Open Access Article
Dinh Ngo Vua,
Van Chau Dinha,
Tien Hoang Nguyenb,
Thi Thao Minha,
Phuoc Cuong Le
c,
Tan Nhat*de and
Dinh Nhi Bui
*a
aElectric Power University, Ha Noi, Vietnam. E-mail: vietnamkz@yahoo.com
bThe University of Danang, University of Science and Education, 459 Ton Duc Thang st., Da Nang 550000, Vietnam
cThe University of Danang, University of Science and Technology, Danang, 550000, Vietnam
dInstitute of Research and Development, Duy Tan University, Da Nang, Vietnam. E-mail: ttnhat@duytan.edu.vn
eSchool of Medicine and Pharmacy, Duy Tan University, Da Nang, Vietnam
First published on 17th September 2025
This study presents a novel heterogeneous catalyst system – CuMnO2@cobalt phthalocyanine immobilized on titanium foam (CuMnO2@CoPc/Ti foam) – that is highly effective in treating three common antibiotics, ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA), through UV-persulfate-based advanced oxidation (UV/PS). Comprehensive characterization (XRD, SEM-EDS, FTIR, BET, XPS, and UV-vis DRS) confirms a stable mesoporous architecture and uniform coating; the specific surface area increases from 55.3 to 60.8 m2 g−1 after CoPc deposition, and the apparent optical band gap narrows from ≈1.41 to ≈1.31 eV, indicating improved visible-light response. Antibiotic decay follows pseudo-first-order kinetics with kapp of 0.0426 ± 0.0015, 0.0410 ± 0.0017, and 0.0393 ± 0.0013 min−1 for OFL, LEVO, and SFA, respectively, delivering >90% removal in 60 min. TOC mineralization reaches ≈69–72%, evidencing deep oxidation beyond parent-compound loss. Mechanistically, XPS verifies persistent Cu2+/Cu+ and Mn4+/Mn3+ surface couples (Co states stable), while EPR with DMPO detects SO4˙− and ˙OH, with sulfate radicals dominating. The monolithic catalyst is reusable over 5 cycles with negligible activity loss and ultra-low metal leaching (Cu/Mn/Co ≪ 0.1 mg L−1 by ICP-OES), underscoring chemical robustness. These results highlight CuMnO2@CoPc/Ti foam as a durable, separable platform for UV/PS treatment of antibiotic-bearing waters and provide a mechanistic blueprint—combining photosensitization and multi-center redox—to guide future reactor-scale implementations.
Meanwhile, Advanced Oxidation Processes (AOPs) technologies based on the generation of strong oxidizing radicals (SO4˙−, ˙OH) have attracted attention due to their ability to thoroughly treat difficult-to-degrade organic compounds such as antibiotics.7–10 The UV/PS method – activating persulfate with ultraviolet light – is particularly notable for its higher oxidation efficiency compared to UV/H2O2 or ozonation, due to the generation of both sulfate radicals and hydroxyl radicals.11,12 The UV/PS system has demonstrated the effectiveness of removing antibiotics such as ciprofloxacin, sulfamethoxazole with a fast reaction rate and the ability to operate under neutral pH conditions.13,14
However, the actual efficiency of the UV/PS system depends greatly on the structure and properties of the catalyst. In many recent studies, the development of heterostructure catalysts has shown the ability to effectively control the electron flow, reduce electron–hole recombination, and enhance the conversion of PS to SO4˙− radicals through active metal redox pairs.15–17 CuMnO2 is a mixed oxide of Cu and Mn, so they have the ability to enhance acid and base resistance, increase surface area and strengthen catalytic stability compared to single metal oxides.18,19 The induction of CuMnO2 when interacting with PS under UV light leads to continuous peroxo activation, increased SO4˙− generation through the conversion of Cu(II)/Cu(I) and Mn(III)/Mn(II).20,21 A study by Chen et al.22 demonstrated that nano-CuMnO2 when combined with PS could remove up to 93% of orange I in just 30 min of reaction, while the efficiency of CuMnO2 alone was significantly lower (below 50%).
The surface of CuMnO2 when coated with cobalt phthalocyanine (CoPc) (via π–π connection) forms a heterogeneous catalyst system CuMnO2@CoPc, expanding the surface area, providing more active sites and enhancing the interaction with PS and UV light. Several similar studies have applied the combination of different materials with CoPc to increase the efficiency of environmental pollution treatment. For example, in the photocatalytic study of MB degradation, Fe3O4@SiO2@TiO2@CoPcS microspheres showed significantly higher activity (86.4%) than Fe3O4@SiO2@TiO2 microspheres (<60%).23 Another study on CoPcTs/TiO2 showed that coating phthalocyanine on the TiO2 surface improved the degradation efficiency of thionine dye significantly compared to the original TiO2 (degradation efficiency of 90% for CoPcTs/TiO2 and 40% for TiO2 after 120 min of treatment).24 However, these systems still have limitations such as the influence of pH, electron–hole recombination rate and suboptimal efficiency when treating different antibiotic mixtures.
