Silver coupling can enhance the photoelectric properties of nanomaterials.54 It acts as an electron trapper, improving charge separation efficiency and reducing recombination, which is typically observed in ferrite nanoparticles. The plasmonic effect of silver can also enhance catalytic performance through the generation of hot electrons. Specifically, under light excitation, the localized surface plasmon resonance of Ag nanoparticles produces energetic hot electrons, which may be transferred into the conduction band of the semiconductor. This process promotes more efficient charge separation and facilitates electron transfer to the catalyst surface, thereby contributing to improved photocatalytic activity. Under visible light, the removal rate was small, but it still was 3.7-fold higher than for particles without silver. Both results indicate that the presence of silver enhances the activity of the nanomaterials. This can be explained by the fact that the incorporation of Ag promotes the generation of reactive species by increasing the internal flow of electrons, thereby inhibiting recombination of charge carriers in photoexcited ferrite. This effect is particularly pronounced under UV irradiation.55,56 On the contrary, the results with SV_AgZnCu sample demonstrated that on SV_ZnCu, silver does not exhibit the same improvements in photocatalytic activity. It appears that silver does not synergize as efficiently with the nanomaterials formed by the solvothermal method as with those synthesized by the sol–gel probably due to the presence of synthesis-related residues on the surface of the SV_ZnCu nanoparticles, which may hinder the effective interaction between silver and the catalyst, thereby reducing its performance. Overall, results elucidate that the nanomaterials synthesized by the sol–gel method presented substantially more solid results.
Adsorption was identified as a key process influencing the performance of the synthesized nanomaterials, with observable variations between batches of the same sample attributed to different levels of nanoparticle aggregation. To ensure the accuracy and reliability of comparisons, it is essential to use nanomaterials from a same batch. Of the two synthesis methods evaluated, the sol–gel method was chosen due to superior performance of nanomaterials synthesised by this method, which was further explored in the subsequent assays. Also, the silver reduction method was optimized (S2), as described in the previous section.
:
Ag (1
:
1, 1
:
2, 1
:
3.5, 1
:
5, 1
:
10) were tested to address the optimal silver content (batch S2). It is known that low silver percentages are insufficient to significantly enhance the photocatalytic process. Conversely, excessive silver amounts can obstruct the nanomaterial surface, thereby inhibiting the catalytic process by blocking ROS formation from the ferrite surface.55 Indeed, the results showed that the optimal silver content ratio for the nanoparticles synthesised in this study was 1
:
5 (Table 2 and Fig. 8). In these 1
:
5 samples, the photocatalytic behaviour was notably enhanced under visible light with a rate constant of 0.0665 h−1 leading to an overall 40.39% MG removal.
Regarding the initial adsorption with various NPs
:
Ag ratios, the nanoparticles exhibited increased adsorption capacity up to a 1
:
2 ratio, likely due to enhanced dispersion of the nanoparticles facilitated by the presence of silver in the structure. However, as the amount of silver increased, exposure of the ferrite surface decreased, thereby suppressing the adsorption capacity of the cationic dye. At a 1
:
10 ratio, the adsorption capacity was entirely suppressed. A 1
:
5 doping ration was selected in the subsequent steps, as it demonstrated the best photocatalytic potential among the tested ratios.
Different concentrations of the nanomaterial Sol_AgZnCu (1
:
5) were tested under visible light to investigate the ideal concentration among the ones tested (Table 2 and Fig. 9). The lowest concentration assessed was 0.18 g L−1, reaching a removal of almost 19.55% (at a rate of 0.0334 h−1). Efficient light penetration enabled a good photon absorption by the nanoparticles, with negligible adsorption (Table 3). Doubling photocatalyst concentration to 0.36 g L−1 resulted in a doubling of the photocatalytic rate (0.0665 h−1) and of the final removal (40.39%). With this concentration, initial adsorption drastically increases due to a greater availability of active sites. As the concentration reached 0.5 g L−1, the final removal obtained was similar to the obtained with 0.36 g L−1, 38.32%, with a similar initial adsorption but at a 1.35-fold lower rate (0.0491 h−1). This decrease in the rate constant may be due to a lower light penetration. The highest concentration tested, 1 g L−1, caused the greater adsorption capacity, 24.26% (Table 3), but light penetration was considerably reduced. The final removal reached 51.08% (at a rate of 0.0982 h−1), representing the best result in terms of MG removal, despite being mainly due to adsorption. Despite the overall improvement, the difference in photocatalytic efficiency between using 0.36 g L−1 and 1 g L−1 of catalyst was only 10%. Therefore, increasing the concentration to 1 g L−1 would raise process financial costs. Consequently, 0.36 g L−1 of Sol_AgZnCu (1
:
5) was considered the optimal concentration among those tested.
