Next Article in Journal
The Role of Electrocatalysts in the Electrochemical Conversion of CO2 into Multi-Carbon Products (C2+): A Review
Previous Article in Journal
Construction of a Cofactor Self-Sufficient Enzyme Cascade System Coupled with Microenvironmental Engineering for Efficient Biosynthesis of Tetrahydrofolate and Its Derivative of L-5-Methyltetrahydrofolate
Previous Article in Special Issue
Eco-Friendly Synthesis of Thiazole Derivatives Using Recyclable Cross-Linked Chitosan Hydrogel Biocatalyst Under Ultrasonic Irradiation as Anti-Hepatocarcinogenic Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalytic Applications of Natural Iron Oxides and Hydroxides: A Review

by
Adriana Jiménez-Vázquez
1,*,
Raciel Jaimes-López
2,
Carlos Mario Morales-Bautista
1,
Samuel Pérez-Rodríguez
2,
Yadira Gochi-Ponce
3 and
Luis Alberto Estudillo-Wong
2,*
1
División Académica de Ciencias Básicas, Universidad Juárez Autónoma de Tabasco, Km. 1 Carretera Cunduacán-Jalpa de Méndez A.P. 24, Cunduacán C.P. 86690, Mexico
2
Laboratorio de Electroquímica Ambiental y de Materiales (LEAyM), Departamento de Biociencias e Ingeniería, CIIEMAD-IPN, Instituto Politécnico Nacional, Calle 30 de Junio de 1520 s/n Barrio la Laguna Ticomán, Alcaldía GAM, Ciudad de México C.P. 07340, Mexico
3
Tecnológico Nacional de México, Instituto Tecnológico de Tijuana, Posgrado en Ciencias de la Ingeniería, Blvd. Alberto Limón Padilla s/n, Mesa de Otay, Tijuana C.P. 22500, Mexico
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(3), 236; https://doi.org/10.3390/catal15030236
Submission received: 2 February 2025 / Revised: 21 February 2025 / Accepted: 21 February 2025 / Published: 28 February 2025
(This article belongs to the Special Issue Catalytic Energy Conversion and Catalytic Environmental Purification)

Abstract

:
Iron oxides and hydroxides (Fe-OH) extracted from natural sources have garnered significant attention for their diverse catalytic applications. This article provides a comprehensive overview of the catalytic potential of naturally occurring Fe-OH, focusing on the influence of natural sources and preparation methods on their morphological characteristics and application in heterogeneous catalysis. The unique physicochemical properties of these catalysts, including their high surface area, redox activity, and tunable surface chemistry, make them promising candidates for various catalytic processes. The review discusses key catalytic reactions facilitated by natural Fe-OH, such as advanced oxidation processes (AOPs), electrochemical applications, catalytic cracking, and biodiesel production. Furthermore, it highlights recent advancements and challenges in utilizing these materials as heterogeneous catalysts. By presenting an analysis of the catalytic potential of natural iron oxides, this review aims to stimulate further research about the use of these materials, which are widely distributed in the Earth’s crust.

1. Introduction

Iron oxides and hydroxides, both referred to here as Fe-OH, are ubiquitous earth materials with high adsorption capacities for toxic elements and degradation ability towards organic contaminants. By mass, iron is the fourth most abundant element in the Earth’s crust, being found in 16 Fe-OH [1]. Iron oxides are widely distributed in nature and can be found across various components of the global system, including the lithosphere, hydrosphere, biosphere, pedosphere, and atmosphere. They participate in the complex interactions between these environmental compartments [2]. Iron’s different oxidation states (ferric (Fe3+) and ferrous (Fe2+)) further influence its interactions within these systems. For instance, iron availability in the oceans affects the growth of phytoplankton, impacting the entire marine food chain.

Geological Abundance of Iron Oxide and Hydroxide Minerals

Iron constitutes a significant portion of minerals on our planet, ranging from 28 to 35% of Earth’s total mass. The Earth’s inner core primarily comprises solid iron (Fe) and nickel (Ni), while the outer core consists of iron alloyed with 10–20% of other elements, such as oxygen (O), sulfur (S), and potassium (K). In the Earth’s structure, the primitive mantle contains roughly 12% iron, whereas the crust holds 4.32% iron, ranking just below oxygen (47.2%), silicon (Si, 28.8%), and aluminum (Al, 7.96%) in abundance. Consequently, iron silicates and iron oxides emerge as the most prevalent minerals in the Earth’s crust [3]. Table 1 enumerates the naturally occurring iron oxide and hydroxide minerals. Goethite, hematite, and magnetite are the most prevalent, often occurring as standalone rock formations in nature. Lepidocrocite, ferrihydrite, and maghemite are found in significantly smaller quantities but are still distributed across many locations. On the other hand, wüstite, akaganéite, feroxyhyte, and bernalite are comparatively rare, both in terms of their distribution and abundance [4].
The primary source of iron is the rocks known as banded iron formations (BIFs). These formations are named for their characteristic composition, featuring alternating layers of iron (hematite or magnetite) and silica. Iron can also be found in geological formations like greenschist, gabbro, basalt, amphibolite, and others [5].
Schwertmann and Fitzpatrick (1993) [6] describe various notable features of iron oxides found on Earth’s surface. These minerals are physicochemical environmental indicators with unique mineral characteristics and strong pigmenting abilities, even at low concentrations. They tend to have fine grain sizes (5–100 nm) and large surface areas, allowing functionalized surfaces to absorb a range of molecules. Iron oxides are also highly insoluble and may include other cations like aluminum in their crystal structure. Due to their limited crystallinity, advanced techniques, such as Mössbauer spectroscopy, X-ray absorption near-edge structure (XANES), thermogravimetric analysis (TGA), X-ray diffraction (XRD), infrared spectroscopy (IR), differential thermal analysis (DTA), transmission electron microscopy (TEM), and selective chemical dissolution, are essential for identification [6].
Over the past decade, numerous researchers have investigated the utilization of naturally derived iron oxides as heterogeneous catalysts, driven by their widespread viability, readily accessible sources, low production costs, appealing physicochemical properties, and inherent biocompatibility [7,8,9]. The success achieved in numerous applications has led to reviews of articles on the subject. However, to the best of our knowledge, these works have focused on specific processes or the use of specific minerals. Therefore, there is a need for a comprehensive and integrative view of the topic. Fenton’s reaction, a well-established catalytic process, utilizes ferrous and/or ferric ions to decompose hydrogen peroxide, generating potent oxidizing agents capable of degrading a wide range of organic and inorganic contaminants. This reaction exemplifies a catalytic application of iron oxides. Thus, it is only tangentially addressed, referring readers to the reviews already published. A search on Scopus with the search terms TITLE (* “iron” AND “Catalyst” AND “review”) yields 35 results, of which only 14 are about natural minerals of iron oxyhydroxides. Three of these address their use in Fenton, photo-Fenton, and electro-Fenton processes [10,11,12], with four on water–gas shift, syngas and Fischer–Tropsch synthesis [13,14,15,16], three on persulfate activation [17,18,19], one on peroxides [20], one on nanoparticles applied in different catalysts [21], and two on advanced oxidation [22,23]. Additionally, a review of the performance of magnetite [24] in Fenton-like processes is identified. In just the last two years, by using the search terms ’iron’, ’oxide’, ’catalysts’, and ’review’ in article titles, the SciFinder® search platform has yielded additional examples focused on Fischer–Tropsch synthesis [25,26], persulfate activation [27], NOx reduction [28,29], oxygen reduction reaction [30], and selective oxidation of hydrogen sulfide [31]. However, none of these address the use of natural materials. However, other catalytic processes involving iron oxides have also been tested. Among these applications are hydrocracking of coal tar, biodiesel production, hydrogenation, ammonia synthesis, reduction of nitrates, heterogeneous catalytic ozonation, oxidation and reduction of inorganic ions, hydrogen production, and water dissociation, which have not been sufficiently disseminated and analyzed in the literature. The main catalytic application of Fe-OH and the iron species used in them are depicted in Figure 1.
To the authors’ knowledge, there has been no previous attempt to present, in a single document, most of the applications proposed in the literature for naturally derived Fe-OH. In this review, chemical industry research professionals will find a comprehensive compilation of the identified applications of Fe-OH, especially in the last two decades. The aim is to provide an integrative approach, emphasizing the abundance and versatility of natural iron catalysts, which makes them economically attractive and environmentally benign. This document presents basic information on the relative abundance of Fe-OH in the Earth’s crust, their occurrence in nature, pretreatments of the catalysts, and both common and uncommon applications. It is divided into three sections: the first one for the most abundant, the second one for those less abundant but distributed globally, and finally, the third one for those considered difficult to obtain.

2. Most Prevalent Iron Oxides and Hydroxides

Magnetite (Fe3O4) and hematite (α-Fe2O3) are among the most prevalent iron oxides, while goethite (α-FeOOH) is the most abundant iron hydroxide. In this section, we will explore principal characteristics and both the primary and less common catalytic applications of these iron oxides and hydroxides.

2.1. Magnetite (Fe3O4)

Magnetite is a widely distributed accessory mineral in numerous ore deposits and ranks among the most abundant oxide minerals in the continental crust [32]. It occurs globally and can form under diverse temperature and pressure conditions in igneous, metamorphic, and sedimentary settings [33].
At ambient conditions, they crystallize in the inverse spinel cubic structure, which belongs to the space group F d 3 ¯ m . Its distinctive structure is characterized by a combination of diverse anions and cations. In a conventional spinel arrangement, divalent cations exclusively occupy tetrahedral positions, while trivalent cations are confined to octahedral coordination. However, within an inverse spinel configuration, the octahedral sites accommodate both divalent and trivalent cations simultaneously [34]. Fe3+ ions are positioned in tetrahedral (A) sites coordinated by oxygen, while octahedral (B) sites contain an equal mixture of Fe3+ and Fe2+ ions. Below the magnetic transition temperature of 850 K, the spins of iron atoms in the A and B sublattices align in a ferrimagnetic configuration [35].
Fe3O4 is the most utilized among various iron oxides, including FeO, γ-Fe2O3, and FeTiO3, because it exhibits a notable degree of spin polarization at the Fermi level, a high Curie temperature, and electrocatalytic effect, stability, and ease of synthesis. Furthermore, nano-sized Fe3O4 shows additional advantageous features beyond those inherited from its bulk form (see Table 2), such as remarkable size and shape controllability, a large specific surface area, and excellent magnetothermal conversion capability. Magnetite nanoparticles have been utilized in a wide range of catalytic applications, such as drug delivery, cancer treatment, organic synthesis, environmental remediation, synthesis of biodiesel, and electrocatalysis [36]. Magnetite is widely used in advanced oxidation processes (AOPs) [24]; however, its catalytic applications include organic synthesis, electrocatalysis, and environmental remediation, among others [36].
The catalytic performance of magnetite nanoparticles is strongly influenced by their size. In general, smaller nanoparticles exhibit enhanced catalytic efficiency due to their higher surface area and greater number of active sites compared to larger particles. Magnetite-based nanocatalysts are gaining prominence as adaptable tools in heterogeneous catalysis, contributing to the advancement of sustainable reaction methodologies [37].
Numerous methods exist for synthesizing bulk and magnetite nanoparticles [38]. However, the extraction of magnetite nanoparticles from natural sources has received limited attention. While some studies have explored this area [39,40,41,42,43], the primary objective of these syntheses typically differs from catalytic applications. Furthermore, it has been reported that different natural sources of magnetite and preparation methods significantly influence the morphology and size of the particles.
Magnetite prepared from iron sand via coprecipitation shows a more uniform morphology and particle size distribution. In contrast, the ball milling method allows control of particle size based on milling time, with longer times resulting in smaller particles but more irregular shapes (Figure 2A). On the other hand, when the natural source is mineral magnetite, the particles have more regular shapes. Nevertheless, the particle size distribution is broader (Figure 2B,C).
Main catalytic applications. Magnetite’s capacity to generate hydroxyl radicals enables its utilization in various advanced oxidation processes (AOPs), including Fenton, electro-Fenton, photo-Fenton, and Fenton-like reactions. These processes share the common principle of involving the interaction between iron-based catalysts and H2O2 to generate highly reactive hydroxyl radicals for degrading organic pollutants in water. Variations arise in activation methods: in electro-Fenton, H2O2 is generated by O2 reduction at the cathode, photo-Fenton utilizes light to generate OH and Fe2+ from water and Fe3+, and Fenton-like reactions use alternative iron species or transition metals [47]. According to catalytic evaluations and findings performed on various iron oxides in Fenton reactions, magnetite is highly active for both H2O2 decomposition and O2 production at neutral and basic pH [48]. However, due to the magnetic nature of magnetite, the reactor design must consider conditions to prevent particle agglomeration, which could inhibit its catalytic activity by limiting proper dispersion in the solution. To address this issue, magnetic stirring is omitted, using centrifugal [49] or peristaltic pumps [50] for the recirculation of contaminants.
Mineral magnetite has been used to obtain catalysts for the degradation of organic pollutants such as p-nitrophenol [1], textile dye [44,49,51], methylparaben [52], and cefotaxime [46]. The natural sources used for the preparation of these catalysts were mineral magnetite and ferruginous sands. Additionally, the preparation of the catalysts commonly involves magnetic and gravitational separation as well as sieving and ball milling. In some cases, chemical treatment with acid, mainly HCl, is preferred to remove impurities present in the natural magnetite source.
Typical ranges of dosage for the Fenton and Fenton-like reactions are between 0.1 and 0.5 g·L−1, and optimal degradation occurs in acidic or neutral conditions, with pH values of 3 to 7. Although H2O2 is the primary oxidizing agent used, other species, such as S2O82− [49,52], can also be employed. Magnetite obtained from natural sources commonly contains other minerals, including Cr, Mn, Ni, Co, and Ti, among others. However, these impurities can enhance catalytic activity. For example, the comparison between synthetic magnetite and mineral magnetite [1] for p-nitrophenol (p-NP) degradation indicates that the combination of nano-sized magnetite and H2O2 does not exhibit significant degradation throughout the entire process, underscoring the superior catalytic efficacy of natural magnetite compared to its synthetic counterpart. This fact was attributed to the presence of exsolved spinel due to the incorporation of transition metals with variable valence.
The preparation method of catalysts from natural sources significantly influences control over particle size, shape, and purity. Mufti et al. [51] and Amiruddin et al. [44] prepared magnetite nanoparticles using different approaches (coprecipitation and ball milling) and tested them in photo-Fenton degradation and Fenton degradation of methylene blue (MB), respectively. The coprecipitation method resulted in smaller particle sizes compared to the ball milling method (11 nm vs. 24–49 nm), with more uniform particle size distribution and a high-purity crystal phase; also, SBET was higher compared to the usual values reported for Fe3O4. However, it involved the use of high molar concentrations of HCl and NH4OH, as well as thermal treatment, which are avoided in the milling method. Although the catalytic performance was evaluated under different conditions (photo-Fenton and Fenton), the results for MB degradation were similar (80% vs. 86.89%). However, the reaction times required to achieve these results differed significantly (30 min vs. 120 min). Similarly, in the Fenton-like degradation of acid orange 7 (AO7) using natural magnetite and persulfate (S2O82−), high degradation efficiency (90%) was achieved within 120 min under optimized conditions, including 0.5 gL−1 of catalyst and 0.2 mM of S2O82− under acidic pH [49]. The catalyst exhibited low iron leaching and the potential to be scalable, while the reaction time was longer compared to systems leveraging photo-Fenton activation.
The catalytic degradation of methylparaben (MeP) using a Fenton-like process demonstrated the effectiveness of persulfate (SPS) activation by natural magnetite as a heterogeneous catalyst. Peroxymonosulfate and peroxydisulfate are common ions in aqueous effluents, whose presence can be leveraged to generate strong oxidants (SO4., OH, and 1O2) through appropriate catalysts [53]. The study achieved a high degradation efficiency of 99.5% and mineralization rates of up to 37% within 60 min under optimized conditions, including a catalyst loading of 0.1–0.3 g·L−1 and SPS concentrations of 1–5 mmol·L−1 under acidic pH. The catalyst exhibited excellent reusability, maintaining over 90% degradation efficiency after five cycles, with minimal decline in performance. This highlights its potential as a cost-effective and robust option for pollutant removal [52].
The preparation of composites of TiO2/magnetite nanoparticles using natural magnetite and P25 as raw materials resulted in particle sizes of about 20–80 nm [46]. No significant increase in SBET was observed. These particles were tested on Fenton, photo-Fenton, and photocatalysis processes. In the photo-Fenton under UV light and photocatalysis processes, the composites achieved complete degradation of cefotaxime, whereas in the Fenton reaction, the degradation was negligible. The higher degradation rate reached by mediated light degradation was attributed to the fact that Fe3+ in the magnetic nanoparticle could capture photo-generated electrons from TiO2, which accelerated Fe2+/Fe3+ cycles in the Fenton reaction, the separation of electron–hole pairs in photocatalysis, the decomposition of H2O2, and the production of OH radical, resulting in the effective removal of cefotaxime. Regarding its application in an electro-Fenton process, it was able to reach a degradation efficiency of about 40% on the degradation of gemcitabine at a pH of 3, which is related to ferrous ions in its structure [54].
In an initial approach to synthesizing nanoparticles from mineral magnetite, Munoz et al. [55] evaluated their catalytic activity in the catalytic wet peroxide oxidation (CWPO) of phenol. In this mode of the Fenton process (heterogeneous), the catalyst used to decompose H2O2 into OH is an insoluble solid [56]. The preparation involved only sieving the raw material, which was obtained from Marphil, Madrid, Spain. This resulted in a Fe3O4 catalyst with an SBET of 8 m2g−1 and a saturation magnetization (Ms) of 77.7 emu g−1. Under optimal reaction conditions (H2O2 = 500 mg·L−1, Fe3O4 = 2 g·L−1, T = 75 °C, pH = 3, and t = 4 h), the enhanced performance of the minerals was confirmed in the CWPO of phenol (100 mg·L−1), achieving complete conversion of the target pollutant with a high degree of mineralization (70–80%). However, even though the process was effective, relatively low oxidation rates were observed, requiring up to 3.5 h of reaction time to reach the complete removal of the pollutant under ambient temperature and circumneutral pH. Therefore, mineral magnetite was modified through thermal treatments of controlled oxidation and reduction to obtain core–shell materials [57]. A slight growth in the magnetization of the material was observed with an increase in the reduction temperature. At a pH of 5, ambient temperature (25 °C), the stoichiometric amount of H2O2 for the complete mineralization of sulfamethoxazole (SMX), and a catalyst load of 1 g·L−1 were selected. The initial concentration of SMX was fixed at 5 mg·L−1. Reduced iron species exhibited greater activity in the CWPO of SMX, leading to complete removal within 1.5 h. The optimal catalyst was reduced at 400 °C. This material remained active even after five reaction cycles.
Recently, the CWPO of azole pesticides, using natural magnetite as a catalyst, has been investigated [50]. The study utilized a fixed-bed reactor packed with natural magnetite powder to eliminate a representative blend of azole pesticides. System performance was assessed by examining the effects of inlet flow rate (0.25–1 mL·min−1), magnetite loading (2–8 g), H2O2 dosage (3.6–13.4 mg·L−1), and initial pollutant concentration (100–1000 µg·L−1) over 300 h of continuous operation. Under optimized conditions (T = 25 °C, pH = 5, flow rate = 0.5 mL·min−1, [H2O2] = 6.7 mg·L−1, and [Fe3O4] = 8 g), azole pesticide removal exceeded 80%. The catalytic system exhibited high stability over 500 h of operation, with limited iron leaching. Also, the efficiency of the catalyst was proven in the treatment of a real wastewater treatment plant (WWTP) effluent sample, spiked with a mixture of azole pesticides at 500 µg·L−1. The process was less efficient in the WWTP effluent; this was attributed to the presence of organic and inorganic compounds in the treated water since they exhibit OH scavenging properties. On CWPO, H2O2 dosage and catalyst loading are the most significant variables that influence catalytic activity; even with low SBET and high particle size, effective degradation can be reached. The reactions are described well by pseudo-first-order rate constants.
Non-common catalytic applications. The use of catalysts based on magnetite obtained from natural sources is not limited to environmental remediation applications. It has also been tested in the conversion of spent engine oil into reusable diesel-like fuel [45]. This process, known as catalytic pyrolysis, involves the use of solid catalysts to break down long organic matter chains into lighter hydrocarbons, promoting specific transformations such as dehydration, dehydrogenation, deoxygenation, and decarboxylation [58]. Natural magnetite was extracted from freshly mined raw ore through magnetic separation. Pyrolysis reactions were conducted at 500 °C with a residence time of 90 min. The overall yield was close to 100% in a thermal non-catalyzed run and decreased linearly with magnetite dosage. However, the results indicated that magnetite was selective to liquid fraction and could promote different product compositions due to cracking and oxidative dehydrogenation of the paraffinic and alkyl aromatics. The resultant catalytically derived liquid product has fuel properties comparable to commercial diesel. Recently, CaO/MgO/Fe3O4 prepared from iron sand and dolomite was tested as a catalyst for biodiesel production from waste cooking oil [57]. The catalyst calcinated at 1000 °C reached 95.64% yield. The produced biodiesel fulfilled the quality standards for density and viscosity.
Pyrolysis is a promising option in regions with limited petroleum and gas resources because it offers a solution by producing char, tar, and gases simultaneously. Char can be used for combustion or gasification, while tar can be processed into liquid fuels and chemicals, improving economic feasibility. However, coal-derived tar often contains heavy compounds that complicate processing and cause operational challenges. To address this, He et al. [59] proposed a coal pyrolysis method combined with iron ore reduction. Iron facilitates the breaking of C-C and C-O bonds. The study used magnetite with a chemical composition primarily consisting of Fe (62.65%) and minor components like SiO2 and MgO. Pyrolysis experiments were conducted using a Pt-filament pyrolyzer, with volatiles adsorbed and later desorbed at 300 °C for GC/MS analysis. Coal and catalyst samples were heated to 700 °C under argon. Magnetite showed a low surface area (3.32 m2·g−1) and primarily produced phenols and CO2 but performed poorly compared to other iron ores like ilmenite and siderite. Also, the nanohybrid catalyst Fe@C, using natural magnetite as an iron source, was evaluated on Fischer–Tropsch synthesis, a collection of reactions for converting syngas (hydrogen and carbon monoxide) into liquid hydrocarbons [60]. The core–shell structure was confirmed by TEM analysis. The results indicated that CO conversion exceeded 98%, with C5 hydrocarbon selectivity of 51.7% and stability over 180 h of steam [61].
Table 3 summarizes the catalytic applications of hematite prepared from natural sources. Although its main applications are Fenton-type processes, its use for fuel production is also notable. Furthermore, the analysis of the elimination based on the main parameters in AOPs (Figure 3) shows that the specific surface is not a determining factor in the catalytic activity of magnetite; likewise, acidic or neutral conditions allow better performance. In addition, specific surface area is not a determining factor in the catalytic activity of magnetite, while acidic or neutral pH conditions enable better performance. Even with small specific surface areas and low catalyst concentrations, magnetite enables removal rates close to 100% in advanced oxidation processes.