To ensure mechanical strength and easy recovery, CuMnO2@CoPc is immobilized on titanium foam (Ti foam). Titanium foam not only has good resistance to corrosive environments and has a 3D porous structure, but also has effective surface interaction with thin catalyst layers, supporting reusability and reducing catalyst loss.25–27
To date, there have been only a limited number of studies combining CuMnO2 and CoPc heterostructures in metal foam structures. However, similar catalytic systems such as MIL 101(Fe)/MoS2, Cu@MoS2/PAAm/CA NCDN, g-C3N4/Mn3O4/ZIF-8 have provided clear evidence of the ability to enhance the photolysis and PS conversion efficiency in the presence of asymmetric electron pairs and multi-component heterostructures.28–30 In particular, the MIL 101(Fe)/MoS2/PS system has a much faster degradation rate of tetracycline (85% in 40 min) than each individual component, while g-C3N4/Mn3O4/ZIF-8/PMS gives rhodamine B (RhB) degradation efficiency up to >96% and the efficiency does not change much (decrease 1.7%) after 5 cycles of use – a remarkable achievement for the development of new generation heterostructures.
Several recent studies have demonstrated that foam-encapsulated heterojunction catalysts have superior performance. For example, hierarchical porous N–TiO2/carbon foam composite demonstrated impressive performance, this material is capable of removing 97% of phenol within 2 hours under simulated sunlight irradiation and its photocatalytic activity is 4.9 times higher than that of N–TiO2 and 10.8 times higher than that of pure TiO2, confirming its superiority in applications related to pollutant degradation.31 Another study on Ti-foam/TiOxHy–NTs/SnO2–Sb showed that this material exhibited excellent ability in effectively decomposing the real and simulated PAM-containing fracturing flowback fluid, achieving a chemical oxygen demand degradation of over 90.0% after 120 min of EO process.32
The expectation of this study is to demonstrate that CuMnO2@CoPc/Ti foam when combined with UV/PS will bring superior antibiotic treatment efficiency, high material durability and practical application in wastewater treatment plants.
After obtaining the dry CuMnO2, about 0.2 g of the material was dispersed in 50 mL of ethanol containing 0.05 g of CoPc. The mixture was ultrasonicated for 2 h to ensure the π–π interaction between the phthalocyanine rings and the metal oxide surface. This process was to form the CuMnO2@CoPc heterojunction, which helps to expand the surface area and provide more active catalytic sites. Finally, the material was filtered, gently dried at 60 °C, and a characteristic dark blue powder was obtained.
The Ti foam was cut into 2 cm × 2 cm pieces and soaked in 0.1 M HCl solution under ultrasonication for 15 min to clean the surface and enhance adhesion. After rinsing with distilled water and ethanol, the Ti foam was dried and immersed in a solution containing CuMnO2@CoPc dispersed in ethanol. The immersion-heat system was repeated 3 times, each immersion was followed by a drying at 80–100 °C for 1 hour. The catalyst layer was formed stably and firmly adhered to the surface of the Titan foam, forming a CuMnO2@CoPc/Ti foam system ready for the photocatalytic process.
Antibiotic concentrations were measured by LC-MS/MS analysis,34 while the extent of mineralization was assessed using TOC. In addition, the role of free radicals ˙OH and SO4˙− in the degradation mechanism was clarified by radical scavenging experiments using ethanol and tert-butanol.
:
2
:
2
:
1 quartet, g ≈ 2.005), and SO4˙− in a 50% v/v MeOH/H2O mixture to stabilize the DMPO–SO4˙− adduct (6-line model, g ≈ 2.006). Controls included no UV, no PS, and no catalyst; parallel radical quenching experiments with EPR using 0.5 M tert-butanol (quenching ˙OH), 1.0 M MeOH (quenching both SO4˙−/˙OH), and 20 mM NaN3 (quenching 1O2). Each condition had n = 3 replicates and the mean ± SD was reported.
The treatment efficiency was evaluated by measuring the remaining pollutant concentration in the solution after the reaction using LC-MS/MS, and determining the degree of mineralization through analysis of total organic carbon (TOC).
To evaluate the stability of the catalyst after cycles of use, the surface morphology and elemental composition were re-analyzed by SEM-EDS.
After CoPc coating, some XRD peaks of CuMnO2 tended to weaken and broaden, especially at 30–40°, indicating the presence of an organic coating on the metal oxide surface. The CuMnO2@CoPc/Ti foam sample still retains the diffraction characteristics of CuMnO2, while the peaks at 38.4°, 40.1° and 53.2° are attributed to the crystal structure of Titan foam.36 The absence of significant XRD peak shift after CoPc attachment indicates that the π–π interaction between CoPc and CuMnO2 does not alter the oxide crystal structure, which helps preserve the electronic properties of the catalyst matrix.
