Next Article in Journal
Targeted Enrichment and Characterization of Diester Diterpenoid Alkaloids in Aconitum Herbs Using Gas–Liquid Microextraction Coupled with High-Resolution Mass Spectrometry
Previous Article in Journal
Isopropanol Electro-Oxidation on PtCu Alloys for Aqueous Organic Redox Chemistry Toward Energy Storage
Previous Article in Special Issue
Plasmonic Effect of Au Nanoparticles Deposited onto TiO2-Impact on the Photocatalytic Conversion of Acetaldehyde
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights of Nanostructured Ferberite as Photocatalyst, Growth Mechanism and Photodegradation Under H2O2-Assisted Sunlight

by
Andarair Gomes dos Santos
1,2,
Yassine Elaadssi
1,
Virginie Chevallier
1,
Christine Leroux
1,
Andre Luis Lopes-Moriyama
1,3 and
Madjid Arab
1,*
1
Université de Toulon, Aix Marseille Univ, CNRS, IM2NP, 13397 Marseille, France
2
CCEN, UFERSA, Campus Mossoró-F. Mota, Costa e Silva, Universidade Federal Rural do Semi-Árido, Mossoró 59625-900, RN, Brazil
3
Campus Universitário, L. Nova, Universidade Federal do Rio Grande do Norte, Natal 59072-970, RN, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(19), 4026; https://doi.org/10.3390/molecules30194026
Submission received: 12 July 2025 / Revised: 3 October 2025 / Accepted: 7 October 2025 / Published: 9 October 2025
(This article belongs to the Special Issue Research on Heterogeneous Catalysis—2nd Edition)

Abstract

In this study, nanostructured ferberites (FeWO4) were synthesized via hydrothermal routes in an acidic medium. It was then investigated as an efficient photocatalyst for degrading organic dye molecules, with methylene blue (MB) as a model pollutant. The formation mechanism of ferberite revealed that the physical form of the precursor, FeSO4·7H2O, acts as a decisive factor in morphological evolution. Depending on whether it is in a solid or dilute solution form, two distinct nanostructures are produced: nanoplatelets and self-organized microspheres. Both structures are composed of stoichiometric FeWO4 (Fe: 49%, W: 51%) in a single monoclinic phase (space group P2/c:1) with high purity and crystallinity. The p-type semiconductor behavior was confirmed using Mott–Schottky model and the optical analysis, resulting in small band gap energies (≈1.7 eV) favoring visible absorption light. Photocatalytic tests under simulated solar irradiation revealed rapid and efficient degradation in less than 10 min under near-industrial conditions (pH 5). This was achieved using only a ferberite catalyst and a low concentration of H2O2 (4 mM) without additives, dopants, or artificial light sources. Advanced studies based on photocurrent measurements, trapping and stability tests were carried out to identify the main reactive species involved in the photocatalytic process and better understanding of photodegradation mechanisms. These results demonstrate the potential of nanostructured FeWO4 as a sustainable and effective photocatalyst for water purification applications.

1. Introduction

In the textile industry, approximately 800,000 tons of dyes are produced annually, with an estimated 15–20% being lost to wash water during the dyeing process. These dyes, predominantly synthetic, are characterized by complex molecular structures, high water solubility, and strong chemical stability, making them difficult to remove through conventional water treatment methods [1,2,3,4]. Traditional approaches include physical, chemical, and biological processes, as well as advanced oxidation processes (AOPs), each with advantages and limitations. While AOPs are effective, they are often costly and may produce harmful by-products.
In recent years, green technologies such as adsorbents and photocatalysis have gained increasing attention. Unlike adsorption, which merely transfers pollutants between phases, the heterogeneous photocatalysis enables the complete mineralization of contaminants without generating toxic intermediates [5,6,7]. It further offers catalyst reusability, high stability, reduced operating costs, minimal solid waste and low energy consumption, making it both environmentally and economically sustainable [8]. Various semiconductors, including TiO2 [9], ZnO [10], hematite and iron oxides [11], g-C3N4 [12], and tungstate or molybdate-based materials [13,14,15,16,17], have been investigated for pollutant degradation. TiO2 and ZnO remain the most studied due to their stability and low cost, but their activity is limited to the UV range unless doped.
Recent advances in photocatalysis for textile and wastewater treatment have therefore focused on heterostructured systems to improve charge separation and extend visible-light response. Examples include S-scheme heterojunctions for antibiotic degradation [18], core–shell CdS@ZnInS4 nanocomposites for dye reduction and stabilization against photocorrosion [19], Mn-doped Cd1−xS nanorods with optimized band gap engineering [20] and MIL-101(Fe)/Bi2MoO6 heterostructures for tetracycline removal [21].
Iron tungstate (FeWO4), also known as ferberite, is a wolframite-type ternary semiconductor [22,23,24,25,26,27] with intrinsic ferromagnetism, efficient visible light absorption, and promising photocatalytic activity [28]. Several studies report its use for degrading dyes such as methylene blue (MB), methyl orange, and rhodamine B under visible or simulated solar irradiation [22,23,24,25,29]. However, pristine FeWO4 usually shows limited catalytic activity, prompting research into dopants or heterostructures to boost performance. Examples include halogen ion doping, coupling with graphene oxide (GO) or forming heterostructures such as FeWO4/g-C3N4 or FeWO4/BiOCl to facilitate electron transfer and reduce electron–hole pair recombination, which significantly accelerate dye degradation—reported rate being several times faster than those obtained with bare FeWO4 [30,31,32].
These approaches confirm the interest in FeWO4 but rely on complex and costly modifications. In contrast, FeWO4 itself remains underexplored, especially regarding its growth mechanism and the role of synthesis parameters in controlling morphology and photocatalytic activity [22,25,33,34,35]. Previous hydrothermal studies [22,23,24,25,29,36,37] have only rarely addressed formation pathways, with notable exceptions being Fang et al. [35], who examined pH-dependent synthesis using iron ammonium sulfate, and Zhang et al. [33], who proposed a mechanism involving Fe–EDTA complexes.
Building on this background, the present work highlights that even pristine FeWO4, obtained from common precursors under mild hydrothermal conditions, can achieve rapid and efficient MB degradation under solar irradiation. A key originality lies in showing that morphological control alone, governed solely by the physical form of the FeSO4·7H2O precursor (solid vs. dilute), is sufficient to tune photocatalytic performance. This simple adjustment drives a dual growth mechanism, leading to two distinct morphologies (nanoplatelets and self-organized). Despite their relatively low surface area, both morphologies achieved near-complete MB degradation within minutes, with only low doses of H2O2. This distinguishes our work from previous reports where FeWO4 performance improvements required doping, heterostructures, or supports.
Here, a combination of structural, optical, electrochemical, kinetic, and radical analyses is employed to elucidate the growth mechanism of FeWO4 and clarify its photocatalytic activity. We examined the influence of pH and varied the oxidant concentration. Radical scavenging tests and electrochemical characterization (photocurrent response and Mott–Schottky analysis) were combined with band alignment to propose a consistent photo-Fenton mechanism. This integrated approach not only provides new insights into ferberite growth and activity, but also highlights its potential as an efficient, cost-effective, and scalable photocatalyst for textile wastewater treatment.

2. Results and Discussion

2.1. Structural and Chemical Characterization of Ferberite

Single-crystal ferberite powders were synthesized by introducing iron (II) sulfate heptahydrate either in solid form or as an aqueous solution, resulting in two distinct morphologies: self-organized microstructures composed of platelets, and isolated nanoplatelets. Figure 1 displays the X-ray diffraction (XRD) patterns and corresponding refinement results for the crystallized samples self-organized (a) and platelets (b). It is important to note that, regardless of the resulting morphology, the formation of the ferberite phase requires the careful optimization of synthesis parameters, including the type and concentration of precursors and solvents, reaction medium pH, and hydrothermal temperature. The synthesis conditions reported herein were specifically optimized to obtain phase-pure ferberite, free from secondary phases or impurities.
According to the XRD patterns shown in Figure 1, both samples exhibit well-crystallized structures with no detectable secondary phases, confirming the formation of a pure monoclinic phase with space group P2/c:1, consistent with the ICSD 15193 reference pattern. The diffraction profiles both in terms of peak intensity and width vary depending on the synthesis conditions, particularly whether iron (II) sulfate was introduced in solid form or as a solution. Differences in peak broadening between the samples reflect variations in crystallite size, as summarized in Table 1. Notably, the self-organized morphology displays broader and less intense peaks compared to the platelets, suggesting either lower crystallinity, smaller crystallite size, or a preferred crystallographic orientation associated with the anisotropic growth of platelets.
The structural and angular parameters, along with the crystallite size, were determined via Rietveld refinement using the MAUD software (version 2.064), based on the identified monoclinic space group P2/c:1. In Figure 1, the solid black line represents the experimental XRD data, the red line corresponds to the calculated pattern obtained from the Rietveld refinement, and the blue line indicates the difference (residual) between the experimental and calculated intensities. The refined crystallographic parameters, lattice constants, and reliability factors are summarized in Table 1.
The analysis of lattice parameters, crystallite size, and theoretical density presented in Table 1 indicates that the results obtained from the Rietveld refinement are consistent with values reported in the literature [28,37,38]. The quality of the refinement is supported by the reliability indicators: Rwp, Rexp, and the goodness of fit (Rwp/Rexp). Based on these values, the Rietveld refinement can be considered statistically reliable.
The refined lattice parameters obtained from the Rietveld analysis were used in the VESTA software (Visualization for Electronic and Structural Analysis, Version 3.90.0a) to construct the geometric model and evaluate interatomic distances and bond angles. The crystal structure of monoclinic FeWO4 is illustrated in Figure 1b. It features a multilayered arrangement composed of alternating polyhedral units and linear atomic chains involving Fe, W, and O atoms. In this representation, both W and Fe atoms are octahedrally coordinated by oxygen, forming WO6 and FeO6 structural units. The coordination parameters derived from the Rietveld refinement enabled the determination of bond lengths, with W–O distances ranging from 1.883 to 2.109 Å and Fe–O distances from 2.003 to 2.170 Å. The associated bond angles [O–W(Fe)–O] vary between 40° and 166°, indicating significant distortions in the polyhedral environment. These distortions are characterized by a stronger contraction of the WO6 octahedra and a relative elongation of the FeO6 units.
The synthesized structures exhibited good structural stability and retained their morphology after ultrasonic dispersion, suggesting strong chemical bonding within the polyhedral framework. TEM-EDX analyses were conducted to assess the local chemical composition and atomic distribution of Fe and W across different microcrystalline regions. Regardless of morphology self-organized or nanoplatelets, the EDX results confirmed a stoichiometry consistent with the expected Fe:W atomic ratio in FeWO4. Specifically, the self-organized samples contained 47% Fe and 53% W, while the platelet-shaped samples exhibited 48% Fe and 52% W. Notably, achieving the correct stoichiometry for phase-pure ferberite using the synthesis method reported herein required an initial iron concentration twice that of tungsten and oxalic acid.
Ferberite has also been synthesized via hydrothermal methods using different precursors, such as ferrous ammonium sulfate [(NH4)2Fe(SO4)2·6H2O] and sodium tungstate [Na2WO4·2H2O], followed by hydrothermal treatment at 190 °C for 24 h across a wide pH range (1–12) [23]. However, ferberite was not formed under all conditions: at pH 1, WO3 was obtained, while at pH 12, Fe2O3 was the predominant phase. In contrast, in our study, the pure ferberite phase was successfully obtained at pH 1. This discrepancy could be primarily attributed to the nature of the iron precursor, as the hydrothermal temperatures used in both cases were comparable (approximately 200 °C). Kovács et al. [39] also reported the critical influence of the iron precursor on the hydrothermal synthesis of ferberite, demonstrating that even under otherwise identical conditions, changing the iron source alone could inhibit the formation of the FeWO4 phase.