2.2. Hematite (α-Fe2O3)

Hematite (α-Fe2O3) is the oldest recognized form of iron oxide, widely found in soils and rocks. It is more commonly found in sedimentary rocks, either as detrital particles or as precipitates. Hematite is abundant in environments with ample oxygen as the ferric oxyhydroxides initially formed age and transform into hematite. It typically coexists with other iron oxides like goethite. Hematite formation is favored by low water content and elevated temperatures [64]. When finely powdered, hematite exhibits a reddish hue, while in bulk form, it appears black or gray. Known for its exceptional stability, hematite often represents the final stage in the transformation of other iron oxides [2]. Its crystalline structure resembles corundum, with lattice parameters a = 5.0340 Å and c = 13.750 Å.
This mineral displays two magnetic transitions, specifically the Morin transition (TM) around T = 260 K and the Curie transition at approximately 950 K. The manifestation of its weak ferromagnetic conduct stems from a slight misalignment in spin antiparallelism. Notably, a correlation is established between the TM and the hematite particle size, with smaller particle sizes correlating to reduced TM values. However, this relationship is heavily contingent upon preparation methodologies, crystalline lattice defects, and the incorporation of OH groups into the hematite structure [65].
Magnetic properties are intimately tied to nanoparticle morphology and size. For example, the magnetization of saturation decreases and coercivity increases with Mn doping, which are attributed to the presence of more manganese atoms at the grain boundaries [66]. The coercivity interval (Hc) spans from 31 to 530 Oe, while the remanent magnetization (Mr) ranges between 0.6 and 16 Am2 kg1 [67,68,69]. Higher Hc and Mr values are closely linked to larger particle sizes. Conversely, smaller hematite particles of varied morphologies exhibit superparamagnetic behavior beyond the blocking temperature [70]. It is deemed a n-type semiconductor, characterized by a bandgap of 2.2 eV. The valence band comprises occupied 3d orbitals of Fe3+ alongside some non-bonding 2p orbitals of oxygen. Meanwhile, the conduction band is composed of vacant 3d orbitals of Fe3+ [71]. Regarding its specific surface area, it is recognized that this parameter is influenced by the synthesis method and calcination temperature. Hematite subjected to thermal treatment between 1173 and 1273 K yields a specific surface area below 5 m2·g1 due to particle sintering. Broadly, hematite produced at low calcination temperatures (<373 K) possesses specific surface areas ranging from 10 to 90 m2·g1, while those generated through Fe3+ hydrolysis are dependent on particle size and shape, encompassing values between 2 and 30 m2·g1 [2].
Main catalytic applications. While numerous publications document the use of hematite for dye removal, the majority of these studies focus on laboratory-synthesized hematite. However, dye removal using hematite-based catalysts from natural sources has been tested by some research groups. Photo-Fenton [72], Fenton [73,74], Fenton-like [75], and photocatalysis reactions [66,76,77,78] are the main applications where hematite has been tested. In photocatalysis, photon absorption by a material excites electrons from the valence band to the conduction band. The resulting electrons and holes react with O2 and H2O, generating O2•− and OH radicals [76]. Although the preparation of hematite catalysts can result in various morphologies, temperature is the key parameter influencing this characteristic. For instance, catalysts prepared from natural hematite (Figure 4A) and from iron ore (Figure 4B) underwent a calcination process at 700 °C, resulting in similar spherical morphologies.
Phenol is naturally found in oilfield-produced waters, and it is commonly used in sectors such as pesticides and pharmaceuticals. Due to its persistence in the environment, much research focuses on its degradation by different catalytic processes. In an interesting approach, a catalyst using natural clay with cassava starch as a porogenic agent and raw material was tested on phenol photodegradation [78]. The catalyst presented around 40% of α-Fe2O3. The maximum phenol photodegradation achieved was 59.15% at a cassava starch concentration of 7%. Even after eight recycling cycles, the degradation efficiency remained above 90%. However, the synthesis method is more complex compared to other preparation methods [74] used to obtain catalysts for similar applications. Iron-bearing mining rejected from the amethyst geode deposit was processed by grinding, ball milling, and sieving to produce a catalyst for the photo-Fenton degradation of phenol. After this processing, the final percentage of Fe2O3 was close to 25%. Although the material had low porosity, it demonstrated significant catalytic efficiency in phenol degradation, achieving mineralization and removal efficiencies of 78% and 96.5%, respectively. This performance is largely attributed to its iron content, which effectively facilitates OH radical generation. Both studies demonstrated innovative approaches for preparing Fe2O3-based catalysts using low-cost raw materials. Despite differences in synthesis methods and reaction conditions, both highlight sustainable strategies for wastewater treatment.
A similar organic compound, namely, the low-molecular-mass 4-nitrophenol (4-NP), which is involved in many chemical processes and commonly present in soils and in surface and ground waters, causes severe environmental impacts and health risks. Fe2O3 [79] and Cu/Fe2O3 [77] catalysts were evaluated on the photocatalytic reduction of 4-NP to 4-aminophenol (4-AP), a less harmful compound. The raw material (iron ore) was crushed, sieved, and thermally treated to prepare the catalysts. Even if goethite was the main phase of iron oxide in the raw material, after heat treatment at 450 °C, only peaks related to hematite were found in an XRD powder pattern. The morphology of the catalyst prepared from natural sources and from iron nitrate is spherical in both cases. TEM analysis revealed a significant difference in particle size, with values of 50 nm (from natural sources) and 20 nm (synthesized). Impregnation with Cu (5 and 20 wt%) and calcination at higher temperatures resulted in a smaller particle size (50 nm) and reduction in SBET (38 m2·g−1), compared to hematite calcinated at 700 °C (34 nm and 10 m2·g−1, respectively). The same reaction conditions were applied to 4-NP reduction, and conversion reached >99% in both cases. The rate constant was higher using Cu/Fe2O3. This was attributed to the presence of the copper ferrite CuFe2O4.
Hematite catalysts from natural sources have also been successfully tested in dye degradation. For the oxidation and mineralization of methyl orange (MO), catalysts were prepared from Fe-sand [72]. The sand underwent a series of purification steps to enhance its catalytic properties. Degradation experiments were conducted in a cylindrical batch photoreactor under UV irradiation. A higher rate constant of 0.048 min−1 was achieved for a dye concentration of 100 mg·L−1. One of the major current challenges is the scaling up of catalysts to enable their use in treatment plants [80]. A first step towards this goal is to test them in the laboratory using wastewater samples from real effluents. Under optimized conditions, including 200 mL of textile effluent, 0.3 g of catalyst, pH of 2.5, and an initial H2O2 concentration of 200 mg·L−1, the system achieved an 88% reduction in optical density and a 79% decrease in chemical oxygen demand (COD) after 60 min. One interesting property of using iron oxide catalysts is their magnetic response, which allows for easy recovery and subsequent reuse. In this case, the catalyst was able to reach a degradation of 89% after four cycles, with no obvious leaching. Also, hematite can be used as a catalyst in combination with clay for the photocatalytic degradation of dye using raw iron ore [76]. α-Fe2O3/bentonite was further evaluated through photocatalytic degradation on indigo carmine (IC) of a composite prepared by mechanical milling, followed by magnetic separation and purification with HCl. Photocatalytic evaluation was performed under UV and solar light. The α-Fe2O3/bentonite composite has high photocatalytic activity on the removal of IC dye compared to extracted α-Fe2O3. The maximum removal efficiency (almost 100%) of IC was achieved with the following optimal conditions: pH = 1, 250 mg of catalyst, and 5 mg·L−1 of IC. The photocatalytic activity under solar light was higher than that of UV irradiation. The removal of 88.88 % of MB was achieved by Mg/Fe2O3 prepared from iron sand in a photo-Fenton reaction [66].
The growing interest in the degradation of emergent pollutants results in the evaluation of novel or well-known catalysts in processes related to them. Among these emerging contaminants are herbicides. The simultaneous catalytic degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) and 2-methyl-4-chlorophenoxyacetic acid (MCPA) herbicides over persulfate (PS) activated by mineral hematite via Fenton-like processes resulted in 36% of mineralization under optimum conditions (0.5 g·L−1 of catalyst; 200 mg·L−1 of 2,4-D and MCPA; 0.025 M of PS; pH = 3; T = 50 °C; and t = 120 min) [73]. The FESEM analysis indicated that the catalyst was almost porous, providing active sites for the adsorption of both pollutants and oxidant molecules. The synergistic effect between PS and catalysts showed better performance in acidic environments and higher temperatures in the degradation of herbicides.
Due to its abundance on Earth’s surface, different sources can be used to obtain hematite. In this line of knowledge, the same research group evaluated the use of Hormuz Red Soil (HRS) as a catalyst for the degradation of diclofenac (DCF) via PS activation [75]. The catalyst beads were prepared by ionic gelation encapsulation using alginate beads. For comparison, HRS in powder form was evaluated too. The catalytic experiments were conducted in two modes: batch experiments for the activation of peroxymonosulfate (PMS) by HRS and supported HRS by alginate beads in flow-through reactors. X-ray fluorescence (XRF) indicated that hematite was the main oxide on HRS (59.30%), with SiO2 as the second most abundant compound (16.72%). Fresh catalyst particles were spherical, while HRS beads displayed a rougher irregular surface due to HRS crosslinking with alginate, enhancing the surface area and active sites for improved catalytic performance. Additionally, 98.2% of removal (HRS = 5 mg·L−1; 75 mg·L−1 of PMS; 50 mg·L−1 of DCF; and t = 10 min) was achieved with HRS in suspension with PMS. Complete degradation was achieved in HRS supported on alginate beads in batch-mode processes (five granules of HRS beads; 75 mg·L−1 of PMS; 50 mg·L−1 of DCF; t = 2 min; and neutral pH). The alginate-supported HRS demonstrated superior performance, achieving complete degradation in significantly shorter reaction times under neutral pH conditions. However, the batch experiment was also efficient, with nearly 100% removal. Also, the catalytic ozonation and peroxone-mediated removal of acetaminophen (ACT) by HRS resulted in the complete degradation of 50 mg·L−1 of ACT at 10 min. The calcinated catalyst exhibited high stability and reusability in consecutive catalytic cycles [81].
At the degradation of gemcitabine via the electro-Fenton process, hematite had poor degradation efficiency (20%) [54]. The analysis of parameters in Fenton and Fenton-like reactions catalyzed by hematite indicates that a narrow interval of pH values is available (Figure 5). This trend indicates that higher values of all these parameters cause a decrease in the removal efficiency of different contaminants in reactions catalyzed by hematite.
Non-common catalytic applications. While Fenton and Fenton-like reactions constitute the most prevalent applications of hematite, its utility has also been explored in hydrogen reactions. However, the literature reveals inconsistent results across different research groups in this area. The preparation of catalysts Mo/Al2O3/Fe2O3 using bauxite was evaluated through the hydrocracking of high-temperature coal tar, which was used as the raw material [82]. Hydrocracking involves breaking long C-C chains using a catalyst, hydrogen, and high pressures or temperatures [83]. The higher percentage of hematite was 16.85 wt%, and its catalytic evaluation results demonstrated varying conversion ratios and product yields among different bauxite-derived catalysts. The superior catalytic activity of the best catalyst was attributed to the leaching of Fe2O3 and the enrichment of Al2O3 in the bauxite modified with HCl, leading to increased acidity and specific surface area. This optimal combination enhanced mass transfer and acidity, resulting in compromised performance in terms of high-temperature catalytic treatment (HTCT) conversion and liquid yield. This result indicates that the presence of Fe2O3 is not favorable to the hydrocracking of high-temperature coal tar. In contrast, Kairbekov and coworkers [84] investigated the use of modified iron-containing catalysts and the preliminary ozonization of coal for its hydrogenation. The study focused on catalysts derived from natural bauxite samples with varying Fe2O3 contents. Among the tested materials, bauxite samples exhibited superior performance with 23.7% Fe2O3, achieving a higher yield of liquid products compared to other bauxite samples and catalysts such as red mud and pyrite concentrate. This result highlights the importance of Fe2O3 content in the catalytic hydrogenation process. In catalytic hydrogenation, hydrogen gas is used to reduce unsaturated organic compounds, such as aromatic rings, alkenes, or carbonyl groups, in the presence of a catalyst [85]. Although the specific iron oxide phase was not identified, it is likely that the Fe2O3 was in the hematite phase as bauxite typically contains hematite, goethite, and ilmenite. Furthermore, the study noted that the conversion of iron oxides into finely dispersed pyrrhotine during the hydrogenation process, particularly after modification with elemental sulfur, enhanced the catalytic activity of the iron-containing materials.
According to Widayat et al. [86], Indonesia’s substantial demand for catalysts is primarily fulfilled through imports, underscoring the critical need to identify and utilize domestic natural resources for catalyst production. Additionally, biofuel, particularly biodiesel, is one of the most widely used energy sources in the country, offering advantages such as being renewable and biodegradable, having lower emission levels, and promoting complete combustion in engines. The preparation of hematite catalysts from sand is of interest to address these issues.
In 2019, Widayat and coworkers prepared magnetic hematite nanoparticles from iron sand and tested them as catalysts for biodiesel production from waste cooking oil. The preparation of catalysts included acid leaching with HCl, chemical coprecipitation, and calcination at varied temperatures (650–800 °C). The optimal catalyst was calcinated at 700 °C. The catalysts were able to support both esterification and transesterification; biodiesel yield reached 86.78%, with 87.88% of fatty acid methyl ester; yield optimization was found to depend on the calcination temperature. Furthermore, Widayat’s group explored the potential of bifunctional α-Fe2O3/CaO2 catalysts derived from iron sand for biodiesel production using waste cooking oil [87]. In the study, the catalysts were synthesized by combining hematite obtained through chemical coprecipitation with CaO2 prepared from sources such as Ca(OH)2, CaCO3, and limestone. The catalytic performance varied depending on the CaO2 source, with the limestone-derived catalyst yielding the highest biodiesel output at 77.17%. However, the CaCO3-derived catalyst met both Indonesian and European biodiesel standards. It was also discovered that the α-Fe2O3 component favored the esterification reaction and formation of diglycerides, while CaO2 favored the transesterification reaction and formation of methyl esters.
Catalytic cracking is a widely employed chemical process that breaks down complex hydrocarbons into simpler molecules using a catalyst. This process facilities the cleavage of carbon–carbon bonds in large hydrocarbons at lower temperatures and pressures compared to non-catalytic methods, resulting in enhanced energy efficiency and cost-effectiveness. In heavy oil, the increase in lighter hydrocarbons and reduced resin–asphaltene components was achieved by natural hematite containing 41% Fe, 58% Si, and impurities of Al, K, S, Ca, and Cu using a hydrothermal catalytic system. The density of the oil decreased as resin–asphaltene content was reduced [88], and the results demonstrated the possibility of increasing the number of lighter hydrocarbons and reducing their density by a regular decrease in the amount of resin–asphaltene components.
Hematite can be obtained via the heat treatment of other iron oxides. The catalytic cracking performance of hematite derived from thermally treated natural limonite was evaluated using toluene as a model compound for biomass tar [89]. The limonite was crushed, sieved, and calcinated at 700 °C for 2 h under an air atmosphere. The as-prepared hematite achieved 95% toluene conversion at 700 °C. However, at temperatures exceeding 700 °C, the excessive consumption of lattice oxygen led to carbon deposit accumulation, negatively impacting the conversion rate. The evaluation of the stability of the catalysts indicated that toluene conversion was maintained above 80% at 600 min. CO2 and CO production confirmed the role of hematite as an oxygen carrier. XRD analysis showed the transformation of hematite into magnetite, which occurred due to the consumption of active lattice oxygen and the deposition of carbon; because of this, the spent catalyst was regenerated by heat treatment in an air atmosphere, with the regenerated catalyst maintaining a toluene conversion rate above 85% for the first three cycles.
Natural hematite has also proven to be effective in the catalytic cracking of coal tar [90]. Fe/Al and Fe/Ni composite oxygen carriers (OCs) were prepared through mechanical mixing. The hematite block was first crushed into a fine powder using a jaw crusher and then mixed with γ-Al2O3 or NiO powder. After extrusion molding and granulation, the mixture was dried in an oven at 100 °C for 12 h. Compared to natural hematite, the addition of γ-Al2O3 helps reduce the carbon deposition, while the inclusion of NiO promotes it. The carbon deposits on the reduced OCs mainly consist of hard coke formed through the graphitization of carbon black at high temperatures. In addition, γ-Al2O3 helps improve and preserve the porosity of the OC particles, significantly reducing sintering.
In recent years, catalysts from natural hematite have been tested on catalytic systems, which had not been evaluated before. Iron ore (72% hematite, 22% magnetite, and 6% goethite) was ball-milled as a precursor for the synthesis of a Ni-Fe catalyst for oxygen evolution reaction (OER) [91]. The OER involves the anodic electrooxidation of water in acidic media or OH in alkaline media to generate O2. This reaction is ubiquitous in aqueous environments and typically requires significant overpotentials to proceed, which has driven extensive research into electrocatalysts [92]. Even at a low overpotential (280 mV), this electrocatalyst reaches a current density of 10 mA cm−2 and retains its catalytic performance over prolonged electrolysis under alkaline pH. Synthetic hematite has been applied in various structures for the oxygen evolution reaction (OER), including metal–organic frameworks, carbon-supported materials, heterojunctions, and metal–nonmetal/anion-doped Fe2O3, primarily in alkaline media (KOH). Catalyst stability varies significantly, with no clear correlation to synthesis methods or composition. One of the cases reporting the lowest stability is Fe2O3/g-C3N4, which lasts only 10 min and is synthesized by a thermal method [93]. Stability durations of 5–12 h are observed for C-doped CoFe2O4/Fe2O3 [94], Co3O4/Fe3O4 [95], Fe/Fe2O3-Fe-N-doped C [96], FeS/Fe2O3 [97], Fe2O3/FeS [98], Fe2O3/CNT [99], and Fe2O3/FeP [100], synthesized by calcination, coprecipitation, pyrolysis, chemical etching/solvothermal, hydrothermal, coprecipitation, and hydrothermal methods, respectively. On the other hand, cases with acceptable operational durations (about 100 h) are reported for RuNi-Fe2O3/IF [101], WO3/Fe2O3-NiO [102], and CoMo/Fe2O3 [103], synthesized via hydrothermal, chemical etching/decomposition, and hydrothermal/electrodeposition techniques, respectively. Finally, examples with extended durability include IrO2–Fe2O3, which maintained its performance for 600 voltammetry cycles via thermal decomposition [104], and Fe2O3–MnO, synthesized via a sol–gel process with 1000 cycles [105].
Twelve minerals, which were washed, dried, milled, and sieved, were evaluated using the catalytic hydrolysis of microcystin-LR (MC-LR). Limonite displayed the best catalytic activity, while hematite resulted in poor yields [106]. Also, Fischer–Tropsch synthesis could be catalyzed by iron oxides. Hematite derived from iron ores allowed CO conversion of up to 80% when H2 was the reducing gas. This conversion is higher than those achieved by commercial catalysts. However, it was achieved using relatively higher temperature and pressure (270 °C vs. 250 °C and 20.9 vs. 1.85 bar, respectively) [107]. Table 4 provides a summary of the catalytic applications of hematite, which can be derived from various sources. It is the most extensively utilized natural iron oxide in heterogeneous catalysis.

2.3. Goethite (α-FeOOH)

Goethite (α-FeOOH) is one of the most thermodynamically stable and abundant crystal structures of iron in minerals, sediments, and soils. It results from various processes (hydroxylation, oxidation, hydration, or decomposition) of iron silicates, magnetite, siderite, and pyrite [108]. Goethite adopts a compact hexagonal arrangement of OH and O2 anions, with Fe3+ at the center of the octahedra aligned parallel to the [001] direction [109]. Goethite is antiferromagnetic, with two Neel temperatures of 63 and 97 °C. Above this, it becomes paramagnetic, transitioning to hematite at temperatures above 400 °C [110]. It often exhibits a weak ferromagnetic component caused by a mismatch in the antiferromagnetic grid, associated with a higher number of defects such as hollows, boundaries, and grain rotations [111].
Main catalytic applications. Goethite has proven to be effective as a catalyst for removing organic contaminants from aqueous effluents, making it a promising natural candidate for water treatment applications due to its widespread abundance in the environment. Liu et al. [109] and Ruan et al. [22] conducted reviews of Fenton-like processes based on the literature on this topic. The direct formation of 1O2 and OH has been identified, either independently or in the presence of H2O2 or UV [112], under slightly acidic pH conditions. Examples such as hydroquinone, 2-chlorophenol, oxalate, succinate, bisphenol A, and benzoate have been treated in this manner [113,114,115,116].
As a Fenton reagent, goethite demonstrates remarkable performance. De la Plata and colleagues [117] conducted a comparison between goethite and TiO2 in heterogeneous photo-Fenton catalysis. The authors employed a spectrophotometric system specifically designed for highly precise measurement of the optical properties of catalysts, demonstrating that goethite has a significantly larger optical path length and the ability to effectively degrade 2-chlorophenol. In a more comprehensive study [118], it was demonstrated that the decomposition of various organic contaminants follows a trend: salicylic acid ≈ m-hydroxybenzoic acid > p-hydroxybenzoic acid ≈ benzoic acid > p-bipthalic acid > phenol > benzenesulfonic acid. This trend is analogous to the sorption pattern of these species on goethite, which, in turn, depends on the structure (substituent groups and number of rings).
Yeh et al. [119] conducted a comparison of the degradation of five chlorinated ethylenes and three aromatic hydrocarbons through Fenton-like processes. They observed that aromatics were more readily degradable than chlorinated compounds. However, removal was achieved within a few minutes in all cases. The type of electrolyte significantly influences OH generation by goethite–H2O2. Lin et al. [120], demonstrated that the maximum generation occurs at pH 3, remaining substantial at pH 6–7. The most suitable electrolyte would be ClO4 as it does not impact radical formation, while other tested inorganic anions showed interference tendencies: H2PO4 > SO42− > Cl > NO3. Cl and SO42− more severely suppress OH generation with an increasing concentration. The reason for this appears to be related to the adsorption and blocking of active sites by these anions. In general, based on the results reported in the literature [22], it can be concluded that for goethite, the photo-Fenton process yields better results than simple Fenton processes, and electro-Fenton is more effective than photo-Fenton. During toluene oxidation, Mn/goethite nanoparticles (Figure 6A) selectively produced CO2 and water because no significant formation of an incomplete combustion product (CO) was detected [121]. In order to avoid changes in the catalyst structure, the catalytic performance was evaluated at a maximum temperature of 450 °C.
Natural goethite (Figure 6B) has been successfully applied to produce nanofibers for water disinfection under visible light, demonstrating remarkable efficiency in both dye degradation and bacterial inactivation. Specifically, it achieved 90% degradation of MB within 5 h. Its recyclability was confirmed by retaining a 65% degradation rate after five reaction cycles. The goethite used was sourced from an abandoned mine in the Taouz region of Errachidia Province, Morocco. Nanoparticles were synthesized by high-energy ball milling of goethite rock and wet milling of metallic iron rocks in a planetary ball mill. The process utilized stainless steel vials and iron balls, with a ball-to-powder mass ratio of 40:1, operating at a rotational speed of 450 rpm for 5 h [122].
More recently, highly dispersed natural goethite/Fe3O4 composites (Figure 6C) were synthesized via a one-step hydrothermal method, demonstrating a potential as catalysts for tetracycline removal. Under optimal conditions (pH of 3.0, tetracycline concentration of 100 mg·L−1, catalyst dosage of 0.3 g·L−1, and H2O2 concentration of 6.0 mM), a removal efficiency of 90.1% was achieved within 30 min. The system also exhibited strong durability, retaining a tetracycline elimination rate of 73.3% after five treatment cycles [123].
Another application of goethite is its role as an efficient dissociator of ozone molecules, making it valuable in heterogeneous catalytic ozonation [124], as demonstrated in the case of nitrobenzene by Zhang, Ma and coworkers [125]. In heterogeneous catalytic ozonation, O3 is added to a contaminated effluent, along with a solid catalyst that decomposes the ozone molecule into reactive oxygen species, which, in turn, degrade the pollutant [126]. Toxic inorganic agents such as AsO33− can also be oxidized (to AsO43−) this way, being more effective in the presence of H2O2 and UV. Zhang et al. [127] utilized goethite to adsorb tartaric acid and reduce dissolved Cr6+ in water through this process. They demonstrated that the presence of goethite is crucial; in its absence, tartaric acid alone promotes only a 12% reduction in Cr6+ in 72 h. However, the reduction is 100% within 24 h and can occur in just 4 h when reducing the acid concentration, allowing for the co-adsorption of Cr6+. The authors suggest that goethite’s role is to provide efficient channels for electron transfer between the reactants. Similar results were obtained for the reduction and stabilization of Cr6+ in soil by using calcium polysulfide catalyzed by goethite and hematite, where the presence of these oxides enhanced the performance. The removal of Cr6+ improved with increasing iron oxide content, from 0 to 9 g kg−1, but then declined as the concentration rose from 9 to 12 g kg−1 [128]. Pelalak and colleagues synthesized plasma-treated goethite nanoparticles, which exhibited enhanced surface area and a higher density of surface hydroxyl groups. These nanoparticles were derived from natural goethite [129]. They enhanced the heterogeneous catalytic ozonation of the sulfasalazine (SSZ) antibiotic.
Non-common catalytic applications. In a study on chemical synthesis, Fernández-Marchante [130] investigated the catalytic activity of goethite in the Westinghouse process for hydrogen production (HP), specifically focusing on the transformation of H2SO4 into SO2. The study compared the performance of goethite to that of other established catalysts. According to their results, goethite demonstrated superior behavior compared to SiC but was lower than that of Fe2O3 and similar to that of CuO. The advantage lies in its resistance to sintering, making catalyst deactivation less likely over the long term. Shehzad et al. [131] used goethite as a catalyst for water dissociation in catalytic bipolar membranes (CBMs), reducing the activation energy (calculated by DFT) of the process by 80% compared to state-of-the-art membranes. This, combined with the natural availability of goethite, significantly enhances the process both economically and energetically. They achieved excellent catalyst stability through a protection strategy involving the in situ growth of conductive polyaniline (PANI). Hu et al. [132] employed goethite nanoflowers as support for Rh nanoparticles in the reduction and hydrogenation of nitrostyrene to vinylaniline. A key advantage over alternative methodologies is the ability to achieve 100% conversion without compromising the integrity of the vinyl group. This is a significant improvement as the co-reduction of other unsaturated groups is a common challenge in heterogeneous catalysts, often necessitating lower conversions to prevent unwanted side reactions. As a hydrogen donor, the use of hydrazine hydrate avoids the challenges of handling H2 directly, generating only water and N2 as byproducts in this process. Regarding stability, the authors conducted six cycles of catalyst recovery and reuse, detecting no loss of activity. Composites of Mn, Co, Ni, Cu, and Zn oxides were employed by Kuncser et al. [133] in the oxidation of cyclooctene. Catalytic oxidation occurs in the presence of O2, using isobutyraldehyde as a reducing agent and CH3CN as a solvent. In all instances, selectivity towards the formation of the corresponding epoxide exceeded 99% in the composites. However, the conversion varied in the order of Mn > Fe > Co > Ni > Zn > Cu, ranging from 55% to 4%. The concentration of basic sites in the composites was determined by measuring the UV–Vis signal of acrylic acid, irreversibly adsorbed by the materials. The results show a complete correlation between conversion and the ratio of the basic sites to the acidic sites of the composites. In microbial fuel cells (MFCs), goethite has demonstrated its ability to enhance catalysis [134]; the mineral was extracted from mining mud, and its performance was compared to stainless steel alone. Raw goethite tripled the electrical power obtained with stainless steel (11 W m−3 compared to 3.5 W m−3). Furthermore, when goethite was treated at 550 °C, the power increased to 17.1 W m−3. Additionally, it substantially improved the current and current efficiency obtained while reducing COD. Goethite exhibited superior performance compared to specularite and hematite as a catalytic bed for the catalytic reforming of lignite-blended co-pyrolysis with corn straw [135]. The process significantly enhanced the formation of light aromatic hydrocarbons. Verma et al. [136] employed goethite for the synthesis of α-aminonitriles from N,N-dimethylaniline using atmospheric oxygen. The authors experimented with various iron salts and minerals, including Fe(OAc)2, FeCl2, FeSO4, FeCl3, Fe(NO3)3, Fe2(SO4)3, Fe(acac)2, Fe(acac)3, FeO@GO, Fe3O4, FeO, Fe@g-C3N4, and goethite. However, none yielded a performance greater than 17%. Remarkably, when goethite assumed a lychee-like morphology, the yield increased to 96%, indicating a high dependence on maintaining this specific morphology.
Figure 6. SEM micrographs of α-FeOOH particles prepared from natural sources: (A) acid mine drainage; (B) goethite rocks; (C) mineral goethite/Fe3O4; and (D) mineral goethite. Reproduced with permission from References (A) [121], (B) [122], (C) [123], and (D) [137]. Copyright © 2012, 2023 American Chemical Society; 2020 MDPI/CC BY 4.0; and 2023 Elsevier, respectively.
Figure 6. SEM micrographs of α-FeOOH particles prepared from natural sources: (A) acid mine drainage; (B) goethite rocks; (C) mineral goethite/Fe3O4; and (D) mineral goethite. Reproduced with permission from References (A) [121], (B) [122], (C) [123], and (D) [137]. Copyright © 2012, 2023 American Chemical Society; 2020 MDPI/CC BY 4.0; and 2023 Elsevier, respectively.
Catalysts 15 00236 g006
A recent study [137] examined the sunlight-driven activity and wavelength dependency of goethite obtained from Bingzhou, Hunan Province (Figure 6D). Before conducting photochemical experiments, the mineral was crushed, milled, and sieved to attain a particle size of less than 48 μm (<300 mesh). Goethite displayed strong light absorption, extending to approximately 570 nm in the visible spectrum, suggesting a narrow band gap favorable for photocarrier generation. Regarding hydroxyl radical (OH) production, goethite achieved a rate of 0.648 nM min−1, surpassing both hematite and magnetite. Its performance in H2O2 production was still lower than that of hematite.
A comparative study on OH photochemical production between natural and synthetic iron minerals revealed much higher rates in natural minerals. Specifically, natural goethite, hematite, and magnetite showed 1.7-, 4.1-, and 4.9-fold increases in OH production, respectively, compared to their synthetic counterparts. Interestingly, the relatively smaller improvement in natural goethite’s performance corresponds to its lower impurity levels compared to the other minerals.
Goethite exhibits optimal catalytic performance under slightly acidic to neutral pH conditions, where the generation of hydroxyl radicals (OH) and reactive oxygen species is enhanced.
Table 5 presents the catalytic applications of goethite prepared from natural sources, with advanced oxidation processes being the most prominent. However, as previously discussed, synthetic goethite has been successfully used in other types of reactions, such as reduction and selective hydrogenation. Therefore, it would be worthwhile to test the efficiency of natural goethite in the same reactions. Nonetheless, its transformation into hematite above 400 °C should be considered.
Although the most abundant oxides are widely used in catalytic reactions, their preparation from natural sources is less common. In applications where high phase purity is required, their use can be detrimental to the reaction due to impurities commonly found in the sources from which they are obtained. However, these impurities can enhance catalytic activity in other processes. Additionally, in highly reducing or oxidizing environments, a phase change may occur during the process, potentially leading to its deactivation. While its magnetic properties are widely known and catalysts are magnetically characterized in some research reports, such characterization is carried out with a focus on the recovery and reuse of the catalysts. In spite of this, the effect of magnetic properties on catalytic activity has rarely been studied.

