Lucas
Santos-Juanes
a,
Noelia
Rodriguez-Sanchez
bc,
Salvador R. G.
Balestra
cde,
Nuria O.
Núñez
f,
Antonio
Arques
a,
A. Rabdel
Ruiz-Salvador
*cd and
Menta
Ballesteros
*cg
aGrupo de Procesos de Oxidación Avanzada, Departamento de Ingeniería Textil y Papelera, Universitat Politècnica de València, Campus de Alcoy, Alcoy, Spain
bBAM Federal Institute for Materials Research and Testing, Richard-Willstatter-Str, Berlin, 11 12489, Germany
cCentro de Nanociencia y Tecnologías Sostenibles (CNATS), Universidad Pablo de Olavide, Carretera de Utrera km. 1, Sevilla, Spain. E-mail: rruisal@upo.es; mmbalmar@upo.es
dDepartamento de Sistemas Físicos, Químicos y Naturales, Universidad Pablo de Olavide, Ctra. Utrera, km. 1, Sevilla, Spain
eDepartamento de Física Atómica, Molecular y Nuclear, Área de Física Teórica, Universidad de Sevilla, Avenida de Reina Mercedes s/n, Sevilla, 41012, Spain
fInstituto de Ciencia de Materiales de Sevilla, ICMS (CSIC-US), c/Américo Vespucio, 49, Sevilla, Spain
gDepartamento de Biología Molecular e Ingeniería Bioquímica, Universidad Pablo de Olavide, Ctra. Utrera km. 1, Sevilla, Spain
First published on 1st May 2025
Metal–organic frameworks (MOFs) have recently been proposed as a plausible solution to the pressing issue of water scarcity and as a means of remediating contaminated water bodies. In light-assisted water treatment, they have so far only been exploited via the hydroxyl radical route, through Fenton-like processes. A new avenue is introduced here by the biomimetic conceptual design of MOF bearing hypervalent metal atoms for photocatalytic water treatment. We report a zeolitic imidazole framework (ZIF) material doped with iron (Fe-ZIF-7-III; UPO-4) synthesized via a novel mild treatment to stabilize photoactive hypervalent ferryl ions for the first time in a MOF for water treatment. The successful synthesis of the 2D material and the adequate incorporation of iron into the structure were demonstrated using X-ray diffraction (XRD), Fourier-transform infrared (FTIR) spectroscopy, and scanning electron microscopy (SEM). A simulation study analyzed the structure and stability of the Fe-ZIF-7-III material as well as the involvement of ferryl ions in the photo-Fenton-type process. Furthermore, the calculated band gap of this material shows its viability for use in photocatalysis using sunlight. This was confirmed by evaluating the photodegradation of caffeine, a model pollutant in water, without the assistance of hydroxyl radicals as indicated by a scavenger test. The recyclability test revealed that Fe-ZIF-7-III could be used continuously with effective catalytic activity, thus opening the door to the field of studying hypervalent metal MOFs not yet explored in water treatment.
Advanced oxidation processes (AOPs) play a central role in water treatment for the removal of hazardous contaminants. Among them, those using environmentally friendly energy sources are being extensively investigated, especially those using solar energy, which have gained attention for their ability to degrade persistent pollutants, often using photo-assisted reactions such as the photo-Fenton process or semiconductor-mediated photocatalysis.20,21 In this context, heterogeneous catalysis is attracting much attention in the field of AOPs, stimulated by its well-known advantages in implementation and reuse.22–25 The versatility and amenable design of MOFs for catalytic applications22,26–28 lead these materials to be used as heterogeneous catalysts in AOPs for water treatment. The work reported to date using MOFs in this area uses Fenton and Fenton-like reactions, which are based on radicals derived from hydrogen peroxide, sulphate, or other oxidants as reactive species. Since AOPs can also be achieved by routes that do not involve these radicals, but rather hypervalent metal ions,23 it is interesting to investigate whether such an approach could be implemented with MOFs.