The scanning electron microscopy (SEM) image in Fig. 2 shows that CuMnO2 has a nano-sized particle morphology. The morphological data from scanning electron microscopy (SEM) provide clear evidence of the nanostructure of the material. The CuMnO2 sample exhibits a discrete, uniformly distributed bulk morphology with a particle size ranging from 50 to 100 nm. After CoPc coating, the particle surface becomes rough and has an uneven thin film coating, which is typical of the phenomenon of organic material coating on metal oxides. When observing the CuMnO2@CoPc sample fixed on the Titan foam, the adhesion of the catalyst layer to the three-dimensional porous structure of the Ti foam matrix can be clearly seen. This porous structure acts as a mechanical support, both facilitating the surface reaction and allowing the efficient flow of the solution through – which recent study with graphite/TiO2/nickel foam has also exploited to improve the photocatalytic efficiency in continuous flow systems.37
Elemental composition analysis by EDS (Fig. 3) showed the simultaneous presence of Cu, Mn, Co, O and Ti elements with a Cu
:
Mn mass ratio of nearly 1
:
1, reaffirming the mixed oxide structure of CuMnO2. The Co content is about 2.4 wt%, which is equivalent to the appropriate CoPc coverage level for the catalytic reaction without blocking the active sites. Many studies showed that too high a capping agent content can reduce the contact between the catalyst and the oxidizing agent, reducing the overall performance.38
The FT IR spectrum obtained for the CuMnO2@CoPc/Ti foam material clearly shows the characteristic absorption bands for each component (Fig. 4). The broad absorption band at around 3450 cm−1 is due to the stretching vibration of the –OH group and the presence of adsorbed water on the oxide surface, which is a common feature in oxide-based catalysts. The bands at around 1330 cm−1 are assigned to the C–C/C–N stretching vibrations in the phthalocyanine ring, while the peak at 723 cm−1 is the out-of-plane C–H vibration characteristic of the α phase of phthalocyanine. The region of 500–700 cm−1 shows two weak peaks at 625 cm−1 and 515 cm−1, corresponding to the stretching vibrations of Mn–O and Cu–O in the mixed oxide. Compared with the FT IR spectrum of pure Cu doped Mn3O4/Mn doped CuO, the metal–oxygen peaks in the CuMnO2@CoPc sample have significantly lower intensities,39 which reflects the coverage of the CoPc layer on the oxide surface, which attenuates the lattice vibration signal. This result is consistent with the report on NiO/CoPc materials by Sheena et al., in which the characteristic vibrational bands of CoPc such as 1330, 1287, 1156, and 723 cm−1 are preserved when CoPc is covered on the oxide.40 Thus, the FT IR spectrum of CuMnO2@CoPc/Ti foam both confirms the presence of the phthalocyanine layer through characteristic vibrational bands, and shows that this coating weakens the metal–oxygen vibrations, demonstrating the interaction between the CoPc layer and the oxide substrate – a phenomenon that has also been observed in other oxide-coated phthalocyanine material systems.
The thermal stability of the catalyst was evaluated by thermogravimetric analysis (TGA) in air (Fig. 5). The TGA spectrum of CuMnO2@CoPc/Ti foam shows three distinct mass loss stages when heated from room temperature to 800 °C. In the first stage, the sample lost about 3% of its mass below 150 °C, mainly due to the evaporation of water and the removal of hydroxyl groups adsorbed on the oxide surface. This is a common phenomenon for many oxide materials, and the small mass loss confirms that the sample was well dried before the measurement. When the temperature increased to the 300–380 °C region, the mass continued to decrease by about 10%. This loss was attributed to the initial decomposition of the phthalocyanine layer. Our results are consistent with the study of substituted metal phthalocyanines, where TGA showed that CoPc and CuPc started to lose mass at 300–309 °C.41 The organic functional groups on the phthalocyanine decompose first, leading to a decrease in initial mass but the macrocyclic structure remains. The strongest mass loss occurs in the range of 420–550 °C, with a loss of about 20% of the initial mass. This stage corresponds to the complete decomposition of the phthalocyanine ring into inorganic products such as cobalt oxide. In the study of the CoPc/hybrid carbon system, the derivative TGA curve of CoPc shows only one decomposition peak with an onset around 440 °C,42 which is consistent with the strong mass loss in our sample. After the loss of most of the organic component, the TGA curve becomes flat – the mass is stable up to 800 °C. This indicates that the CuMnO2 framework and the Ti foam layer are thermally stable; In fact, studies on CuMnO2 show that this material does not decompose but even absorbs oxygen, leading to an increase in mass when calcined in an oxidizing environment.43 Compared with other works, the mass loss of CuMnO2@CoPc/Ti foam is smaller than that of CuMnO2@CoPc without Ti foam matrix (usually losing 30–40% of mass), due to the large mass fraction of Ti foam matrix reducing the proportion of CoPc in the whole material. At the same time, the observed decomposition temperatures (∼300 °C and 440–500 °C) coincide with the decomposition milestones recorded for CoPc and its derivatives, indicating that the coating of CoPc onto CuMnO2 did not significantly change the pyrolysis behavior of phthalocyanine. The TGA spectra also confirmed that after the decomposition of CoPc, the CuMnO2 framework and Ti foam remained stable, similar to previous reports on the thermal stability of CuMnO2.
The BET results show a clear difference in surface area and pore volume between the materials (Table 1). The pure CuMnO2 sample has a surface area of 55.3 m2 g−1 and a pore volume of 0.25 cm3 g−1, this value is very close to the 54.9 m2 g−1 previously reported for CuMnO2/NF44 and is within the usual range for mesoporous transition oxides. When cobalt phthalocyanine is added, the surface area increases slightly to 60.8 m2 g−1 and the pore volume to 0.30 cm3 g−1. This increase is consistent with the trend observed in similar material systems: in the CuMnO2–rGO nanocomposite, the voids between the rGO and the CuMnO2 nanoplates increased the surface area from 54.9 to 90.3 m2 g−1, and in the CoPc/Ti3C2 system, the CoPc layer increased the SBET from 118.5 to 126.1 m2 g−1.45 These studies show that introducing an organic phase into the oxide framework can create new pores or expand the diffusion channel without clogging the existing micropores. Compared to the gCN system, in which CuMnO2-gCN achieved a surface area of 82.29 m2 g−1 and a pore volume of 0.429886 cm3 g−1,46 the CoPc layer produced a more moderate increase – perhaps because the bulky structure of phthalocyanine did not create a pore network as large as gCN. However, when the CuMnO2@CoPc material was anchored onto the titanium foam framework, the surface area and pore volume decreased sharply to 20.4 m2 g−1 and 0.14 cm3 g−1, respectively. This decrease reflects the influence of the large mass fraction and low surface area of the Ti foam. The Ti foam matrix itself has a surface area of only 0.055–0.12 m2 g−1,47 when coated with the active material, the specific surface area of the entire composite is diluted by the foam mass. This result emphasizes that compared to light matrices such as gCN or rGO, the presence of a heavy metal matrix will significantly reduce the surface area per unit mass, thereby affecting the catalytic activity of the material.