3. Less Abundant but Widely Distributed Fe-OH

While less abundant than magnetite and hematite, iron oxide such as maghemite (γ-Fe2O3), lepidocrocite (γ-FeOOH), and ferrihydrite (Fe8.2O8.5(OH)7.4 + 3H2O) also exhibit notable catalytic properties. This section will explore some of their applications in catalysis.

3.1. Maghemite (γ-Fe2O3)

Maghemite is widely recognized as a common component of various soils, particularly in highly weathered toxic soils found in tropical and subtropical regions. Recent environmental magnetic studies have shown that maghemite is distributed globally, having been detected in tropical, subtropical, arid, and tundra regions. The formation of ultra-fine maghemite during pedogenesis is considered the primary factor behind magnetic enhancement in paleosols. Studies have detailed the pedogenic processes responsible for maghemite formation, revealing that pedogenic maghemite, often associated with hematite, acts as an intermediate phase in the transformation of ferrihydrite into hematite. Experimental findings demonstrate that when phosphated ferrihydrite is subjected to a temperature of 150 °C for 120 days, it forms hydromaghemite (with grain sizes between 10 and 30 nm). Over time, the grain size of hydromaghemite grows, transitioning from the superparamagnetic to the single-domain range. After four months, hydromaghemite further transforms into the more thermodynamically stable hematite [138]; at nanosizes (36–40 nm), it transforms into hematite at 576 °C [139].
Maghemite (γ-Fe2O3) possesses an inverse spinel cubic structure with a lattice parameter a = 8.3500 Å. Its nomenclature derives from the amalgamation of magnetite (Fe3O4) and hematite (α-Fe2O3), owing to substantial congruence in both structural attributes and chemical composition. Operating as a ferrimagnetic substance at ambient temperature, it exhibits saturation magnetization values as impressive as 90 emu g−1. However, its stability is inferior to that of hematite. Characterized as an n-type semiconductor, it displays a band gap of 2.0 eV. Similarities with magnetite (Fe3O4) are clearly observable. The distinctive demarcation lies in the exclusive presence of the Fe3+ oxidation state within maghemite, while magnetite encompasses both Fe2+ and Fe3+. The distinction between maghemite and magnetite is often complex when using XRD and X-ray photoelectron spectroscopy (XPS). However, some methods have been proposed to differentiate them from one another [140,141]. Determining the Curie temperature of maghemite has been proven to be challenging due to its tendency to transform into hematite, even at temperatures such as 413 K. This transformational behavior is contingent upon precursor selection, synthesis methodology, and external pressure exerted [142]. In addition, its Curie temperature is estimated to be within the range of 820 to 986 K [143,144]; nanoparticles present a considerably lower value of about 545 K [145]. Notwithstanding, since this analysis was conducted via the thermogravimetric analysis (TGA) technique and the observed transition temperature closely corresponds to the maghemite-to-hematite phase transition, assigning it as the Curie temperature remains a significant challenge.
Concerning its band structure, ab initio calculations elucidate that the upper segment of the valence band is dominated by oxygen 2p orbitals, with occupied 3d levels of Fe situated 6 to 7 eV below the Fermi level. The lower tier of the conduction band is populated by unoccupied 3d levels of Fe (octahedrally coordinated). This classification designates maghemite as a charge transfer semiconductor, with the foremost excitation potentially corresponding to electron transfer from O2− anions to octahedral Fe3+ cations [146].
Synthetic maghemite has been successfully tested in hydrogenation reactions, with remarkable yields, especially when used in conjunction with other oxides [85,147,148]. However, to the best of our knowledge, maghemite prepared from natural sources has been poorly explored as a catalyst. Vasanthakumar and Karvembu [149] reported the catalytic evaluation of maghemite, obtained from river sand (Figure 7A), based on the base-free transfer hydrogenation of furfural, levulinic acid, and o-vanillin. River sand was collected from the Cauvery River in Kallanai, Tamil Nadu, India, and γ-Fe2O3 was separated using simple magnetic separation with a bar magnet. The γ-Fe2O3 powder was then washed with acetone, ethanol, and water to remove impurities and was dried at 80 °C for 3 h. To compare the catalytic activity of natural γ-Fe2O3, synthetic γ-Fe2O3 was also prepared following existing literature protocols. Catalytic tests were performed under a vacuum at 80 °C for 3 h. The activated catalyst was added to a sealed tube containing 1 mmol of substrate in 2-propanol and placed in a preheated oil bath at 150 °C for a specific duration. After 30–90 min, depending on the substrate, the catalyst was separated with an Nd magnet, and the solution was centrifuged. Magnetic characterization showed that the γ-Fe2O3 from river sand had higher saturation and remanent magnetization than synthetic γ-Fe2O3, exhibiting stronger magnetic behavior. Catalytic tests revealed that river sand γ-Fe2O3 displayed superior catalytic activity, which remained effective even after 10 cycles. The base-free catalytic transformation of levulinic acid to γ-valerolactone, via the 4-hydroxypentanoic acid intermediate, was attributed to the presence of silica in the γ-Fe2O3 from river sand.
Iron ore raw materials obtained through the catalytic reduction of NO with NH3 resulted in excellent selective catalytic reduction (SCR) activity (above 80% at 170–350 °C) and N2 selectivity (above 90% up to 250 °C) at low temperatures. However, the addition of H2O and SO2 in the feed gas showed some adverse effects on SCR activity.
For the preparation of catalysts, the iron ore raw materials were dried at 105 °C for 4 h and ground. Then, these ground samples were dried again at 105 °C for 5 h. Finally, they were calcinated at 250, 350, and 450 °C, respectively, for 6 h [152]. The characterization results indicated that the best catalytic performance was achieved by a catalyst comprised mainly of α-Fe2O3 and γ-Fe2O3. The utilization of natural maghemite as a catalyst presents challenges due to its inherent impurities. Natural maghemite typically contains substitutional metal cations such as Ti4+, Mg2+, or Al+3, among others, which can significantly influence its catalytic properties. Since γ-Fe2O3 is a metastable phase, the methods used to prepare catalysts based on this compound must be adequate to avoid the formation of the most thermodynamically stable hematite (α-Fe2O3) [153]. The summary results are presented in Table 6.

3.2. Lepidocrocite (γ-FeOOH)

Lepidocrocite is named after its platy crystal shape and orange coloration [2]. It is commonly found in hydromorphic soils with alternating reducing and oxidizing conditions and can biogenically be obtained [154]. It is metastable, serving as a precursor to more magnetic phases like maghemite and magnetite [155], and plays a key role in the geochemical behavior of nutrients, heavy metals, and organic pollutants. It presents an orthorhombic crystal structure with lattice parameters a = 3.072(2), b = 12.516(3), c = 3.873(2) Å, and space group Cmcm [156]. In the lepidocrocite structure, ferric iron is coordinated in edge-shared octahedra, which form layers within [001] linked by hydrogen bridges. In contrast to other oxyhydroxides, lepidocrocite exhibits a low Neel temperature, which ranges between 50 and 70 K [157].
Up to the time of writing this article and to the best of the authors’ knowledge, lepidocrocite has only been used in its synthetic form in Fenton [158] and Fenton-like processes [159,160,161], photodegradation [162,163], water splitting [164], CO oxidation [154], arsenic adsorption [165], and the combination of H2O2 and PMS to produce ROS and remove chloramphenicol (CAP) [166]. However, no reports were found in the literature regarding its use in mineral form, likely because it is rare to find it in ore deposits [167]. The unique way to form lepidocrocite is the oxidation of Fe(OH)2 by air in suspension at pH ≈ 7 [2]. When it is prepared in the presence of citric acid, the crystallization of lepidocrocite can be controlled by its concentration, which decreases with its increasing concentration [162]. Related to photocatalytic activity, visible light and the presence of ethylenediaminetetraacetic acid (EDTA) could accelerate the reaction rate of lepidocrocite [163].
Experimental laboratory results indicate that the morphology of catalysts influences the exposed facets and interaction with OH groups of iron oxides, having a direct effect on their reactivity. For example, the adsorption and degradation of Orange G (OG) on lath-like particles were mediated by the interaction with the µ-OH groups of the (010) facet, while in rod morphology, it was adsorbed laterally by the interactions with the µ-OH and µ3-OH groups of the (010) and (001) facets [161].
When lepidocrocite VOOH hollow nanospheres were prepared by the hydrothermal method, a variation in the reaction temperature allowed the tuning of the surface area, modifying the catalytic activity. Also, it exhibited stability in alkaline media and achieved good performance in water splitting at a potential of 270 mV for OER [164].
The findings from the evaluation of the adsorption of arsenic suggest that pH affects it remarkably and the co-occurrence of aqueous Fe2+ and Fe3+ solids is beneficial for arsenic removal because Fe2+ can induce As3+ oxidation to As5+. Above pH = 8, As5+ can be more mobile than As3+; at low pH, the radical scavengers can compete with As3+ for the oxidants [165].
Theoretical results comparing goethite and lepidocrocite in the elimination of 5-bromosalicylic acid (BSA) via the Fenton process show that electrons are more easily transferred from the H2O2 molecule to the Fe atoms of α-FeOOH (goethite), enabling it to perform better in catalyzing the decomposition of H2O2. However, free radicals are more likely to desorb from γ-FeOOH (lepidocrocite), making the γ-FeOOH/H2O2 system more efficient in degrading BSA [168]. Lepidocrocite’s catalytic applications are versatile, but its metastability, crystal morphology, and interaction with functional groups affect its reactivity and performance.

3.3. Ferrihydrite (Fe8.2O8.5(OH)7.4 + 3H2O)

Ferrihydrite is a hydrated Fe(III) nano-oxide commonly found in various surface environments on Earth. Its presence has a significant impact on soil properties. Ferrihydrite was officially recognized as a mineral by the International Mineralogical Association (IMA) in 1975. In nature, it is commonly associated with silicates (with an atomic ratio of Si/Fe ranging from 0.17 to 0.37), and these silicates seem to directly influence its low crystallinity. Due to its small particle size and poor crystallinity, ferrihydrite exhibits crystal and compositional properties that are challenging to characterize [169]. The complexity of ferrihydrite’s structural analysis is also influenced by the presence of various other species in its composition, including 0.5–31.5 wt% SiO2, 0.5–9.6 wt% Al2O3, and up to 7.2 wt% PO4, along with As(V), which ranges from 0.7 to 28.3 wt% [170]. Initially, its formula was identified as 5Fe2O3 + 9H2O, featuring a hexagonal unit cell with parameters of a = 0.508 nm and c = 0.94 nm and containing FeO6 octahedra [171]. Eggleton and Fitzpatrick [172] proposed a trigonal cell structure for ferrihydrite, with two-thirds of the iron in octahedral coordination and one-third in tetrahedral coordination. However, other studies have questioned the presence of tetrahedral Fe [173]. Additional formulas such as Fe5HO8 + 4H2O [174] and Fe2O3 + 2FeOOH + 6H2O [175] have also been suggested. The number of X-ray diffraction peaks varies from two to five in natural ferrihydrites, while synthetic forms can show up to seven well-defined peaks, reflecting the degree of Fe atom ordering in the lattice.
Currently, one of the most accepted crystalline structures for two-line ferrihydrite is Fe8.2O8.5(OH)7.4 + 3H2O [176], which consists of 20% tetrahedral iron and approximately 80% octahedral iron, featuring a local δ-Keggin structure. Additionally, the structural depletion (SD) model proposed by Hiemstra in 2013 suggests a distinction between the nanoparticle’s core and surface. While the core is largely free of defects, the surface is depleted of tetrahedral structures. The ideal structure of ferrihydrite consists of three types of Fe: two hexacoordinated and one tetracoordinated, arranged in a δ-Keggin framework with proportions of 60%, 20%, and 20%, respectively. The general structure includes three layers of polyhedra, with the first hexacoordinated Fe located in the outer layers, the second hexacoordinated Fe in the middle (surrounding the tetracoordinated Fe), and the third octahedral Fe, which is highly distorted, located in the inner layer [177].
Ferrihydrite, owing to its widespread presence, common occurrence as a coating on particles, and reactive surface with a small particle size (1–7 nm), is widely recognized for its role in influencing the mobility of both inorganic and organic pollutants via sorption processes. It typically forms under near-neutral pH conditions in a variety of redox-active environments as well as in large sedimentary deposits where there is a significant supply of ferrous iron from groundwater and high oxidation rates [170]. Its chemical composition and crystal structure have been debated for over five decades, largely due to its extremely small particle size and the high degree of structural defects it incorporates [178]. Ab initio calculation results suggest that its reaction sites are mainly disposed along rows at the edges of sheets of iron octahedra. Molecular dynamics studies on nanoparticles of up to 10 nm show that highly reactive hydroxo groups on the surface are typically unbound, but they participate as hydrogen bond acceptors in a network with less reactive groups [179]. Magnetic properties reveal that ferrihydrite exhibits antiferromagnetism with a ferromagnetic-like moment at lower temperatures (100 K and 10 K), but it is paramagnetic at room temperature. As the crystallite size of ferrihydrite increases, both magnetization and coercivity decrease, a trend attributed to the pronounced surface effects found in fine-grained materials [180].
The main catalytic applications of ferrihydrite are related to the Fenton reaction [181], primarily conducted with synthetic ferrihydrite. In their review, Zhu et al. [181] highlighted that ferrihydrite consistently exhibits higher reactivity with H2O2 compared to other iron minerals. Specifically, they noted performance hierarchies, such as ferrihydrite > hematite > goethite, ferrihydrite > feroxyhyte ≈ magnetite > maghemite > goethite > hematite, and ferrihydrite > hematite > goethite. These comparisons underscore ferrihydrite’s superior catalytic potential. The use of ferrihydrite in the Fischer–Tropsch (FT) reaction has been studied by Kyu Seomoon [182]. By incorporating Al, Cu, and K into the synthesis of the catalyst and analyzing the FT reaction using gas chromatography–mass spectrometry (GC-MS) over 100 h, Seomoon observed a 25% increase in paraffin selectivity, attributed to reduced catalyst basicity. Increasing the H2/CO ratio from 1 to 3 reduced CO2 emissions by 40% and enhanced hydrocarbon formation. Similarly, Sumit Bali et al. [183] doped ferrihydrite with Al and Cu, achieving slightly better FT performance than a commercial fixed-bed catalyst, securing a patent for their synthesis method. However, there are some reports on the use of natural minerals as catalysts. The decomposition of hydrogen peroxide was tested using ferrihydrite obtained directly from acid mine drainage in the Zlaté Hory underground mine district (Czech Republic). In this case, the only further treatment was vacuum drying of the collected solid. The dried sample was gently powdered to minimize mechanical disruption (Figure 7B) [150]. Physicochemical characterization revealed a composition of mainly iron oxyhydroxides (approximately 96 wt%), with ferrihydrite as the principal phase, along with traces of goethite and possibly schwertmannite. The constant rate (6.59 min−1) was slightly higher than that of commercial goethite, with the advantage of using extremely low-cost precursors or even industrial waste.
Ferrihydrite has been extensively used to adsorb heavy metals and other ions. The adsorption mechanisms for metals such as lead [184], copper [185], molybdenum [186], chromium [187,188], and zinc [189] occur via inner-sphere complexation, typically forming bidentate or bidentate–monodentate complexes depending on the pH and concentration. For radium and barium, tetradentate bonds with a single surface functional group are feasible [190]. Actinides like thorium [191], plutonium [192], and uranium [193] also prefer bidentate adsorption, with the latter two being extensively studied for removal from industrial systems through ferrihydrite nanoparticle precipitation.
Sorption capacity is strongly influenced by temperature and pH. Anion adsorption can significantly influence cation sorption by competing for available binding sites on the surface, thereby reducing the capacity for cation uptake. Carbonates alter ferrihydrite solubility [194], while phosphate addition enhances Cd2+ sorption by shifting from bidentate to ternary complex formation [195]. Zhu et al. [196] observed that ligands’ ability to inhibit arsenic (As3+ and As5+) sorption follows the sequence SeO4 ≈ SO4 < oxalic acid ≈ malic acid < citric acid < SeO3 << PO4. Additionally, the content of other metals in ferrihydrite’s lattice, such as aluminum, affects the sorption of chromates, selenates, and sulfates, as demonstrated by Johnston and Chrysochoou [187].
Maghemite, lepidocrocite, and ferrihydrite are considered less abundant oxides. Maghemite can only be distinguished from magnetite through characterizations that identify the oxidation states of iron. However, in the absence of such results, its catalytic activity may be mistakenly attributed to magnetite. Additionally, it is common to find traces of maghemite during hematite synthesis, making the evaluation of its performance as a catalyst a challenge.
Lepidocrocite, on the other hand, is rarely found in ore deposits as it is a metastable phase that serves as a precursor to other magnetic phases. Nevertheless, its synthesis can be achieved under controlled laboratory conditions by adjusting factors such as pH and the presence of additives. The morphology of the obtained particles is highly related to their catalytic activity.
As for ferrihydrite, it is found in nanometric sizes and, compared to other oxides and hydroxides, its recognition as a mineral is relatively recent. Its nanometric size and lack of crystallinity make it difficult to even determine its chemical formula. However, its capacity to adsorb heavy metals has been well established.

4. Rare Iron Oxides and Hydroxides

This section explores the catalytic applications of several rare iron oxides, including wüstite (FeO), akaganéite (β-FeOOH, feroxyhyte (δ′-FeOOH), and bernalite (Fe(OH)3). The unique properties of these iron oxides and their potential catalytic applications will be discussed.

4.1. Wüstite (FeO)

Under ambient conditions, wüstite (FeO) crystallizes in a highly defective form of the ideal NaCl-type structure, with a nonstoichiometric formula of Fe1−xO [197]. It is metastable and tends to decay into a two-phase mixture of α-Fe and magnetite Fe3O4. In addition, cubic Fe1−xO can be synthesized at ambient pressure for iron deficiencies ranging from 0.05 < x < 0.15 at 560 °C [198,199]. The number of vacant octahedral sites is twice the amount of iron deficiency (x), according to the neutron diffraction data, implying a fraction approximately equal to x of iron located at interstitial tetrahedral sites. Further analysis of several samples reveals a direct linear relationship between the stoichiometry of Fe1-xO and its lattice parameter, given by the equation a (Å) = 3.856 + 0.478(1 − x).
Wüstite exhibits antiferromagnetic ordering as temperatures descend below approximately 195 K. The nature of the magnetic order–disorder transition undergoes modulation in response to variations in the composition of Fe1−xO, displaying heightened cooperativity in the presence of nearly stoichiometric FeO. The Fe2+ ion spins align antiferromagnetically along the (111) plane, with the predominant magnetic vector aligned parallel to the [111] direction. However, this axial component (3.8 µB for Fe0.99O) is accompanied by a less perpendicular component (1.3 µB for Fe0.99O). The magnitude of the ordered magnetic moment experiences a marked reduction in samples with decreased iron content. The paramagnetic phase assumes the structural attributes of a NaCl-type arrangement (space group F m 3 ¯ m ), while the lower temperature phase exhibits a rhombohedral distortion (space group R3) [200].
Liu and colleagues [201] have conducted a review of the applications of wüstite in ammonia synthesis. The most successful ammonia synthesis method is the Haber–Bosch process, in which atmospheric nitrogen is converted into ammonia in the presence of a catalyst, high pressures, and high temperatures [202]. According to their findings, wüstite demonstrates ideal performance in terms of stability, activity, and cost. Its two industrial-level competitors are catalysts based on magnetite and ruthenium. The Ru/C catalyst can be up to 20 times more active than the magnetite-based catalyst (Haber catalyst), but in comparison to wüstite, it lacks competitiveness. Pernicone et al. [203] compared the performance of wüstite A301 to that of the best Ru/C catalyst [204], obtaining equivalent results. This positions wüstite as the most attractive option due to its abundance and cost-effectiveness. Its main challenges still lie in its production as wüstite does not exist naturally. In the laboratory, it is obtained by the decomposition of ferrous oxalate, resulting in a mixture with Fe3O4 and Fe that lacks the high-purity properties of wüstite. Industrially, it is obtained by reducing natural magnetite, using Fe as a reducing agent, directly in the smelting furnace at T > 834.5 K, incurring a significant energy cost.
Wüstite, with a bandgap of 2.1 eV (Figure 7C), has been utilized as a photocatalyst for removing fluoroquinolones from water [151]. When combined with H2O2, it functions as a photo-Fenton heterogeneous catalytic system, facilitating the generation of reactive oxygen species (ROS), including O2•−, 1O2, and OH. The complete elimination of microbial activity is achieved within 3 h, accompanied by the mineralization of 21% of enrofloxacin. The catalyst’s performance proves reproducible for at least four treatment cycles following the recovery and drying of wüstite in an oven at 105 °C. The reduction of nitrates (NO3) is another application in which wüstite has been implicated [205]. Nearly stoichiometric conversion from NO3 to NH4+ is attained, with NO2 identified as an intermediate. In this case, it is not a catalysis but rather a heterogeneous redox reaction as FeO serves as a reactant, forming Fe3O4 as a product. Advanced oxidation processes represent a field where iron oxides and hydroxides have been extensively applied [206]. In a study by Expósito and colleagues [207], the performance of wüstite was compared to that of other iron oxides in the electro-Fenton removal of aniline. Wüstite and magnetite, along with hematite, goethite, and FeSO4, exhibited approximately 80% total organic carbon (TOC) removal at a rate similar to FeSO4 (homogeneous phase). Furthermore, the decay pattern and final concentrations of H2O2 and Fe2+ also mirrored the homogeneous phase process. This suggests that the mechanism of the degradation of wüstite aligns with the homogeneous phase, initiated by dissolved iron ions. The advantage lies in the minerals’ ability to self-regulate the supply of iron ions to the solution based on the process’s requirements. Also, FeO catalysis based on the catalytic degradation of chloramphenicol (CAP) by water falling film dielectric barrier discharge (WFFDBD) enhanced the degradation from 69.5% to 97.7% in 8 min by just using WFFDBD. The reaction conditions are shown in Table 6. DFT studies of the catalytic mechanism showed that the key step in the catalytic reactions was the electron transfer from FeO to the antibonding orbitals of dissolved oxygen, H2O2, and O3, which facilitated their transformation and decomposition. This process converted dissolved oxygen into O2•−, which was then adsorbed onto FeO through electron transfer, crucial to the catalytic reaction. Additionally, oxygen vacancies increased the charge density of Fe, enhancing electron transfer [208].
The metastable nature of wüstite, its tendency to decompose into magnetite and hematite, and the difficulty in obtaining high-purity forms are the main limitations to its widespread use. Here, wüstite is effective for ammonia synthesis. However, its effectiveness in AOPs is dependent on acidic conditions (pH~3).