We have noted that the use of ferrihydrite iron oxydroxide with hypervalent ferryl ions as a key oxidant in organic matter degradation reactions has been described.29 Similar to Long et al., we have chosen a biomimetic molecular environment to design our material with stable isolated hypervalent metal sites for water treatment. Learning from nature, it is known that iron coordinated with multiple nitrogen atoms, as in porphyrinoid systems, can form ferryl ions with significant catalytic activity.30,31 Such an N-rich molecular environment could be replicated in Fe-bearing MOFs, making Fe-ZIFs (ZIF stands for zeolitic imidazolate framework) a natural choice for exploration. ZIFs, with topologies similar to those of zeolites, are a subclass of MOFs composed of tetrahedrally coordinated transition metal ions and organic imidazole ligands.32 They have excellent thermal and hydrolytic stability, high surface area, and abundant catalytic sites.33,34 Previous studies have primarily used specific ZIFs, such as ZIF-9 and ZIF-67, to activate ˙OH or other reactive species for the degradation of environmental contaminants. For example, ZIF-9 and ZIF-12 have been used to activate peroxymonosulfate (PMS) for the degradation of contaminants such as rhodamine B and tetracycline.35,36 Cu-doped ZIF-67 and CuCo-ZIFs have shown promise as visible light-driven photocatalysts for dye degradation via Fenton-related reactions and α-Fe2O3/ZIF-67 was able to completely degrade ciprofloxacin in 30 min via SO4˙− and ˙OH radicals.37
In designing the targeted hypervalent ferryl ion-containing MOF for water remediation, we would need a very stable material as a proof of concept. According to Amombo Noa et al.,38 the order of stability increases from dot, rod to sheet or 2-D MOFs. In the versatile ZIF-7 family, the 2-D ZIF-7-III member emerges as the most stable phase.39 It should be noted that the 2-D ZIF-L solid, derived from ZIF-8, is also the most stable phase of this system.40 The choice of ZIF-7-III over ZIF-L is due to the presence of benzene chromophores, which enhance its stability and light absorption properties, making it effective for photocatalysis.39,41 The incorporation of benzimidazole lowers the conduction band energy compared to pure imidazole and methylimidazole, further improving the photocatalytic efficiency.42 In addition, the incorporation of transition metals can optimize light absorption by enhancing the valence band.42
Building on these insights, our study aims to explore the potential of Fe-doped ZIF-7-III (UPO-4) as a heterogeneous catalyst for photocatalytic AOPs at circumneutral pH under simulated solar irradiation. Caffeine was selected as the target pollutant due to its stability and inability to complex with iron, which ensures that it does not interfere with the Fenton process.43,44 We further investigated the nature of photoreactivity using radical scavengers and conducted DFT-based simulations to better understand the structural and electronic properties of Fe-ZIF-7-III in relation to its photocatalytic performance.
The irradiation was performed with a solar simulator (Oriel Instruments, Model 81160 equipped with a 300W xenon lamp). Glass filters were used to cut off the transmission of wavelengths λ < 300 nm. UV-A irradiance in the 315–380 nm region was 32 W m−2. The irradiation started simultaneously with the addition of hydrogen peroxide and was kept for up to 180 min. Samples were periodically taken from the solution for analysis; those to be analysed by HPLC were diluted 1:
2 with methanol to quench the excess peroxide. For the reuse of the catalyst test, three cycles of oxidation were performed with the same catalyst, adding the corresponding amount of caffeine to achieve the same initial concentration. In a complementary experiment, tert-butyl alcohol (100 μM) was added to study if ˙OH radicals contribute to caffeine degradation, as expected in Fenton-like processes. To check the possible involvement of hypervalent iron ions, tests were performed in the presence of dimethyl sulfoxide (DMSO) at a concentration of 100 μM.
Dark controls were performed with the catalyst and with and without hydrogen peroxide and irradiation of the caffeine solution with H2O2 and in the absence of the catalyst was also performed.
All experiments were performed in duplicate.
The DFT-simulated structures of ZIF-7-III with and without Fe atoms were very similar (Fig. 2). This suggests that the layers are sufficiently flexible to host Fe atoms without a large overall deformation. Note that both Fe2+ and Fe3+ were considered, as they are expected to appear in the prepared materials. A closer look at the structures shows that the local environment is modified by the presence of Fe, shown in Fig. 2, which induces orientation of the benzimidazole ligands, along with small displacement and together lead to reduce the atomic overlap. This view helps rationalize the partial loss of crystallinity observed in X-ray diffraction (Fig. 1, ∼ 16–24°). Due to the high cost of DFT calculations, our simulated systems contained only 466 atoms.