2.2. Morphological Characterization of Ferberite

Figure 2 presents the morphological characterization of the ferberite samples, which exhibit self-organized desert rose-like structures and isolated nanoplatelet morphologies. These features were analyzed by scanning electron microscopy (SEM, images (a)–(c) and transmission electron microscopy (TEM, image (d)). The electron diffraction patterns of the self-organized/platelet morphologies are shown in Figure S1 (TEM, Supplementary Material). The SEM images were acquired at different magnifications to highlight structural details: 1.2 k× for image (a), and 70 k× for images (b) and (c).
As observed in the SEM images, the self-organized ferberite sample primarily consists of microstructures resembling desert roses, characterized by radially arranged platelets forming rosette-like architectures (Figure 2a,b). These results are consistent with those reported by Kovács et al. [39]. In contrast, the second sample, obtained using diluted iron(II) sulfate heptahydrate, primarily exhibits irregular rectangular and square-shaped platelets with varying sizes (Figure 2c,d). While the self-organized morphology shows a more uniform size distribution, the individual platelets in the second sample are considerably smaller, despite their irregular shape. The morphological contrast between Figure 2a,c clearly highlights the influence of the physical form of the iron precursor solid or dissolved on the resulting ferberite morphology, even in the presence of oxalic acid. This simple variation in precursor introduction leads to a transition from well-organized rosette-like assemblies to disordered and isolated nanoplatelets. Selected-area electron diffraction (SAED) patterns confirm that the plates for both samples are single crystals with a (100) zone axis aligned along the a-axis. This indicates preferential two-dimensional growth along the b and c directions, with growth suppression along the large surface area. This anisotropic growth behavior is primarily driven by the presence of oxalic acid and further modulated by whether the iron precursor is added in diluted solution or solid form.

2.3. Growth Mechanism of Ferberite Catalyst

The formation mechanism of ferberite was investigated by analyzing samples collected at different stages of growth using microscopy and infrared spectroscopy. The hydrothermal synthesis produced two distinct morphologies: isolated nanoplatelets and self-organized aggregates resembling desert rose-like architectures. Experimental evidence suggests that the physical form in which iron (II) sulfate heptahydrate is introduced either as a solid or in solution plays a key role in directing the morphological evolution toward either individual platelets or their self-organized into rosette-like structures. Overall, both morphologies appear to originate from the initial formation of single-crystalline platelets.

2.3.1. Sodium Tungstate Dihydrate Dissolution and Role of Oxalic Acid

Initially, sodium tungstate dihydrate was dissolved as described in Equation (1), resulting in a clear solution. Subsequent acidification with HCl to pH 1 leads to the protonation of tungstate ions (Equation (2)). At this first stage, the solution undergoes a visible color change from colorless to pale green, indicating the formation of hydrated tungstic acid species (H2WO4·nH2O), as represented in Equation (3). In the second stage, the WO5(H2O) units become further hydrated and undergo dimerization through oxygen bridges, forming crystalline WO3·2H2O, in which tungsten centers are octahedrally coordinated. The WO3·2H2O structure consists of layers of corner-sharing WO5(H2O) octahedra separated by interlayer water molecules. Each WO5(H2O) octahedron contains four W–O single bonds, one terminal W=O double bond, and one W–OH2 bond, where the tungsten center is coordinated in relation to a water molecule. The interlayer water molecules form hydrogen bonds with the WO5(H2O) units, contributing to the layered architecture of the WO3·2H2O phase. Upon the removal of the interlayer water, orthorhombic H2WO4 is formed, as previously reported in the literature [33]. This transformation step is further supported by the infrared spectroscopy data presented in Figure 3a,a’. The vibrational spectrum displays a broad absorption band in the 3100–3550 cm−1 region, the characteristics of O–H stretching vibrations associated with the hydrated phase. Additionally, a distinct band corresponding to the W=O stretching mode is observed at 872 cm−1, while O–W–O bending vibrations appear at 932 cm−1. These features confirm the presence of tungstate-based hydrated species in the sample.
2 N a 2 W O 4 · 2 H 2 O 4 N a + + 2 W O 4 2 + n + 2 H 2 O
2 N a + + W O 4 2 + 2 H + + 2 C l 2 N a C l + W O 5 ( O H 2 )
2 N a + + W O 4 2 + 2 H C l + n H 2 O 2 N a + + 2 C l + H 2 W O 4 · 2 H 2 O
Upon the addition of oxalic acid (H2C2O4), the solution becomes colorless. This change reflects the strong interaction between oxalate ions (C2O42−) and the interlayer water molecules in WO3·2H2O. The oxalate species disrupt the hydrogen bonding network between the water molecules and the WO5(H2O) octahedra, thereby accelerating the dehydration of WO3·2H2O into the intermediate WO3·H2O phase. A similar oxalate-induced dehydration mechanism has been reported by Li et al. [40]. This transformation is further supported by FTIR analysis (Figure 3), which reveals characteristic vibrational bands. These include a broad absorption band in the 2100–3550 cm−1 range, attributed to O–H stretching vibrations from structural water; a ν(W=O) stretching band at 872 cm−1; and O–W–O bending modes at 689 and 800 cm−1. Additional bands corresponding to oxalic acid are observed at 909, 1270, 1406, 1684, and 1711 cm−1, associated with ν(C–O), ν(C=O), and ν(O–C–O) vibrational modes [34].
Oxalic acid, which contains two carboxyl groups, is known to form bidentate chelates with tungsten species through O–W–O bridges, thereby stabilizing acidic tungstate solutions and suppressing premature polycondensation [41,42]. This chelation effect promotes the selective stabilization of specific crystal growth planes, thereby influencing both the nucleation and anisotropic growth of WO3·nH2O during the early stages of the hydrothermal process. Notably, Fang et al. [43] demonstrated that oxalic acid dihydrate can act as both a reactant and a lamellar template in the mechanochemical synthesis of WO3·2H2O ultrathin nanosheets. Although their solvent-free approach differs from the present hydrothermal method, their findings reinforce the critical role of oxalate species in directing anisotropic growth and controlling the morphology of tungsten oxide hydrates. A greater detailed description is provided in Section 2.3 for stage 3.

2.3.2. Nucleation and Growth Mechanism: Insights from TEM and FTIR Analyses

Transmission electron microscopy (TEM) analyses of the morphology and composition of aliquots collected before the hydrothermal treatment revealed clear morphological differences, as shown in Figure 4a,d. Additional analyses were carried out to observe the behavior of iron sulfate alone under the same growth conditions. This resulted in the formation of iron or iron hydroxide platelets, as reported in Figure 4e. Some goethite-like platelets were also identified (Figure 4e). TEM analysis conducted during the early stages of synthesis (Figure 4) revealed that the addition of iron sulfate heptahydrate induced the formation of fusiform amorphous aggregates, along with a film-like amorphous phase (as verified by the absence of the electronic diffraction pattern of these observed areas).
Energy-dispersive X-ray spectroscopy (EDS) confirmed that the amorphous film (Figure 4a) contained approximately 50% W, 40% Fe, and 10% S (oxygen not quantified). The aggregates (Figure 4c,d) exhibited an Fe- and S-rich core surrounded by a tungsten-rich shell, supporting the hypothesis of core–shell-type nucleation followed by interdiffusion processes.
As the reaction proceeded, these mixed amorphous precursors progressively reorganized and crystallized into platelet-like structures. TEM images recorded at later stages (not presented) show that the interdiffusion and dehydration processes led to the development of well-defined lamellar morphologies, consistent with the platelet architecture observed in the final product. This sequence of transformations highlights the critical role of the initial heterogeneous nucleation in promoting the formation of crystalline nanoplatelets from amorphous intermediates. FTIR spectra collected during this stage (Stage 1, Figure 3a,a’ displayed broad bands around 2100–3500 cm−1, corresponding to O–H stretching vibrations from structural water. Vibrations associated with C–O and C=O bonds of oxalic acid were detected near 1300 and 1650 cm−1, further confirming the presence of oxalate complexes species.

2.3.3. Proposed Dual-Pathway Growth Mechanism

Both observed morphologies individual nanoplatelets and self-organized rosette-like architectures are monoclinic, single-crystalline FeWO4, as confirmed by XRD, TEM, and EDS analyses. The main difference lies in how iron sulfate heptahydrate is introduced into the reaction medium, which decisively influences nucleation and aggregation phenomena. This observation suggests that the two morphologies originate from a common set of chemical transformations (Equations (4)–(6)), diverging only at the stage of iron incorporation and particle assembly.
2 C 2 O 4 2 + W O 4 2 + 6 H + C 4 H 4 O 4 W O 4 · 2 H 2 O
C 4 H 4 O 4 W O 4 · 2 H 2 O + F e ( O H ) 2 F e 2 + + W O 4 2 + 2 H + + C 2 O 4 2
F e 2 + + W O 4 2 + 2 H + + C 2 O 4 2 2 H + + C 2 O 4 2 + F e W O 4
Scheme 1 illustrates this dual-pathway mechanism, consisting of two parallel reaction routes. The first step corresponds to the tungsten oxalate complexation (a1) and (a2), as reported by Li et al. [40]. We then suggested the following steps based on our observations.
Pathway A1 → A2 (platelet growth):
In this route, iron sulfate heptahydrate was introduced in a dilute aqueous form. Sodium tungstate reacts with oxalic acid under acidic conditions, leading first to the formation of a tungsten–oxalate complex. Subsequent hydrothermal treatment and progressive dehydration result in the transformation of this precursor into hydrated WO3 platelets. Iron ions interact via hydrogen bonds with structural water. This promotes the nucleation and growth of FeWO4 nanocrystals while preserving platelet morphology.
In this scenario, oxalic acid plays a critical role in promoting 2D growth by stabilizing specific faces of the WO3 nuclei through hydrogen bonding. The carboxyl groups on oxalate interact with tungsten precursors, effectively isolating the nuclei and inhibiting their aggregation, as previously reported [41,44,45].
Pathway A1 → B1 (self-organized):
When iron sulfate heptahydrate is added in solid form (stage 3), a localized excess of iron and sulfate ions is generated in the immediate vicinity of tungstate–oxalate complexes. TEM images (Figure 4c,d) show that iron sulfate acts locally as a nucleation center, being rapidly coated by a shell of tungsten oxalate and hydrated WO3. This heterogeneous nucleation yields core–shell aggregates enriched in Fe and S (as confirmed by EDS). Under these conditions, the combination of strong hydrogen bonding between oxalate ions and structural water, and high local supersaturation of Fe2+, promotes the aggregation of nascent platelets into interconnected disks. As hydrothermal ripening proceeds, these assemblies evolve into rosette-like superstructures. FTIR spectra recorded at this stage exhibit characteristic O–H and C–O vibrations (bands at ~1300 cm−1 and ~1650 cm−1, as shown in Figure 3c,c’), further confirming the presence of oxalate ligands mediating the assembly process.