4.2. Akaganéite (β-FeOOH)

Akaganeite (β-FeOOH) derives its name from the Akagané mine in Japan, where its initial discovery occurred in 1962 [209]. Widely acknowledged as a corrosion byproduct of steel [210], it stands as the predominant iron oxide in both soils and geothermal brines [211]. Showing a hollandite-like structure, its empirical formula is Fe7.6Ni0.4O6.35(OH)9.65Cl1.25, discerned from the corrosion crust formed on the Camp del Cielo meteorite. Exhibiting monoclinic symmetry (space group I2/m), its lattice parameters are characterized by a = 10.600(2) Å, b = 3.0339(5) Å, c = 10.513(2) Å, and β = 90.24(2)°. Within akaganeite, Fe3+ ions occupy octahedral sites, forming a double-chain structure that shares corners, thereby allowing ample space for the inclusion of large tunnels featuring square cross-sections, suitable for hosting Cl and H atoms [212]. This compound is recognized as an antiferromagnetic material, manifesting a Neel temperature of 259 K, coupled with ferromagnetic-type hysteresis [213].
Akaganeite is particularly characterized by its ability to remove contaminants through adsorption. According to Zhao et al.’s review [214], the successful adsorption of As3+, As5+, Cr6+, Zn2+, U6+, Cd2+, Cs+, and PO43− has been explored previously (for further details on the removal of these contaminants, please refer to that work and the references cited therein). Briefly, the morphology (size and shape) of akaganeite largely determines its properties. Nanocrystals with various shapes, such as rods, bundles, needles, fibrils, and cigars, have been obtained. These depend on the preparation method, usually involving hydrolysis and the use of different surfactant templates such as polymers, zeolites, activated carbon, and clays. Using ATR-FTIR and liquid chromatography, Marsac et al. [215] demonstrated that in the case of oxolinic acid removal, the removal occurs via the adsorption of the contaminant and not by oxidation on akaganeite. This adsorption occurred via metal complexes under slightly acidic pH and via hydrogen-bonded complexes under alkaline pH. They even determined that the most reactive crystal plane is (010).
Additionally, degradation applications have been found. In the laboratory, starting from FeCl3, in the presence of glucose or the surfactant Brij30 at 60 °C for 3 days, Xiong et al. [216] determined that it can remove 100% of methyl orange in the presence of H2O2 and when pH = 4.5 using photo-Fenton reactions. This capacity is reduced to 86% after four working cycles. Polyaniline degrades acidic orange II even better than akaganeite alone [217] in Fenton processes. The β-FeOOH@GO catalyst, evaluated in a photo-Fenton-like system using UV irradiation, achieved 99.7% decolorization of MB after 60 min. The pseudo-first-order rate constant was 0.632 min−1. The proposed degradation pathway of MB predominantly proceeded with the rupture of phenothiazine ring oxides with OH, HO2•−, and singlet oxygen (1O2) radicals [218].

4.3. Feroxyhyte (δ′-FeOOH)

It is a relatively uncommon iron oxide mineral and one of the few phases in the Fe2O3-H2O system [219]. The feroxyhyte structure is composed of O2− and OH ions arranged in the closest packing, with the Fe3+ ions statistically distributed in half of the octahedral positions. It has a hexagonal cell, with parameters a = 2.93 Å and c = 4.6 Å. It is completely converted to goethite in 6 h at a temperature of 60 °C and is partially transformed into hematite at 80 °C. Feroxyhyte is ferrimagnetic with a low Curie temperature (155 °C). The widely varying degree of the order leads to the formation of both magnetic and nonmagnetic varieties. The specific magnetic susceptibility reached 5000 × 10−6 cm3·g−1 and only 400 × 10−6 cm3·g−1 in the studied coarse crystalline and fine crystalline samples of synthetic feroxyhyte, respectively. Magnetic susceptibility was even lower in the finer crystals. In natural occurrences, it is always fine-grained, although samples with larger particle sizes and better crystallinity can be prepared in the laboratory [220].
Feroxyhite has demonstrated strong catalytic performance in various applications, although reported uses primarily involve laboratory-synthesized materials, often as nanoparticles. Among FeOOH polymorphs, feroxyhite stands out for its sufficient magnetic properties, enabling easy recovery in aqueous media. It can be synthesized through a single-step reaction—a critical factor for scalability in laboratory chemical processes [80]—and features a significantly higher proportion of surface OH groups compared to common magnetic iron oxides [221]. Recent studies highlight its use in photo-Fenton reactions, water splitting, and metal adsorption [222,223,224].
Weiping Du [225] studied the photodegradation of orange II using synthetic hematite, maghemite, magnetite, goethite, lepidocrocite, and feroxyhite. According to their data, hydroxides outperformed iron oxides, contrary to the authors’ expectations. Among hydroxides, feroxyhite showed the lowest performance. The mechanism involved dye adsorption, significantly enhanced by H2O2, AgNO3, and NaF through different pathways. Pinto [226] utilized synthetic feroxyhite to decompose H2O2 into radicals via the Haber–Weiss mechanism, effectively degrading MB and IC. They demonstrated that the adsorption of anionic dyes is more effective than that of cationic dyes. Here, the degradation reaction occurred almost entirely on the surface, primarily through the generation of OH radicals. Li et al. [227] developed δ-FeOOH/BiOBrxI1-x for the visible-light photodegradation of tetracycline, demonstrating improved charge transfer, synergy with H2O2, and enhanced recyclability. Radical species O2•− and OH were identified as key players.
Chen Wang [228], in a review of FeOOH applications, highlighted that δ-FeOOH exhibits the smallest band gap (1.94 eV) and the largest specific surface area among FeOOH crystalline phases, making its band edge positions ideal for oxidizing H2O to OH radicals. Pereira [229] first applied synthetic feroxyhite for water splitting and H2 production, citing its 2.2 eV band gap as being sufficient for water dissociation while absorbing visible light. The photocurrent generated was tenfold that of P25 (TiO2). Da Silva Rocha et al. [230] demonstrated that Ni2+ doping in synthetic δ-FeOOH enhanced photocatalytic activity, conductivity, and charge transfer, while surface Ni(OH)2 improved charge separation.
When modified with other materials, one of the most common applications of δ-FeOOH is the removal of heavy metals from water. Jing Hu [231] used δ-FeOOH-coated γ-Fe2O3 for Cr6+ adsorption via an outer-sphere complexation mechanism. In a Fe/Mn oxy-hydroxide (δ-Fe0.76Mn0.24OOH), Tresintsi et al. [232] found it to be highly efficient for As3+ and As5+ adsorption. Pinakidou [233] employed amorphous tetravalent manganese feroxyhyte nanoparticles (TMFx) to remove As5+. Later, Kokkinos et al. [234] reported TMFx as being effective in eliminating Cd, Hg, and Ni.
The versatility of these materials is also evident in their ability to immobilize phosphates with the aid of Fe2+, achieving up to 94% removal under acidic pH [221]. Additionally, they have biomedical applications, such as controlled heat release under an AC magnetic field [235]. Recently, Lacerda and coworkers proposed, using ab initio calculations, that feroxyhyte doped with niobium can function as a bifunctional catalyst. Their results suggest that doping δ-FeOOH layers leads to stronger van der Waals interactions, enhancing the catalyst’s thermal stability. Additionally, the incorporation of Nb enhances Brönsted acid characteristics and introduces Lewis acidic sites on the catalyst’s surface [236].

4.4. Bernalite (Fe(OH)3)

Bernalite, Fe(OH)3, is one of the more recently identified iron oxyhydroxide minerals, recognized in 1992. It is the scarcest iron mineral and has been documented in only a handful of locations [4], including New South Wales, Australia, and the central Black Forest, Germany [237]. McCammon et al. deduced [238] its empirical formula to be [(H2O)0.04(CO2)0.03Pb0.01]Fe3+0.93Si0.06Zn0.01)(OH)2.95O0.04. This mineral exhibits a protonated octahedral framework (POF), adopting a distorted perovskite-type structure (ReO3) within the Pmmn space group [239]. Its tilt system of a + b + c arises from the topology of hydrogen bonds, resulting in alternating layers of Fe(1) and Fe(2) octahedra with an inverted tilt along the [001] direction, n-glides passing through the Fe atoms, and notable anisotropy in hydrogen bond topology [239]. With a slight deviation from the collinear antiferromagnet, bernalite manifests a weak ferromagnetic moment [240]. To date, there exists scant specific information regarding its application as a catalyst, neither within Scopus nor Google Scholar. Scopus records a mere seven references with “bernalite” in the title, with only two in the last two decades, the remainder hailing from the 1990s. Of these, four discuss structural or magnetic properties.
Recently, a potential use for this mineral was serendipitously uncovered. Under conditions of pH = 10, in the presence of arsenates, goethite undergoes partial transformation into bernalite. Remarkably, bernalite exhibits 2.18 times greater arsenate sorption than goethite, achieved through the formation of monodentate mononuclear (MM) complexes. The goethite used in these studies was sourced from rod-shaped nano-goethite (US3162) crystals obtained from US Research Nanomaterials (USA) [241] or Sigma-Aldrich (St. Louis, MO, USA) [242]. In another possible application, a composite comprising bernalite, Fe2O3, and Ru was synthesized through microwave-assisted heating in deionized water. This resultant material demonstrates the potential for oxygen reduction and hydrogen oxidation reactions. In spite of that, discerning the catalytic role of each component remains challenging [243].
Wüstite, akaganéite, feroxyhyte, and bernalite are among the least abundant iron oxyhydroxides, limiting their use in large-scale applications due to the rarity of natural sources. Their catalytic applications are less explored compared to more common iron oxides. Nevertheless, these materials exhibit properties suitable for advanced oxidation processes (AOPs), particularly in acidic conditions and systems like Fenton, photo-Fenton, and electro-Fenton reactions. Their optical band gaps, i.e., 1–2.1 eV for wüstite [151,198], 1.94 eV for feroxyhyte, and 2.1 eV for akaganéite [244], make them suitable for solar light-driven photocatalysis. Excluding wüstite, their crystal structure and chemical composition provide a high affinity for adsorbing heavy metals and organic pollutants. Notable differences include akaganéite’s superior chemical stability and wüstite’s potential in ammonia synthesis. Bernalite remains underexplored due to its rarity, although recent studies hint at promising environmental and electrochemical applications.
Production challenges include synthesis complexity and stability, especially for large-scale use. With magnetic properties, feroxyhyte is easier to handle in aqueous environments, while wüstite’s instability and bernalite’s scarcity hinder broader adoption. These contrasts highlight the need for innovative applications that optimize cost-effectiveness.

5. Future Directions and Perspectives on Catalytic Systems

The reaction mechanism of Fe-OH catalysts from natural sources depends on the evaluated reaction, conditions, and impurities. For example, in photocatalytic processes with Fe2O3 over gelcasting porous ceramic [78] (cf. Figure 8A), UV-induced electron excitation produces ROS that attack phenolic compounds. In contrast, persulfate (PS) activation (cf. Figure 8D), used for the catalytic degradation of 2,4-D and MCPA [73], occurs via Fe2+-generating SO4•− radicals both in homogeneous and heterogeneous phases. This dual-phase reaction provides enhanced degradation efficiency, with higher pH values compared to the more pH-dependent process photocatalytic, Fenton, and photo-Fenton processes.
The mechanism of the catalytic ozonation process mediated by HRS [81] (Figure 8B) involves three main pathways for the degradation of acetaminophen (ACT): (i) the direct oxidation of ACT by dissolved ozone (O3) molecules in the bulk solution; (ii) the radical-based degradation of ACT by hydroxyl radicals (OH) generated in the solution, and (iii) the indirect oxidation through the interaction of O3 with the catalyst surface, which produces OH and superoxide radicals (O2•−). The presence of ions like phosphate and carbonate can hinder the process by occupying active sites on the catalyst or reacting with OH. In turn, two key mechanisms drive SSZ ozonation using goethite as a catalyst [129] (cf. Figure 8C): (i) the reductive degradation on the catalyst surface and (ii) the direct oxidation by non-adsorbed O2•− radicals. Additionally, plasma treatment on α-FeOOH enhances the production of OH and O2•−, improving the ozonation process.
Figure 8E shows the mechanism of the photocatalytic reduction of 4-NP over Cu/α-Fe2O3 [77]. Here, we can observe that this follows three main steps: (i) the diffusion and adsorption of 4-NP onto the metal catalyst surface, (ii) electron transfer from BH4 to 4-NP, mediated by Fe and Cu metal particles, and (iii) the reduction of the nitro group by hydride transfer from a metal–hydride complex. This leads to the formation of 4-hydroxylaminophenol as a stable intermediate, which undergoes three hydro-deoxygenation reactions to produce 4-AP. Finally, 4-AP desorbs from the catalyst surface.
In transfer hydrogenation (cf. Figure 8F), the γ-Fe2O3 catalyst facilitates the initial activation of levulinic acid (LA) by interacting with its carbonyl carbon [149]. This interaction enhances the reactivity of LA, allowing it to interact with 2-propanol, which donates hydrogen to form 4-hydroxypentanoic acid as an intermediate. The iron catalyst promotes hydrogen transfer. For example, 4-hydroxypentanoic acid undergoes dehydration, which leads to the formation of Gamma-valerolactone (GVL). The Fe-OH minerals from natural sources can act as catalysts for biodiesel production. The mechanism proposed in Figure 8G [87] initiates with the adsorption of the carbonyl group in free fatty acids (FFAs) onto the acidic site and the adsorption of methanol onto the basic site of the catalyst. These interactions generate a carbocation and an oxygen anion, respectively. A nucleophilic attack occurs; the carbocation reacts with the hydroxyl group in glycerol on the acidic site, while the oxygen anion in methanol reacts with the carbonyl group in triglycerides on the basic site. This leads to the transesterification reaction, breaking down the three fatty acid chains of the triglyceride, yielding biodiesel molecules and glycerol as a byproduct. Simultaneously, the esterification of glycerol with FFAs on the acidic site produces water and diglycerides. Finally, the reaction products desorb, and the catalytic cycle restarts.
The separation of Fe-OH from other minerals and their purification can be difficult to achieve due to their textural intergrowths. However, the impurities present in Fe-OH catalysts can either inhibit or enhance catalytic reactions. There are few studies comparing synthetic catalysts to those prepared from natural sources. Based on existing comparative results [1,149], Fe-OHs from natural sources are more effective than their synthetic counterparts, but their catalytic activity is much more complex. For example, catalysts from different sources of mineral Fe3O4 with similar impurity content showed marked differences in the degradation of p-NP, indicating that the purity of the mineral might not be the determining factor for its catalytic activity. Here, the key factor in the performance is related to the surface site density of the parameter.
Maghemite prepared from river sand exhibited higher biomass conversion by transfer hydrogenation [149]. This was attributed to its crystallinity and higher pore size, allowing more interaction with the reactants. The presence of impurities of silica and metal carbonates also promotes the Lewis acidic sites. Therefore, one of the challenges for the use of these types of catalysts is understanding the effect of impurities on catalytic activity. A greater number of comparative studies between synthetic catalysts and those prepared from natural sources are needed, focusing on both their physicochemical characteristics and catalytic performances.
On the other hand, catalytic evaluation requires an appropriate experimental design as magnetic stirring could cause poor dispersion of particles in the solution, while mechanical stirring could hinder the scaling up of reactors to pilot or large-scale levels. Figure 9 shows different catalytic systems where Fe-OH prepared from natural sources were used as catalysts. At the laboratory scale, magnetic stirring (Figure 9C,F) can be used when there is a decrease in the magnetic response of catalysts. However, the use of pumps (Figure 9A,B,F) or gas carriers (Figure 9D,E) allows the continuous flow of solutions passing through the catalyst bed. To facilitate real-world scaling, we propose adapting and optimizing proven experimental reactor designs from analogous catalytic processes [80,245,246].
The scalability of catalysts largely depends on their stability, lifespan, and resistance to poisoning. One of the most remarkable characteristics of Fe-OH is their ability to be recovered using magnetic fields, enabling their reuse. Figure 10 depicts conversion versus the number of cycles reached by natural Fe-OH. There is more information available on the reuse cycles of the most abundant iron oxide (magnetite (Figure 10A) and hematite (Figure 10B)) and hydroxide catalysts (goethite, Figure 10C). In contrast, information about less abundant and rare Fe-OH (maghemite and wüstite, Figure 10D) is rarely found.
According to reported results, magnetite prepared from natural sources is a stable catalyst in the photo-Fenton degradation of cefotaxime [46], CWPO degradation of phenol [62], and Fenton-like degradation of MeP reactions [52], showing less than 10% of deactivation until after five reused cycles. The primary cause of deactivation appears to be the partial oxidation of the catalyst surface, leading to a progressive decrease in the Fe2+/Fe3+ ratio [57]. Nevertheless, to understand the decrease in reaction performance, it is necessary to characterize the catalysts after being recovered from the reaction. Comparisons between the physicochemical properties before and after the catalytic performance would help elucidate the mechanism of catalyst deactivation. The reuse results of hematite indicate that its conversion capacity decreases only slightly in the photocatalytic reduction of 4-NP to 4-AP [77], the ozonation of ACT [81], and the photo-Fenton degradation of phenol [74]. It also demonstrates stability in the cracking of toluene [89], where the reaction conditions are more aggressive than in AOPs. In this case, the catalysts’ deactivation is attributed to carbon deposits, the aggregation of active metal particles, and poisonous impurities. This decline could be due to (i) the loss of Fe mass and/or iron leaching from the surface after each usage cycle, (ii) the reduction in the activation rate of PS molecules by the catalyst, and, as a result, (iii) the reduction in the efficiency of the heterogeneous catalytic degradation process. However, the iron loss was minimal, which confirmed the stability of the catalyst. Further characterization is needed to better understand the deactivation mechanism.
Figure 10. Conversion vs. recycled use of iron oxides and hydroxides: (A) magnetite, (B) hematite, (C) goethite, and (D) maghemite and wüstite. Elaborated with data from References [46,49,51,52,55,57,62], [72,73,74,77,81,89], [122,123,129], and [149,151], respectively. Dash lines are guides for the eye.
Figure 10. Conversion vs. recycled use of iron oxides and hydroxides: (A) magnetite, (B) hematite, (C) goethite, and (D) maghemite and wüstite. Elaborated with data from References [46,49,51,52,55,57,62], [72,73,74,77,81,89], [122,123,129], and [149,151], respectively. Dash lines are guides for the eye.
Catalysts 15 00236 g010
The decrease in catalytic conversion mediated by goethite is more evident (cf. Figure 10C) compared to magnetite and hematite, being more stable on the ozonation of SSZ using plasma-treated goethite (shown by pink triangles) [129]. Natural goethite (shown by blue triangles) interestingly exhibited a higher drop in conversion. Using plasma treatment, the surface enhancement of α-FeOOH can increase the density of hydroxyl groups at the surface.
Related to less abundant and rare iron oxides, the scant available results about reuse indicated that they are highly stable during transfer hydrogenation [149] and the photo-Fenton degradation of ENR [151]. The recovered catalyst (maghemite) from transfer hydrogenation remained stable and did not lose crystallinity, phase purity, and activity, even after extensive usage. Meanwhile, the surface area of the goethite decreased compared to that of the fresh catalyst [149]. This is consistent with the discussion throughout this review, which suggests that surface area is not the determining factor in the activity of these catalysts. In the photo-Fenton degradation of ENR, the reuse of wüstite resulted in the release of iron, but after the first cycle, iron leaching decreases, and the macroscopic solid undergoes disaggregation. Additionally, the first cycle shows a reduction in catalyst particle size. Furthermore, the observed disaggregation and particle size reduction suggest alterations to the solid surface throughout reuse cycles, potentially impacting the catalyst’s interaction with light and H2O2 [151].
Generally, the reuse of Fe-OH catalysts does not require thermal treatments to reactivate the recovered catalysts, which reduces energy consumption and preparation costs. Additionally, the fact that they can be pretreated (or used directly without the need for thermal or chemical treatments) from inexpensive, accessible, and widely available sources may contribute to their scalability to real-world applications. One of the challenges to overcome in AOPs is the need for acidic pH conditions for these catalysts to remain active. One potential approach is exploring the use of oxidizing agents other than H2O2, such as PS, which has been shown to require nearly neutral pH conditions for effective catalytic reactions. In catalytic cracking, high temperatures are required, which can lead to the excessive consumption of network oxygen, causing catalyst deactivation. A thermal treatment in an oxidizing atmosphere could help reactivate it. However, this would increase energy consumption, potentially hindering its large-scale application.
Natural Fe-OH catalysts present challenges that must be addressed to enable their practical application. Their composition and mineralogical structure can vary depending on the source, affecting catalytic activity and reproducibility. Also, impurities may either inhibit or enhance catalytic performance, making it crucial to understand their role. In addition, AOP processes usually require acidic pH conditions, which limit natural Fe-OH applications in environmental samples or industrial waters. Another key challenge is the stability and reusability of these catalysts as iron leaching, surface fouling, or structural changes can lead to deactivation over multiple cycles. Furthermore, natural catalysts are widely available and cost-effective, but their efficiency must be comparable to synthetic catalysts to justify large-scale implementation without excessive energy-consuming treatments. Proper experimental design is also essential since factors such as stirring methods can impact catalyst dispersion and overall performance. To overcome these challenges, more comparative studies between natural and synthetic catalysts are needed, along with strategies to enhance stability, efficiency, and operational conditions.

6. Conclusions

While iron oxides are abundantly available in the Earth’s crust, their utilization as catalysts has been largely overshadowed by the prevalence of laboratory-synthesized materials. A key advantage of iron oxide catalysts lies in their magnetic properties, facilitating easy recovery and potential for reuse. However, crucial knowledge gaps remain. First, the impact of repeated reuse cycles on the magnetic properties of these catalysts and their subsequent influence on catalytic performance requires further investigation. Second, a comprehensive comparative cost analysis between the synthesis of iron oxide catalysts and their extraction and preparation from natural sources is currently lacking. The main applications of iron oxides are AOPs, electrochemical applications (water splitting and OER), FTS, and ammonia synthesis. Hematite (α-Fe2O3) is the most widely used naturally sourced iron oxide, probably because it usually represents the final stage in the transformation of other iron oxides. However, AOPs are operative only in a narrow interval of pH compared to other iron oxides. The next most used iron oxides are magnetite (Fe3O4) and goethite (α-FeOOH). The catalytic activity of magnetite has been related to the presence of Fe3+ and Fe2+ in its structure, where the oxidizing conditions can transform it into hematite. On the other hand, goethite can participate in redox reactions through the interconversion of Fe3+; and Fe2+, facilitating electron transfer in catalytic processes. This results in reactions where reactive species such as hydroxyl radicals (OH) are necessary. Additionally, goethite exhibits the highest specific surface area among the three most used iron oxides.
Although conventional catalyst preparation often involves acid purification and thermal treatments, simpler approaches utilizing magnetic separation, grinding, sieving, and milling have demonstrated promising catalytic activity in some evaluations. Despite that, these methods may have limitations when precise control over particle size and morphology is critical as achieving consistent shape and size distribution through these techniques can be challenging.
The preparation and evaluation of catalytic applications of less abundant natural iron oxides, such as maghemite (γ-Fe2O3), lepidocrocite (γ-FeOOH), and ferrihydrite (5Fe2O3⋅9H2O), face diverse challenges, as follows: natural maghemite is not readily available in a pure form, and lepidocrocite is rarely found in natural ore deposits, limiting its accessibility for catalytic applications. On the one hand, ferrihydrite exhibits poor crystallinity, a small particle size, and structural defects that make its physicochemical characterization and reproducibility in catalytic processes difficult. On the other hand, the so-called rare iron oxides, i.e., wüstite (FeO), akaganeite (β-FeOOH), feroxyhyte (δ′-FeOOH), and bernalite (Fe(OH)3), have been employed in their synthetic forms in few catalytic reactions. This may be related to their metastability, synthesis complexity, and dependence on specific operational conditions.
Despite the potential complexities associated with the preparation and evaluation of catalysts derived from natural sources, their abundance on the Earth’s surface, coupled with their unique physicochemical properties and low-cost availability, presents a compelling opportunity to address environmental and energy-related challenges. Regardless, optimizing their catalytic performance necessitates a thorough understanding of the specific characteristics of each iron oxide to enable the development of effective strategies for scaling up these processes to pilot and industrial levels.

Author Contributions

Writing—original draft preparation and conceptualization, supervision, investigation, and methodology, A.J.-V.; Writing—original draft preparation, investigation, and conceptualization, R.J.-L.; Funding acquisition and methodology, S.P.-R.; investigation and methodology, C.M.M.-B. and Y.G.-P.; writing—review and editing, writing—original draft preparation and conceptualization, and funding acquisition, L.A.E.-W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially funded by Instituto Politécnico Nacional through the SIP projects (Reference No. SIP-20240967, SIP-20242804, and SIP-20240049).