In real systems, many different configurations are likely to appear, giving rise to a variety of local deformations, which explains the partial loss of crystallinity. It is also noted that the main periodicity perpendicular to the basal plane, retrieved from the diffraction peak at 9.6° in Fig. 1, is apparent from the left panel of Fig. 2, as the layers are nicely parallel. Using infrared spectroscopy, as a local probe, the structural similarity of ZIF-7-III with and without Fe was confirmed (Fig. 3). The bands observed in the FTIR spectra agree with those found by other authors for ZIF-7-III,63 where most of the observed peaks were obtained below 2000 cm−1 due to the vibration energy of the different types of bonds in the structure. Therefore, the characteristic peaks from 550 to 900 cm−1 are due to aromatic sp2 C–H bending, those from 1000 to 1300 cm−1 are related to in-plane bending of the ring, and those at 1600 and 1800 cm−1 correspond to CN and C
O stretching, respectively. The incorporation of Fe atoms was confirmed by the large peak appearing at 615 cm−1, which agrees with that measured in Fe-doped ZIF-8.64
Although diffraction and FTIR did not show large differences in ZIF-7-III before and after Fe doping, SEM images did (Fig. 4). Therefore, iron affects the crystallization of the synthesized solids. The Fe-ZIF-7-III powder exhibited an almost 2D dense stacked layer structure, with several micrometres in size (around 2.0–2.5 μm) and a micrometre size planar shape and a smooth surface (Fig. 4), which is consistent with the previously reported ZIF-7-III.61,63 Under our synthesis conditions, the ZIF-7-III powder exhibited a whisker morphology with the presence of a minor amount of lamellar particles.
While iron strongly modifies the morphology and generates disorder in the layer stacking, it contrasts with its low incorporation into the material. ICP revealed that the Fe concentration was only 0.263 ± 0.007 ppm (0.16 wt%). Nevertheless, this is sufficient to provide measurable visible-light absorption, as detected by diffuse-reflectance UV-vis spectroscopy (Fig. 5). To understand the electronic and optical properties of MOFs, it is necessary to consider that electron transitions can occur within MOFs involving the movement of electrons from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO).65 To achieve optimal excitation of the MOFs, incident light (Elight = 1240/λ > EHOMO–LUMO) was limited to the UV region, characterized by shorter wavelengths (λ < 365 nm). It is therefore evident that MOFs capable of responsively absorbing visible light (λ > 400 nm, or Elight < 3.1 eV) are to be preferred for the effective utilization of solar energy, as is commonly the case with Fe-MOFs.65,66
Grau-Crespo et al.42 showed that Zn-ZIFs have band gaps that are determined by the HOMO–LUMO of the ligands. In the case of ZIF-7, the calculated band gap was 4.50 eV due to the benzimidazole ligand. This explains the presence of a high band in the experimental UV-vis spectrum of ZIF-7-III at 280 nm, since this MOF is formed by the same ligand (Fig. 5). We turn to the first-principles computed bandgap of ZIF-7-III, obtaining a value of 4.39 eV (282 nm), which confirms that indeed this band arises from tetrahedral zinc and benzimidazole. When iron was incorporated into ZIF-7-III the experimental spectrum showed an additional band at 375 nm (3.30 eV), and a shoulder at 550 nm (2.25 eV). To confirm that iron is related to observed differences in UV-vis spectra, the band gaps of Fe2+-ZIF-7-III and Fe3+Cl-ZIF-7-III were calculated. The calculated band gaps shift from 4.39 eV to 3.60 eV (344 nm) and 2.45 eV (505 nm) for Fe2+-ZIF-7-III and Fe3+Cl−-ZIF-7-III, respectively, which is in reasonable agreement with the experimental light absorption bands. Other authors have also found that the incorporation of iron in ZIF-8-NH2 reduces the band gap from 5.11 eV to 2.17 eV.67 We therefore expect photochemical activity in Fe-ZIF-7-III that could be exploited directly for solar heterogeneous photo-Fenton catalysis or for the photo-activation of hydrogen peroxide to form ferryl ions.
Considering that the presence of iron modifies the electronic structure of ZIF-7-III materials, it is expected that the reactivity would also be altered. To predict the reactivity potential of these materials, Fukui functions were analysed because they relate the change in chemical potential as a function of the variation in the number of electrons.68 A high value of f+(r) indicates that an atom is more likely to be attacked in a nucleophilic reaction, or that these regions will readily donate charge to the electron acceptor. In contrast, a high value of f−(r) indicates that the atom is more likely to undergo an electrophilic reaction or that these regions will stabilize the uptake of charge from electron donors.65,69 Therefore, Fig. 6 illustrates the clear differences among the three materials. In Fig. 6a, we demonstrate how ZIF-7-III (without Fe atoms) exhibits low reactivity, except for the metallic centers (Zn atoms), which are difficult to access due to steric hindrance. The presence of Fe2+ atoms (Fig. 6b) enhances the reactivity of the metallic center. Calculations revealed that within the three-dimensional lattice, ferrous ions can act as electrophilic agents, decomposing H2O2, which can result in two different scenarios (˙OH generation via a Fenton-like reaction or the formation of hypervalent ferryl ions). The presence of Cl atoms (Fig. 6c) significantly increased the reactivity of the material. Most surface aromatic groups become susceptible to electrophilic and nucleophilic attacks, such as those that initiate the Criegee mechanism.