| Samples | Surface area (m2 g−1) | Capillary volume (cm3 g−1) |
|---|---|---|
| CuMnO2 | 55.3 | 0.25 |
| CuMnO2@CoPc | 60.8 | 0.3 |
| CuMnO2@CoPc/Ti foam | 20.4 | 0.14 |
| ln(Ct) = ln(Co) − kappt |
![]() | ||
| Fig. 6 Kinetic model of the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam/UV/PS: (a) zero order; (b) first order; (c) second order. | ||
From there, the apparent reaction rate coefficient kapp was determined from the slope of the trendline. This is a common feature of advanced oxidation processes (AOPs) involving free radicals (˙OH, SO4˙−), which follow assumed first-order kinetics because the concentration of the oxidant (PS or free radical) is much larger than that of the initial substrate. According to Oturan & Aaron,48 free radical-mediated oxidation reactions mostly follow an assumed first-order mechanism because the concentration of the oxidant remains almost constant throughout the process. Compared with other studies, this result is consistent with Milh et al.13 when studying the degradation of ciprofloxacin (CIP) in the UV/PS and UV/PMS systems, in which both processes exhibited a strong fit to a pseudo-first-order kinetic model. Chen et al.49 also reported first-order kinetics for the degradation of sulfamethoxazole in the sunlight/Fe(II)/Cit/PS system, with k values of ∼0.0363 min−1 at room temperature. In this study, the kapp values were determined to be 0.0426 ± 0.0015 hours−1 (OFL), 0.0410 ± 0.0017 hours−1 (LEVO), and 0.0393 ± 0.0013 minutes−1 (SFA), and the corresponding semi-automatic values were 16.3, 16.9, and 17.6 minutes.
In contrast, the zero- and second-order models had significantly lower R2 coefficients (0.92–0.98), indicating that the reaction did not proceed at a linear rate in C (zero-order) or 1/C (second-order). Specifically: With the zero-order model (Fig. 6a), the reaction was assumed to be independent of the initial concentration, which is not consistent with experimental observations, especially at low concentrations. The second-order model (Fig. 6c) – commonly seen in co-catalysis – shows a clear deviation at long times, which may be due to strong adsorption at the catalyst surface or free radical competition.
In summary, the assumed first-order kinetic model best reflects the degradation of OFL, LEVO and SFA on the CuMnO2@CoPc/Ti foam catalyst under UV/PS. This is the basis for further calculations of kinetic parameters such as activation energy (Ea) and rate coefficients in future studies.
![]() | ||
| Fig. 7 Effect of UV/PS on the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam. | ||
When exposed to a single 365 nm UV light, the degradation efficiency improved but was still limited. Specifically, OFL achieved an efficiency of about 18.4%, LEVO was 15.9%, while SFA only achieved 8.3%. This reflects the difference in UV absorption capacity and molecular structure of each compound. Fluoroquinolone compounds such as OFL and LEVO have better UV light absorption capacity due to the aromatic ring with conjugated double bonds and attached fluorine atoms, thereby leading to a higher direct photolysis process.50 However, this process is still not strong enough to completely break the bonds in the molecular structure.
The addition of PS without UV irradiation increased the degradation efficiency to 19–25%, due to the spontaneous decomposition of PS in an aqueous environment to produce SO4˙− radicals. However, this reaction was slow and inefficient without a catalyst or external activation energy source. Compared to the study by Liu et al.,51 in which PS was used to treat SFA, the degradation efficiency was about 0% after 60 min without UV, UV irradiation significantly boosted the efficiency of sulfanilamide decomposition by PS, achieving 72.6% within 60 minutes, indicating the need for PS activation to accelerate the reaction.
When UV was combined with PS, the degradation efficiency of all three compounds increased significantly. OFL and LEVO reached 82.3% and 79.5% after 60 min, respectively, while SFA reached 68.4%. The mechanism of PS activation by UV leads to the cleavage of the O–O bond in S2O82− to generate strong oxidants such as SO4˙− and ˙OH.
| S2O82− + hv (λ = 254 nm) → 2SO4˙− |
| SO4˙− + H2O → ˙OH + HSO4− |
These radicals have very high oxidation potentials, strong enough to break the C–N, C–F and C–S bonds in the antibiotic structure.
What is particularly remarkable is that when adding the CuMnO2@CoPc/Ti foam catalyst to the UV/PS system, the degradation efficiency is clearly superior. Specifically, OFL reached 96.8%, LEVO reached 94.5%, and SFA also reached 91.3% after only 30 minutes of reaction. This increase in efficiency is explained by the catalytic role of the Cu2+/Cu+ and Mn3+/Mn2+ redox pairs in the activation of persulfate.
| Cu2+ + S2O82− → Cu+ + SO4˙− + SO42− |
| Mn3+ + S2O82− → Mn2+ + SO4˙− + SO42− |
Experimental results show that the UV/PS system with CuMnO2@CoPc/Ti foam can decompose over 90% of OFL, LEVO and SFA compounds after only 60 minutes of reaction. This confirms the synergistic effect between UV light, persulfate and catalytic materials in creating strong oxidants and decomposing stable antibiotic compounds in the aquatic environment.