2.3.4. Influence of Iron Precursor Physical Form on Growth Mechanism and Morphology

The final morphology of the ferberite particles was found to be strongly dependent on the physical form in which iron sulfate heptahydrate was introduced into the oxalate–tungstate reaction medium. When added as a diluted aqueous solution, homogeneous nucleation and controlled interdiffusion promoted the growth of isolated, well-dispersed nanoplatelets. Under these conditions, oxalic acid acted as a stabilizing ligand, selectively adsorbing onto specific crystal facets and inhibiting lateral aggregation through hydrogen bonding interactions. This pathway favored a controlled two-dimensional growth of lamellar structures, consistent with previous reports on oxalate-assisted morphological control in tungsten oxide nanostructures [40]. In contrast, the direct addition of solid iron sulfate heptahydrate created localized supersaturation and heterogeneous nucleation sites, leading to amorphous mixed aggregates with core–shell-like composition. Subsequent interdiffusion, dehydration, and recrystallization transformed these precursors into interconnected platelets that progressively self-organized into rosette-like hierarchical architectures. Oxalate ions further facilitated this organization by bridging adjacent nanoplatelets through hydrogen bonds with interlayer water.
Overall, the morphology emerged from the interplay between oxalic acid complexation and the nucleation environment dictated by the iron precursor form. This dual mechanism accounts for the transition from isolated nanoplatelets to hierarchical structures, as summarized in Scheme 1.

2.4. Textural Properties: Surface Area and Porosity

The specific surface area and porosity of the synthesized FeWO4 materials, with distinct self-organized and platelet morphologies, were investigated using nitrogen (N2) adsorption–desorption process. Figure 5 shows the N2 adsorption–desorption isotherms of each morphology. The FeWO4 self-organized and platelet morphologies exhibited Type IV isotherms with an H3-type hysteresis loop, in accordance with the IUPAC classification [46].
The specific surface area (SBET) was calculated using the Brunauer–Emmett–Teller (BET) method [47]. The average pore diameter and the total pore volume were determined using the Barrett–Joyner–Halenda (BJH) method from the desorption branch of the isotherms [48]. The FeWO4 with platelet morphology exhibited a specific surface area of 12.7 m2g−1, almost double that of self-organized FeWO4 (6.80 m2g−1). Similarly, the platelet morphology exhibited a total pore volume that was 17% larger (0.041 cm3g−1) than that of the self-organized morphology (0.035 cm3g−1). However, the average pore diameters for both samples were nearly similar, around 19 Å.
These textural characteristics, particularly the specific surface area and pore volume, are known to be influenced by the synthesis method and its parameters [49]. The larger surface area and pore volume observed for the platelet FeWO4 are generally considered advantageous for photocatalysis, as they can provide more active sites for reactant adsorption and facilitate the diffusion of reactants and products [22,50].

2.5. Optical Properties—UV-Vis Diffuse Reflectance

The electronic band structure of ferberite arises from the overlap of atomic orbitals of its constituents. As reported in the literature, Fe 3d orbitals dominate the top of the valence band, while the bottom of the conduction band is primarily composed of hybridized W5d and O2p states [51]. Structural variations, including those induced by synthesis conditions and morphology, can reduce the band gap energy, typically attributed to an enhanced hybridization between Fe2+ 3d and O 2p orbitals.
To assess the light-harvesting capabilities of FeWO4 with both distinct morphologies (self-organized vs. isolated platelets), UV–visible diffuse reflectance spectra (DRS) were recorded (Figure 6). Band gap energies (Eg) were estimated using the Tauc relation Equation (S1) (in the Supplementary Materials).
The nature of the band gap in FeWO4 remains debated. Some band structure studies predict a direct transition [52,53] while others support an indirect one [23,35]. These discrepancies are often linked to differences in synthesis method, grain size, morphology, and microstructuring, as confirmed in DFT-based studies. For self-organized FeWO4, the band gap was found to be 1.62 eV, assuming a direct transition, and 1.69 eV, assuming an indirect transition. In contrast, the platelet morphology exhibited a slightly lower value of 1.47 eV for the direct transition, and 1.77 eV for the indirect one. These values fall within the broad range of band gaps reported for FeWO4, which vary depending on synthesis methods, e.g., 1.7 eV (microemulsion) [24], 2.96 eV (precipitation) [25], 2.1 eV (solvothermal), and up to 3.17 eV (sol–gel) [26]. In hydrothermal conditions, reported band gaps range from 1.53 eV to 2.4 eV [22], with intermediate values such as 1.98 eV also reported for amine-assisted synthesis [54]. Overall, the literature reports indicate that the Eg of FeWO4 varies significantly depending on the assumed transition type: 1.6–2.3 eV for direct transitions and 1.3–2.0 eV for indirect ones, reflecting the influence of synthesis routes and microstructural factors.
The results in Figure 6 show the Tauc graph of FeWO4 with the band gap of 1.69 eV and 1.77 eV for self-organized and platelets. Based on the literature and our experimental spectra (Figure 6a), which show a single dominant linear region, we consider a direct band gap model to be most appropriate for our hydrothermally synthesized ferberite [24,25,29,55]. This assumption will be used in subsequent energy band diagram discussions.

2.6. Photocatalytic Activity of Ferberite with Self-Organized and Platelet Morphologies

2.6.1. Adsorption Studies

Before testing photocatalytic performance, the interaction between methylene blue (MB) molecules and the catalyst surface in the absence of light was investigated to distinguish adsorption effects from true photocatalytic degradation. Adsorption–desorption equilibrium was assessed through dark experiments. Figure S2 presents the evolution of the UV-Vis absorption spectra of MB over time in the Supplementary Materials. Regarding platelet morphology, equilibrium was reached within 10 min, with no significant variation in MB observed up to 60 min. During this period, approximately 30% of the MB was removed via adsorption (Figure S2a, in the Supplementary Materials). Based on these observations, a 20 min dark stirring step was systematically applied before initiating all photocatalytic degradation tests to ensure equilibrium. It is noteworthy, however, that under irradiation, the photodegradation of MB by both ferberite morphologies-platelets or self-organized structures, remained limited. These results are detailed in the next section.