Once the synthesized material was characterized and the presence of iron in the material was assessed, different experiments were carried out to check its activity in photo-degradation processes. First, caffeine irradiation alone, with H2O2 or with Fe-ZIF-7-III (UPO-4) in parallel experiments, resulted in negligible degradation of the pollutant after 90 min, which allowed us to rule out direct photolysis, oxidation by H2O2, or photocatalytic activity of Fe-ZIF-7-III. The caffeine concentration also remained constant in dark controls consisting of caffeine with Fe-ZIF-7-III, indicating that the adsorption of the pollutant on the catalyst surface was negligible.
Next, the ability of the material to drive a (photo)-catalytic decontamination process at pH = 6.4 was tested. For this purpose, experiments were carried out in the presence of Fe-ZIF-7-III (200 mg L−1) and hydrogen peroxide (twice the stoichiometric amount) to oxidize 4 mg L−1 caffeine in the dark and under solar simulated irradiation. Approximately 80% caffeine removal was reached after 90 min of irradiation and more than 90% abatement after 180 min (Fig. 7) and the dark experiment resulted in negligible pollutant degradation (data not shown). These results agree with the improved photoreduction of Fe(III) present in the solid material, which cannot be reached in the dark. Thus, light irradiation is an essential component in the catalytic activity of Fe-ZIF-7-III, at near neutral pH, at least for the removal of very stable molecules such as caffeine.
We have noted that there are no studies reporting caffeine degradation by photo-Fenton-like processes using ZIF catalysts. To provide a comparative view of the performance of our material, we show here some prominent examples of photocatalytic caffeine degradation. Refluxed MIL-125 over NTU-9 shows 85% of caffeine degradation under 105 min of visible light.70 A nano-TiO2 photocatalyst was used to remove 94% of caffeine within 220 min at pH 3 and pH 1, while 88% and 80% of caffeine were removed within 220 min at pH 6 and pH 9, respectively, under UV light.71 Additionally, the application of a three-layer perovskite Dion-Jacobson phase CsBa2M3O10 (M = Ta and Nb) and oxynitride crystals was applied for the removal of caffeine by photocatalysis under UVA and visible light irradiation, achieving 89% degradation under UVA light after 60 min.72 Also, ZnO nanoparticles achieved 98.6% degradation of a 1 ppm caffeine solution under sunlight in 120–140 min.73 Iron oxide magnetic nanoparticles (MH0.2) combined with H2O2 achieved significant caffeine removal (5 mg L−1) after 180 min.74 Similar to this photocatalyst, degradation of caffeine was achieved under UV-vis radiation using magnetic nanoparticles (NPs) based on iron oxides after 60 min.75 Fe-ZIF-7-III (UPO-4) exhibits significantly better performance under simulated sunlight, achieving higher degradation rates and superior catalytic activity while maintaining structural stability without detectable iron leaching, thus ensuring a truly heterogeneous process, unlike other systems that often suffer from stability issues and iron leaching, leading to a homogeneous-like mechanism.
It should be noted that in some cases solids have been reported to act as simple iron reservoirs, which are leached into the solution, thus leading to a predominating homogeneous photo-Fenton process.76 To determine the extent of this mechanism, dissolved iron was measured according to the o-phenanthroline colorimetric method and ICP. In none of the measurements, dissolved iron was detected, which indicates that the role, if any, of leached iron cations in the observed caffeine removal was very marginal. It was confirmed by EDX before and after the photocatalytic reactions, as the atomic percentage of Fe (Fig. S1, ESI†) showed only a slight decrease from 3.06% to 2.82%. The X-ray diffraction pattern and SEM also reveal the robustness of the material after photocatalysis (Fig. S2, ESI†). XPS analysis (Fig. S3, ESI†) further supports the stability of the material, showing that prior to the reaction, Fe was predominantly in the Fe3+ oxidation state with a small contribution of Fe2+. Before the reaction, a main peak was observed in the Fe 2p3/2 region, with a predominant contribution from Fe3+ (712.5 eV) and a smaller fraction of Fe2+ (710.9 eV). However, after photocatalysis, no clear Fe2+ or Fe3+ signals were detected, suggesting that the remaining iron has transitioned to a higher oxidation state, such as a ferryl species, which is not readily identified by XPS. This interpretation is consistent with previous reports indicating that XPS is unable to directly confirm the presence of high valence iron species.77
To determine whether the process was indeed controlled by the targeted ferryl ion or mediated by ˙OH radicals, as in photo-Fenton processes, photocatalytic experiments were repeated in the presence of t-butanol, a known scavenger of hydroxyl radicals. Fig. 7 shows that caffeine removal coincided with and without t-butanol. Based on these results, it seems that the role of ˙OH is not predominant in this case, in contrast to most information reported on homogeneous and heterogeneous photo-Fenton reactions.44,66 Therefore, these results support the hypothesis of the presence of ferryl ions in Fe-ZIF-7-III.