![]() | ||
| Fig. 8 Effect of pH on the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam/UV/PS. | ||
The PS reaction mechanism in acidic medium is more favorable because HSO5− can be catalyzed or irradiated to form SO4˙− radicals with high yield. The SO4˙− radical has a high oxidation potential (2.5–3.1 V), is stable in acidic solutions, and readily reacts with organics via an electron-attack mechanism. Meanwhile, in alkaline environment, HSO5− is easily decomposed into SO42− and cannot form highly active free radicals. In addition, SO4˙− at high pH is also converted into weaker OH˙ radicals, reducing the overall decomposition efficiency. These phenomena have been demonstrated in the work of Kemmou et al.,52 when treating sulfamethoxazole by biochar-activated persulfate, the decomposition efficiency decreased as the pH increased from acidic to near-neutral and then to alkaline conditions.
In addition, pH affects the ionization state of antibiotic compounds in solution. Compounds such as OFL and LEVO have pKa in the range of 5.5–6.5, so at pH 4, they mainly exist in the protonated form. This increases the ability to interact with the negatively charged catalyst surface, thereby improving the initial adsorption efficiency – an important prerequisite for the oxidation reaction. In contrast, at high pH, most of the molecules exist in the anionic or neutral form, reducing the adsorption capacity due to the electrostatic repulsion between the catalyst surface and the pollutant. SFA was more strongly affected, due to its molecular structure with less conjugated aromatic rings, so its reaction with oxidant radicals was also lower in alkaline environments. These results are consistent with the study of Jiang et al.,53 in which the ZVI/BC/PS system treated atrazine with an efficiency of 68% at pH 3 but only 46% at pH 8 after the same reaction time.
From the above analysis, it can be concluded that the environmental pH has a significant influence on the efficiency of antibiotic treatment by the CuMnO2@CoPc/Ti foam/UV/PS system, and the most optimal condition is in a slightly acidic environment (pH ∼4). This conclusion has practical value in the design and adjustment of pharmaceutical wastewater treatment systems, especially in the pretreatment stage or neutralization of the influent pH to maximize the efficiency of pollutant removal.
![]() | ||
| Fig. 9 Effect of temperature on the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam/UV/PS. | ||
The mechanism of temperature impact on treatment efficiency can be explained in three aspects. First, according to the Arrhenius law, high temperature increases the average energy of molecules in the reaction system, thereby increasing the probability of effective collisions and promoting the reaction rate between the oxidant radical and the antibiotic molecule. Second, high temperature supports the decomposition of PS to form SO4˙− radicals without completely depending on UV irradiation or catalyst, thereby increasing the density of free radicals in the solution. Third, temperature also increases the diffusion and adsorption rate of antibiotics onto the catalyst surface, shortening the required contact time and improving the surface reaction efficiency. These results are consistent with the observation from the study of Ma et al.,54 in which the treatment of ciprofloxacin by US/E/PS system also showed that increasing the reaction temperature significantly improved ciprofloxacin removal. However, it should be noted that excessive temperature increase can cause undesirable effects. Specifically, PS can decompose rapidly at high temperatures without generating effective oxidizing radicals, leading to the formation of products such as SO42−, HSO4− or O2 – substances that are not strong oxidizing agents. At the same time, high temperatures also increase the rate of self-decomposition of SO4˙− and ˙OH radicals, reducing the lifetime of these radicals in the solution, thereby affecting the overall treatment efficiency if the optimal temperature threshold is exceeded.
![]() | ||
| Fig. 10 Effect of PS concentration on the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam/UV/PS. | ||
The decrease in yield at high PS concentrations is explained by the competitive reactions between PS molecules and the SO4˙− radicals themselves. When the PS concentration is too high, the SO4˙− radicals may react undesirably with excess PS to form less active radicals or even unreacted products, according to the equation:
| SO4˙− + S2O82− → SO5˙− + SO42− |
The SO5˙− radical has been shown to have a significantly lower oxidation potential than SO4˙−, and is incapable of breaking the strong bonds in antibiotic molecules. Furthermore, the SO4˙− radical can also react with each other to form neutral SO42−:
| SO4˙− + SO4˙− → S2O82− |
These reactions reduce the concentration of active radicals in the solution and lead to a decrease in reaction efficiency.
In addition, PS is a strong oxidant, when present at high concentrations, it can also damage the catalyst structure or cause aging of the catalyst surface through excessive oxidation, reducing the adsorption capacity and promoting subsequent reactions. A similar phenomenon was also observed in the study by Dong et al.,55 where an excess amount of PS in the Fe(II)/PS system caused a sharp decrease in iopamidol (IPM) treatment efficiency, likely due to the scavenging of oxidative radicals by the excess persulfate (PS).
From the obtained data, it can be determined that the optimal PS concentration for the CuMnO2@CoPc/UV/PS reaction system is about 2.0 mM. This is the level that gives the highest treatment efficiency without causing side effects, which is in line with the practical operational requirements of chemical costs and environmental safety.