2.6.2. Influence of pH and H2O2 on MB Degradation

The photocatalytic degradation of MB activated by FeWO4 with both self-organized and platelet morphologies was investigated under visible light, with and without hydrogen peroxide (H2O2), and at varying initial pH values (3, 5 and 10). Figure 7 shows the evolution of MB relative concentration (C/C0) over time under various conditions: photolysis (MB alone) without catalyst, FeWO4 alone, and the FeWO4 combined with H2O2 (photo-Fenton system) for both morphologies (Figure 7a–f. It should also be noted that, in the absence of light, the H2O2-assisted catalysts show no additional catalytic activity compared with the adsorption stage.
  • Blank tests under visible light (photolysis, H2O2 alone, FeWO4 alone)
The direct photolysis of MB or degradation using H2O2 alone led to negligible removal (ca. 9% after 30 min with H2O2 alone). In contrast, FeWO4 alone promoted partial degradation: self-organized FeWO4 achieved 25–45% degradation, while platelet-shaped FeWO4 reached 45–65%, depending on pH. In both cases, photocatalytic activity increased with rising pH, suggesting more efficient degradation pathways in basic media. Nonetheless, the overall performance of FeWO4 alone remained moderate, likely limited by low surface area, electron–hole pair recombination, and insufficient ROS radical production.
b.
Effect of H2O2 addition (photo-Fenton conditions)
A marked improvement was observed upon H2O2 addition. Under photo-Fenton conditions, MB degradation increased, reaching 75% to 100% within 15 min, as the pH rose from acidic to basic, regardless of morphology. This enhancement reflects the strong synergistic effect between FeWO4 and H2O2 (Figure 7), especially in acidic or neutral media. This synergy arises from the dual action within the photo-Fenton mechanism: photogenerated electrons in the FeWO4 conduction band that can react with H2O2 to form hydroxyl groups (Equations (7) and (8), respectively).
F e W O 4 + h v e C B + h V B +
e C B + H 2 O 2 O H * + O H
In parallel, Fe2+/Fe3+ surface sites catalyze the classical Fenton reactions (Fe2+ + H2O2 → Fe3+ + *OH + OH), further boosting the generation of highly oxidizing O H * radicals.
c.
pH influence and surface charge
Solution pH also plays a crucial role in the catalytic performance of ferberite. An overall increase in MB degradation was consistently observed as the pH increased from acidic to basic conditions. This effect was particularly pronounced for the self-organized FeWO4. This trend can be attributed to pH-dependent electrostatic interactions between the catalyst surface and the cationic dye MB, as revealed by the catalyst’s point of zero charge (pHpzc) determined for both morphologies. The pHpzc of the ferberite materials was determined to be approximately 4.8 for the self-organized structures and 4.5 for the platelets (Figure 8c). At pH value above pHpzc, the catalyst surface becomes negatively charged, thereby promoting electrostatic attraction with the cationic MB molecules. This enhances dye adsorption and subsequent degradation.
d.
Morphology effect
The comparison of the two morphologies (Figure 8) shows that platelet-shaped ferberite generally outperformed the self-organized form, particularly in the absence of H2O2. For example, at pH 3, the platelets achieved ca. 40% higher degradation than their self-organized counterpart. However, under neutral or alkaline conditions, both morphologies exhibited similar efficiencies. Under photo-Fenton conditions, degradation increased to 77% and 90% for self-organized and platelet morphologies, respectively, at pH 3, and nearly 100% at higher pH, regardless of the morphology. This enhanced photodegradation is attributed to the low band gap values of both samples (see Figure 6), which facilitate visible-light absorption. In addition, the superior performance of the platelet’s morphology can be linked to its favorable textural properties, namely its higher specific surface area and greater pore volume, which promote reactant adsorption and light harvesting.
As mentioned above, morphology has a direct impact on catalytic performance. Compared to the self-organized structure, the platelet morphology features larger crystallites (87 nm) and higher crystallinity. This is confirmed by the narrower diffraction peaks in the XRD pattern (Figure 1b), contrasting with the smaller crystallites (68 nm) found in the self-organized morphology. Consequently, platelets have a larger specific surface area, creating a greater number of active surface sites and a larger light absorption cross-section than self-assembled samples. Additionally, the larger surface area and pore volume of the platelets promote better adsorption of methylene blue (MB), improving the diffusion of H2O2 and O2. This results in the greater exposure of active sites, and the localized generation of OH radicals. Furthermore, the platelet morphology has a more negative conduction band potential of –0.05 V (as detailed in Section 2.7). This property, combined with a higher photocurrent, ensures greater availability of charge carriers, which effectively promotes the reduction of H2O2 to OH radicals. This explains the superior degradation activity observed under acidic conditions, even when the catalyst surface is positively charged.
e.
Influence of H2O2 concentration
Based on the analysis of the MB photodegradation results as a function of pH presented in Figure 8, a second study is presented to verify the influence of the H2O2 concentration present in the reaction medium (4 and 8 mM), as shown in Figure 9. Figure 9c shows the photodegradation rate (%) of the two different ferberite morphologies (platelets in blue and self-organized in green) as a function of the addition of H2O2 (0, 4 and 8 mM) in the photodegradation of MB. At low concentrations (4 mM), both perform well, achieving yields over 90% and indicating limited radical recombination. Both morphologies show almost similar degradation rates, with a slight drop of 2%, within the margin error. This shows that the 4mM concentration of H2O2 is close to the optimum concentration required for the photo-Fenton reaction. The result indicates that surface area (SA) is not the determining factor. In comparison to self-organized FeWO4, platelets possess nearly double the number of active sites based on SA.
f.
Kinetic analysis
From a kinetic standpoint, the degradation mechanism of MB involves not only photogenerated charge carriers, but also the transport and diffusion of reactive oxygen species, such as hydroxyl radicals OH and hydrogen peroxide H2O2. Despite the rather low specific surface area of the FeWO4 sample, catalytic activity in the presence of H2O2 remained remarkably fast. Regardless of pH, MB degradation consistently exceeded 80% and reached 100% under acidic conditions. To obtain further insight into the reaction kinetics, additional experiments were conducted at a fixed H2O2 concentration of 4 mM. Figure 9d illustrates the color change between the initial MB solution, the solution after 30 min in the absence of irradiation (adsorption–desorption equilibrium), and the solution after 30 min of irradiation.
As shown in Figure 9a,b, the self-organized ferberite appeared to be slightly more efficient than its platelet-shaped counterpart, achieving a 10% higher uptake prior to illumination, but much slower degradation during the first 10 min of light exposure. A detailed examination of catalytic activity reveals different kinetic behaviors. For example, MB degradation by platelets occurs in one-step, whereas self-assembled samples require several steps. Moreover, kinetic constant representation revealed a difference in apparent reaction order between the two morphologies. The self-organized form followed pseudo-first-order kinetics, while the platelet morphology exhibited pseudo-second-order behavior. This is confirmed by the linear fits of the apparent rate constants for each material (platelets: k = 3.086 min−1; self-organized structures: k = 0.129 min−1) as reported in Figure 9a,b. The difference likely arises from the quantities and types of reactive oxygen species (ROS) involved, particularly between OH and H2O2. The structuring of ferberite influences the transport of H2O2 and intermediates (-OH). The presence of microporosity and extended platelet surfaces favors diffusion limited processes and accelerates surface reactions. Although these initial findings are promising, further investigation is needed to confirm the reaction orders and identify the dominant factors controlling the kinetics and reaction mechanism.
Increasing the catalyst dosage enhances the number of exposed active sites, improves MB adsorption, and promotes the generation of greater numbers of e/h+ pairs. However, beyond an optimal threshold, the photocatalytic activity tends to plateau or even decrease due to the increased turbidity of the suspension (light scattering), particle shadowing, and agglomeration, which collectively reduce the effective active surface area. Therefore, in this study, we adopted a typical photocatalyst dosage of 1 g·L−1, in agreement with the literature [5]. Regarding the oxidizing agent, at low [H2O2] (4 mM) (Figure 9c), the system operates near its optimum condition, providing efficient OH generation without excess oxidant. Increasing [H2O2] to 8 mM leads to only a slight improvement in degradation, suggesting that the system approaches saturation in OH production.
g.
Comparison with literature
In comparison, other studies in the literature using either pure or modified FeWO4 typically require longer reaction times, higher concentrations of H2O2, or additional conditions such as UV radiation, metal doping, or heterostructure formation. Qian et al. [22] employed FeWO4 nanorods synthesized via a simple hydrothermal method. This achieved approximately 90% degradation of MB in 60 min under visible light with 6 mM H2O2 at pH 3 demonstrating lower activity and a slower degradation rate compared to the present work. Similarly, Liu et al. [29] utilized a microemulsion hydrothermal route to synthesize FeWO4 nanorods, reaching only 75% degradation after 120 min, and notably without the use of H2O2.
Doped and hybrid systems have also been reported. In the first case, ferberite was doped with either nickel [24] or with iodide/graphene oxide (I/GO) [30], achieving degradation efficiencies between 95% and 98%. However, these results were obtained under more restrictive conditions such as UV irradiation or extremely acidic media (pH 3), and required chemical modification steps that may compromise reproducibility and increase production costs. In the second case, the FeWO4/FeS2 hybrid system [36] achieved 97% degradation in 30 min, which is a value comparable to the present work. However, this involved the synthesis of a binary composite with more complex morphological control. The most efficient system reported among the reviewed studies was that of Sun et al. [49], which used sol–gel synthesized nanofibers combined with ultrasound irradiation, achieving 99% degradation in just 15 min. Nevertheless, the combined use of visible light and ultrasound significantly increases both the energy demand and operational complexity of the process. Table 2 provides a comparative overview of the materials and experimental conditions used in some of the works reported in the literature and in this study.
It is worth noting that, in addition to the conditions listed in Table 2 such as the type of material, irradiation source, pH, and reaction time, the photodegradation performance of the pollutant is also strongly influenced by other parameters of the photocatalytic system. These include the initial concentration of methylene blue (C0), the amount of catalyst used, the reaction temperature, the intensity and distance of the light source, and the geometry of the photoreactor.
h.
Recyclability and stability
Figure 9e,f display the recycling performance of ferberite catalysts with platelets and self-organized morphologies. The results reveal that the platelets catalyst exhibits superior cyclic stability, retaining higher activity up to the third cycle. In contrast, the self-organized catalyst undergoes a more pronounced deactivation during the early cycles. The catalysts stability was also investigated by FTIR (Figure S3a) and XRD (Figure S3b) analyses after using samples for photodegradation testing. The IR absorption bands and Bragg diffraction positions remain the same as before use, showing that the catalysts remain stable without any structural modification.
To assess chemical stability and leaching of the catalyst, the reaction solution was analyzed by Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES) after the final degradation cycle. Measurements were taken for 100 g of catalyst dispersed in 100 mL of solution.
Concentrations of iron cations measured at pH 5 and 10 show a mean value of 14 and 95 µg/L, respectively, for platelets and self-assembled samples. These values remain relatively negligible, particularly for platelets. However, at pH 3, in an acidic medium, the concentrations were high, in excess of 500 µg/L, reflecting the instability of both catalysts in acidic media. The low concentrations of iron ions in the solution indicate that the leaching phenomenon through the cycles is negligible for platelets. The ability to maintain high efficiency over multiple cycles is due to the high stability of the catalyst and demonstrates the Ferberite efficiency for practical photodegradation applications.
i.
Radical scavenging experiments
To identify the main reactive chemical species responsible for MB degradation, radical scavenging experiments were performed. Disodium ethylenediaminetetraacetic acid (EDTA), isopropyl alcohol (IPA) and benzoquinone (BQ) were used as specific quenchers for photogenerated holes (h+), hydroxyl radicals (OH), and superoxide radicals ( O 2 ) , respectively [59]. A significant reduction in degradation rate upon the addition of a given scavenger indicates the involvement of the corresponding species in the photodegradation mechanism.
Figure 10 demonstrates first that there is no significant difference between the two morphologies (self-organized and nanoplatelets), then highlights the central role of OH in MB degradation mechanism under photo-Fenton conditions. The addition of IPA led to a substantial decrease in degradation activity, approximately 39% for the self-organized form and 35% for the platelet’s morphology indicating that hydroxyl radicals are the main reactive species with both catalysts. In contrast, the addition of EDTA and BQ resulted in only minor reductions (less than 10%), indicating that removing holes or superoxide radicals does not substantially affect photodegradation.
The involvement of OH radicals is consistent with classical Fenton-type reactions occurring at Fe2+/Fe3+ sites on the catalyst surface (Equations (9) and (10)):
F e 2 + + H 2 O 2 F e 3 + + O H + O H
F e 2 + + O H F e 3 + + O H
Although morphology influences the specific surface and number of active sites, as well as surface accessibility, both FeWO4 forms enable similar OH generation through this redox cycle, as confirmed by the similar drop in degradation activity upon IPA addition (39% vs. 35%). These results agree with previous studies. For example, Zhang et al. investigated photo-Fenton systems involving Fe3O4/α-FeOOH and FeWO4/GO [33]. They concluded that the photo-Fenton reactions involved are also controlled by OH radicals, even with other sources of charge (electrons or holes) [27].
It is worth emphasizing that the enhanced degradation observed at higher pH values cannot be directly ascribed to the increased generation of reactive oxygen species (ROS). The radical scavenging experiments (Figure 10) clearly demonstrated that hydroxyl radicals are the dominant species in both morphologies, independently of the solution pH. Thus, the improved photocatalytic activity under alkaline conditions is mainly related to electrostatic effects between the negatively charged catalyst surface (pH > pHpzc) and the cationic MB molecules, which favor adsorption and subsequent reaction, rather than to a higher intrinsic ROS yield.

2.7. Photocatalysis Mechanism Based on Photoelectrochemical Investigations

2.7.1. Transient Photocurrent Response

Transient photocurrent measurements were conducted under chopped wavelength illumination (390 nm) at an applied potential of +0.5 V vs. Ag/AgCl, optimized values to evaluate the generation and separation of photogenerated charge carriers. Figure 11 presents the current density versus time for the FeWO4 electrodes. The current density increased upon illumination and rapidly decreased when the light was turned off for both samples. The response was stable over the multiple light on/off cycles shown, indicating good photostability under these conditions and confirming its ability to generate hole–electron pairs. A significant difference in the photocurrent intensity was observed between the two morphologies. The platelet-shaped FeWO4 consistently generated a much higher photocurrent density, reaching peak values of approximately 1.48 µA/cm under illumination at this potential. By contrast, the self-organized FeWO4 achieved a peak photocurrent density of only about 0.78 µAcm−1. Beyond size effect, this difference is primarily due to the way the two samples are structured: the individual platelets possess twice the specific surface area of the self-organized ones, providing a larger surface for light absorption. Howover, the self-organized structures contains interstices and cavities which, through internal light reflection, reduce photoactivation. As reported in the literature, a higher photocurrent density generally implies more efficient generation and separation of electron–hole pairs, along with more effective transport of these carriers to the catalyst/electrolyte interface where they participate in redox reactions. Accordingly, based on their current density, platelets demonstrated enhanced photocatalytic activity in both the presence and absence of H2O2, compared to the self-organized ferberite.
In general, photocurrent measurement assesses the performance of a thin-film electrode under the influence of an external voltage bias. This setup favors materials that exhibit efficient long-range charge transport through the film to the collector. In this system, catalytic performance is governed by other factors owing to the use of isolated particles. This includes the conversion of reactant adsorption onto the particle surface and charge separation at the local surface level. The high photocurrent of the platelets demonstrates their catalytic activity when used as electrodes. Nevertheless, this parameter alone does not fully predict their performance in a suspended photocatalytic system, particularly since the self-organized material, while exhibiting lower current densities, shows a superior adsorption capacity.

2.7.2. Mott–Schottky Analysis

To further investigate the semiconductor properties of the synthesized FeWO4 materials, Mott–Schottky analysis was carried out. Figure 12a displays the Mott–Schottky plots, i.e., 1/C2 vs. V, where C is the space charge capacitance and V is the applied potential. Both self-organized and platelets-like FeWO4 electrodes were evaluated at a frequency of 300 Hz. They exhibit negative slopes in the linear region of plots, which is indicative of p-type semiconductor behavior [60].
The flat-band potential (Vfb) was determined by extrapolating the linear portion of the 1/C2 vs. V plot to the potential axis (i.e., where 1/C2 = 0). These experimentally determined Vfb values, combined with the optical band gap energies (1.69 eV for self-organized and 1.77 eV for platelet-shaped FeWO4), were used to calculate the conduction band (ECB) and valence band (EVB) edge position, as illustrated in Figure 12b.
For a p-type semiconductor, the Fermi level (EF) is close to EVB. Equation (11) is assumed as a common approximation for p-type material; the EVB for self-organized FeWO4 is estimated at +1.97 V vs. NHE, and for plates at +1.72 V vs. NHE. Consequently, the ECB positions (Equation (12)) are calculated to be +0.28 V vs. NHE for self-organized FeWO4 and −0.05 V vs. NHE for platelets-like FeWO4.
E V B = V f b + 0.2   V
E C B = E V B E g
The resulting band structure diagram, shown in Figure 12b, also includes the redox potentials of key ROS involved in the degradation mechanism. It provides a useful framework for interpreting the photocatalytic behavior of FeWO4. By correlating these energy levels with the results of radical scanvenging experiments, a coherent and plausible MB degradation mechanism can be proposed, as described in the following section.