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

R. Jaimes-López acknowledges the postdoctoral scholarship from SECIHTI “Estancias Posdoctorales por México para la Formación y Consolidación de las y los Investigadores por México 2022”.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of this study, the collection, analysis, or interpretation of the data, the writing of the manuscript, or the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
2,4-D2,4-dichlorophenoxyacetic acid
4AP4-aminophenol
4-NP4-nitrophenol
ACTAcetaminophen
AO7Acid orange 7
AOPAdvanced oxidation process
ATR-FTIRAttenuated total reflection–Fourier transform infrared spectroscopy
BIFsBanded iron formations
BSA5-bromosalicylic acid
CAPChloramphenicol
CGWContaminated groundwater
CODChemical oxygen demand
CWPOCatalytic wet peroxide oxidation
DCFDiclofenac
DFTDensity functional theory
DTADifferential thermal analysis
EDTAEthylenediaminetetracetic acid
ENREnrofloxacin
FBRFixed-bed reactor
Fe-OHIron oxides and iron hydroxides
GC/MSGas chromatography–mass spectrometry
GHSVGas hourly space velocity
GVLGamma-valerolactone
HcCoercivity
HERHydrogen evolution reaction
HRSHormuz red soil
HTCTHigh-temperature catalytic treatment
ICIndigo carmine
IMAInternational Mineralogical Association
IRInfrared spectroscopy
LALevulinic acid
LEVLevofloxacin
MBMethylene blue
MC-LRMicrocystin-LR
MCPA2-methyl-4-chlorophenoxyacetic acid
MePMethylparaben
MFCsMicrobial fuel cells
MMMonodentate mononuclear
MOMethyl orange
MrRemanent magnetization
NPsNanoparticles
OCsOxygen carriers
OEROxygen evolution reaction
ORROxygen reduction reaction
OGOrange G
PANIPolyaniline
p-NPp-nitrophenol
PSPersulfate
ROSReactive oxygen species
SEMScanning electron microscopy
SMXSulfamethoxazole
SPSPersulfate
SSZSulfasalazine antibiotic
TCTetracycline
TEMTransmission electron microscopy
TGAThermogravimetric analysis
TMTemperature of Morin transition
WFFDBDWater falling film dielectric barrier discharge
WWTPWastewater treatment plant
XANESX-ray absorption near-edge structure
XPSX-ray photoelectron spectroscopy
XRDX-ray diffraction
XRFX-ray fluorescence