:
5). A concentration of 0.75 g L−1 was also tested to determine if a higher nanomaterial concentration in prolonged visible light exposure and under UV would enhance the overall photodegradation. The results are presented in Table 2 and Fig. 10. With 0.36 g L−1, under visible irradiation, increasing time from 6 h to 18 h did not cause significant differences in the final dye removal, and the rate was even lower (Table 3). This behaviour could be due to the suppression of active sites by the degradation products of the dye that remains adsorbed. Increasing the catalyst to 0.75 g L−1 increased the extent and the rate of dye removal: circa 62%, at a rate of 0.0381 h−1. When comparing the results for 0.75 g L−1 with those for 1 g L−1 of catalyst, the decolorization rate was lower at the lowest catalyst concentration. Still, the overall degree of decolorization was slightly higher. This slightly lower concentration may enable more efficient light penetration.
![]() | ||
| Fig. 10 Assays using the best conditions for different photocatalyst concentration of (a) 0.36 g L−1 and (b) 0.75 g L−1; (c) final removal of the performed assays using the best conditions. | ||
Under UV irradiation, a final removal of 71.35% was achieved after 12 hours of irradiation (rate of 0.1033 h−1), with a catalyst at a concentration of 0.36 g L−1. Under the same concentration but using visible light, the final removal was 43.05% over 18 h (rate of 0.0218 h−1). At a catalyst concentration of 0.75 g L−1, the final removal under UV light was 58.52% (rate = 0.0730 h−1), which is lower than the results obtained under lower concentrations. Thus, increasing the concentration does not offer any additional advantage. Contrary, an increase in activity is observed under visible light, with the increase in catalyst concentration to 0.75 g L−1, reaching 61.54% (at a rate of 0.052 h−1).
In the first assays, using nanomaterials from batch S1, a 90% removal was achieved, a result not observed in the current assays. This discrepancy is attributed to the lower contribution of adsorption, which is likely due to a stronger binding of silver to the nanoparticle in the later batch (S2).
Recent studies on photocatalytic degradation of MG have reported various materials, such as MnFe2O4 and TiO2 composites, showing promising activity under visible light when combined with additives like H2O2 to enhance oxidative species generation.59 However, the present study introduces a silver-modified zinc-copper ferrite nanoparticle (AgZn0.5Cu0.5Fe2O4) that shows effective MG degradation under exclusive visible light irradiation without the need for external oxidants. This offers both economic and operational advantages, especially in low-resource settings.
A mechanism for the photocatalytic degradation of MG dye using Ag-doped Zn0.5Cu0.5Fe2O4 nanoparticles functionalized with silver is proposed based in the obtained results and supported by the works of Guandao et al.61 and Zetra et al.62 As illustrated in Fig. 11, the degradation of the dye is mainly driven by the action of photogenerated holes (h+), superoxide anions (˙O2−) and hydroxyl radicals (˙OH). The photogenerated holes (h+) can directly attack the dye molecules or react with OH− or water molecules, forming ˙OH radicals. However, according to the reduction potentials of ˙OH/−OH (1.99 eV) and ˙OH/H2O (2.31 eV) pairs, this reaction is unlikely to occur under these conditions. According to literature Zn0.5Cu0.5Fe2O4 conduction band edge is estimated to be between −1.10 eV for ZnFe2O4 and -0.92 eV for CuFe2O4, which renders this process ineffective.63,64 The valence band edge is reported to be between 1.08 eV for ZnFe2O4 and 1.03 eV for CuFe2O4 (ref. 63 and 64). Excited electrons in the conduction band tend to form superoxide anions (˙O2−) through reaction with the dissolved oxygen, which is favorable because the conduction band edge is more negative than the redox potential of O2/˙O2− (0.33 eV).
![]() | ||
| Fig. 11 Representation of the proposed mechanisms for the degradation of MG using Ag doped Zn0.5Cu0.5Fe2O4. | ||
Reactive species attacking the dye may generate various degradation products depending on the degradation pathways. For example, ˙OH radicals can specifically target the central carbon atoms in MG structure, as described by Guandao et al.61 This can lead to the formation of species like N,N-dimethylbenzylamine. Subsequent cleavage of the C–C bonds can produce 4-(N,N-dimethylamino)methylbenzylone. The photocatalytic reaction can further yield compounds such as benzoic acid, p-(dimethylamino)-benzoic acid and p-dimethylaminoethanol, ultimately leading to the formation of inert species like water and carbon dioxide, resulting in complete degradation.