2.7.3. Photo-Fenton Mechanism and OH Generation Pathways in FeWO4/H2O2 System

The degradation mechanism in the FeWO4/H2O2 system under light irradiation proceeds mainly through the generation of hydroxyl radicals (OH), as confirmed by scavenger experiments. Four potential pathways for OH production can be considered, each involving distinct redox processes. Their feasibility is discussed below based on the electronic band positions of FeWO4 and thermodynamic considerations.
  • Reduction of H2O2 by Photogenerated Electrons
H 2 O 2 + e O H + O H   ( E 0 = + 0.32   V   vs .   NHE )
This is the most plausible and dominant route. Both FeWO4 morphologies posses conduction band (CB) potentials more negative than +0.32 V (i.e., +0.28 V for self-organized and −0.05 V for platelets), making this reduction thermodynamically favorable. Therefore, photogenerated electrons can efficiently reduce H2O2 to OH. This is consistent with trapping experiments showing that OH radicals are the primary reactive species involved in pollutant degradation.
b.
Oxidation of OH by Photogenerated Holes
O H + h V B + O H   ( E 0 = + 1.90   V   vs .   NHE )
This pathway is only possible for the self-organized FeWO4, whose valence band (VB) lies at +1.97 V vs. NHE, slightly above the required potential. In contrast, the platelet morphology, with a VB at +1.72 V, cannot oxidize hydroxide to OH. Thus, OH generation via hole-mediated OH oxidation is morphology-dependent and limited to the self-organized form.
c.
Oxidation of H2O2 by Photogenerated Holes (Equation (13))
H 2 O 2 + h V B + O H + H +   ( E 0 + 2.31   V   vs .   NHE )
Neither morphology has a VB potential high enough to drive this reaction. Therefore, this pathway is thermodynamically inaccessible under the given conditions.
d.
Classical Fenton Reaction via Fe2+/Fe3+ Cycling
F e 2 + + H 2 O 2 F e 3 + + O H + O H
FeWO4 inherently contains Fe2+ species that can participate in the conventional Fenton reaction with H2O2 to generate OH. Moreover, the photogenerated electrons in the CB are sufficiently reducing to convert Fe3+ back to Fe2+:
F e 3 + + e F e 2 +   ( E 0 = + 0.77   V   vs .   NHE )
This redox cycle is thermodynamically favorable for both morphologies, reinforcing the self-sustained nature of OH production through both photo- and Fenton-type mechanisms.
Among the four possible pathways, pathway (1) (H2O2 reduction by electrons) is active in both morphologies and is likely dominant. Pathway (4) (classical Fenton reaction) also contributes, aided by photogenerated electron-driven Fe3+/Fe2+ cycling. Pathway (2) is only feasible for the self-organized material due to its more positive VB, but contributes marginally. Pathway (3) is ruled out for both.
Interestingly, although the self-organized FeWO4 exhibits a more positive VB potential theoretically compatible with OH oxidation, it shows lower photocatalytic performance than the platelet morphology. This is consistent with our scavenging experiments, which revealed no significant contribution of photogenerated holes to the degradation activity in either morphology. These observations confirm that electron-driven processes—particularly H2O2 reduction and Fe3+ regeneration—are the dominant pathways in the photo-Fenton mechanism of FeWO4.

3. Experimental Section

3.1. Ferberite Hydrothermally Synthesis

In the syntheses, all chemicals were used as received with no further purification (sodium tungstate dihydrate (Aldrich, St. Quentin Fallavier, France, 99.0%), iron sulfate heptahydrate (Aldrich, France, 99.5%), oxalic acid (Aldrich, France, 99.5%), and hydrochloric acid (Aldrich, France, 37.5%)). FeWO4 powders were prepared by a hydrothermally route, according to the following procedure: first, sodium tungstate dihydrate (12 mM) was diluted in 100 mL of water and stirred for 10 min at room temperature (26 °C); the pH was adjusted by adding a 3 M HCl solution until the pH was 1. At this stage, there was a change in color, the transparent solution became light green; while stirring, we added oxalic acid (25 mM) in solid form and 150 mL of water, and the solution became transparent again. Finally, 14 mM iron sulfate hephydrate was added and transferred to the autoclave, which was heated to 200 °C for 24 h and then cooled naturally; the precipitate was collected and washed with water several times. Finally, the precipitate was dried at 80 °C for 24 h. During the synthesis process, approximately 10% of the ferberite mass was lost during the washing and drying steps of the precipitate obtained after the hydrothermal treatment. After the treatment process, the recovered powder was ground for appropriate physicochemical characterization and the evaluation of the properties of interest. To investigate the influence of iron (II) sulfate addition during synthesis, two experimental conditions were employed: (i) the direct addition of solid iron (II) sulfate heptahydrate to the final solution and (ii) the prior dilution of iron (II) sulfate in water before addition. As a result, two distinct types of ferberite samples were obtained, herein referred to as self-organized and platelet morphologies, respectively.

3.2. Structural, Morphological, Chemical, Textural Properties and Optical Investigation

The FeWO4 powders with different morphologies were structurally investigated by an Empyrean Panalytical diffractometer, equipped with a Cu anticathode. Powder X-ray diffraction (XRD) patterns were collected in the classic θ-2θ mode, from 10° to 80° (step size 0.003°, scan speed 0.004°s−1). To obtain the lattice parameters, crystal size and crystallographic data, the XRD patterns were used and refined by employing the Material Analysis Using Diffraction (MAUD) software version 2.064 [61]. The minimization was carried out using the reliability factors as numerical criteria of the refinement quality calculations (Rwp, Rexp and Rwp/Rexp(S)). To represent the refined structures, Carine Crystallography software (version 3.1) was used for a three-dimensional representation of the ferberite, illustrating the coordination polyhedral, and bonds between tungsten, iron, and oxygen.
The morphologies of the different powders were observed by transmission electron microscopy (TEM) and electron diffraction (ED) characterization were carried out using a conventional 200 kV Tecnai microscope equipped with a LaB6 source. The chemical composition of ferberite was determined by electron dispersive spectroscopy (EDS). Cation concentrations were quantified using standards. The statistical study of the chemical homogeneity of the powder consisted of between 10 and 15 analyses on randomly selected grains. Surface morphology was studied using a GEMINI Zeiss SUPRA 40Vp scanning electron microscope (SEM). Observations were made in InLens mode at a working distance between 2 and 3 mm and at a voltage of 5 kV without any prior treatment.
To verify the formation of products and intermediates during the synthesis of ferberite, aliquots of the solution were collected at each stage of the synthesis methodology: the addition of hydrochloric acid (1), the addition of oxalic acid (2), and the addition of iron sulfate hephydrate (3) and analyzed by Fourier transform infrared spectroscopy (FTIR) and MET.
Nitrogen adsorption–desorption isotherms (N2) were measured at 77 K using a Quantachrome Autosorb iQ AG surface area and porosity analyzer. The Brunauer–Emmett–Teller (BET) method was used to calculate the specific surface area from these isotherms [40]. Specific surface area, pore size and total pore volume were determined from the desorption branch of the isotherm using the Barrett–Joyner–Halenda (BJH) method [48]. Before analysis, the samples were degassed under vacuum at 200 °C for 24 h to remove any possibly adsorbed species. Finally, the optical properties (absorption, reflectance and band gap) of the obtained ferberites were investigated by UV-Vis spectroscopy in normal and diffuse reflectance modes with respect to photoactivated applications. The measured UV-Vis spectra of different samples were recorded using a Shimadzu 2600 UV/Vis spectrophotometer with a 150 mm integrating sphere and using BaSO4 as the baseline reference. Measurements were made in the range of 250–800 nm at room temperature with a resolution of 0.08 nm. The background was determined using a calibrated reflectance standard with an accuracy of 0.005.

3.3. Photoelectrochemical (PEC) Measurements

Photoelectrochemical characterizations (Mott–Schottky plot and transient photocurrent responses) were conducted using an OrigaLys (OrigaLys ElectroChem SAS, Rillieux-la-Pape, France) potentiostat/galvanostat in a standard three-electrode configuration. Working electrodes (WE) were prepared by depositing the synthesized FeWO4 materials (self-organized and plateles) onto conductive indium tin oxide (ITO)-coated glass substrates (1 cm × 1 cm). A suspension of 5 mg of FeWO4 powder in 1 mL of ethanol was prepared by ultrasonication for 30 min. A controlled volume of this suspension (1 mL) was drop-cast onto the active area of the ITO substrate and dried at 80 °C for 180 min. The reference electrode (RE) and the counter electrode (CE) were Ag/AgCl and platinum foil, respectively. The PEC measurements were performed using a 0.1 M aqueous solution of KOH (pH 14) and Na2 SO4 (pH 2) Mott–Schottky plot and transient photocurrent responses measures, respectively.
Transient photocurrent responses were measured under chopped illumination conditions. The light source was a Revoler Instytut Fotonowy LED system, providing monochromatic light at a wavelength of (390 nm). Measurements were typically recorded at an applied potential of +0.5 V vs. Ag/AgCl. Mott–Schottky analysis was used to determine the flat-band potential (Vfb) of the FeWO4 semiconductor electrodes (both for self-organized and platelets) for a p-type semiconductor. The Mott–Schottky relationship is given by Equation (18).
1 C 2 = 2 ε ε 0 A 2 e N A V V F b k B T e
where ε is the dielectric constant of the semiconductor, ε0 is the permittivity of free space, e is the elementary charge, A is the electrode area, V is the applied potential, kB is the Boltzmann constant, and T is the absolute temperature. The flat-band potential (Vfb) was determined from the intercept of the linear region of the 1/C2 vs. V plot.
Potentials measured against Ag/AgCl were converted to the Reversible Hydrogen Electrode (RHE) scale using the Nernst equation, shown in Equation (19):
E R H E = E A g / A g c l + 0.197   V + 0.059   p H
where EAg/AgCl is 0.197 V vs. NHE at 25 °C of the Na2 SO4 (pH 2) and KOH (pH 14) electrolytes.

3.4. Photocatalytic Action Evaluation

The photocatalytic performance of the synthesized FeWO4 materials (self-organized and platelets) was evaluated by monitoring the degradation of methylene blue (MB), a model organic pollutant, under simulated solar irradiation. It is common to find concentrations around 5 ppm for MB in the literature, as it is a practical experimental range that avoids spectral saturation and maintains the linearity and efficiency of the process [57,62]. The photocatalytic system adopted basically consists of a 250 mL quartz tube with a water-cooling system coupled under a 300 W xenon lamp to simulate solar radiation. The spectral width of radiation from Philips lamps (300 W) simulating solar radiation, with maximum intensity near the visible region (400–800 nm).
First, the adsorption–desorption equilibrium was established. In a typical experiment, 100 mg of the FeWO4 catalyst was dispersed in 100 mL of an aqueous methylene blue (MB) solution with a concentration of 5 ppm. The study was conducted without pH adjustment (initial pH ≈ 5), with sampling at 10 min intervals over a period of 60 min. Once equilibrium was reached, preliminary photocatalytic experiments were carried out at the same pH (5) for 210 min, with measurements taken every 30 min.
To investigate the role of reactive oxygen species (ROS) and possible photo-Fenton contributions, comparative experiments were conducted under identical conditions using (i) hydrolysis (control), (ii) FeWO4 alone, (iii) H2O2 (8 mM) alone, and (iv) a combined FeWO4/H2 O2 (8 mM) system. Additionally, the influence of pH on MB degradation was systematically evaluated by adjusting the initial solution pH to 3, 5, and 10 using dilute nitric acid (HNO3) or sodium hydroxide (NaOH). These values were selected based on the point of zero charge (pHpzc) of the catalyst. When H2O2 was used, it was added after reaching the adsorption–desorption equilibrium time.
Based on the methylene blue degradation results as a function of pH variation, a second study was performed to investigate the effect of H2O2 concentration. Three conditions were evaluated: no H2O2, 4 mM H2O2, and 8 mM H2O2. Subsequently, to quantify the degradation kinetics of methylene blue, the experimental data were fitted to pseudo-first-order and pseudo-second-order kinetic models, as appropriate. The apparent rate constants (k1 and k2) were then determined according to Equations (20) and (21), respectively.
l n C C o = k t
1 C 1 C o = k 2 t
At predetermined time intervals during the irradiation, approximately 3 mL of the suspension were withdrawn, centrifuged at 13000 rpm for 30 min to remove catalyst particles, and the supernatant was analyzed. MB concentrations were determined by measuring the characteristic absorbance peak at λmax = 665 nm using a Shimadzu UV-1800 UV–visible spectrophotometer with version 2.71 of the software. Degradation activity was expressed as C/C0, where C0 is the initial MB concentration and C is the MB concentration at irradiation time t.
To evaluate the stability and reusability of FeWO4 (platelets and self-organized), recycling tests were carried out for the photodegradation of MB under optimized conditions (pH 5, 4 mM H2O2, under visible light). The catalysts were tested over four consecutive cycles. After each cycle, they were separated from the reaction mixture by centrifugation, washed with double-distilled water and dried overnight in an oven at 80 °C before. Additionally, XRD and FTIR analyses are performed to confirm the structural stability of the samples.