![]() | ||
| Fig. 12 Effect of catalyst mass on the degradation of ofloxacin (OFL), levofloxacin (LEVO) and sulfanilamide (SFA) by CuMnO2@CoPc/Ti foam/UV/PS. | ||
The phenomenon of performance saturation with increasing catalyst loading has been explained by many previous studies. On the one hand, as the catalyst loading increases, the active surface area and the number of catalyst sites also increase, leading to higher PS activation and pollutant adsorption. However, when the catalyst loading is too high, the catalyst particles tend to aggregate together (agglomeration), reducing the total actual contact area. At the same time, excess catalyst can also lead to inefficient consumption of free radicals through side reactions such as:
| SO4˙− + Mn+ → Mn++ + SO42− |
From these XPS observations, the mechanism can be described as follows. Under UV irradiation, CoPc absorbs light to form an excited state CoPc* as a local electron/hole source:
| CoPc + hν → CoPc* |
| CoPc* → CoPc˙+ + e− |
Photogenerated electrons are injected to transition metal pairs on CuMnO2, rapidly re-reducing the higher oxidation states:
| Cu2+ + e− → Cu+ |
| Mn4+ + e− → Mn3+ |
| Co3+ + e− → Co2+ |
| S2O82− + e− → SO4˙− + SO42− |
| Cu+/Mn3+/Co2+ + S2O82− → Cu2+/Mn4+/Co3+ + SO4˙− + SO42− |
The SO4˙− radical can be further converted to ˙OH in water, expanding the spectrum of oxidizing agents:
| SO4˙− + H2O → ˙OH + HSO4− |
The role of surface oxygen is evident in O 1s: Olattice provides a stable M–O–M framework, while Odefect/OH anchors persulfate and contaminants, lowering the electron transfer barrier; these defects are also the site of formation of active species near the surface. The conservation of the oxygen structure (no BE shift, stable distribution) is consistent with the fact that the Cu/Mn/Co redox pairs maintain the same ratio after cycling, meaning that the heterojunction is capable of self-recharging electrons: electrons from CoPc* and organic intermediates continuously close the cycles Cu(I) ↔ Cu(II), Mn(III) ↔ Mn(IV), Co(II) ↔ Co(III), while persulfate SO4˙−/˙OH to mineralize the substrate. This “role assignment” – CoPc captures light and donates electrons, Cu/Mn/Co serves as the polycentric PS activation site, Odefect assists in adsorption and electron transfer – explains why the peak shapes and valence ratios remain nearly constant over multiple cycles, and is consistent with the observed stable degradation efficiency and high mineralization. In other words, XPS not only confirms the surface composition and valence, but also provides direct evidence that the UV/PS mechanism in this system is based on a cooperative redox network that is stable over time.
The EPR data directly demonstrated the free radical mechanism of the UV/PS system with CuMnO2@CoPc/Ti foam and quantified the contribution of each active agent. In H2O with DMPO (50 mM), the DMPO–˙OH spectrum appeared immediately with a 1
:
2
:
2
:
1 tetralinear pattern, g ≈ 2.005 and aN ≈ aH ≈ 14.9 G; while in 50% MeOH/H2O to stabilize the sulfate adduct, the DMPO–SO4˙− spectrum clearly showed a hexagonal pattern (triplet × doublet) with g ≈ 2.006, aN ≈ 13.2 G and aH ≈ 9.6 G (Fig. S2a, S2b and Table S2). All signals disappeared in the no UV/no PS/no catalyst controls, confirming that radicals are only generated when all three factors are present. Radical quenching experiments showed that 0.5 M TBA strongly reduced the ˙OH signal (∼72 ± 3%), 1.0 M MeOH also suppressed both SO4˙−/˙OH (∼85 ± 4%), while 20 mM NaN3 had almost no effect (<8%), ruling out a dominant role for 1O2. Combining radical quenching with the quadratic integration of the EPR spectrum, the estimated relative contributions were SO4˙− = 58 ± 6%, ˙OH = 34 ± 5%, others ≤ 8% (Table S2). The time course (0–120 s) shows that the EPR intensity increases rapidly in the first 20–40 s and then reaches near equilibrium, consistent with the continuous activation of PS on the heterojunction surface (Fig. S2c). Depending on the concentration, PS exhibits a saturation pattern with a maximum at 2.0 mM and a slight decrease at 2.5–3.0 mM (Fig. S2d), possibly due to the “self-quenching”/radical capture reaction by excess PS and radical recombination: SO4˙− + S2O82− → SO42− + S2O8˙− (forming less active radicals) and SO4˙− + SO4˙− → S2O82−. These observations are consistent with XPS: the Cu(II)/Cu(I) and Mn(IV)/Mn(III) pairs (along with Co(III)/Co(II) of CoPc) remained almost unchanged after five cycles, indicating that the sustained redox cycling on the surface is the PS activator.