References

  1. He, H.; Zhong, Y.; Liang, X.; Tan, W.; Zhu, J.; Wang, C.Y. Natural Magnetite: An Efficient Catalyst for the Degradation of Organic Contaminant. Sci. Rep. 2015, 5, 10139. [Google Scholar] [CrossRef] [PubMed]
  2. Cornell, R.M.; Schwertmann, U. The Iron Oxides; Wiley: Hoboken, NJ, USA, 2003; ISBN 9783527302741. [Google Scholar]
  3. Lagroix, F.; Banerjee, S.K.; Jackson, M.J. Geological Occurrences and Relevance of Iron Oxides. In Iron Oxides; Faivre, D., Ed.; Wiley: Hoboken, NJ, USA, 2016; pp. 7–30. [Google Scholar]
  4. Jambor, J.L.; Dutrizac, J.E. Occurrence and Constitution of Natural and Synthetic Ferrihydrite, a Widespread Iron Oxyhydroxide. Chem. Rev. 1998, 98, 2549–2586. [Google Scholar] [CrossRef] [PubMed]
  5. Misra, K.C. Understanding Mineral Deposits; Springer: Dordrecht, The Netherlands, 2000; ISBN 978-94-010-5752-3. [Google Scholar]
  6. Schwertmann, U.; Fitzpatrick, R. Iron Minerals in Surface Environments. Catena Suppl. 1992, 21, 7–30. [Google Scholar]
  7. Dowlath, M.J.H.; Musthafa, S.A.; Mohamed Khalith, S.B.; Varjani, S.; Karuppannan, S.K.; Ramanujam, G.M.; Arunachalam, A.M.; Arunachalam, K.D.; Chandrasekaran, M.; Chang, S.W.; et al. Comparison of Characteristics and Biocompatibility of Green Synthesized Iron Oxide Nanoparticles with Chemical Synthesized Nanoparticles. Environ. Res. 2021, 201, 111585. [Google Scholar] [CrossRef]
  8. Patil, R.M.; Shete, P.B.; Thorat, N.D.; Otari, S.V.; Barick, K.C.; Prasad, A.; Ningthoujam, R.S.; Tiwale, B.M.; Pawar, S.H. Superparamagnetic Iron Oxide/Chitosan Core/Shells for Hyperthermia Application: Improved Colloidal Stability and Biocompatibility. J. Magn. Magn. Mater. 2014, 355, 22–30. [Google Scholar] [CrossRef]
  9. Demirer, G.S.; Okur, A.C.; Kizilel, S. Synthesis and Design of Biologically Inspired Biocompatible Iron Oxide Nanoparticles for Biomedical Applications. J. Mater. Chem. B 2015, 3, 7831–7849. [Google Scholar] [CrossRef]
  10. Pereira, M.C.; Oliveira, L.C.A.; Murad, E. Iron Oxide Catalysts: Fenton and Fentonlike Reactions—A Review. Clay Min. 2012, 47, 285–302. [Google Scholar] [CrossRef]
  11. Rahim Pouran, S.; Abdul Raman, A.A.; Wan Daud, W.M.A. Review on the Application of Modified Iron Oxides as Heterogeneous Catalysts in Fenton Reactions. J. Clean. Prod. 2014, 64, 24–35. [Google Scholar] [CrossRef]
  12. Wu, Q.; Siddique, M.S.; Wang, H.; Cui, L.; Wang, H.; Pan, M.; Yan, J. Visible-Light-Driven Iron-Based Heterogeneous Photo-Fenton Catalysts for Wastewater Decontamination: A Review of Recent Advances. Chemosphere 2023, 313, 137509. [Google Scholar] [CrossRef]
  13. Gholami, Z.; Gholami, F.; Tišler, Z.; Hubáček, J.; Tomas, M.; Bačiak, M.; Vakili, M. Production of Light Olefins via Fischer-Tropsch Process Using Iron-Based Catalysts: A Review. Catalysts 2022, 12, 174. [Google Scholar] [CrossRef]
  14. Okoye-Chine, C.G.; Mubenesha, S. The Use of Iron Ore as a Catalyst in Fischer–Tropsch Synthesis—A Review. Crystals 2022, 12, 1349. [Google Scholar] [CrossRef]
  15. Ramadhani, B.; Kivevele, T.; Kihedu, J.H.; Jande, Y.A.C. Catalytic Tar Conversion and the Prospective Use of Iron-Based Catalyst in the Future Development of Biomass Gasification: A Review. Biomass Convers. Biorefinery 2022, 12, 1369–1392. [Google Scholar] [CrossRef]
  16. Zhu, M.; Wachs, I.E. Iron-Based Catalysts for the High-Temperature Water–Gas Shift (HT-WGS) Reaction: A Review. ACS Catal. 2016, 6, 722–732. [Google Scholar] [CrossRef]
  17. Cao, Y.; Yuan, X.; Zhao, Y.; Wang, H. In-Situ Soil Remediation via Heterogeneous Iron-Based Catalysts Activated Persulfate Process: A Review. Chem. Eng. J. 2022, 431, 133833. [Google Scholar] [CrossRef]
  18. Cui, B.; Tian, T.; Duan, L.; Rong, H.; Chen, Z.; Luo, S.; Guo, D.; Naidu, R. Towards Advanced Removal of Organics in Persulfate Solution by Heterogeneous Iron-Based Catalyst: A Review. J. Environ. Sci. 2024, 146, 163–175. [Google Scholar] [CrossRef]
  19. Sang, W.; Xu, X.; Zhan, C.; Lu, W.; Jia, D.; Wang, C.; Zhang, Q.; Gan, F.; Li, M. Recent Advances of Antibiotics Degradation in Different Environment by Iron-Based Catalysts Activated Persulfate: A Review. J. Water Process Eng. 2022, 49, 103075. [Google Scholar] [CrossRef]
  20. Lai, L.; He, Y.; Zhou, H.; Huang, B.; Yao, G.; Lai, B. Critical Review of Natural Iron-Based Minerals Used as Heterogeneous Catalysts in Peroxide Activation Processes: Characteristics, Applications and Mechanisms. J. Hazard. Mater. 2021, 416, 125809. [Google Scholar] [CrossRef]
  21. Roy, S.D.; Das, K.C.; Dhar, S.S. Conventional to Green Synthesis of Magnetic Iron Oxide Nanoparticles; Its Application as Catalyst, Photocatalyst and Toxicity: A Short Review. Inorg. Chem. Commun. 2021, 134, 109050. [Google Scholar] [CrossRef]
  22. Ruan, Y.; Kong, L.; Zhong, Y.; Diao, Z.; Shih, K.; Hou, L.; Wang, S.; Chen, D. Review on the Synthesis and Activity of Iron-Based Catalyst in Catalytic Oxidation of Refractory Organic Pollutants in Wastewater. J. Clean. Prod. 2021, 321, 128924. [Google Scholar] [CrossRef]
  23. Azfar Shaida, M.; Verma, S.; Talukdar, S.; Kumar, N.; Salim Mahtab, M.; Naushad, M.; Haq Farooqi, I. Critical Analysis of the Role of Various Iron-Based Heterogeneous Catalysts for Advanced Oxidation Processes: A State of the Art Review. J. Mol. Liq. 2023, 374, 121259. [Google Scholar] [CrossRef]
  24. Munoz, M.; de Pedro, Z.M.; Casas, J.A.; Rodriguez, J.J. Preparation of Magnetite-Based Catalysts and Their Application in Heterogeneous Fenton Oxidation—A Review. Appl. Catal. B 2015, 176–177, 249–265. [Google Scholar] [CrossRef]
  25. Buthelezi, A.S.; Tucker, C.L.; Heeres, H.J.; Shozi, M.L.; van de Bovenkamp, H.H.; Ntola, P. Fischer-Tropsch Synthesis Using Promoted, Unsupported, Supported, Bimetallic and Spray-Dried Iron Catalysts: A Review. Results Chem. 2024, 9, 101623. [Google Scholar] [CrossRef]
  26. Jia, J.-Y.; Shan, Y.-L.; Tuo, Y.-X.; Yan, H.; Feng, X.; Chen, D. Review of Iron-Based Catalysts for Carbon Dioxide Fischer–Tropsch Synthesis. Trans. Tianjin Univ. 2024, 30, 178–197. [Google Scholar] [CrossRef]
  27. Zhang, M.; Li, Y.; Tian, X.; Dai, L.; Wang, G.; Lei, Z.; Ma, G.; Zuo, Q.; Li, M.; Zhao, M.; et al. A Review of Iron-Based Catalysts for Persulfate Activation to Remove PFAS in Water: Catalytic Effects of Various Iron Species, Influencing Factors and Reaction Pathways. Water Air Soil. Pollut. 2025, 236, 42. [Google Scholar] [CrossRef]
  28. Wu, H.; Li, J.; Gao, M.; Chen, Y.; Ren, S.; Yang, J.; Liu, Q. Deactivation Mechanisms and Strategies to Mitigate Deactivation of Iron-Based Catalysts in NH3-SCR for NO Reduction: A Comprehensive Review. Sep. Purif. Technol. 2025, 358, 130268. [Google Scholar] [CrossRef]
  29. Al-Anazi, A.; Newair, E.F.; Han, C. The Development of Iron-Based Catalysts for the Reduction of Nitrogen and per- and Polyfluoroalkyl Substances in Wastewater Treatment: A Review. Curr. Opin. Chem. Eng. 2024, 44, 101024. [Google Scholar] [CrossRef]
  30. Xu, H.; Ma, Z.; Wan, Z.; An, Z.; Wang, X. Recent Advances and Prospects of Iron-Based Noble Metal-Free Catalysts for Oxygen Reduction Reaction in Acidic Environment: A Mini Review. Int. J. Hydrogen Energy 2024, 59, 697–714. [Google Scholar] [CrossRef]
  31. Tarek, M.; Santos, J.S.; Márquez, V.; Fereidooni, M.; Yazdanpanah, M.; Praserthdam, S.; Praserthdam, P. A Critical Review towards the Causes of the Iron-Based Catalysts Deactivation Mechanisms in the Selective Oxidation of Hydrogen Sulfide to Elemental Sulfur from Biogas. J. Energy Chem. 2024, 90, 388–411. [Google Scholar] [CrossRef]
  32. Peng, H.; Hou, L.; Sun, C.; Zou, H.; Wang, T.-R.; Ma, Z.-Z. Geochemistry of Magnetite from the Hongniu–Hongshan Cu Skarn Deposit in Yunnan Province, SW China. Ore Geol. Rev. 2021, 136, 104237. [Google Scholar] [CrossRef]
  33. Deditius, A.P.; Reich, M.; Simon, A.C.; Suvorova, A.; Knipping, J.; Roberts, M.P.; Rubanov, S.; Dodd, A.; Saunders, M. Nanogeochemistry of Hydrothermal Magnetite. Contrib. Mineral. Petrol. 2018, 173, 46. [Google Scholar] [CrossRef]
  34. Usman, M.; Byrne, J.M.; Chaudhary, A.; Orsetti, S.; Hanna, K.; Ruby, C.; Kappler, A.; Haderlein, S.B. Magnetite and Green Rust: Synthesis, Properties, and Environmental Applications of Mixed-Valent Iron Minerals. Chem. Rev. 2018, 118, 3251–3304. [Google Scholar] [CrossRef] [PubMed]
  35. Kozlenko, D.P.; Dubrovinsky, L.S.; Kichanov, S.E.; Lukin, E.V.; Cerantola, V.; Chumakov, A.I.; Savenko, B.N. Magnetic and Electronic Properties of Magnetite across the High Pressure Anomaly. Sci. Rep. 2019, 9, 4464. [Google Scholar] [CrossRef]
  36. Liu, M.; Ye, Y.; Ye, J.; Gao, T.; Wang, D.; Chen, G.; Song, Z. Recent Advances of Magnetite (Fe3O4)-Based Magnetic Materials in Catalytic Applications. Magnetochemistry 2023, 9, 110. [Google Scholar] [CrossRef]
  37. Gawande, M.B.; Branco, P.S.; Varma, R.S. Nano-Magnetite (Fe3O4) as a Support for Recyclable Catalysts in the Development of Sustainable Methodologies. Chem. Soc. Rev. 2013, 42, 3371. [Google Scholar] [CrossRef] [PubMed]
  38. Niculescu, A.-G.; Chircov, C.; Grumezescu, A.M. Magnetite Nanoparticles: Synthesis Methods—A Comparative Review. Methods 2022, 199, 16–27. [Google Scholar] [CrossRef]
  39. Nengsih, S.; Madjid, S.N.; Mursal; Jalil, Z. Synthesis and Characterization of Magnetite Particles from Syiah Kuala Iron Sand Prepared by Co-Precipitation Method. J. Phys. Conf. Ser. 2023, 2582, 012005. [Google Scholar] [CrossRef]
  40. Safitri, I.; Wibowo, Y.G.; Rosarina, D.; Sudibyo. Synthesis and Characterization of Magnetite (Fe3O4) Nanoparticles from Iron Sand in Batanghari Beach. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1011, 012020. [Google Scholar] [CrossRef]
  41. Heryanto, H.; Tahir, D. The Correlations between Structural and Optical Properties of Magnetite Nanoparticles Synthesised from Natural Iron Sand. Ceram. Int. 2021, 47, 16820–16827. [Google Scholar] [CrossRef]
  42. Darvina, Y.; Yulfriska, N.; Rifai, H.; Dwiridal, L.; Ramli, R. Synthesis of Magnetite Nanoparticles from Iron Sand by Ball-Milling. J. Phys. Conf. Ser. 2019, 1185, 012017. [Google Scholar] [CrossRef]
  43. Morel, M.; Martínez, F.; Mosquera, E. Synthesis and Characterization of Magnetite Nanoparticles from Mineral Magnetite. J. Magn. Magn. Mater. 2013, 343, 76–81. [Google Scholar] [CrossRef]
  44. Amiruddin, E.; Awaluddin, A.; Hariani, I.; Sihombing, R.; Angraini, R. The Influence of Milling Ball Size on the Structural, Morphological and Catalytic Properties of Magnetite (Fe3O4) Nanoparticles toward Methylene Blue Degradation. J. Phys. Conf. Ser. 2020, 1655, 012006. [Google Scholar] [CrossRef]
  45. Khan, R.; Ahmad, I.; Khan, H.; Ismail, M.; Gul, K.; Yasin, A.; Ahmad, W. Production of Diesel-like Fuel from Spent Engine Oil by Catalytic Pyrolysis over Natural Magnetite. J. Anal. Appl. Pyrolysis 2016, 120, 493–500. [Google Scholar] [CrossRef]
  46. Jiang, Q.; Zhu, R. Facile Synthesis of Highly Efficient and Cost-Effective Photo-Fenton Catalyst by Ball Milling Commercial TiO2 and Natural Magnetite. J. Alloys Compd. 2021, 862, 158670. [Google Scholar] [CrossRef]
  47. Priyadarshini, M.; Das, I.; Ghangrekar, M.M.; Blaney, L. Advanced Oxidation Processes: Performance, Advantages, and Scale-up of Emerging Technologies. J. Environ. Manag. 2022, 316, 115295. [Google Scholar] [CrossRef] [PubMed]
  48. Vafaei Molamahmood, H.; Geng, W.; Wei, Y.; Miao, J.; Yu, S.; Shahi, A.; Chen, C.; Long, M. Catalyzed H2O2 Decomposition over Iron Oxides and Oxyhydroxides: Insights from Oxygen Production and Organic Degradation. Chemosphere 2022, 291, 133037. [Google Scholar] [CrossRef]
  49. Aghdasinia, H.; Arehjani, P.; Vahid, B.; Khataee, A. Fluidized-Bed Fenton-like Oxidation of a Textile Dye Using Natural Magnetite. Res. Chem. Intermed. 2016, 42, 8083–8095. [Google Scholar] [CrossRef]
  50. Lopez-Arago, N.; Munoz, M.; de Pedro, Z.M.; Casas, J.A. Natural Magnetite as an Effective and Long-Lasting Catalyst for CWPO of Azole Pesticides in a Continuous up-Flow Fixed-Bed Reactor. Environ. Sci. Pollut. Res. 2024, 31, 29148–29161. [Google Scholar] [CrossRef]
  51. Mufti, N.; Maryam, S.; Setiyanto, H.; Taufiq, A.; Sunaryono, S. Recyclable Natural Magnetite Nanoparticles for Effective Degradation of Methylene Blue in Water under UV Light Irradiation. Key Eng. Mater. 2020, 855, 315–321. [Google Scholar] [CrossRef]
  52. Rostamifasih, Z.; Pasalari, H.; Mohammadi, F.; Esrafili, A. Heterogeneous Catalytic Degradation of Methylparaben Using Persulfate Activated by Natural Magnetite; Optimization and Modeling by Response Surface Methodology. J. Chem. Technol. Biotechnol. 2019, 94, 1880–1892. [Google Scholar] [CrossRef]
  53. Guo, R.; Xi, B.; Guo, C.; Cheng, X.; Lv, N.; Liu, W.; Borthwick, A.G.L.; Xu, J. Persulfate-Based Advanced Oxidation Processes: The New Hope Brought by Nanocatalyst Immobilization. Environ. Funct. Mater. 2022, 1, 67–91. [Google Scholar] [CrossRef]
  54. Hajiahmadi, M.; Zarei, M.; Khataee, A. Introducing an Effective Iron-Based Catalyst for Heterogeneous Electro-Fenton Removal of Gemcitabine Using Three-Dimensional Graphene as Cathode. J. Ind. Eng. Chem. 2021, 96, 254–268. [Google Scholar] [CrossRef]
  55. Munoz, M.; Conde, J.; de Pedro, Z.M.; Casas, J.A. Antibiotics Abatement in Synthetic and Real Aqueous Matrices by H2O2/Natural Magnetite. Catal. Today 2018, 313, 142–147. [Google Scholar] [CrossRef]
  56. Rueda Márquez, J.J.; Levchuk, I.; Sillanpää, M. Application of Catalytic Wet Peroxide Oxidation for Industrial and Urban Wastewater Treatment: A Review. Catalysts 2018, 8, 673. [Google Scholar] [CrossRef]
  57. Álvarez-Torrellas, S.; Munoz, M.; Mondejar, V.; de Pedro, Z.M.; Casas, J.A. Boosting the Catalytic Activity of Natural Magnetite for Wet Peroxide Oxidation. Environ. Sci. Pollut. Res. 2020, 27, 1176–1185. [Google Scholar] [CrossRef]
  58. Lyu, L.; Bagchi, M.; Markoglou, N.; An, C.; Peng, H.; Bi, H.; Yang, X.; Sun, H. Towards Environmentally Sustainable Management: A Review on the Generation, Degradation, and Recycling of Polypropylene Face Mask Waste. J. Hazard. Mater. 2024, 461, 132566. [Google Scholar] [CrossRef]
  59. He, L.; Hui, H.; Li, S.; Lin, W. Production of Light Aromatic Hydrocarbons by Catalytic Cracking of Coal Pyrolysis Vapors over Natural Iron Ores. Fuel 2018, 216, 227–232. [Google Scholar] [CrossRef]
  60. Schulz, H. Short History and Present Trends of Fischer–Tropsch Synthesis. Appl. Catal. A Gen. 1999, 186, 3–12. [Google Scholar] [CrossRef]
  61. Zhang, Q.; Gu, J.; Chen, J.; Qiu, S.; Wang, T. Facile Fabrication of Porous Fe@C Nanohybrids from Natural Magnetite as Excellent Fischer–Tropsch Catalysts. Chem. Commun. 2020, 56, 4523–4526. [Google Scholar] [CrossRef]
  62. Munoz, M.; Domínguez, P.; de Pedro, Z.M.; Casas, J.A.; Rodriguez, J.J. Naturally-Occurring Iron Minerals as Inexpensive Catalysts for CWPO. Appl. Catal. B 2017, 203, 166–173. [Google Scholar] [CrossRef]
  63. Widayat; Satriadi, H.; Setyojati, P.W.; Shihab, D.; Buchori, L.; Hadiyanto, H.; Nurushofa, F.A. Preparation CaO/MgO/Fe3O4 Magnetite Catalyst and Catalytic Test for Biodiesel Production. Results Eng. 2024, 22, 102202. [Google Scholar] [CrossRef]
  64. Roberts, A.P.; Zhao, X.; Heslop, D.; Abrajevitch, A.; Chen, Y.-H.; Hu, P.; Jiang, Z.; Liu, Q.; Pillans, B.J. Hematite (α-Fe2O3) Quantification in Sedimentary Magnetism: Limitations of Existing Proxies and Ways Forward. Geosci. Lett. 2020, 7, 8. [Google Scholar] [CrossRef]
  65. Manjunatha, M.; Kumar, R.; Anupama, A.V.; Khopkar, V.B.; Damle, R.; Ramesh, K.P.; Sahoo, B. XRD, Internal Field-NMR and Mössbauer Spectroscopy Study of Composition, Structure and Magnetic Properties of Iron Oxide Phases in Iron Ores. J. Mater. Res. Technol. 2019, 8, 2192–2200. [Google Scholar] [CrossRef]
  66. Amiruddin, E.; Awaluddin, A.; Sinuraya, S.; Hadianto, H.; Noferdi, M.D.; Fitri, A.S. Study of Iron Oxide Nanoparticles Doped with Manganese for Catalytic Degradation of Methylene Blue. J. Phys. Conf. Ser. 2021, 2049, 012021. [Google Scholar] [CrossRef]
  67. Jalil, Z.; Rahwanto, A.; Mulana, F.; Ismail, I.; Anggreini Akbar, I.; Budi Aulia, T. Effet of Hematite (Fe2O3) from Natural Iron Ore as a Catalyst in MgH2 as Hydrogen Storage Materials. J. Phys. Conf. Ser. 2023, 2498, 012048. [Google Scholar] [CrossRef]
  68. Mandal, S.; Müller, A.H.E. Facile Route to the Synthesis of Porous α-Fe2O3 Nanorods. Mater. Chem. Phys. 2008, 111, 438–443. [Google Scholar] [CrossRef]
  69. Zhang, Y.C.; Tang, J.Y.; Hu, X.Y. Controllable Synthesis and Magnetic Properties of Pure Hematite and Maghemite Nanocrystals from a Molecular Precursor. J. Alloys Compd. 2008, 462, 24–28. [Google Scholar] [CrossRef]
  70. Malaidurai, M.; Thangavel, R. Study of Structural and Magnetic Properties of Co-Precipitated α-Fe2O3 Nanocrystals. Superlattices Microstruct. 2018, 120, 553–560. [Google Scholar] [CrossRef]
  71. Lohaus, C.; Klein, A.; Jaegermann, W. Limitation of Fermi Level Shifts by Polaron Defect States in Hematite Photoelectrodes. Nat. Commun. 2018, 9, 4309. [Google Scholar] [CrossRef]
  72. Omri, A.; Hamza, W.; Benzina, M. Photo-Fenton Oxidation and Mineralization of Methyl Orange Using Fe-Sand as Effective Heterogeneous Catalyst. J. Photochem. Photobiol. A Chem. 2020, 393, 112444. [Google Scholar] [CrossRef]
  73. Kermani, M.; Mohammadi, F.; Kakavandi, B.; Esrafili, A.; Rostamifasih, Z. Simultaneous Catalytic Degradation of 2,4-D and MCPA Herbicides Using Sulfate Radical-Based Heterogeneous Oxidation over Persulfate Activated by Natural Hematite (α-Fe2O3/PS). J. Phys. Chem. Solids 2018, 117, 49–59. [Google Scholar] [CrossRef]
  74. Hollanda, L.R.; de Souza, J.A.B.; Dotto, G.L.; Foletto, E.L.; Chiavone-Filho, O. Iron-Bearing Mining Reject as an Alternative and Effective Catalyst for Photo-Fenton Oxidation of Phenol in Water. Environ. Sci. Pollut. Res. 2024, 31, 21291–21301. [Google Scholar] [CrossRef] [PubMed]
  75. Mohammadi, S.; Moussavi, G.; Shekoohiyan, S.; Marín, M.L.; Boscá, F.; Giannakis, S. A Continuous-Flow Catalytic Process with Natural Hematite-Alginate Beads for Effective Water Decontamination and Disinfection: Peroxymonosulfate Activation Leading to Dominant Sulfate Radical and Minor Non-Radical Pathways. Chem. Eng. J. 2021, 411, 127738. [Google Scholar] [CrossRef]
  76. Lubis, S.; Sheilatina; Murisna. Synthesis, Characterization and Photocatalytic Activity of α-Fe2O3/Bentonite Composite Prepared by Mechanical Milling. J. Phys. Conf. Ser. 2018, 1116, 042016. [Google Scholar] [CrossRef]
  77. Elfiad, A.; Galli, F.; Boukhobza, L.M.; Djadoun, A.; Boffito, D.C. Low-Cost Synthesis of Cu/α-Fe2O3 from Natural HFeO2: Application in 4-Nitrophenol Reduction. J. Environ. Chem. Eng. 2020, 8, 104214. [Google Scholar] [CrossRef]
  78. Putri, S.E.; Pratiwi, D.E.; Tjahjanto, R.T.; Hasri; Andi, I.; Rahman, A.; Ramadani, A.I.W.S.; Ramadhani, A.N.; Subaer; Fudholi, A. The Renewable of Low Toxicity Gelcasting Porous Ceramic as Fe2O3 Catalyst Support on Phenol Photodegradation. Int. J. Des. Nat. Ecodynamics 2022, 17, 503–511. [Google Scholar] [CrossRef]
  79. Elfiad, A.; Galli, F.; Djadoun, A.; Sennour, M.; Chegrouche, S.; Meddour-Boukhobza, L.; Boffito, D.C. Natural α-Fe2O3 as an Efficient Catalyst for the p-Nitrophenol Reduction. Mater. Sci. Eng. B 2018, 229, 126–134. [Google Scholar] [CrossRef]
  80. Jaimes-López, R.; Jiménez-Vázquez, A.; Pérez-Rodríguez, S.; Estudillo-Wong, L.A.; Alonso-Vante, N. Catalyst for the Generation of OH Radicals in Advanced Electrochemical Oxidation Processes: Present and Future Perspectives. Catalysts 2024, 14, 703. [Google Scholar] [CrossRef]
  81. Kohantorabi, M.; Moussavi, G.; Oulego, P.; Giannakis, S. Heterogeneous Catalytic Ozonation and Peroxone-Mediated Removal of Acetaminophen Using Natural and Modified Hematite-Rich Soil, as Efficient and Environmentally Friendly Catalysts. Appl. Catal. B 2022, 301, 120786. [Google Scholar] [CrossRef]
  82. Yue, Y.; Niu, P.; Jiang, L.; Cao, Y.; Bao, X. Acid-Modified Natural Bauxite Mineral as a Cost-Effective and High-Efficient Catalyst Support for Slurry-Phase Hydrocracking of High-Temperature Coal Tar. Energy Fuels 2016, 30, 9203–9209. [Google Scholar] [CrossRef]
  83. Robinson, P.R. Hydroconversion Processes and Technology for Clean Fuel and Chemical Production. In Advances in Clean Hydrocarbon Fuel Processing; Elsevier: Amsterdam, The Netherlands, 2011; pp. 287–325. [Google Scholar]
  84. Kairbekov, Z.K.; Maloletnev, A.S.; Dzheldybaeva, I.M.; Sabitova, A.N.; Ermoldina, E.T. Application of Modified Iron-Containing Catalysts and Preliminary Ozonization of Coal from the Shubarkol Deposit to the Hydrogenation of This Coal. Solid. Fuel Chem. 2017, 51, 365–369. [Google Scholar] [CrossRef]
  85. Kretzschmar, N.; Busse, O.; Seifert, M. Impact of Geometric and Electronic Factors on Selective Hydro-Deoxygenation of Guaiacol by Surface-Rich Metal/Silica Catalysts. Catalysts 2023, 13, 425. [Google Scholar] [CrossRef]
  86. Widayat; Putra, D.A.; Nursafitri, I. Synthesis and Catalytic Evaluation of Hematite (α-Fe2O3) Magnetic Nanoparticles from Iron Sand for Waste Cooking Oil Conversion to Produce Biodiesel through Esterification-Transesterification Method. IOP Conf. Ser. Mater. Sci. Eng. 2019, 509, 012035. [Google Scholar] [CrossRef]
  87. Prameswari, J.; Widayat, W.; Buchori, L.; Hadiyanto, H. Novel Iron Sand-Derived α-Fe2O3/CaO2 Bifunctional Catalyst for Waste Cooking Oil-Based Biodiesel Production. Environ. Sci. Pollut. Res. 2022, 30, 98832–98847. [Google Scholar] [CrossRef]
  88. Abdrafikova, I.M.; Kayukova, G.P.; Petrov, S.M.; Ramazanova, A.I.; Musin, R.Z.; Morozov, V.I. Conversion of Extra-Heavy Ashal’chinskoe Oil in Hydrothermal Catalytic System. Pet. Chem. 2015, 55, 104–111. [Google Scholar] [CrossRef]
  89. Zou, X.; Chen, T.; Liu, H.; Zhang, P.; Chen, D.; Zhu, C. Catalytic Cracking of Toluene over Hematite Derived from Thermally Treated Natural Limonite. Fuel 2016, 177, 180–189. [Google Scholar] [CrossRef]
  90. Li, Y.; Gong, J.; Huang, F.; Bai, H.; Wang, F.; Wang, C. Carbon Deposition and Sintering Characteristics on Iron-Based Oxygen Carriers in the Catalytic Cracking Process of Coal Tar. Energy Fuels 2017, 31, 6501–6506. [Google Scholar] [CrossRef]
  91. Sayeed, M.A.; Fernando, J.F.S.; O’Mullane, A.P. Conversion of Iron Ore into an Active Catalyst for the Oxygen Evolution Reaction. Adv. Sustain. Syst. 2018, 2, 1800019. [Google Scholar] [CrossRef]
  92. Fabbri, E.; Schmidt, T.J. Oxygen Evolution Reaction—The Enigma in Water Electrolysis. ACS Catal. 2018, 8, 9765–9774. [Google Scholar] [CrossRef]
  93. Alduhaish, O.; Ubaidullah, M.; Al-Enizi, A.M.; Alhokbany, N.; Alshehri, S.M.; Ahmed, J. Facile Synthesis of Mesoporous α-Fe2O3@g-C3N4-NCs for Efficient Bifunctional Electro-Catalytic Activity (OER/ORR). Sci. Rep. 2019, 9, 14139. [Google Scholar] [CrossRef]
  94. Farooq, A.; Khalil, S.; Basha, B.; Habib, A.; Al-Buriahi, M.S.; Warsi, M.F.; Yousaf, S.; Shahid, M. Electrochemical Investigation of C-Doped CoFe2O4/Fe2O3Nanostructures for Efficient Electrochemical Water Splitting. Int. J. Hydrogen Energy 2024, 51, 1318–1332. [Google Scholar] [CrossRef]
  95. Ng, K.-L.; Kok, K.-Y.; Ong, B.-H. Facile Synthesis of Self-Assembled Cobalt Oxide Supported on Iron Oxide as the Novel Electrocatalyst for Enhanced Electrochemical Water Electrolysis. ACS Appl. Nano Mater. 2018, 1, 401–409. [Google Scholar] [CrossRef]
  96. Zang, Y.; Zhang, H.; Zhang, X.; Liu, R.; Liu, S.; Wang, G.; Zhang, Y.; Zhao, H. Fe/Fe2O3 Nanoparticles Anchored on Fe-N-Doped Carbon Nanosheets as Bifunctional Oxygen Electrocatalysts for Rechargeable Zinc-Air Batteries. Nano Res. 2016, 9, 2123–2137. [Google Scholar] [CrossRef]
  97. Guo, W.H.; Zhang, Q.; Fu, H.C.; Yang, Y.X.; Chen, X.H.; Luo, H.Q.; Li, N.B. Semi-Sacrificial Template Growth of FeS/Fe2O3 Heterogeneous Nanosheets on Iron Foam for Efficient Oxygen Evolution Reaction. J. Alloys Compd. 2022, 909, 164670. [Google Scholar] [CrossRef]
  98. Khan, N.A.; Rashid, N.; Ahmad, I.; Zahidullah; Zairov, R.; Rehman, H.U.; Nazar, M.F.; Jabeen, U. An Efficient Fe2O3/FeS Heterostructures Water Oxidation Catalyst. Int. J. Hydrogen Energy 2022, 47, 22340–22347. [Google Scholar] [CrossRef]
  99. Bandal, H.A.; Jadhav, A.R.; Chaugule, A.A.; Chung, W.-J.; Kim, H. Fe2O3 Hollow Nanorods/CNT Composites as an Efficient Electrocatalyst for Oxygen Evolution Reaction. Electrochim. Acta 2016, 222, 1316–1325. [Google Scholar] [CrossRef]
  100. Ahmad, I.; Ahmed, J.; Batool, S.; Zafar, M.N.; Hanif, A.; Zahidullah; Nazar, M.F.; Ul-Hamid, A.; Jabeen, U.; Dahshan, A.; et al. Design and Fabrication of Fe2O3/FeP Heterostructure for Oxygen Evolution Reaction Electrocatalysis. J. Alloys Compd. 2022, 894, 162409. [Google Scholar] [CrossRef]
  101. Cui, T.; Zhai, X.; Guo, L.; Chi, J.-Q.; Zhang, Y.; Zhu, J.; Sun, X.; Wang, L. Controllable Synthesis of a Self-Assembled Ultralow Ru, Ni-Doped Fe2O3 Lily as a Bifunctional Electrocatalyst for Large-Current-Density Alkaline Seawater Electrolysis. Chin. J. Catal. 2022, 43, 2202–2211. [Google Scholar] [CrossRef]
  102. Xiao, X.; Ma, R.; Li, D.; Yang, M.; Tong, Y.; Qiu, Z.; Jing, B.; Yang, T.; Hu, R.; Yang, Y.; et al. Tungsten Oxide Decorating-Regulated Iron-Nickel Oxide Heterostructure as Electrocatalyst for Oxygen Evolution Reaction. J. Alloys Compd. 2023, 968, 171910. [Google Scholar] [CrossRef]
  103. Tong, H.; Zheng, X.; Qi, M.; Li, D.; Zhu, J.; Jiang, D. Synergistically Coupled CoMo/Fe2O3 Electrocatalyst for Highly Efficient and Stable Overall Water Splitting. J. Colloid. Interface Sci. 2024, 676, 837–846. [Google Scholar] [CrossRef]
  104. Yang, X.; Li, Y.; Deng, L.; Li, W.; Ren, Z.; Yang, M.; Yang, X.; Zhu, Y. Synthesis and Characterization of an IrO2 –Fe2O3 Electrocatalyst for the Hydrogen Evolution Reaction in Acidic Water Electrolysis. RSC Adv. 2017, 7, 20252–20258. [Google Scholar] [CrossRef]
  105. Kim, J.; Heo, J.N.; Do, J.Y.; Chava, R.K.; Kang, M. Electrochemical Synergies of Heterostructured Fe2O3-MnO Catalyst for Oxygen Evolution Reaction in Alkaline Water Splitting. Nanomaterials 2019, 9, 1486. [Google Scholar] [CrossRef] [PubMed]