4. Conclusions

Nanostructured ferberite was successfully synthesized via hydrothermal routes under acidic conditions. The resulting powders crystallized in a single-phase monoclinic structure, exhibiting high purity, good crystallinity, and a near-stoichiometric composition (Fe: 49%, W: 51%). The investigation of the formation mechanisms revealed that the physical form of the FeSO4 precursor, whether introduced as a solid or in diluted solution, plays a key role in directing crystal growth and morphology. This led to the formation of two distinct structures: isolated platelets and self-organized desert rose-like microstructures. Optical and electrochemical measurements confirmed the p-type semiconductor nature of both morphologies, with band gap energies of 1.69 eV and 1.77 eV for self-organized form and platelets, respectively, indicating their sensitivity to visible light radiation. Photocatalytic tests under simulated sunlight revealed high degradation activity via photo-Fenton processes. The presence of H2O2 significantly enhanced the degradation rate of MB, achieving complete removal within 30 min. Each morphology exhibited optimal performance under different pH conditions, with a notable synergistic effect between the catalysts and H2O2, especially in alkaline environments. While both morphologies were active, kinetic analysis revealed a pseudo-second order behavior for platelet-shaped FeWO4 consistent with its enhanced textural properties. In contrast, the self-organized microstructure followed a pseudo-first-order kinetic model. The combined analysis of radical scavenging experiments and band structure diagrams identified hydroxyl radicals as the only reactive oxygen species (ROS) responsible for MB degradation. These results demonstrate the potential of nanostructured FeWO4 as a sustainable and efficient photocatalyst for water purification applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30194026/s1, Figure S1. Electronic diffraction pattern (TEM) of ferberite: Platelets (a) and self-organized (b) Section 2.5 dedicated to the optical properties and UV-vis diffuse reflectance. Figure S2. Corresponds to the UV-vis absorption spectra of MB solution (5 ppm) in presence of platelets ferberite morphology at pH 5; Figure S3. Concerns FTIR (a) and XRD (b) of powders after two cycles of photodegradation.