The UV-vis DRS spectrum shows that the CuMnO2@CoPc/Ti foam sample absorbs more strongly and broadly than CuMnO2, with the characteristic bands of CoPc (shoulder at ∼340 nm and “Q-bands” at ∼610–690 nm) overlapping the absorption background of CuMnO2; the absorption edges are clearly red-shifted and the overall intensity increases (Fig. S3a). The Kubelka–Munk transform F(R) = (1 − R)2/(2R) combined with the Tauc plot [F(R)·hν]1/2 − hν allows extrapolation of the linear region to determine the optical band gap. The results show that the apparent Eg of CuMnO2 ≈ 1.41 eV, decreases to ≈ 1.31 eV for CuMnO2@CoPc (fresh sample) and remains almost unchanged after 5 cycles (≈1.32 eV), demonstrating the optical stability of the CoPc layer (Fig. S3b). The narrowing of Eg ∼0.10 eV together with the appearance/emphasis of the Soret–Q bands of CoPc indicate that CoPc acts as a photosensitizer, creating additional effective electron–hole transition channels in the visible region. Mechanistically, under illumination excited states of CoPc (CoPc*) form and transfer electrons across the interface to the acceptor (CuMnO2 or directly to persulfate), while the remaining holes on CoPc/CuMnO2 oxidize water/–OH. This increases the excited carrier density and explains the resonance with persulfate activation. The “red shift” of the absorption edge and the smaller apparent Eg may also reflect the IFCT/Urbach tail transition at the CoPc–CuMnO2 interface, which is commonly observed in phthalocyanine/oxide systems and induces a “light-induced electron conduction” effect in the lower energy region. On the other hand, the fact that the two Tauc curves of the fresh and after 5 cycles samples are almost identical indicates no degradation or photoabsorption peeling after reuse—consistent with XPS (Cu2+/Cu+, Mn4+/Mn3+, Co3+/Co2+ ratios are almost preserved) and EPR (SO4˙− radical intensity is dominant, optimal at PS = 2.0 mM).
In the first pathway, the OFL molecule is attacked by strong free radicals such as SO4˙− or ˙OH generated from the activation of PS under the action of the catalyst, leading to quinolone ring opening and the formation of product P1 (m/z = 305). The ring system was then further cleaved to form P2 (m/z = 279) and finally P3 (m/z = 168), indicating a deep and comprehensive mineralization of the quinolone structure. This ring opening mechanism is similar to that reported in the CoFe2O4/PS catalyst system by Fan et al., in which the quinolone ring was also considered to be a vulnerable starting point due to the high electron density near the carbonyl position.58 The second pathway was reported to involve the reduction of the carboxylic group (–COOH) to a methyl group (–CH3), reducing the molecular weight to m/z = 332 (P4). The free radicals then further cleaved the single rings – especially the piperazinyl ring – to form product P5 (m/z = 223). Compared with the mechanism reported by Liu et al. reported that during the UV/PS treatment of ofloxacin, decarboxylation is one of the characteristic pathways in the initial detoxification process and a precursor to further destruction.59 The third pathway begins with the removal of the fluorine (F) atom from the aromatic position, forming P6 (m/z = 344). This is a common reaction in fluoroquinolone treatment, since the F atom, although stable, is easily replaced by the SO4˙− radical under transition metal catalytic conditions. Next, the COOH group is removed to form P7 (m/z = 286), and the remaining single rings are broken, forming P8 (m/z = 139). The F and COOH elimination mechanism was previously observed in the UV/PDS system by Zhu et al., and was found to be the pathway that significantly reduced the biological activity of OFL.60 The fourth pathway involves the loss of peripheral methyl groups, typically from the piperazinyl ring and aromatic substituents, leading to intermediate P9 (m/z = 318). This is followed by the combined loss of F and COOH to form P10 (m/z = 258), before the remaining rings are completely oxidized, resulting in P11 (m/z = 138). This process clearly demonstrates the preference of the free radical system to attack easily oxidizable groups before completely destroying the ring framework. This result is consistent with the observation of Xing et al. when treating ofloxacin by magnetic CuFe2O4 coupled PMS system, in which the removal of the methyl group of the piperazinyl substituent was noted to be the initial step for the radical degradation.61
In addition to determining the degradation pathways via LC-MS/MS, TOC analysis was used to assess the overall mineralization level of the system. The results showed that the TOC concentration decreased from 8.62 mg L−1 initially to 2.65 mg L−1 after 60 minutes of UV irradiation, corresponding to a mineralization efficiency of 69.3% (Table 2). This is direct evidence of not only the destruction of the chemical structure but also the conversion to CO2 and H2O, which significantly reduces the risk of bioresidues of the antibiotic in the aquatic environment.
| Time (min) | TOC (mg L−1) | Degree of mineralization (%) |
|---|---|---|
| 0 | 8.62 | 0 |
| 15 | 6.98 | 19.1 |
| 30 | 4.87 | 43.5 |
| 45 | 3.56 | 58.7 |
| 60 | 2.65 | 69.3 |
The levofloxacin molecule (m/z = 362) is oxidized via SO4˙− and ˙OH radical reactions, generating at least four main degradation pathways. In the first pathway, the free radical attacks the C
C double bond at position 2–3 of the quinolone ring, causing ring opening and forming intermediate product L1 (m/z = 305). This process continues with the cleavage of C–C, C–N bonds in the open ring system, leading to L2 (m/z = 279), and finally L3 (m/z = 168) – a small, highly polar product, representing the complete destruction of the quinolone ring framework. This mechanism was previously described in the study of Liu et al. when treating LEVO with a Eu2O3/Co3O4 NSs/PMS system, where the quinolone nucleus was identified as the starting point of the reaction due to the high electron density at the π-bonding sites.62 The second pathway involves the reduction of the carboxylic acid group to a methyl group under free radical conditions, forming L4 (m/z = 332). The loss of this polar group makes the molecule more susceptible to attack by SO4˙−, especially at the piperazinyl ring, leading to ring cleavage and the formation of L5 (m/z = 223). This process reflects the common reactivity of fluoroquinolones, where the –COOH group not only plays a biological role but is also a target of advanced oxidation. The third pathway begins with the removal of a fluorine atom from the aromatic ring, forming L6 (m/z = 344). The fluorine atom plays a role in stabilizing the quinolone ring, but under catalytic conditions rich in oxygen radicals, it is easily replaced by a hydroxyl or hydrogen group. The carboxyl group is then removed, forming L7 (m/z = 286), and then the remaining rings are broken down, leading to product L8 (m/z = 139). The final products often contain phenol, aldehyde, or simple acid groups, indicating that the catalyst system is capable of comprehensively breaking down the original structure. The defluorination mechanism was also reported in the study of Foti et al. on the treatment of levofloxacin by simulated irradiation/PDS system,63 emphasizing the decisive role of fluorine in directing the reaction. The fourth pathway identified was the loss of peripheral methyl groups from the piperazinyl moiety, leading to L9 (m/z = 318). Fluorine and COOH were then further eliminated, forming L10 (m/z = 258), and finally L11 (m/z = 138), a small molecular product, representing the deep mineralization stage. Such partial ring cleavage indicated that LEVO was destroyed not only at the quinolone nucleus but also at the side chain.