| Sample | Nanomaterial concentration (g L−1) | Light source | Time (h) | Luminescence inhibition (%) | Toxicity interpretation | |
|---|---|---|---|---|---|---|
| Controls | MG (10 mg L−1) | — | — | 0 | 99.45 ± 0.12 | Very toxic |
| AgZn0.5Cu0.5Fe2O4 | 0.75 | Visible | 12 | 23.9 ± 7.7 | Slightly toxic | |
| Effect of photocatalyst concentrations | Sol_ZnCu | 0.36 | Visible | 0 | 99.25 ± 0.08 | Very toxic |
| Sol_ZnCu | 0.36 | Visible | 6 | 99.30 ± 0.40 | Very toxic | |
| Sol_ZnCu | 0.36 | UV | 0 | 99.72 ± 0.14 | Very toxic | |
| Sol_ZnCu | 0.36 | UV | 6 | 99.41 ± 0.33 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.18 | Visible | 0 | 99.69 ± 0.13 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.18 | Visible | 6 | 97.50 ± 0.23 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.36 | Visible | 0 | 99.72 ± 0.04 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.36 | Visible | 6 | 99.78 ± 0.10 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.50 | Visible | 0 | 95.15 ± 4.22 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.50 | Visible | 6 | 99.15 ± 0.82 | Very toxic | |
Sol_AgZnCu (1 : 5) |
1.00 | Visible | 0 | 99.98 ± 0.00 | Very toxic | |
Sol_AgZnCu (1 : 5) |
1.00 | Visible | 6 | 99.71 ± 0.21 | Very toxic | |
| Best conditions and prolongated irradiation | Sol_AgZnCu (1 : 5) |
0.36 | UV | 0 | 99.73 ± 0.24 | Very toxic |
Sol_AgZnCu (1 : 5) |
0.36 | UV | 12 | 96.41 ± 1.31 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.36 | Visible | 0 | 99.93 ± 0.03 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.36 | Visible | 18.3 | 98.76 ± 0.76 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.75 | UV | 0 | 99.88 ± 0.01 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.75 | UV | 12 | 98.65 ± 1.50 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.75 | Visible | 0 | 98.87 ± 0.48 | Very toxic | |
Sol_AgZnCu (1 : 5) |
0.75 | Visible | 18.3 | 99.83 ± 0.12 | Very toxic |
The initial MG solution at a concentration of 10 mg L−1 caused a luminescent inhibition of 99.45%, indicating a level of toxicity considered “very toxic” towards Vibrio fischeri.67,68 The initial MG solution at a concentration of 10 mg L−1 caused a luminescent inhibition of 99.45%, indicating a level of toxicity considered “very toxic” towards Vibrio fischeri. Berberidou et al.,69 similarly reported that a 10 mg L−1 MG solution is extremely ecotoxic, completely inhibiting the bioluminescence of V. fischeri. Hernando et al.68 reported an EC50 value of 0.031 mg L−1 for MG after 30 min of exposure to V. fischeri, highlighting its high toxicity. The toxic character of MG may be attributed to its lipophilic nature, which enables it to easily pass through the bacterial phospholipid bilayer. This same characteristic, in humans may potentially lead to carcinogenesis, mutagenesis, teratogenicity and respiratory toxicity.70 After photocatalytic treatment, all the samples remained highly toxic, with luminescence inhibition values exceeding 98% (Table 4), despite the high MG removal percentage observed in the photocatalytic assays (Fig. 9 and 10). This could be explained by the toxicity of the dye, which persists even at low concentrations in the treated solution (0.93 mg L−1), but also by the potentially more toxic degradation by-products formed during the degradation process. This is consistent with previous published works observing the formation of degradation by-products from the photocatalytic degradation of pollutants, that are more toxic than the parent compound.70,71 For instance, Pérez-Estrada et al.71 conducted the photolytic degradation of MG, resulting in a large number of degradation by-products that were more persistent and toxic than the parent compound. They identified 28 possible by-products, with the molecule 4-(dimethylamino)benzophenone (C15H16NO) being detected at the highest concentration after 220 h of exposure to natural sunlight. After photocatalytic treatment, the concentration of this compound was still over its EC50 value towards V. fischeri (EC50 of 0.061 mg L−1), indicating that it strongly contributed to the toxicity of the treated effluent.71 Therefore, extending the irradiation assays to promote complete mineralization of the compounds is crucial to reduce the toxicity of the treated sample.
The study identified several key areas for further development that will help elevate the process to a higher level of efficiency and applicability. An optimized cleaning process is essential to enhance the photocatalytic activity of SV_AgZnCu and render the nanomaterials suitable for potential industrial applications. The Sol_AgZnCu nanomaterial is well-suited for degrading low residual concentrations of dye, though it exhibits some saturation due to the accumulation of adsorbed degradation products. This makes it effective in the final stages of wastewater treatment, where the organic load is lower and micropollutants are targeted. Refining the silver reduction methodology could further boost photocatalytic activity, reduce dye toxicity, and shorten the required irradiation time. Addressing these aspects will enhance both the efficiency and industrial relevance of the developed nanomaterials.
| This journal is © The Royal Society of Chemistry 2025 |