Author Contributions

Conceptualization, C.L. and M.A.; Formal analysis, A.G.d.S.; Investigation, A.G.d.S., Y.E., V.C., and M.A.; Data curation, A.G.d.S., Y.E., V.C., C.L., A.L.L.-M. and M.A.; Writing—original draft, A.G.d.S. and Y.E.; Writing—review & editing, V.C., C.L., A.L.L.-M. and M.A.; Supervision, C.L. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the Departmental Council of Var (CD83), the urban community of Toulon Provence Mediterranean and University of Toulon for their financial supports in the framework of the “NanoCat”, “Disolar” and “NSPEC” projects.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors gratefully acknowledge the Regional Council of Provence-Alpes Cote d’Azur, Departmental Council of Var (CD83), the urban community of Toulon Provence Mediterranean and University of Toulon for their financial supports in the framework of the “NanoCat”, “Disolar” and “NSPEC” projects. A. GOMES DOS SANTOS thanks UFERSA/CCEN/DCME to take charge of her delegation in the PhD framework.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. How Is Fast Fashion Degrading Our Water Resources? | Bosaq. Available online: https://bosaq.com/water-and-fashion-industry/ (accessed on 7 July 2025).
  2. Niinimäki, K.; Peters, G.; Dahlbo, H.; Perry, P.; Rissanen, T.; Gwilt, A. The Environmental Price of Fast Fashion. Nat. Rev. Earth Environ. 2020, 1, 189–200. [Google Scholar] [CrossRef]
  3. Bailey, K.; Basu, A.; Sharma, S. The Environmental Impacts of Fast Fashion on Water Quality: A Systematic Review. Water 2022, 14, 1073. [Google Scholar] [CrossRef]
  4. Chavan, R.B. Thirsty Textile and Fashion Industry PART I: Water Distribution on Earth and Virtual Water, Water Footprint Concepts. Latest Trends Text. Fash. Des. 2018, 2, 1–22. [Google Scholar] [CrossRef]
  5. Ahmed, S.; Rasul, M.G.; Brown, R.; Hashib, M.A. Influence of Parameters on the Heterogeneous Photocatalytic Degradation of Pesticides and Phenolic Contaminants in Wastewater: A Short Review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef] [PubMed]
  6. Khan, I.; Saeed, K.; Ali, N.; Khan, I.; Zhang, B.; Sadiq, M. Heterogeneous Photodegradation of Industrial Dyes: An Insight to Different Mechanisms and Rate Affecting Parameters. J. Environ. Chem. Eng. 2020, 8, 104364. [Google Scholar] [CrossRef]
  7. Chengli, Z.; Ronghua, M.; Qi, W.; Mingrui, Y.; Rui, C.; Xiaonan, Z. Photocatalytic Degradation of Organic Pollutants in Wastewater by Heteropolyacids: A Review. J. Coord. Chem. 2021, 74, 1751–1764. [Google Scholar] [CrossRef]
  8. Antonopoulou, M. Homogeneous and Heterogeneous Photocatalysis for the Treatment of Pharmaceutical Industry Wastewaters: A Review. Toxics 2022, 10, 539. [Google Scholar] [CrossRef]
  9. Van Thuan, D.; Ngo, H.L.; Thi, H.P.; Chu, T.T.H. Photodegradation of Hazardous Organic Pollutants Using Titanium Oxides -Based Photocatalytic: A Review. Environ. Res. 2023, 229, 116000. [Google Scholar] [CrossRef]
  10. Masuleh, M.T.; Hasheminiasari, M.; Ashiri, R. Enhanced Photocatalytic Efficiency of Eco-Friendly Synthesized ZnO for Rapid Full Degradation of Methylene Blue Dye. Mater. Adv. 2025, 6, 2611–2621. [Google Scholar] [CrossRef]
  11. Zong, M.; Song, D.; Zhang, X.; Huang, X.; Lu, X.; Rosso, K.M. Facet-Dependent Photodegradation of Methylene Blue by Hematite Nanoplates in Visible Light. Environ. Sci. Technol. 2021, 55, 677–688. [Google Scholar] [CrossRef]
  12. Wang, T.; Nie, C.; Ao, Z.; Wang, S.; An, T. Recent Progress in G-C3N4 Quantum Dots: Synthesis, Properties and Applications in Photocatalytic Degradation of Organic Pollutants. J. Mater. Chem. A 2020, 8, 485–502. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Yu, H.; Zhai, R.; Zhang, J.; Gao, C.; Qi, K.; Yang, L.; Ma, Q. Recent Progress in Photocatalytic Degradation of Water Pollution by Bismuth Tungstate. Molecules 2023, 28, 8011. [Google Scholar] [CrossRef] [PubMed]
  14. Alam, U.; Verma, N. Direct Z-Scheme-Based Novel Cobalt Nickel Tungstate/Graphitic Carbon Nitride Composite: Enhanced Photocatalytic Degradation of Organic Pollutants and Oxidation of Benzyl Alcohol. Colloids Surf. Physicochem. Eng. Asp. 2021, 630, 127606. [Google Scholar] [CrossRef]
  15. Dirany, N.; Arab, M.; Leroux, C.; Villain, S.; Madigou, V.; Gavarri, J.R. Effect of WO3 Nanoparticles Morphology on the Catalytic Properties. Mater. Today Proc. 2016, 3, 230–234. [Google Scholar] [CrossRef]
  16. El Aouni, A.; El Ouardi, M.; Arab, M.; Saadi, M.; Haspel, H.; Kónya, Z.; Ben Ali, A.; Jada, A.; BaQais, A.; Ait Ahsaine, H. Design of Bismuth Tungstate Bi2WO6 Photocatalyst for Enhanced and Environmentally Friendly Organic Pollutant Degradation. Materials 2024, 17, 1029. [Google Scholar] [CrossRef]
  17. Dirany, N.; Hallaoui, A.; Valmalette, J.C.; Arab, M. Effect of morphology and temperature treatment control on the photocatalytic and photoluminescence properties of SrWO4 crystals. Photochem Photobiol Sci. 2020, 19, 235–250. [Google Scholar] [CrossRef] [PubMed]
  18. Ma, D.; Xue, Q.; Liu, Y.; Liang, F.; Li, W.; Liu, T.; Zhuang, C.; Zhao, Z.; Li, S. Manipulating interfacial charge redistribution in Mn0.5Cd0.5S/N-rich C3N5 S-scheme heterojunction for high-performance photocatalytic removal of emerging contaminants. J. Mater. Sci. Technol. 2026, 243, 265–274. [Google Scholar] [CrossRef]
  19. Chen, W.; Yan, R.; Chen, G.; Chen, M.; Huang, G.; Liu, X. Hydrothermal route to synthesize helical CdS@ZnIn2S4 core-shell heterostructures with enhanced photocatalytic hydrogeneration activity. Ceram. Int. 2019, 45, 1803–1811. [Google Scholar] [CrossRef]
  20. Pan, Z.; Wang, S.; Yan, R.; Song, C.; Jin, Y.; Huang, G.; Huang, J. Enhanced photocatalytic properties of Mn doped CdS catalysts by decomposition of complex precursors. Opt. Mater. 2020, 109, 110324. [Google Scholar] [CrossRef]
  21. Zhao, Z.; Wang, K.; Chan, L.; Yan, R.; Zhang, J.; Zhang, M.; Wang, L.; Chen, W.; Huang, G. Construction of S-scheme MIL-101(Fe)/Bi2MoO6 heterostructures for enhanced catalytic activities towards tetracycline hydrochloride photodegradation and nitrogen photofixation. Sol. Energy. 2023, 264, 112042. [Google Scholar] [CrossRef]
  22. Qian, J.; Shen, L.; Wang, Y.; Li, L.; Zhang, Y. Photo-Fenton Catalytic and Photocatalytic Performance of FeWO4 Nanorods Prepared at Different pH. Mater. Lett. 2023, 334, 133705. [Google Scholar] [CrossRef]
  23. Mahendran, N.; Udayakumar, S.; Praveen, K. PH-Controlled Photocatalytic Abatement of RhB by an FeWO4/BiPO4 p–n Heterojunction under Visible Light Irradiation. New J. Chem. 2019, 43, 17241–17250. [Google Scholar] [CrossRef]
  24. Abdelbasir, S.M.; Elseman, A.M.; Harraz, F.A.; Ahmed, Y.M.Z.; El-Sheikh, S.M.; Rashad, M.M. Superior UV-Light Photocatalysts of Nano-Crystalline (Ni or Co) FeWO4: Structure, Optical Characterization and Synthesis by a Microemulsion Method. New J. Chem. 2021, 45, 3150–3159. [Google Scholar] [CrossRef]
  25. Narendhran, S.; Shakila, P.B.; Manikandan, M.; Vinoth, V.; Rajiv, P. Spectroscopic Investigation on Photocatalytic Degradation of Methyl Orange Using Fe2O3/WO3/FeWO4 Nanomaterials. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 232, 118164. [Google Scholar] [CrossRef]
  26. Jansi Rani, B.; Ravi, G.; Yuvakkumar, R.; Praveenkumar, M.; Ravichandran, S.; Muthu Mareeswaran, P.; Hong, S.I. Bi2WO6 and FeWO4 Nanocatalysts for the Electrochemical Water Oxidation Process. ACS Omega 2019, 4, 5241–5253. [Google Scholar] [CrossRef] [PubMed]
  27. Guo, J.; Zhou, X.; Lu, Y.; Zhang, X.; Kuang, S.; Hou, W. Monodisperse Spindle-like FeWO4 Nanoparticles: Controlled Hydrothermal Synthesis and Enhanced Optical Properties. J. Solid State Chem. 2012, 196, 550–556. [Google Scholar] [CrossRef]
  28. Almeida, M.A.P.; Cavalcante, L.S.; Morilla-Santos, C.; Filho, P.N.L.; Beltrán, A.; Andrés, J.; Gracia, L.; Longo, E. Electronic Structure and Magnetic Properties of FeWO4 Nanocrystals Synthesized by the Microwave-Hydrothermal Method. Mater. Charact. 2012, 73, 124–129. [Google Scholar] [CrossRef]
  29. Liu, C.; Lü, H.; Yu, C.; Wu, X.; Wang, P. Hydrothermal-Assisted Microemulsion Synthesis of FeWO4 Nanorods and Their Superior Visible-Light- Driven Photocatalytic Activity. Mater. Lett. 2019, 257, 126707. [Google Scholar] [CrossRef]
  30. Irfan, M.; Tahir, N.; Zahid, M.; Noreen, S.; Yaseen, M.; Shahbaz, M.; Mustafa, G.; Shakoor, R.A.; Shahid, I. The Fabrication of Halogen-Doped FeWO4 Heterostructure Anchored over Graphene Oxide Nanosheets for the Sunlight-Driven Photocatalytic Degradation of Methylene Blue Dye. Molecules 2023, 28, 7022. [Google Scholar] [CrossRef]
  31. Dadigala, R.; Bandi, R.; Gangapuram, B.R.; Guttena, V. Construction of in Situ Self-Assembled FeWO4/g-C3N4 Nanosheet Heterostructured Z-Scheme Photocatalysts for Enhanced Photocatalytic Degradation of Rhodamine B and Tetracycline. Nanoscale Adv. 2019, 1, 322–333. [Google Scholar] [CrossRef]
  32. Shi, X.; Wang, L.; Zuh, A.A.; Jia, Y.; Ding, F.; Cheng, H.; Wang, Q. Photo-Fenton Reaction for the Degradation of Tetracycline Hydrochloride Using a FeWO4/BiOCl Nanocomposite. J. Alloys Compd. 2022, 903, 163889. [Google Scholar] [CrossRef]
  33. Zhang, J.; Wang, Y.; Li, S.; Wang, X.; Huang, F.; Xie, A.; Shen, Y. Controlled Synthesis, Growth Mechanism and Optical Properties of FeWO4 Hierarchical Microstructures. CrystEngComm 2011, 13, 5744–5750. [Google Scholar] [CrossRef]
  34. He, D.; Liu, X.; Li, X.; Lyu, P.; Chen, J.; Rao, Z. Regulating the Polysulfide Redox Kinetics for High-Performance Lithium-Sulfur Batteries through Highly Sulfiphilic FeWO4 Nanorods. Chem. Eng. J. 2021, 419, 129509. [Google Scholar] [CrossRef]
  35. Yu, F.; Cao, L.; Huang, J.; Wu, J. Effects of pH on the Microstructures and Optical Property of FeWO4 Nanocrystallites Prepared via Hydrothermal Method. Ceram. Int. 2013, 39, 4133–4138. [Google Scholar] [CrossRef]
  36. Parasuraman, B.; Kandasamy, B.; Murugan, I.; Alsalhi, M.S.; Asemi, N.; Thangavelu, P.; Perumal, S. Designing the Heterostructured FeWO4/FeS2 Nanocomposites for an Enhanced Photocatalytic Organic Dye Degradation. Chemosphere 2023, 334, 138979. [Google Scholar] [CrossRef] [PubMed]
  37. Sun, B.; Liu, Y.; Chen, P. Room-Temperature Multiferroic Properties of Single-Crystalline FeWO4 Nanowires. Scr. Mater. 2014, 89, 17–20. [Google Scholar] [CrossRef]
  38. 38. Variation of Unit Cell Parameters in Wolfranite Series. Available online: https://www.jstage.jst.go.jp/article/minerj1953/2/6/2_6_375/_article (accessed on 8 July 2025).
  39. Kovács, T.N.; Pokol, G.; Gáber, F.; Nagy, D.; Igricz, T.; Lukács, I.E.; Fogarassy, Z.; Balázsi, K.; Szilágyi, I.M. Preparation of Iron Tungstate (FeWO4) Nanosheets by Hydrothermal Method. Mater. Res. Bull. 2017, 95, 563–569. [Google Scholar] [CrossRef]
  40. Li, L.; Zhao, J.; Wang, Y.; Li, Y.; Ma, D.; Zhao, Y.; Hou, S.; Hao, X. Oxalic Acid Mediated Synthesis of WO3·H2O Nanoplates and Self-Assembled Nanoflowers under Mild Conditions. J. Solid State Chem. 2011, 184, 1661–1665. [Google Scholar] [CrossRef]
  41. Patil, V.B.; Adhyapak, P.V.; Suryavanshi, S.S.; Mulla, I.S. Oxalic Acid Induced Hydrothermal Synthesis of Single Crystalline Tungsten Oxide Nanorods. J. Alloys Compd. 2014, 590, 283–288. [Google Scholar] [CrossRef]
  42. Lassner, E.; Schubert, W.-D.; Lüderitz, E.; Wolf, H.U. Tungsten, Tungsten Alloys, and Tungsten Compounds. In Ullmann’s Encyclopedia of Industrial Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2000; ISBN 978-3-527-30673-2. [Google Scholar]
  43. Fang, Z.; Jiao, S.; Wang, B.; Yin, W.; Liu, S.; Gao, R.; Liu, Z.; Pang, G.; Feng, S. Synthesis of Reduced Cubic Phase WO3−x Nanosheet by Direct Reduction of H2WO4·H2O. Mater. Today Energy 2017, 6, 146–153. [Google Scholar] [CrossRef]
  44. Sun, M.; Xu, N.; Cao, Y.W.; Yao, J.N.; Wang, E.G. Nanocrystalline Tungsten Oxide Thin Film: Preparation, Microstructure, and Photochromic Behavior. J. Mater. Res. 2000, 15, 927–933. [Google Scholar] [CrossRef]
  45. Nayak, A.K.; Lee, S.; Choi, Y.I.; Yoon, H.J.; Sohn, Y.; Pradhan, D. Crystal Phase and Size-Controlled Synthesis of Tungsten Trioxide Hydrate Nanoplates at Room Temperature: Enhanced Cr(VI) Photoreduction and Methylene Blue Adsorption Properties. ACS Sustain. Chem. Eng. 2017, 5, 2741–2750. [Google Scholar] [CrossRef]
  46. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  47. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  48. Barrett, E.P.; Joyner, L.G.; Halenda, P.P. The Determination of Pore Volume and Area Distributions in Porous Substances. I. Computations from Nitrogen Isotherms. J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar] [CrossRef]