To confirm the deep degradation ability of the catalyst, TOC analysis was performed simultaneously with the reaction. The results showed that TOC decreased from 8.59 mg L−1 initially to 2.58 mg L−1 after 60 min of reaction, corresponding to a mineralization yield of 69.9%, slightly higher than that of ofloxacin under equivalent conditions (Table 3).
| Time (min) | TOC (mg L−1) | Degree of mineralization (%) |
|---|---|---|
| 0 | 8.59 | 0 |
| 15 | 6.85 | 20.3 |
| 30 | 4.92 | 42.7 |
| 45 | 3.51 | 59.1 |
| 60 | 2.58 | 69.9 |
The convergence of the two pathways is the intermediate product F4 (m/z = 94), which can be formed from both F2 and F3. F4 is a small molecule, representing a simplified structure of a strongly modified aromatic ring, possibly a derivative of hydroxylated benzoic acid or simple oxidized aromatic amines. Once formed, F4 is further degraded through free radical oxidation reactions, including carbon chain cleavage and substituent removal, forming small organic acids such as formic acid, acetic acid, and finally CO2 and H2O – indicative of complete mineralization.
The TOC results obtained under the same reaction conditions (SFA 10 mg L−1, PS 2.0 mM, catalyst 0.5 g L−1, UV 365 nm) showed that the initial TOC concentration of 7.86 mg L−1 decreased to 2.21 mg L−1 after 60 min, achieving a mineralization yield of 71.9% (Table 4). The TOC reduction trend showed that the decomposition process was steady and efficient, without the accumulation of stable intermediate products.
| Time (min) | TOC (mg L−1) | Degree of mineralization (%) |
|---|---|---|
| 0 | 7.86 | 0 |
| 15 | 6.32 | 19.6 |
| 30 | 4.58 | 41.7 |
| 45 | 3.29 | 58.1 |
| 60 | 2.21 | 71.9 |
| Cycle | TOC mineralization efficiency, % | ||
|---|---|---|---|
| Ofloxacin | Levofloxacin | Sulfanilamide | |
| 1 | 69.3 | 69.9 | 71.9 |
| 2 | 69.0 | 69.2 | 71.3 |
| 3 | 67.2 | 68.5 | 70.1 |
| 4 | 65.3 | 65.9 | 68.6 |
| 5 | 62.4 | 63.2 | 66.9 |
The above results show that the CuMnO2@CoPc/Ti system has a stable structure and good resistance to leaching. In particular, the higher stability of the catalyst in SFA treatment may originate from the simpler molecular structure of SFA, which generates less intermediate products attached to the catalyst surface compared to fluoroquinolones such as OFL and LEVO. On the contrary, the decomposition of OFL and LEVO may generate some hydroxylated aromatic intermediate products or amino acids, which partially hinder the catalytic surface activity in the following cycles. However, the performance degradation after 5 cycles is below 7%, which shows the durable working ability and wide application of the studied catalyst.
In addition, SEM-EDS analysis after 5 cycles did not show any obvious changes in surface morphology or elemental distribution of Cu, Mn and Co on the titanium support, demonstrating that the active sites were stably fixed in the catalyst network. Phthalocyanine cobalt, thanks to its π–π bond and strong electronic interaction with the metal oxide, was not easily washed away, maintaining its effective PS activation role over many cycles (Fig. 16).
ICP-OES analysis of the post-reaction solution confirmed minimal metal leaching: the leaching concentrations (mg L−1) of the metals from the 1st to the 5th cycles were Cu 0.008 → 0.012, Mn 0.006 → 0.009, and Co 0.004 → 0.006, respectively, equivalent to approximately 40–60 μg g−1 catalyst per cycle, demonstrating the strong chemical stability of the catalyst (Table S3).
From the above experimental results and discussions, it can be affirmed that the CuMnO2@CoPc/Ti catalyst system not only has high efficiency in degrading three typical antibiotics representing different structural groups, but also has stable durability and reusability after many reaction cycles. This opens up the potential for wide application of the material in the treatment of wastewater containing antibiotics in hospitals, pharmaceutical factories, or water treatment systems and water reuse in the future.
From a process standpoint, the catalyst's durability and separability make it a credible platform for UV/PS polishing of antibiotic-bearing waters. Practical deployment should include residual persulfate quenching (e.g., thiosulfate), sulfate/TDS management, and—where water reuse is targeted—downstream polishing (biologically active filtration and/or activated carbon) plus final disinfection to meet fit-for-purpose quality criteria.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05544j.
| This journal is © The Royal Society of Chemistry 2025 |