To determine the participation of hypervalent iron ions, that is, the ferryl ions, experiments were carried out in the presence of a compound that deactivates this transient state of iron. This compound was dimethyl sulfoxide (DMSO). In Fig. 7, the experiment in the presence of DMSO is presented and an important inhibition of caffeine removal can be observed. These results would indicate an important role of the ferryl ion in the degradation of the model pollutant. The reactions involved could be represented, in a simplified way, as follows:
Fe(II) + H2O2 → [FeIVO]2+ + H2O |
[FeIVO]2+ + pollutant → Fe (II) + pollutantox |
Fe in N-rich environments has been reported to be easily activated to adopt hypervalent states, such as ferryl ions.30,31 One distinctive feature from a structural point of view is the short Fe–O distance of ca. 1.63 Å.30,31 On the basis of this background and the above catalytic experiments, to confirm our hypothesis, we performed DFT calculations to elucidate whether hydrogen peroxide could lead to the formation of stable ferryl groups in Fe-ZIF-7-III. Fig. 8 shows the energy minimized structure of a hydrogen peroxide molecule and a water molecule adsorbed near the Fe atom in Fe-ZIF-7-III and a ferryl group with two adsorbed water molecules in Fe-ZIF-7-III. The FeO distance in the ferryl group is 1.646 Å, which is in good agreement with the experimental distances measured by EXAFS in N-rich environments. The formation of the Fe
O bond leads to an increase in the Fe atomic charge from 0.875e to 1.102e, while the Zn charges remain unchanged at ca. 0.948e, which is also evidence of the formation of the ferryl group. It is useful to note that the energy of these structures only differs by 1.44 eV per Fe atom, an energy value that is compatible with visible light. The computed energy difference suggests that once the ferryl groups are formed, they can be stabilized in the material, and thus they can be available as alternative active sites for photocatalysis. The calculated band gap of the ferryl group bearing Fe-ZIF-7-III is 2.116 eV, indicating that the material retains the ability to absorb light in the visible region (below 586 nm). The degradation experiments shown above indicated that light activation is a key point in the catalytic process. We acknowledge that the elucidation of the mechanism is a difficult task, as it would be needed to consider excited state reactions. Thus, the mechanism of the photocatalytic reactions involved in the degradation of the caffeine molecule in the presence of the ferryl ion anchored in ZIF-7-III is beyond this study and will be treated in a separate study.
To check the reusability of the catalyst, we used it in three consecutive cycles of photocatalytic degradation of caffeine. After 90 min of irradiation, the initial conditions were reset by adding the amount of caffeine and H2O2 consumed. To examine better the changes among cycles, caffeine degradation was fitted to a first-order kinetics and pseudo-rate constants (k) were calculated in each case (Fig. 9). For the first cycle, a k value of 0.0073 min−1 was obtained; there was a decrease in the second cycle to 0.047 min−1 and then remained constant in the third cycle (0.049 min−1). Despite a slight initial loss of efficiency, the material appeared to maintain its activity as a photocatalyst, an important result in terms of its re-use.
The activity of the synthesized material as a photocatalyst was demonstrated through the successful degradation in water of the model contaminant caffeine using simulated sunlight irradiation. The presence of a specific ferryl ion scavenger clearly affected the behaviour of the process indicating a mechanism without the major intervention of hydroxyl radicals, supporting the hypothesis of the presence of ferryl ions. To support this point, DFT calculations were performed. Innovations include the successful stabilization of hypervalent iron within the MOF, marking a novel approach in the design of photocatalysts for water treatments as there is no previous work in this area using a hypervalent metal MOF. Therefore, Fe-ZIF-7-III (UPO-4) represents a significant advancement by effectively leveraging solar energy for environmental remediation, while maintaining catalytic activity across multiple cycles. Future work should focus on the effect of photo-Fenton parameters on degradation kinetics, optimizing iron incorporation, broadening the scope of target pollutants, and elucidating the mechanisms behind photocatalytic activity, ultimately contributing to sustainable water treatment solutions. Further work is underway to design and test other hypervalent metals and MOFs for water decontamination and disinfection.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ma01217h |
This journal is © The Royal Society of Chemistry 2025 |