  106. Fang, Y.; Cao, X.; Zhou, W.; Li, Y.; Johnson, D.M.; Huang, Y. Catalytic Hydrolysis of Microcystin-LR Peptides on the Surface of Naturally Occurring Minerals. Res. Chem. Intermed. 2020, 46, 1141–1152. [Google Scholar] [CrossRef]
  107. Ramutsindela, F.K.; Okoye-Chine, C.G.; Mbuya, C.O.L.; Mubenesha, S.; Gorimbo, J.; Okonye, L.U.; Liu, X.; Hildebrandt, D. The Effect of Reducing Gases on Raw Iron Ore Catalyst for Fischer-Tropsch Synthesis. J. Taiwan. Inst. Chem. Eng. 2022, 131, 104163. [Google Scholar] [CrossRef]
  108. Liu, H.; Chen, T.; Xie, Q.; Zou, X.; Qing, C.; Frost, R.L. Kinetic Study of Goethite Dehydration and the Effect of Aluminium Substitution on the Dehydrate. Thermochim. Acta 2012, 545, 20–25. [Google Scholar] [CrossRef]
  109. Liu, H.; Chen, T.; Frost, R.L. An Overview of the Role of Goethite Surfaces in the Environment. Chemosphere 2014, 103, 1–11. [Google Scholar] [CrossRef]
  110. Strangway, D.W.; Honea, R.M.; McMahon, B.E.; Larson, E.E. The Magnetic Properties of Naturally Occurring Goethite. Geophys. J. Int. 1968, 15, 345–359. [Google Scholar] [CrossRef]
  111. Valezi, D.F.; Carneiro, C.E.A.; Costa, A.C.S.; Paesano, A.; Spadotto, J.C.; Solórzano, I.G.; Londoño, O.M.; Di Mauro, E. Weak Ferromagnetic Component in Goethite (α-FeOOH) and Its Relation with Microstructural Characteristics. Mater. Chem. Phys. 2020, 246, 122851. [Google Scholar] [CrossRef]
  112. Han, S.K.; Hwang, T.-M.; Yoon, Y.; Kang, J.-W. Evidence of Singlet Oxygen and Hydroxyl Radical Formation in Aqueous Goethite Suspension Using Spin-Trapping Electron Paramagnetic Resonance (EPR). Chemosphere 2011, 84, 1095–1101. [Google Scholar] [CrossRef]
  113. Shindo, H.; Huang, P.M. Catalytic Effects of Manganese (IV), Iron(III), Aluminum, and Silicon Oxides on the Formation of Phenolic Polymers. Soil. Sci. Soc. Am. J. 1984, 48, 927–934. [Google Scholar] [CrossRef]
  114. Cunningham, K.M.; Goldberg, M.C.; Weiner, E.R. Mechanisms for Aqueous Photolysis of Adsorbed Benzoate, Oxalate, and Succinate on Iron Oxyhydroxide (Goethite) Surfaces. Environ. Sci. Technol. 1988, 22, 1090–1097. [Google Scholar] [CrossRef]
  115. Lin, K.; Ding, J.; Wang, H.; Huang, X.; Gan, J. Goethite-Mediated Transformation of Bisphenol A. Chemosphere 2012, 89, 789–795. [Google Scholar] [CrossRef] [PubMed]
  116. Lu, M.-C. Oxidation of Chlorophenols with Hydrogen Peroxide in the Presence of Goethite. Chemosphere 2000, 40, 125–130. [Google Scholar] [CrossRef] [PubMed]
  117. Ortiz de la Plata, G.B.; Alfano, O.M.; Cassano, A.E. Optical Properties of Goethite Catalyst for Heterogeneous Photo-Fenton ReactionsComparison with a Titanium Dioxide Catalyst. Chem. Eng. J. 2008, 137, 396–410. [Google Scholar] [CrossRef]
  118. He, J.; Ma, W.; Song, W.; Zhao, J.; Qian, X.; Zhang, S.; Yu, J.C. Photoreaction of Aromatic Compounds at α-FeOOH/H2O Interface in the Presence of H2O2: Evidence for Organic-Goethite Surface Complex Formation. Water Res. 2005, 39, 119–128. [Google Scholar] [CrossRef]
  119. Yeh, C.K.-J.; Hsu, C.-Y.; Chiu, C.-H.; Huang, K.-L. Reaction Efficiencies and Rate Constants for the Goethite-Catalyzed Fenton-like Reaction of NAPL-Form Aromatic Hydrocarbons and Chloroethylenes. J. Hazard. Mater. 2008, 151, 562–569. [Google Scholar] [CrossRef]
  120. Lin, Z.-R.; Zhao, L.; Dong, Y.-H. Quantitative Characterization of Hydroxyl Radical Generation in a Goethite-Catalyzed Fenton-like Reaction. Chemosphere 2015, 141, 7–12. [Google Scholar] [CrossRef]
  121. Andersen, S.L.F.; Flores, R.G.; Madeira, V.S.; José, H.J.; Moreira, R.F.P.M. Synthesis and Characterization of Acicular Iron Oxide Particles Obtained from Acid Mine Drainage and Their Catalytic Properties in Toluene Oxidation. Ind. Eng. Chem. Res. 2012, 51, 767–774. [Google Scholar] [CrossRef]
  122. Aziz, F.; El Achaby, M.; Aziz, K.; Ouazzani, N.; Mandi, L.; Ghazzal, M.N. Nanocomposite Fiber Based on Natural Material for Water Disinfection under Visible Light Irradiation. Nanomaterials 2020, 10, 1192. [Google Scholar] [CrossRef]
  123. Kuang, J.; Guo, H.; Si, Q.; Guo, W.; Ma, F. Highly Dispersed Natural Goethite/Fe3O4 Heterogeneous Fenton-like Catalyst for Efficient Tetracycline Degradation: Rapid Accumulation of Hydroxyl Radicals by Fe–O–Si Bonds. J. Clean. Prod. 2023, 420, 138262. [Google Scholar] [CrossRef]
  124. Muruganandham, M.; Wu, J.J. Granular α-FeOOH—A Stable and Efficient Catalyst for the Decomposition of Dissolved Ozone in Water. Catal. Commun. 2007, 8, 668–672. [Google Scholar] [CrossRef]
  125. Zhang, T.; Li, C.; Ma, J.; Tian, H.; Qiang, Z. Surface Hydroxyl Groups of Synthetic α-FeOOH in Promoting OH Generation from Aqueous Ozone: Property and Activity Relationship. Appl. Catal. B 2008, 82, 131–137. [Google Scholar] [CrossRef]
  126. Li, X.; Fu, L.; Chen, F.; Zhao, S.; Zhu, J.; Yin, C. Application of Heterogeneous Catalytic Ozonation in Wastewater Treatment: An Overview. Catalysts 2023, 13, 342. [Google Scholar] [CrossRef]
  127. Zhang, Y.; Yang, J.; Du, J.; Xing, B. Goethite Catalyzed Cr(VI) Reduction by Tartaric Acid via Surface Adsorption. Ecotoxicol. Environ. Saf. 2019, 171, 594–599. [Google Scholar] [CrossRef]
  128. Zhang, T.; Wang, T.; Wang, W.; Liu, B.; Li, W.; Liu, Y. Reduction and Stabilization of Cr(VI) in Soil by Using Calcium Polysulfide: Catalysis of Natural Iron Oxides. Environ. Res. 2020, 190, 109992. [Google Scholar] [CrossRef]
  129. Pelalak, R.; Alizadeh, R.; Ghareshabani, E. Enhanced Heterogeneous Catalytic Ozonation of Pharmaceutical Pollutants Using a Novel Nanostructure of Iron-Based Mineral Prepared via Plasma Technology: A Comparative Study. J. Hazard. Mater. 2020, 392, 122269. [Google Scholar] [CrossRef]
  130. Fernández-Marchante, C.M.; Raschitor, A.; Mena, I.F.; Rodrigo, M.A.; Lobato, J. Evaluation of Goethite as a Catalyst for the Thermal Stage of the Westinghouse Process for Hydrogen Production. Catalysts 2021, 11, 1145. [Google Scholar] [CrossRef]
  131. Shehzad, M.A.; Yasmin, A.; Ge, X.; Ge, Z.; Zhang, K.; Liang, X.; Zhang, J.; Li, G.; Xiao, X.; Jiang, B.; et al. Shielded Goethite Catalyst That Enables Fast Water Dissociation in Bipolar Membranes. Nat. Commun. 2021, 12, 9. [Google Scholar] [CrossRef]
  132. Hu, Z.; Ai, Y.; Liu, L.; Zhou, J.; Zhang, G.; Liu, H.; Liu, X.; Liu, Z.; Hu, J.; Sun, H.; et al. Hydroxyl Assisted Rhodium Catalyst Supported on Goethite Nanoflower for Chemoselective Catalytic Transfer Hydrogenation of Fully Converted Nitrostyrenes. Adv. Synth. Catal. 2019, 361, 3146–3154. [Google Scholar] [CrossRef]
  133. Kuncser, A.C.; Vlaicu, I.D.; Pavel, O.D.; Zavoianu, R.; Badea, M.; Radu, D.; Culita, D.C.; Rostas, A.M.; Olar, R. Soft Synthesis and Characterization of Goethite-Based Nanocomposites as Promising Cyclooctene Oxidation Catalysts. RSC Adv. 2021, 11, 27589–27602. [Google Scholar] [CrossRef]
  134. Jadhav, D.A.; Ghadge, A.N.; Ghangrekar, M.M. Enhancing the Power Generation in Microbial Fuel Cells with Effective Utilization of Goethite Recovered from Mining Mud as Anodic Catalyst. Bioresour. Technol. 2015, 191, 110–116. [Google Scholar] [CrossRef]
  135. Zhao, H.; Li, Y.; Song, Q.; Liu, S.; Ma, Q.; Ma, L.; Shu, X. Catalytic Reforming of Volatiles from Co-Pyrolysis of Lignite Blended with Corn Straw over Three Different Structures of Iron Ores. J. Anal. Appl. Pyrolysis 2019, 144, 104714. [Google Scholar] [CrossRef]
  136. Verma, S.; Baig, R.B.N.; Nadagouda, M.N.; Varma, R.S. Oxidative C-H Activation of Amines Using Protuberant Lychee-like Goethite. Sci. Rep. 2018, 8, 2024. [Google Scholar] [CrossRef] [PubMed]
  137. Zheng, X.; Wu, B.; Zhou, C.; Liu, T.; Wang, Y.; Zhao, G.; Chen, B.; Chu, C. Sunlight-Driven Production of Reactive Oxygen Species from Natural Iron Minerals: Quantum Yield and Wavelength Dependence. Environ. Sci. Technol. 2023, 57, 1177–1185. [Google Scholar] [CrossRef]
  138. Liu, X.; Shaw, J.; Jiang, J.; Bloemendal, J.; Hesse, P.; Rolph, T.; Mao, X. Analysis on Variety and Characteristics of Maghemite. Sci. China Earth Sci. 2010, 53, 1153–1162. [Google Scholar] [CrossRef]
  139. Chen, Y.H. Thermal Properties of Nanocrystalline Goethite, Magnetite, and Maghemite. J. Alloys Compd. 2013, 553, 194–198. [Google Scholar] [CrossRef]
  140. Kim, W.; Suh, C.-Y.; Cho, S.-W.; Roh, K.-M.; Kwon, H.; Song, K.; Shon, I.-J. A New Method for the Identification and Quantification of Magnetite–Maghemite Mixture Using Conventional X-Ray Diffraction Technique. Talanta 2012, 94, 348–352. [Google Scholar] [CrossRef]
  141. Yamashita, T.; Hayes, P. Analysis of XPS Spectra of Fe2+ and Fe3+ Ions in Oxide Materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  142. Gnanaprakash, G.; Mahadevan, S.; Jayakumar, T.; Kalyanasundaram, P.; Philip, J.; Raj, B. Effect of Initial PH and Temperature of Iron Salt Solutions on Formation of Magnetite Nanoparticles. Mater. Chem. Phys. 2007, 103, 168–175. [Google Scholar] [CrossRef]
  143. Murad, E. Properties and Behavior of Iron Oxides as Determined by Mössbauer Spectroscopy. In Iron in Soils and Clay Minerals; Springer: Dordrecht, The Netherlands, 1988; pp. 309–350. [Google Scholar]
  144. Dunlop, D.J.; Özdemir, Ö. Magnetizations in Rocks and Minerals. In Treatise on Geophysics; Elsevier: Amsterdam, The Netherlands, 2007; pp. 277–336. [Google Scholar]
  145. Nikiforov, V.N.; Goldt, A.E.; Gudilin, E.A.; Sredin, V.G.; Irhin, V.Y. Magnetic Properties of Maghemite Nanoparticles. Bull. Russ. Acad. Sci. Phys. 2014, 78, 1075–1080. [Google Scholar] [CrossRef]
  146. Grau-Crespo, R.; Al-Baitai, A.Y.; Saadoune, I.; De Leeuw, N.H. Vacancy Ordering and Electronic Structure of γ-Fe2O3(Maghemite): A Theoretical Investigation. J. Phys. Condens. Matter 2010, 22, 255401–255408. [Google Scholar] [CrossRef]
  147. Ramirez, A.; Ould-Chikh, S.; Gevers, L.; Chowdhury, A.D.; Abou-Hamad, E.; Aguilar-Tapia, A.; Hazemann, J.; Wehbe, N.; Al Abdulghani, A.J.; Kozlov, S.M.; et al. Tandem Conversion of CO2 to Valuable Hydrocarbons in Highly Concentrated Potassium Iron Catalysts. ChemCatChem 2019, 11, 2879–2886. [Google Scholar] [CrossRef]
  148. Xu, D.; Lai, X.; Guo, W.; Dai, P. Microwave-Assisted Catalytic Degradation of Methyl Orange in Aqueous Solution by Ferrihydrite/Maghemite Nanoparticles. J. Water Process Eng. 2017, 16, 270–276. [Google Scholar] [CrossRef]
  149. Vasanthakumar, P.; Karvembu, R. Unmodified Maghemite from River Sand as a Selective Catalyst for Base-Free Transfer Hydrogenation of Furfural, Levulinic Acid, and o-Vanillin: A Pathway for Sustainable Biomass Conversions. ACS Sustain. Chem. Eng. 2020, 8, 17069–17078. [Google Scholar] [CrossRef]
  150. Filip, J.; Zboril, R.; Schneeweiss, O.; Zeman, J.; Cernik, M.; Kvapil, P.; Otyepka, M. Environmental Applications of Chemically Pure Natural Ferrihydrite. Environ. Sci. Technol. 2007, 41, 4367–4374. [Google Scholar] [CrossRef]
  151. Jojoa-Sierra, S.D.; Herrero-Albillos, J.; Ormad, M.P.; Serna-Galvis, E.A.; Torres-Palma, R.A.; Mosteo, R. Wüstite as a Catalyst Source for Water Remediation: Differentiated Antimicrobial Activity of by-Products, Action Routes of the Process, and Transformation of Fluoroquinolones. Chem. Eng. J. 2022, 435, 134850. [Google Scholar] [CrossRef]
  152. Husnain, N.; Wang, E.; Fareed, S. Low-Temperature Selective Catalytic Reduction of NO with NH3 over Natural Iron Ore Catalyst. Catalysts 2019, 9, 956. [Google Scholar] [CrossRef]
  153. Cabrera, A.F.; Rodríguez Torres, C.E.; Stewart, S.J. Nanostructured Al-doped maghemite: A low-cost and eco-friendly material tested for methylene blue removal from water and as an accelerator in ammonium nitrate decomposition. J. Phys. Chem. Solids 2024, 185, 111784. [Google Scholar] [CrossRef]
  154. Shopska, M.; Paneva, D.; Kolev, H.; Kadinov, G.; Briančin, J.; Fabián, M.; Cherkezova-Zheleva, Z.; Mitov, I. Characterization and Catalytic Activity in CO Oxidation of Biogenic Lepidocrocite Layered on Anodic Alumina. Catal. Today 2020, 357, 436–441. [Google Scholar] [CrossRef]
  155. Guyodo, Y.; Bonville, P.; Till, J.L.; Ona-Nguema, G.; Lagroix, F.; Menguy, N. Constraining the Origins of the Magnetism of Lepidocrocite (γ-FeOOH): A Mössbauer and Magnetization Study. Front. Earth Sci. 2016, 4, 28. [Google Scholar] [CrossRef]
  156. Zhukhlistov, A.P. Crystal Structure of Lepidocrocite FeO(OH) from the Electron-Diffractometry Data. Crystallogr. Rep. 2001, 46, 730–733. [Google Scholar] [CrossRef]
  157. Hirt, A.M.; Lanci, L.; Dobson, J.; Weidler, P.; Gehring, A.U. Low-temperature Magnetic Properties of Lepidocrocite. J. Geophys. Res. Solid. Earth 2002, 107, EPM-5. [Google Scholar] [CrossRef]
  158. Zhong, D.; Feng, W.; Ma, W.; Liu, X.; Ma, J.; Zhou, Z.; Du, X.; He, F. Goethite and Lepidocrocite Catalyzing Different Double-Oxidant Systems to Degrade Chlorophenol. Environ. Sci. Pollut. Res. 2022, 29, 72764–72776. [Google Scholar] [CrossRef]
  159. Matta, R.; Hanna, K.; Chiron, S. Fenton-like Oxidation of 2,4,6-Trinitrotoluene Using Different Iron Minerals. Sci. Total Environ. 2007, 385, 242–251. [Google Scholar] [CrossRef] [PubMed]
  160. Chou, S.; Huang, C. Application of a Supported Iron Oxyhydroxide Catalyst in Oxidation of Benzoic Acid by Hydrogen Peroxide. Chemosphere 1999, 38, 2719–2731. [Google Scholar] [CrossRef]
  161. Qin, M.; Lu, B.; Feng, S.; Zhen, Z.; Chen, R.; Liu, H. Role of Exposed Facets and Surface OH Groups in the Fenton-like Reactivity of Lepidocrocite Catalyst. Chemosphere 2019, 230, 286–293. [Google Scholar] [CrossRef]
  162. Chen, R.; Zhao, S.; Liu, H.; Song, X.; Wei, Y. Preparation and Photocatalytic Activity of Lepidocrocites Obtained Byphotocatalytic Oxidation of Fe(II) in the Presence of Citric Acid. J. Photochem. Photobiol. A Chem. 2015, 312, 73–80. [Google Scholar] [CrossRef]
  163. Lin, Y.; Wei, Y.; Sun, Y. Room-Temperature Synthesis and Photocatalytic Properties of Lepidocrocite by Monowavelength Visible Light Irradiation. J. Mol. Catal. A Chem. 2012, 353–354, 67–73. [Google Scholar] [CrossRef]
  164. Shi, H.; Liang, H.; Ming, F.; Wang, Z. Efficient Overall Water-Splitting Electrocatalysis Using Lepidocrocite VOOH Hollow Nanospheres. Angew. Chem. 2017, 129, 588–592. [Google Scholar] [CrossRef]
  165. Wang, L.; Giammar, D.E. Effects of PH, Dissolved Oxygen, and Aqueous Ferrous Iron on the Adsorption of Arsenic to Lepidocrocite. J. Colloid. Interface Sci. 2015, 448, 331–338. [Google Scholar] [CrossRef]
  166. He, F.; Zhong, D.; Ma, W.; Yuan, Y.; Li, K.; Dai, C. Activation of the Combined Hydrogen Peroxide and Peroxymonosulphate by Lepidocrocite for Chloramphenicol Removal: Kinetics and Mechanisms. Environ. Technol. 2023, 44, 1936–1946. [Google Scholar] [CrossRef]
  167. Liu, H.; Li, X.; Zhang, X.; Coulon, F.; Wang, C. Harnessing the Power of Natural Minerals: A Comprehensive Review of Their Application as Heterogeneous Catalysts in Advanced Oxidation Processes for Organic Pollutant Degradation. Chemosphere 2023, 337, 139404. [Google Scholar] [CrossRef] [PubMed]
  168. Huang, W.; Yuan, Y.; Zhong, D.; Zhang, P.; Liangdy, A.; Lim, T.-T.; Ma, W.; Yuan, Y. Catalytic Activity of H2O2 by Goethite and Lepidocrocite: Insight from 5-Bromosalicylic Acid Removal Mechanism and Density Functional Theory Calculation (ID:CHEM114760). Chemosphere 2023, 329, 138551. [Google Scholar] [CrossRef] [PubMed]
  169. Childs, C.W. Ferrihydrite: A Review of Structure, Properties and Occurrence in Relation to Soils. Z. Pflanzenernährung Bodenkd. 1992, 155, 441–448. [Google Scholar] [CrossRef]
  170. Cismasu, A.C.; Michel, F.M.; Tcaciuc, A.P.; Tyliszczak, T.; Brown, J.G.E. Composition and Structural Aspects of Naturally Occurring Ferrihydrite. Comptes Rendus. Géoscience 2011, 343, 210–218. [Google Scholar] [CrossRef]
  171. Fleischer, M.; Chao, G.Y.; Cabri, L.J. New Minerals Names. Am. Mineral. J. Earth Planet. Mater. 1975, 60, 736–739. [Google Scholar]
  172. Eggleton, R.A.; Fitzpatrick, R.W. New Data and a Revised Structural Model for Ferrihydrite. Clays Clay Min. 1988, 36, 111–124. [Google Scholar] [CrossRef]
  173. Cardile, C.M. Tetrahedral Fe3+ in Ferrihydrite: 57 Fe Mössbauer Spectroscopic Evidence. Clays Clay Min. 1988, 36, 537–539. [Google Scholar] [CrossRef]
  174. Towe, K.M.; Bradley, W.F. Mineralogical Constitution of Colloidal “Hydrous Ferric Oxides”. J. Colloid. Interface Sci. 1967, 24, 384–392. [Google Scholar] [CrossRef]
  175. Russell, J.D. Infrared Spectroscopy of Ferrihydrite: Evidence for the Presence of Structural Hydroxyl Groups. Clay Min. 1979, 14, 109–114. [Google Scholar] [CrossRef]
  176. Barrón, V.; Torrent, J. Diffuse Reflectance Spectroscopy of Iron Oxides. Encycl. Surf. Colloid. Sci. 2002, 1, 1438–1446. [Google Scholar]
  177. Hiemstra, T. Surface and Mineral Structure of Ferrihydrite. Geochim. Cosmochim. Acta 2013, 105, 316–325. [Google Scholar] [CrossRef]
  178. Funnell, N.P.; Fulford, M.F.; Inoué, S.; Kletetschka, K.; Michel, F.M.; Goodwin, A.L. Nanocomposite Structure of Two-Line Ferrihydrite Powder from Total Scattering. Commun. Chem. 2020, 3, 22. [Google Scholar] [CrossRef] [PubMed]
  179. Boily, J.-F.; Song, X. Direct Identification of Reaction Sites on Ferrihydrite. Commun. Chem. 2020, 3, 79. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, X.; Zhu, M.; Koopal, L.K.; Li, W.; Xu, W.; Liu, F.; Zhang, J.; Liu, Q.; Feng, X.; Sparks, D.L. Effects of Crystallite Size on the Structure and Magnetism of Ferrihydrite. Environ. Sci. Nano 2016, 3, 190–202. [Google Scholar] [CrossRef]
  181. Zhu, Y.; Xie, Q.; Deng, F.; Ni, Z.; Lin, Q.; Cheng, L.; Chen, X.; Qiu, R.; Zhu, R. The Differences in Heterogeneous Fenton Catalytic Performance and Mechanism of Various Iron Minerals and Their Influencing Factors: A Review. Sep. Purif. Technol. 2023, 325, 124702. [Google Scholar] [CrossRef]
  182. Seomoon, K. On-Line GC and GC–MS Analyses of the Fischer–Tropsch Products Synthesized Using Ferrihydrite Catalyst. J. Ind. Eng. Chem. 2013, 19, 2108–2114. [Google Scholar] [CrossRef]
  183. Bali, S.; Bali, G.; Huggins, F.E.; Seehra, M.S.; Singh, V.; Hancock, J.M.; Harrison, R.; Huffman, G.P.; Pugmire, R.J.; Ernst, R.D.; et al. Synthetic Doped Amorphous Ferrihydrite for the Fischer–Tropsch Synthesis of Alternative Fuels. Ind. Eng. Chem. Res. 2012, 51, 4515–4522. [Google Scholar] [CrossRef]
  184. Trivedi, P.; Dyer, J.A.; Sparks, D.L. Lead Sorption onto Ferrihydrite. 1. A Macroscopic and Spectroscopic Assessment. Environ. Sci. Technol. 2003, 37, 908–914. [Google Scholar] [CrossRef]
  185. Scheinost, A.C.; Abend, S.; Pandya, K.I.; Sparks, D.L. Kinetic Controls on Cu and Pb Sorption by Ferrihydrite. Environ. Sci. Technol. 2001, 35, 1090–1096. [Google Scholar] [CrossRef]
  186. Brinza, L.; Vu, H.P.; Neamtu, M.; Benning, L.G. Experimental and Simulation Results of the Adsorption of Mo and V onto Ferrihydrite. Sci. Rep. 2019, 9, 1365. [Google Scholar] [CrossRef]
  187. Johnston, C.P.; Chrysochoou, M. Mechanisms of Chromate, Selenate, and Sulfate Adsorption on Al-Substituted Ferrihydrite: Implications for Ferrihydrite Surface Structure and Reactivity. Environ. Sci. Technol. 2016, 50, 3589–3596. [Google Scholar] [CrossRef] [PubMed]
  188. Mamun, A.A.; Morita, M.; Matsuoka, M.; Tokoro, C. Sorption Mechanisms of Chromate with Coprecipitated Ferrihydrite in Aqueous Solution. J. Hazard. Mater. 2017, 334, 142–149. [Google Scholar] [CrossRef] [PubMed]
  189. Trivedi, P.; Dyer, J.A.; Sparks, D.L.; Pandya, K. Mechanistic and Thermodynamic Interpretations of Zinc Sorption onto Ferrihydrite. J. Colloid. Interface Sci. 2004, 270, 77–85. [Google Scholar] [CrossRef]
  190. Sajih, M.; Bryan, N.D.; Livens, F.R.; Vaughan, D.J.; Descostes, M.; Phrommavanh, V.; Nos, J.; Morris, K. Adsorption of Radium and Barium on Goethite and Ferrihydrite: A Kinetic and Surface Complexation Modelling Study. Geochim. Cosmochim. Acta 2014, 146, 150–163. [Google Scholar] [CrossRef]
  191. Rojo, I.; Seco, F.; Rovira, M.; Giménez, J.; Cervantes, G.; Martí, V.; de Pablo, J. Thorium Sorption onto Magnetite and Ferrihydrite in Acidic Conditions. J. Nucl. Mater. 2009, 385, 474–478. [Google Scholar] [CrossRef]
  192. Smith, K.F.; Morris, K.; Law, G.T.W.; Winstanley, E.H.; Livens, F.R.; Weatherill, J.S.; Abrahamsen-Mills, L.G.; Bryan, N.D.; Mosselmans, J.F.W.; Cibin, G.; et al. Plutonium(IV) Sorption during Ferrihydrite Nanoparticle Formation. ACS Earth Space Chem. 2019, 3, 2437–2442. [Google Scholar] [CrossRef]
  193. Winstanley, E.H.; Morris, K.; Abrahamsen-Mills, L.G.; Blackham, R.; Shaw, S. U(VI) Sorption during Ferrihydrite Formation: Underpinning Radioactive Effluent Treatment. J. Hazard. Mater. 2019, 366, 98–104. [Google Scholar] [CrossRef]
  194. Mendez, J.C.; Hiemstra, T. Carbonate Adsorption to Ferrihydrite: Competitive Interaction with Phosphate for Use in Soil Systems. ACS Earth Space Chem. 2019, 3, 129–141. [Google Scholar] [CrossRef]
  195. Tiberg, C.; Gustafsson, J.P. Phosphate Effects on Cadmium(II) Sorption to Ferrihydrite. J. Colloid. Interface Sci. 2016, 471, 103–111. [Google Scholar] [CrossRef]
  196. Zhu, J.; Pigna, M.; Cozzolino, V.; Caporale, A.G.; Violante, A. Sorption of Arsenite and Arsenate on Ferrihydrite: Effect of Organic and Inorganic Ligands. J. Hazard. Mater. 2011, 189, 564–571. [Google Scholar] [CrossRef]
  197. Mao, H.; Shu, J.; Fei, Y.; Hu, J.; Hemley, R.J. The Wüstite Enigma. Phys. Earth Planet. Inter. 1996, 96, 135–145. [Google Scholar] [CrossRef]
  198. Schrettle, F.; Kant, C.; Lunkenheimer, P.; Mayr, F.; Deisenhofer, J.; Loidl, A. Wüstite: Electric, Thermodynamic and Optical Properties of FeO. Eur. Phys. J. B 2012, 85, 164. [Google Scholar] [CrossRef]
  199. Yin, M.; Chen, Z.; Deegan, B.; O’Brien, S. Wüstite Nanocrystals: Synthesis, Structure and Superlattice Formation. J. Mater. Res. 2007, 22, 1987–1995. [Google Scholar] [CrossRef]
  200. Fjellvag, H.; Hauback, B.; Vogt, T.; Stolen, S. Monoclinic Nearly Stoichiometric Wüstite at Low Temperatures. Am. Mineral. 2002, 87, 347–349. [Google Scholar] [CrossRef]
  201. Liu, H.; Han, W.; Huo, C.; Cen, Y. Development and Application of Wüstite-Based Ammonia Synthesis Catalysts. Catal. Today 2020, 355, 110–127. [Google Scholar] [CrossRef]
  202. Humphreys, J.; Lan, R.; Tao, S. Development and Recent Progress on Ammonia Synthesis Catalysts for Haber–Bosch Process. Adv. Energy Sustain. Res. 2021, 2, 2000043. [Google Scholar] [CrossRef]
  203. Pernicone, N. Wustite as a New Precursor of Industrial Ammonia Synthesis Catalysts. Appl. Catal. A Gen. 2003, 251, 121–129. [Google Scholar] [CrossRef]
  204. Rossetti, I.; Pernicone, N.; Forni, L. Promoters Effect in Ru/C Ammonia Synthesis Catalyst. Appl. Catal. A Gen. 2001, 208, 271–278. [Google Scholar] [CrossRef]
  205. Rakshit, S.; Matocha, C.J.; Haszler, G.R. Nitrate Reduction in the Presence of Wüstite. J. Environ. Qual. 2005, 34, 1286–1292. [Google Scholar] [CrossRef]
  206. Poza-Nogueiras, V.; Rosales, E.; Pazos, M.; Sanromán, M.Á. Current Advances and Trends in Electro-Fenton Process Using Heterogeneous Catalysts—A Review. Chemosphere 2018, 201, 399–416. [Google Scholar] [CrossRef]
  207. Expoósito, E.; Saánchez-Saánchez, C.M.; Montiel, V. Mineral Iron Oxides as Iron Source in Electro-Fenton and Photoelectro-Fenton Mineralization Processes. J. Electrochem. Soc. 2007, 154, E116. [Google Scholar] [CrossRef]
  208. Ren, J.; Yao, Z.; Wei, Q.; Wang, R.; Wang, L.; Liu, Y.; Ren, Z.; Guo, H.; Niu, Z.; Wang, J.; et al. Catalytic Degradation of Chloramphenicol by Water Falling Film Dielectric Barrier Discharge and FeO Catalyst. Sep. Purif. Technol. 2022, 290, 120826. [Google Scholar] [CrossRef]
  209. Mackay, A.L. β-Ferric Oxyhydroxide—Akaganéite. Mineral. Mag. J. Mineral. Soc. 1962, 33, 270–280. [Google Scholar] [CrossRef]
  210. Jegdić, B.V.; Ristić, S.S.; Polić-Radovanović, S.R.; Alil, A.B. Corrosion and Conservation of Weapons and Military Equipment. Vojnoteh. Glas. 2012, 60, 169–182. [Google Scholar] [CrossRef]
  211. Holm, N.G.; Dowler, M.J.; Wadsten, T.; Arrhenius, G. β-FeOOH · Cln (Akaganéite) and Fe1-XO (Wüstite) in Hot Brine from the Atlantis II Deep (Red Sea) and the Uptake of Amino Acids by Synthetic β-FeOOH · Cln. Geochim. Cosmochim. Acta 1983, 47, 1465–1470. [Google Scholar] [CrossRef]
  212. Post, J.F.; Buchwald, V.F. Crystal Structure Refinement of Akaganeite. Am. Mineral. 1991, 76, 272–277. [Google Scholar]
  213. Tadic, M.; Milosevic, I.; Kralj, S.; Saboungi, M.-L.; Motte, L. Ferromagnetic Behavior and Exchange Bias Effect in Akaganeite Nanorods. Appl. Phys. Lett. 2015, 106, 183706. [Google Scholar] [CrossRef]
  214. Zhao, J.; Lin, W.; Chang, Q.; Li, W.; Lai, Y. Adsorptive Characteristics of Akaganeite and Its Environmental Applications: A Review. Environ. Technol. Rev. 2012, 1, 114–126. [Google Scholar] [CrossRef]
  215. Marsac, R.; Martin, S.; Boily, J.-F.; Hanna, K. Oxolinic Acid Binding at Goethite and Akaganéite Surfaces: Experimental Study and Modeling. Environ. Sci. Technol. 2016, 50, 660–668. [Google Scholar] [CrossRef]
  216. Xiong, H.; Shen, X.; Cui, C.; Cheng, L.; Xu, Y. Synthesis of Nanospindle-Akaganéite and Its Photocatalytic Degradation for Methyl Orange. Water Sci. Technol. 2020, 82, 481–491. [Google Scholar] [CrossRef]
  217. Xiong, H.; Shi, K.; Han, J.; Cui, C.; Liu, Y.; Zhang, B. Synthesis of β-FeOOH/Polyaniline Heterogeneous Catalyst for Efficient Photo-Fenton Degradation of AOII Dye. Environ. Sci. Pollut. Res. 2023, 30, 59366–59381. [Google Scholar] [CrossRef] [PubMed]
  218. Su, S.; Liu, Y.; Liu, X.; Jin, W.; Zhao, Y. Transformation Pathway and Degradation Mechanism of Methylene Blue through β-FeOOH@GO Catalyzed Photo-Fenton-like System. Chemosphere 2019, 218, 83–92. [Google Scholar] [CrossRef] [PubMed]
  219. Majzlan, J.; Koch, C.B.; Navrotsky, A. Thermodynamic Properties of Feroxyhyte (Δ′-FeOOH). Clays Clay Min. 2008, 56, 526–530. [Google Scholar] [CrossRef]
  220. Vodyanitskii, Y.N. Iron Hydroxides in Soils: A Review of Publications. Eurasian Soil. Sci. 2010, 43, 1244–1254. [Google Scholar] [CrossRef]