  49. Sun, D.; Iqbal, N.; Liao, W.; Lu, Y.; He, X.; Wang, K.; Ma, B.; Zhu, Y.; Sun, K.; Sun, Z.; et al. Efficient Degradation of MB Dye by 1D FeWO4 Nanomaterials through the Synergistic Effect of Piezo-Fenton Catalysis. Ceram. Int. 2022, 48, 25465–25473. [Google Scholar] [CrossRef]
  50. Ali, M.A.; Maafa, I.M.; Qudsieh, I.Y. Photodegradation of Methylene Blue Using a UV/H2O2 Irradiation System. Water 2024, 16, 453. [Google Scholar] [CrossRef]
  51. Diaz-Anichtchenko, D.; Aviles-Coronado, J.E.; López-Moreno, S.; Turnbull, R.; Manjón, F.J.; Popescu, C.; Errandonea, D. Electronic, Vibrational, and Structural Properties of the Natural Mineral Ferberite (FeWO4): A High-Pressure Study. Inorg. Chem. 2024, 63, 6898–6908. [Google Scholar] [CrossRef]
  52. Bera, S.; Rawal, S.B.; Kim, H.J.; Lee, W.I. Novel Coupled Structures of FeWO4/TiO2 and FeWO4/TiO2/CdS Designed for Highly Efficient Visible-Light Photocatalysis. ACS Appl. Mater. Interfaces 2014, 6, 9654–9663. [Google Scholar] [CrossRef]
  53. Wang, H.; Wang, C.; Cui, X.; Qin, L.; Ding, R.; Wang, L.; Liu, Z.; Zheng, Z.; Lv, B. Design and Facile One-Step Synthesis of FeWO4/Fe2O3 Di-Modified WO3 with Super High Photocatalytic Activity toward Degradation of Quasi-Phenothiazine Dyes. Appl. Catal. B Environ. 2018, 221, 169–178. [Google Scholar] [CrossRef]
  54. Ojha, D.P.; Karki, H.P.; Song, J.H.; Kim, H.J. Decoration of G-C3N4 with Hydrothermally Synthesized FeWO4 Nanorods as the High-Performance Supercapacitors. Chem. Phys. Lett. 2018, 712, 83–88. [Google Scholar] [CrossRef]
  55. Wang, C.; Wang, G.; Zhang, X.; Dong, X.; Ma, C.; Zhang, X.; Ma, H.; Xue, M. Construction of G-C3N4 and FeWO4 Z-Scheme Photocatalyst: Effect of Contact Ways on the Photocatalytic Performance. RSC Adv. 2018, 8, 18419–18426. [Google Scholar] [CrossRef]
  56. Lau, A.; Goh, C.Y.; Guo, Y.; Alsultan, A.G.; Hin, T.-Y.Y.; Nurhadi, M.; Lai, S.Y. Visible-Light Degradation of Methylene Blue Using Energy-Efficient Carbon-Doped TiO2: Kinetic Study and Mechanism. Bull. Chem. React. Eng. Catal. 2025, 20, 177–192. [Google Scholar] [CrossRef]
  57. Mohammed, W.; Matalkeh, M.; Soubaihi, R.M.A.; Elzatahry, A.; Saoud, K.M. Visible Light Photocatalytic Degradation of Methylene Blue Dye and Pharmaceutical Wastes over Ternary NiO/Ag/TiO2 Heterojunction. ACS Omega 2023, 8, 40063. [Google Scholar] [CrossRef]
  58. Albiss, B.; Abu-Dalo, M. Photocatalytic Degradation of Methylene Blue Using Zinc Oxide Nanorods Grown on Activated Carbon Fibers. Sustainability 2021, 13, 4729. [Google Scholar] [CrossRef]
  59. Mahlaule-Glory, L.M.; Mapetla, S.; Makofane, A.; Mathipa, M.M.; Hintsho-Mbita, N.C. Biosynthesis of Iron Oxide Nanoparticles for the Degradation of Methylene Blue Dye, Sulfisoxazole Antibiotic and Removal of Bacteria from Real Water. Heliyon 2022, 8, e10536. [Google Scholar] [CrossRef] [PubMed]
  60. Lin, R.; Chen, H.; Cui, T.; Zhang, Z.; Zhou, Q.; Nan, L.; Cheong, W.-C.; Schröck, L.; Ramm, V.; Ding, Q.; et al. Optimization of P-Type Cu2O Nanocube Photocatalysts Based on Electronic Effects. ACS Catal. 2023, 13, 11352–11361. [Google Scholar] [CrossRef]
  61. Ferrari, M.; Lutterotti, L. Method for the Simultaneous Determination of Anisotropic Residual Stresses and Texture by X-ray Diffraction. J. Appl. Phys. 1994, 76, 7246–7255. [Google Scholar] [CrossRef]
  62. Kalaycıoğlu, Z.; Uysal, B.Ö.; Pekcan, Ö.; Erim, F.B. Efficient Photocatalytic Degradation of Methylene Blue Dye from Aqueous Solution with Cerium Oxide Nanoparticles and Graphene Oxide-Doped Polyacrylamide. ACS Omega 2023, 8, 13004–13015. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of synthesized ferberite: (a) self-organized morphology and (b) nanoplatelets. (c) Coordination polyhedral (WO6 and FeO6 octahedral units) and (d) three-dimensional representations of the monoclinic unit cell of FeWO4.
Figure 1. X-ray diffraction patterns of synthesized ferberite: (a) self-organized morphology and (b) nanoplatelets. (c) Coordination polyhedral (WO6 and FeO6 octahedral units) and (d) three-dimensional representations of the monoclinic unit cell of FeWO4.
Molecules 30 04026 g001
Figure 2. SEM/TEM of ferberite (a,b) self-organized and (c,d) platelets.
Figure 2. SEM/TEM of ferberite (a,b) self-organized and (c,d) platelets.
Molecules 30 04026 g002
Figure 3. FTIR spectra corresponding to (a) stage 1 (acidification of a sodium tungstate solution to pH 1 with 3 M HCl), (a’) zoom section of spectra (a,b) stage 2 (addition of oxalic acid), (c) stage 3 (addition of iron sulfate heptahydrate) of the synthesis methodology and (c’) zoom section of spectra (c).
Figure 3. FTIR spectra corresponding to (a) stage 1 (acidification of a sodium tungstate solution to pH 1 with 3 M HCl), (a’) zoom section of spectra (a,b) stage 2 (addition of oxalic acid), (c) stage 3 (addition of iron sulfate heptahydrate) of the synthesis methodology and (c’) zoom section of spectra (c).
Molecules 30 04026 g003
Figure 4. TEM images at different stages of ferberite growth (a) after mixing the precursors corresponding to step 3. (b) An amorphous zone of the sample. (c) Disk-forming stage and (d) growth form and formation of platelets. (e) Goethite-like platelets formation only with iron precursor.
Figure 4. TEM images at different stages of ferberite growth (a) after mixing the precursors corresponding to step 3. (b) An amorphous zone of the sample. (c) Disk-forming stage and (d) growth form and formation of platelets. (e) Goethite-like platelets formation only with iron precursor.
Molecules 30 04026 g004
Scheme 1. Growth mechanism of ferberite. (a1) Tungstate oxalate. (a2) Tungstate hydrate. (b1) Iron–tungsten oxalic complexation. (b2) Iron–tungsten hydrate and (c) final ferberite formation step. Reproduced with permission from Ref. [33]. Copyright Elsevier. 2011.
Scheme 1. Growth mechanism of ferberite. (a1) Tungstate oxalate. (a2) Tungstate hydrate. (b1) Iron–tungsten oxalic complexation. (b2) Iron–tungsten hydrate and (c) final ferberite formation step. Reproduced with permission from Ref. [33]. Copyright Elsevier. 2011.
Molecules 30 04026 sch001
Figure 5. N2 adsorption/desorption isotherm of ferberite self-organized and platelets.
Figure 5. N2 adsorption/desorption isotherm of ferberite self-organized and platelets.
Molecules 30 04026 g005
Figure 6. Tauc plots and absorption spectra of FeWO4: (a) nanoplatelets and (b) self-organized.
Figure 6. Tauc plots and absorption spectra of FeWO4: (a) nanoplatelets and (b) self-organized.
Molecules 30 04026 g006
Figure 7. Catalytic activities of FeWO4 under various conditions for MB dye photodegradation: photolysis without catalyst, with H2O2 addition, with ferberite and ferberite/H2O2 system. Self-organized (ac) and platelets (df), at pH 3, 5 and 10, respectively.
Figure 7. Catalytic activities of FeWO4 under various conditions for MB dye photodegradation: photolysis without catalyst, with H2O2 addition, with ferberite and ferberite/H2O2 system. Self-organized (ac) and platelets (df), at pH 3, 5 and 10, respectively.
Molecules 30 04026 g007
Figure 8. Effect of pH on photodegradation of MB with and without H2O2. (a) self-organized, (b) platelets, and (c) pHpzc of self-organized and platelets with MB molecule.
Figure 8. Effect of pH on photodegradation of MB with and without H2O2. (a) self-organized, (b) platelets, and (c) pHpzc of self-organized and platelets with MB molecule.
Molecules 30 04026 g008
Figure 9. Photocatalytic degradation of MB at pH 5 with 4 mM H2O2: (a) self-organized microstructures, and (b) platelet morphology. Inserts: corresponding kinetic model fits, respectively, for self-organized (insert (a)) and platelet morphology (insert (b)). (c) Effect of H2O2 concentration on MB degradation. (d) Photograph color change in MB solution, after adsorption–desorption equilibrium and after 30 min of irradiation recyclability tests of (e) self-organized and (f) platelets samples.
Figure 9. Photocatalytic degradation of MB at pH 5 with 4 mM H2O2: (a) self-organized microstructures, and (b) platelet morphology. Inserts: corresponding kinetic model fits, respectively, for self-organized (insert (a)) and platelet morphology (insert (b)). (c) Effect of H2O2 concentration on MB degradation. (d) Photograph color change in MB solution, after adsorption–desorption equilibrium and after 30 min of irradiation recyclability tests of (e) self-organized and (f) platelets samples.
Molecules 30 04026 g009
Figure 10. Radicals trapping test by ferberite (self-organized and platelet morphologies) (measurements after 30 min).
Figure 10. Radicals trapping test by ferberite (self-organized and platelet morphologies) (measurements after 30 min).
Molecules 30 04026 g010
Figure 11. The transient photocurrent curves of self-organized and platelet-shaped FeWO4.
Figure 11. The transient photocurrent curves of self-organized and platelet-shaped FeWO4.
Molecules 30 04026 g011
Figure 12. (a) Mott–Schottky plots (1/C2 vs. V) of both FeWO4 electrodes measured at 300 Hz. (b) Energy band diagram with VB and CB edge positions in the NHE scale with the key ROS redox potentials involved in photodegradation mechanism.
Figure 12. (a) Mott–Schottky plots (1/C2 vs. V) of both FeWO4 electrodes measured at 300 Hz. (b) Energy band diagram with VB and CB edge positions in the NHE scale with the key ROS redox potentials involved in photodegradation mechanism.
Molecules 30 04026 g012
Table 1. Lattice parameters and crystallographic data of FeWO4 synthesized via hydrothermal method under acidic conditions using iron (II) sulfate heptahydrate as precursor.
Table 1. Lattice parameters and crystallographic data of FeWO4 synthesized via hydrothermal method under acidic conditions using iron (II) sulfate heptahydrate as precursor.
Compound FeWO4Self-OrganizedPlatelets
Space groupP2/c:1
Cell parameters (Å)a = 4.733a = 4.731
b = 5.710b = 5.708
c = 4.977c = 4.976
Angle (degrees)β = 90.334β = 90.292
Density (g/cm3)7.497.50
Crystal size (nm)67.7887.28
Reliability factors
Rwp (%)14.912.8
Rexp (%)12.910.5
Rwp/Rexp(S)1.161.22
Table 2. Comparison of photocatalytic performance of FeWO4 catalysts synthesized in this work and other representative reported FeWO4-based catalysts for MB degradation.
Table 2. Comparison of photocatalytic performance of FeWO4 catalysts synthesized in this work and other representative reported FeWO4-based catalysts for MB degradation.
Synthesis
Method
Morphology/
Composite
Volume
(mL)
Mass
Masse
(mg)
Irradiation[H2O2] (mM)pHTime (min)MB
Removal (%)
Reference
HydrothermalPlatelets/
FeWO4
100100Simulated sunlight453094–100This work
HydrothermalSelf-organized
/FeWO4
100100Simulated sunlight453092–99This work
HydrothermalNanorods/FeWO4300300Visible light66560~90[22]
Microemulsion/
Hydrothermal
Nanorods/FeWO410025Visible light512075[29]
MicroemulsionNi–FeWO4-100UV light36098[24]
Co-precipitation-assisted hydrothermalNanosheets/FeWO4–GO (I-doped)10015–40Sunlight72595[30]
HydrothermalNanofibers/
FeWO4
50100Visible light + ultrasound271499.5[49]
HydrothermalNI */(30wt%-TiO2-250)500100UV light51012089.53[56]
CoprecipitationSpherical/
(NiO/Ag/TiO2)
802/4/8UV lightNI6093.15[57]
Precipitation/
Hydrothemal
Nanorods/ZnO-NR/ACF10050UV light6.712099[58]
* NI—not informed by the authors.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gomes dos Santos, A.; Elaadssi, Y.; Chevallier, V.; Leroux, C.; Lopes-Moriyama, A.L.; Arab, M. Insights of Nanostructured Ferberite as Photocatalyst, Growth Mechanism and Photodegradation Under H2O2-Assisted Sunlight. Molecules 2025, 30, 4026. https://doi.org/10.3390/molecules30194026

AMA Style

Gomes dos Santos A, Elaadssi Y, Chevallier V, Leroux C, Lopes-Moriyama AL, Arab M. Insights of Nanostructured Ferberite as Photocatalyst, Growth Mechanism and Photodegradation Under H2O2-Assisted Sunlight. Molecules. 2025; 30(19):4026. https://doi.org/10.3390/molecules30194026

Chicago/Turabian Style

Gomes dos Santos, Andarair, Yassine Elaadssi, Virginie Chevallier, Christine Leroux, Andre Luis Lopes-Moriyama, and Madjid Arab. 2025. "Insights of Nanostructured Ferberite as Photocatalyst, Growth Mechanism and Photodegradation Under H2O2-Assisted Sunlight" Molecules 30, no. 19: 4026. https://doi.org/10.3390/molecules30194026

APA Style

Gomes dos Santos, A., Elaadssi, Y., Chevallier, V., Leroux, C., Lopes-Moriyama, A. L., & Arab, M. (2025). Insights of Nanostructured Ferberite as Photocatalyst, Growth Mechanism and Photodegradation Under H2O2-Assisted Sunlight. Molecules, 30(19), 4026. https://doi.org/10.3390/molecules30194026

Article Metrics

Back to TopTop