  221. Li, Y.; Fu, F.; Cai, W.; Tang, B. Synergistic Effect of Mesoporous Feroxyhyte Nanoparticles and Fe(II) on Phosphate Immobilization: Adsorption and Chemical Precipitation. Powder Technol. 2019, 345, 786–795. [Google Scholar] [CrossRef]
  222. Xu, Z.; Yu, Y.; Fang, D.; Xu, J.; Liang, J.; Zhou, L. Microwave–ultrasound assisted synthesis of β-FeOOH and its catalytic property in a photo-Fenton-like process. Ultrason. Sonochem. 2015, 27, 287–295. [Google Scholar] [CrossRef]
  223. Liu, B.; Wang, Y.; Peng, H.Q.; Yang, R.; Jiang, Z.; Zhou, X.; Lee, C.-S.; Zhao, H.; Zhang, W. Iron vacancies induced bifunctionality in ultrathin feroxyhyte nanosheets for overall water splitting. Adv. Mater. 2018, 30, 1803144. [Google Scholar] [CrossRef]
  224. Kokkinos, E.; Soukakos, K.; Kostoglou, M.; Mitrakas, M. Cadmium, mercury, and nickel adsorption by tetravalent manganese feroxyhyte: Selectivity, kinetic modeling, and thermodynamic study. Environ. Sci. Pollut. Res. 2018, 25, 12263–12273. [Google Scholar] [CrossRef]
  225. Du, W.; Xu, Y.; Wang, Y. Photoinduced Degradation of Orange II on Different Iron (Hydr)Oxides in Aqueous Suspension: Rate Enhancement on Addition of Hydrogen Peroxide, Silver Nitrate, and Sodium Fluoride. Langmuir 2008, 24, 175–181. [Google Scholar] [CrossRef]
  226. Pinto, I.S.X.; Pacheco, P.H.V.V.; Coelho, J.V.; Lorençon, E.; Ardisson, J.D.; Fabris, J.D.; de Souza, P.P.; Krambrock, K.W.H.; Oliveira, L.C.A.; Pereira, M.C. Nanostructured δ-FeOOH: An Efficient Fenton-like Catalyst for the Oxidation of Organics in Water. Appl. Catal. B 2012, 119–120, 175–182. [Google Scholar] [CrossRef]
  227. Li, J.; Ding, Y.; Chen, K.; Li, Z.; Yang, H.; Yue, S.; Tang, Y.; Wang, Q. δ-FeOOH Coupled BiOBr0.5I0.5 for Efficient Photocatalysis-Fenton Synergistic Degradation of Organic Pollutants. J. Alloys Compd. 2022, 903, 163795. [Google Scholar] [CrossRef]
  228. Wang, C.; Shi, P.; Wang, Z.; Guo, R.; You, J.; Zhang, H. Efficient Wastewater Disinfection through FeOOH-Mediated Photo-Fenton Reaction: A Review. J. Environ. Chem. Eng. 2023, 11, 111269. [Google Scholar] [CrossRef]
  229. Pereira, M.C.; Garcia, E.M.; Cândido Da Silva, A.; Lorenon, E.; Ardisson, J.D.; Murad, E.; Fabris, J.D.; Matencio, T.; De Castro Ramalho, T.; Rocha, M.V.J. Nanostructured δ-FeOOH: A Novel Photocatalyst for Water Splitting. J. Mater. Chem. 2011, 21, 10280–10282. [Google Scholar] [CrossRef]
  230. da Silva Rocha, T.; Nascimento, E.S.; da Silva, A.C.; dos Santos Oliveira, H.; Garcia, E.M.; de Oliveira, L.C.A.; Monteiro, D.S.; Rodriguez, M.; Pereira, M.C. Enhanced Photocatalytic Hydrogen Generation from Water by Ni(OH)2 Loaded on Ni-Doped δ-FeOOH Nanoparticles Obtained by One-Step Synthesis. RSC Adv. 2013, 3, 20308. [Google Scholar] [CrossRef]
  231. Hu, J.; Lo, I.; Chen, G. Performance and Mechanism of Chromate (VI) Adsorption by δ-FeOOH-Coated Maghemite (γ-Fe2O3) Nanoparticles. Sep. Purif. Technol. 2007, 58, 76–82. [Google Scholar] [CrossRef]
  232. Tresintsi, S.; Simeonidis, K.; Estradé, S.; Martinez-Boubeta, C.; Vourlias, G.; Pinakidou, F.; Katsikini, M.; Paloura, E.C.; Stavropoulos, G.; Mitrakas, M. Tetravalent Manganese Feroxyhyte: A Novel Nanoadsorbent Equally Selective for As(III) and As(V) Removal from Drinking Water. Environ. Sci. Technol. 2013, 47, 9699–9705. [Google Scholar] [CrossRef]
  233. Pinakidou, F.; Katsikini, M.; Paloura, E.C.; Simeonidis, K.; Mitraka, E.; Mitrakas, M. Monitoring the Role of Mn and Fe in the As-Removal Efficiency of Tetravalent Manganese Feroxyhyte Nanoparticles from Drinking Water: An X-Ray Absorption Spectroscopy Study. J. Colloid. Interface Sci. 2016, 477, 148–155. [Google Scholar] [CrossRef]
  234. Kokkinos, E.; Kellartzis, I.; Diamantopoulou, I.; Stavropoulos, G.; Vourlias, G.; Mitrakas, M. Study of Elemental Mercury Removal from Flue Gases Using Tetravalent Manganese Feroxyhyte. Chem. Eng. J. 2017, 315, 152–158. [Google Scholar] [CrossRef]
  235. Chagas, P.; Da Silva, A.C.; Passamani, E.C.; Ardisson, J.D.; De Oliveira, L.C.A.; Fabris, J.D.; Paniago, R.M.; Monteiro, D.S.; Pereira, M.C. δ-FeOOH: A Superparamagnetic Material for Controlled Heat Release under AC Magnetic Field. J. Nanoparticle Res. 2013, 15, 1544. [Google Scholar] [CrossRef]
  236. Lacerda, L.C.T.; Pires, M.S.; Oliveira, I.S.S.; Silva, T.C.; de Castro, A.A.; Corrêa, S.; Vaiss, V.S.; Ramalho, T.C. Bulk and Surface Theoretical Investigation of Nb-Doped δ-FeOOH as a Promising Bifunctional Catalyst. J. Mol. Model. 2021, 27, 249. [Google Scholar] [CrossRef]
  237. Kolitsch, U. Bernalite from the Clara Mine, Germany, and the Incorporation of Tungsten in Minerals Containing Ferric Iron. Can. Mineral. 1998, 36, 1211–1216. [Google Scholar]
  238. McCammon, C.A.; Pring, A.; Keppler, H.; Sharp, T. A Study of Bernalite, Fe(OH)3, Using Mössbauer Spectroscopy, Optical Spectroscopy and Transmission Electron Microscopy. Phys. Chem. Min. 1995, 22, 11–20. [Google Scholar] [CrossRef]
  239. Welch, M.D.; Crichton, W.A.; Ross, N.L. Compression of the Perovskite-Related Mineral Bernalite Fe(OH) 3 to 9 GPa and a Reappraisal of Its Structure. Min. Mag. 2005, 69, 309–315. [Google Scholar] [CrossRef]
  240. McCammon, C.A.; De Grave, E.; Pring, A. The Magnetic Structure of Bernalite, Fe(OH)3. J. Magn. Magn. Mater. 1996, 152, 33–39. [Google Scholar] [CrossRef]
  241. Han, J.; Ro, H.-M. Identification of Bernalite Transformation and Tridentate Arsenate Complex at Nano-Goethite under Effects of Drying, PH and Surface Loading. Sci. Rep. 2018, 8, 8369. [Google Scholar] [CrossRef]
  242. Han, J.; Ro, H.-M. Interpreting competitive adsorption of arsenate and phosphate on nanosized iron (hydr)oxides: Effects of pH and surface loading. Environ. Sci. Pollut. Res. 2018, 25, 28572–28582. [Google Scholar] [CrossRef]
  243. Su-Gallegos, J.; Magallón-Cacho, L.; Borja-Arco, E.; García-Valdés, J.; Sebastian, P.J. Oxygen Reduction and Hydrogen Oxidation Reactions on Ru-Fe Electrocatalyst Synthesized by a Microwave-Assisted Synthesis. Mater. Res. Express 2020, 7, 035505. [Google Scholar] [CrossRef]
  244. Nur’aeni; Chae, A.; Jo, S.; Choi, Y.; Park, B.; Park, S.Y.; In, I. Synthesis of β-FeOOH/Fe3O4 Hybrid Photocatalyst Using Catechol-Quaternized Poly(N-Vinyl Pyrrolidone) as a Double-Sided Molecular Tape. J. Mater. Sci. 2017, 52, 8493–8501. [Google Scholar] [CrossRef]
  245. Sciscenko, I.; Mestre, S.; Climent, J.; Valero, F.; Escudero-Oñate, C.; Oller, I.; Arques, A. Magnetic Photocatalyst for Wastewater Tertiary Treatment at Pilot Plant Scale: Disinfection and Enrofloxacin Abatement. Water 2021, 13, 329. [Google Scholar] [CrossRef]
  246. Mahy, J.G.; Wolfs, C.; Vreuls, C.; Drot, S.; Dircks, S.; Boergers, A.; Tuerk, J.; Hermans, S.; Lambert, S.D. Advanced Oxidation Processes for Waste Water Treatment: From Laboratory-Scale Model Water to on-Site Real Waste Water. Environ. Technol. 2021, 42, 3974–3986. [Google Scholar] [CrossRef]
Figure 1. Scheme of main catalytic applications of Fe-OH from natural sources and natural iron species.
Figure 1. Scheme of main catalytic applications of Fe-OH from natural sources and natural iron species.
Catalysts 15 00236 g001
Figure 2. SEM micrographs of magnetite particles prepared from natural sources: (A) iron sand via ball milling, (B) raw ore after washing and heat treatment, and (C) mineral magnetite after magnetic separation and 20 min of ball milling. Reproduced with permission from References (A) [44], (B) [45], and (C) [46]. Copyright© 2020 IOPScience/CC BY 3.0, 2016 and 2021 Elsevier.
Figure 2. SEM micrographs of magnetite particles prepared from natural sources: (A) iron sand via ball milling, (B) raw ore after washing and heat treatment, and (C) mineral magnetite after magnetic separation and 20 min of ball milling. Reproduced with permission from References (A) [44], (B) [45], and (C) [46]. Copyright© 2020 IOPScience/CC BY 3.0, 2016 and 2021 Elsevier.
Catalysts 15 00236 g002
Figure 3. Removal as a function of magnetite catalyst concentration (A), pH (B), oxidant concentration (C), and specific surface area (SBET, (D)). Elaborated with data from References [1,46,50,52,55,62].
Figure 3. Removal as a function of magnetite catalyst concentration (A), pH (B), oxidant concentration (C), and specific surface area (SBET, (D)). Elaborated with data from References [1,46,50,52,55,62].
Catalysts 15 00236 g003
Figure 4. SEM micrographs of α-Fe2O3 nanoparticles prepared from natural sources: (A) natural hematite via purification and calcination and (B) iron ore via magnetic separation, ball milling, and calcination. Reproduced with permission from References (A) [73] and (B) [76]. Copyright© 2018 Elsevier and 2018 IOPScience/CC BY 3.0, respectively.
Figure 4. SEM micrographs of α-Fe2O3 nanoparticles prepared from natural sources: (A) natural hematite via purification and calcination and (B) iron ore via magnetic separation, ball milling, and calcination. Reproduced with permission from References (A) [73] and (B) [76]. Copyright© 2018 Elsevier and 2018 IOPScience/CC BY 3.0, respectively.
Catalysts 15 00236 g004
Figure 5. Removal as a function of hematite catalyst concentration (A), pH (B), oxidant concentration (C), and specific surface area (SBET, (D)). Elaborated with data from References [72,73,74,75].
Figure 5. Removal as a function of hematite catalyst concentration (A), pH (B), oxidant concentration (C), and specific surface area (SBET, (D)). Elaborated with data from References [72,73,74,75].
Catalysts 15 00236 g005
Figure 7. Micrographs of catalysts from natural sources of less abundant and rare oxides: (A) γ-Fe2O3 from river sand, (B) TEM image of ferrihydrite from AMD, and (C) wüstite from Probus reagents. Reproduced with permission from References (A) [149], (B) [150], and (C) [151]. Copyright© 2020, 2007 American Chemical Society and 2022 Elsevier, respectively.
Figure 7. Micrographs of catalysts from natural sources of less abundant and rare oxides: (A) γ-Fe2O3 from river sand, (B) TEM image of ferrihydrite from AMD, and (C) wüstite from Probus reagents. Reproduced with permission from References (A) [149], (B) [150], and (C) [151]. Copyright© 2020, 2007 American Chemical Society and 2022 Elsevier, respectively.
Catalysts 15 00236 g007
Figure 8. Schemes of proposed mechanism for different processes catalyzed by iron oxides and hydroxides from natural sources: (A) Photodegradation of phenol by α-Fe2O3 catalyst impregnated on gelcasting porous ceramic, (B) ozonation of ACT over HRS, (C) ozonation of SSZ over goethite, (D) Fenton-like oxidation of 2,4-D and MCPA over hematite, (E) photocatalytic reduction of 4 NP to 4 AP over hematite, (F) transfer hydrogenation of LA to form GVL over maghemite, and (G) waste cooking oil-based biodiesel production over α-Fe2O3/CaO2. Reproduced with permission from References (A) [78], (B) [81], (C) [129], (D) [73], (E) [77], (F) [149], and (G) [87]. Copyright© 2022 IIETA/CC BY 4.0, 2022, 2020, 2018, and 2020 Elsevier; 2020 American Chemical Society; and 2022 Springer Nature, respectively.
Figure 8. Schemes of proposed mechanism for different processes catalyzed by iron oxides and hydroxides from natural sources: (A) Photodegradation of phenol by α-Fe2O3 catalyst impregnated on gelcasting porous ceramic, (B) ozonation of ACT over HRS, (C) ozonation of SSZ over goethite, (D) Fenton-like oxidation of 2,4-D and MCPA over hematite, (E) photocatalytic reduction of 4 NP to 4 AP over hematite, (F) transfer hydrogenation of LA to form GVL over maghemite, and (G) waste cooking oil-based biodiesel production over α-Fe2O3/CaO2. Reproduced with permission from References (A) [78], (B) [81], (C) [129], (D) [73], (E) [77], (F) [149], and (G) [87]. Copyright© 2022 IIETA/CC BY 4.0, 2022, 2020, 2018, and 2020 Elsevier; 2020 American Chemical Society; and 2022 Springer Nature, respectively.
Catalysts 15 00236 g008
Figure 9. Schemes of reactors used in different processes catalyzed by Fe-OH from natural sources: (A) FBR packed with powdered magnetite catalyst used in the continuous CWPO experiments, (B) photocatalytic reactor continuous flow system for MB degradation, (C) photocatalytic reactor system for oxidation and mineralization of MO, (D) pyrolysis reactor to produce diesel-like fuel, (E) catalytic cracking of toluene, and (F) ozonation and peroxone-mediated removal of ACT. Reproduced with permission from References (A) [50], (B) [122], (C) [72], (D) [45], (E) [89], and (F) [81]. Copyright© 2024 Springer Nature, 2020 MDPI/CC BY, and 2020, 2016, 2016, and 2022 Elsevier, respectively.
Figure 9. Schemes of reactors used in different processes catalyzed by Fe-OH from natural sources: (A) FBR packed with powdered magnetite catalyst used in the continuous CWPO experiments, (B) photocatalytic reactor continuous flow system for MB degradation, (C) photocatalytic reactor system for oxidation and mineralization of MO, (D) pyrolysis reactor to produce diesel-like fuel, (E) catalytic cracking of toluene, and (F) ozonation and peroxone-mediated removal of ACT. Reproduced with permission from References (A) [50], (B) [122], (C) [72], (D) [45], (E) [89], and (F) [81]. Copyright© 2024 Springer Nature, 2020 MDPI/CC BY, and 2020, 2016, 2016, and 2022 Elsevier, respectively.
Catalysts 15 00236 g009
Table 1. Naturally occurring iron oxide minerals. Reproduced with permission from Reference [3]. Copyright© 2016, John Wiley and Sons.
Table 1. Naturally occurring iron oxide minerals. Reproduced with permission from Reference [3]. Copyright© 2016, John Wiley and Sons.
NameFormulaCrystal Symmetry
Magnetite (Mt)Fe3O4Cubic spinel
Hematite (Ht)α-Fe2O3Rhombohedral
Goethite (Gt)α-FeOOHOrthorhombic
Lepidocrociteγ-FeOOHOrthorhombic
Ferrihydrite (Fh)5Fe2O3⋅9H2OHexagonal
Maghemite (Mh)γ-Fe2O3Cubic spinel
WüstiteFeOIsometric-hexoctahedral
Akaganéiteβ-FeOOHMonoclinic
Feroxyhyteδ′-FeOOHHexagonal
BernaliteFe(OH)3Orthorhombic
Table 2. Characteristics of bulk Fe3O4 and Fe3O4 NPs. Reproduced with permission from Reference [36]. Copyright© 2023 MDPI/CC BY 4.0.
Table 2. Characteristics of bulk Fe3O4 and Fe3O4 NPs. Reproduced with permission from Reference [36]. Copyright© 2023 MDPI/CC BY 4.0.
CharacteristicsBulk Fe3O4Fe3O4 NPs
MagnetismFerromagnetismSuperparamagnetism or
ferrimagnetism
Saturation magnetization (Ms, 300 K, emu/g)~(84–100)Depend on size, shape, and coating:
~(0.5–92)
Size controllabilityUnachievablePrecisely controllable:
~(2–100) nm
Shape controllability
(Spherical, cubic, rod, hollow, and 2D nanoplate)
UnachievablePrecisely controllable
Specific surface area (m2/g)0.34 *Depend on size, shape, and coating:
~(20–300)
Magnetothermal conversion (W/g)Absent~(100–2500)
Electrocatalytic activityAchievableAchievable
* For an average crystallite size between 38 and 62 µm.
Table 3. Summary of catalytic applications of natural magnetite (Fe3O4).
Table 3. Summary of catalytic applications of natural magnetite (Fe3O4).
Catalytic ReactionNatural
Source
Pretreatment MethodSBET
(m2·g−1)
Reaction ConditionsResults of Catalytic EvaluationRate ConstantRef.
Fenton p-NP degradationOresCrushing, grinding, and magnetic and gravity separation0.54–2.46Cat = 1.0 g·L−1; p-nitrophenol = 10 mg·L−1; pH = 7; H2O2 = 0.05 mol·L−1; NaOH = 0.1 mol·L−117.8–95.0% degradation0.24, 0.19, and 0.12 μg·L−1·min−1[1]
Photo-Fenton degradation of MBNatural iron sandsMagnetic separation and coprecipitation147.12Cat = 2 g·L−1; MB = 10 mL of 20 ppm; UV irradiation68.52–76.32% degradation0.000194, 0.000262, and 0.000336 g·mg−1·min−1[51]
Fenton degradation of MBBeach sandDrying, magnetic separation, and ball millingNDCat = 1 g·L−1; MB = 15 mg·L−1; pH = 4.8; H2O2 = 150 mL L−186.79% degradationND[44]
Photo-Fenton degradation of cefotaxime *Natural magnetiteBall milling method2.92–5.64Cat = 0.2 g·L−1; cefotaxime = 0.4 mmol·L−1; pH = 5.6; H2O2 = 10 mmol·L−1100% degradation after 60 min0.09 min−1[46]
Fenton-like degradation of AO7Natural magnetiteNDNDCat = 0.5 mg·L−1; AO7 = 15 mg·L−1; pH = 5; S2 O 8 2 = 0.2 mM75% degradation after 120 min0.0019 min−1[49]
Fenton-like degradation of MePNatural magnetiteAs received62.5Cat = 0.3 g·L−1; MeP = 10 µmol·L−1; pH = 6.5; SPS = 5 mmol·L−190.2–99.5% degradation0.0066 min−1[52]
Electro-Fenton removal of GemcitabineNatural mineralCrushing, sieving, and millingNDPt anode; 140 mL·min−1 of air; 1 cm separation between electrodes; 0.05 mol·L−1 of Na2SO425–35% degradation63 M−1s−1 min−1[54]
CWPO of phenolNatural iron mineralsSieving8.0Cat = 1 g·L−1; phenol = 100 mg·L−1; pH = 3; H2O2 = 500 mg·L−1; T = 75 °C; t = 4 h70–80% mineralization and 100% of conversion11 L2 mg−1 gcat−1 min−1[62]
CWPO of azole pesticidePristine magnetiteAs received7.5Cat = 8 g; H2O2 = 6.7 mg·L−1; T = 25 °C; pH = 5; flow rate = 0.5 mL·min−1Conversion of the pollutants > 95%0.22 mL·gcat−1·min−1[50]
CWPO of SMXPristine magnetiteAs received7.5Cat = 1 g·L−1; H2O2 = 25 mg·L−1; T = 25 °C; pH = 550% mineralization and 100% removal0.0263 mg·L−1·min−1[55]
Fischer–Tropsch synthesis **Natural magnetiteSol–gel and high-temperature pyrolysis73.4–157.7Cat = 1 g; 2 g of quartz sand; T = 300 °C; P = 2 MPa; GHSV of 3000 mL·g−1·h−1; t = 180 h93–98.5% conversion of COND[61]
Production of diesel by catalytic pyrolysisMagnetite oreMagnetic separation, crushing, grinding, and ball millingNDCat = 3 wt%; T = 500 °C; t = 90 minSelectivity towards the formation of hydrocarbons having fuel value in conformity with diesel fuelND[45]
Cracking of coalMineral magnetiteAs received3.32Cat = 1 mg; coal; 1 mg; T = 700 °C; Ar atmosphereSelectivity to aliphaticsND[59]
Biodiesel production from waste cooking oil ***Iron sandCoprecipitation and wet impregnationNDRatio of methanol to oil = 9:1; Cat = 1 wt%; T = 60 °C; t = 6 h; 1100 rpm95.64% yieldND[63]
* TiO2/Fe3O4, ** Fe@C, and *** CaO/MgO/Fe3O4; ND = No data; Cat = Catalyst.
Table 4. Summary of catalytic applications of natural hematite (α-Fe2O3).
Table 4. Summary of catalytic applications of natural hematite (α-Fe2O3).
Catalytic ReactionCatalystNatural SourcePretreatment MethodSBET
(m2·g−1)
Reaction ConditionsResults of Catalytic EvaluationRate ConstantRef.
Photocatalytic degradation of MBMg/α-Fe2O3Natural sand Magnetic separation, drying, and ball millingNDUV irradiation; t = 300 minRemoval of 88.8% of MB ND[66]
Photocatalytic degradation of phenolα-Fe2O3/gelcasting porous ceramicNatural clayGrinding, sieving, and extraction by 3 M of H2SO4 at 80 °C31.92UV irradiation; T = 23 °C; pH = 8; t = 3 h; phenol = 10 mg·L−1Removal of 57% of phenol; catalytic activity remains without changes after 8 cyclesND[78]
Photo-Fenton degradation of MOα-Fe2O3/SiO2Iron sandSieving, immersion in HCl solution, washing, and drying11.16Cat = 1.5 g·L−1; MO = 100 mg·L−1; H2O2 = 200 mg·L−1; pH = 3; UV irradiation93% of degradation of MO. 89% after 4 cycles of reaction0.048 min−1[72]
Fenton degradation of (2,4-D) and MCPAα-Fe2O3Natural mineralCrushing, sieving, washing, sonication, and calcination60.4Cat = 0.5 g·L−1; 2,4-D and MCPA = 200 mg·L−1; PS = 0.025 M; pH = 3; T = 50 °C; t = 120 min36% of mineralizationMCPA = 0.0064; 2,4-D = 0.0059 min−1[73]
Fenton and photo-Fenton oxidation of phenol in waterMining rejectMining rejectBall milling and sieving16.35Cat = 0.75 g·L−1; phenol = 50 mg·L−1; H2O2 = 0.75 g·L−1; pH = 396.5% degradation at 180 min0.0411 min−1[74]
Degradation of DCF via PMS activationα-Fe2O3HRSAs received5.17Cat = 5 mg·L−1; DCF = 50 mg·L−1; PMS = 75 mg·L−1; t = 10 min; neutral pH98.2% of degradation in 10 min0.334 min−1[75]
Photocatalytic degradation of ICα-Fe2O3/bentoniteIron oreMagnetic separation, grinding, ball milling, and coprecipitationNDCat = 250 mg; IC = 5 mg·L−1; pH = 1; solar light100% of degradation after 2 hND[76]
Photocatalytic reduction of 4 NP to 4 APCu/Fe2O3Natural iron ore rockCrushing, sieving, calcination, and impregnation10–42Cat = 33.3 mg·L−1; 4 NP = 5 × 10−5 mol·L−1; NaBH4 = 12.5 mL (0.5 M); pH = 11.5; λ = 200–500 nm>99% conversion in less than 1 min2.34, 3.36, and 5.4 min−1[77]
Photocatalytic reduction of 4 NP to 4 APα-Fe2O3Natural iron ore rockCrushing, sieving, heat treatment, and calcination18–85Cat = 33.3 mg·L−1; 4 NP = 100 mL (0.1 mM); λ = 220–550 nm; NaBH4 = 12.5 mL (0.5 M) >99% conversion in less than 3 min1.38 min−1[79]
Catalytic ozonation and peroxone-mediated removal of ACTα-Fe2O3
HRSCalcination in air atmosphere3.63Cat = 1 g·L−1; ACT = 50 mg·L−1; O3 = 1.2 mg/min; t= 10 min; pH = 7100% degradation0.40 min−1[81]
Hydrocracking of high-temperature coal tarMo/Al2O3-Fe2O3Natural bauxiteCrushing, calcination, purification, washing, and wet impregnation126.9–237.9Cat = 1.3 g; sulfur powder = 0.4 g; HTCT = 42 g; T = 430 °C; P = 12.5 MPa; t = 90 minThe presence of Fe2O3 is not favorable to the hydrocracking of high-temperature coal tarND[82]
Hydrogenation of coalFe2O3–SiO2–Al2O3–TiO2–MnO2Natural bauxiteNDNDT = 380–440 °C; P = 3–5 MPa; sulfur additive = 0–2%; t = 90 minSelectivity to liquid productsND[84]
Biodiesel production from waste cooking oilα-Fe2O3Iron sandMagnetic separation, coprecipitation, and calcination10.5–22.9Esterification: methanol: waste cooking oil molar ratio of 5:1; H2SO4 = 1 wt%; T = 70 °C; t = 300 min. Transesterification: waste cooking oil: methanol molar ratio of 15:1; Cat = 1 wt%; T = 65 °C; t = 3 h87.88% biodiesel yieldND[86]
Biodiesel production from waste cooking oilFe2O3/CaO2Iron sandMagnetic separation, coprecipitation, and calcinationNDWaste cooking oil: methanol molar ratio of 1:15; Cat = 1 wt%; T = 65 °C; t = 3 h 97.04% biodiesel yield with a Fe/Ca ratio of 1:4ND[87]
Hydrothermal catalytic conversion of extra-heavy Ashal’chinskoe oilα-Fe2O3Natural hematiteAs receivedNDT = 210, 230, and 300 °C; P = 2–18 MPa; t = 2 hThe possibility of increasing the number of lighter hydrocarbons in heavy oil and reducing its density by a regular decrease in the amount of resin–asphaltene componentsND[88]
Catalytic cracking of tolueneα-Fe2O3Natural limoniteCrushing, sieving, and calcination12.4Cat = 0.5–1.5 g; toluene = 1000 ppm; T = 500–800 °C; t = 60 min95% of toluene conversionND[89]
Catalytic cracking of coal tarα-Fe2O3/γ-Al2O3
α-Fe2O3/NiO
Mineral hematiteCrushing, drying, and mechanical mixingNDT = 700–900 °C; consumption of coal tar = 0.3 kg·h−1; mass flow rate of coal tar = 5 g·min−1Poor performance of hematite/NiO; the addition of γ-Al2O3 could effectively inhibit carbon depositionND[90]
OERNi/iron oreIron oreBall milling and solution-assisted electrode preparationNDCat = 150 mg; NaOH = 1 MAchieves a current density of 10 mA cm−2 at a low overpotential of 280 mV; potentially scalable to industrial applicationsND[91]
Catalytic hydrolysis of microcystin-LR peptidesα-Fe2O3Mineral hematiteWashing, drying, milling, and sievingNDCat = 20 mg; MC-LR = 1 mL (10 ppm); T = 60 °C20.7% hydrolysis yieldND[106]
Fischer–Tropsch synthesisα-Fe2O3Raw iron oreGrinding and sieving59.0Syngas at space velocity of 60 Nml·gcat−1; T = 270 °C; P = 20 barThe CO-reduced catalyst exhibited the highest CO conversion of 94.1%, followed by the H2-reduced catalyst with a CO conversion of 80.1%, while the syngas-reduced catalyst showed the least CO conversion of 54.1%ND[107]
Table 5. Summary of catalytic applications of natural goethite (α-FeOOH). Unless otherwise stated, the catalyst is goethite.
Table 5. Summary of catalytic applications of natural goethite (α-FeOOH). Unless otherwise stated, the catalyst is goethite.
Catalytic ReactionNatural SourcePretreatment MethodSBET
(m2·g−1)
Reaction ConditionsResults of Catalytic EvaluationRate ConstantRef.
Oxidation of toluene *Acid mine drainage Sequential precipitation method and wetness impregnation115.70Cat = 1 g; T = 250–450 °C; toluene = 0.9 vol%; flow rate = 500 L·min−1; GHSV = 18,000 h−130% of toluene conversion; selectivity to CO2ND[121]
Photocatalytic degradation of MB **Goethite rocksBall milling, stirring, sonication, and electro-spinning processNDCat = 0.5 g; MB = 100 mL (10 ppm); t = 5 h; visible light and UV90% of bleaching after 5 h of illumination0.0141 min−1; similar in UV and visible light[122]
Fenton-like TC degradation ***Mineral goethiteHydrothermal method20.4Cat = 0.3 g·L−1; pH = 3; TC = 100 mg·L−1; H2O2 = 6.0 mM90.1% removal after 30 min of treatmentND[123]
Reduction and stabilization of Cr(VI) in soilNatural goethiteAs receivedNDCat = 3, 6, 9, and 12 g; pH = 8.6 to 9Goethite increased the apparent rate constant 3.33 × 10−6 kg·mg−1·min−1 to 8.33 × 10−6 kg·mg−·min−1[128]
Ozonation of SSZ ****Natural goethiteCrushing, milling, washing, and plasma process29.65, 73.24 and 77.31, respectivelyCat = 1–9 mg·L−1; T = 25 °C; P = 1 atm; inlet flow = 1 L·h−1; Na2SO3 = 1 mL (0.01 M) 96.5% degradation efficiency reached by N2-goethite after 40 min of reaction0.076 min−1[129]
Catalytic reforming of volatiles from co-pyrolysisMineral goethiteDehydration18.05–140.2815–20 mg of a mixture of raw lignite and corn straw; T = room temperature to 800 °C; Ar flow = 100 mL·min−1Goethite significantly promotes the production of light aromatic hydrocarbonsND[135]
Production of ROSNatural goethiteMechanical crushing, milling, and sievingNDCat = 0.1 g·L−1; LI = 10 mW cm−2; pH = 7; T = 25 °CNatural goethite ROS production was 4.9-fold higher than the standard0.648 nM min−1[137]
* Mn/goethite; ** PAN/goethite nanofibers; *** Goethite/Fe3O4; **** Natural N2-treated goethite; ND = No data; LI = Light intensity.
Table 6. Summary of catalytic applications of less abundant and rare Fe-OH prepared from natural sources.
Table 6. Summary of catalytic applications of less abundant and rare Fe-OH prepared from natural sources.
Catalytic ReactionCatalystMineral SourcePretreatment MethodSBET
(m2·g−1)
Reaction ConditionsResults of Catalytic EvaluationRef.
Transfer hydrogenation of furfural, levulinic acid, and o-vanillinγ-Fe2O3River sandMagnetic separation, grinding, washing, and drying at 80 °C22.3Cat = 20 mg (7 mol %); T = 150 °C; 1 mmol of substrate in 2-propanol (3 mL); t = 30–90 minγ-Fe2O3 recovered from the natural source exhibited superior activity compared to the synthetic counterpart under base-free conditions[149]
Catalytic reduction of NO with NH3α-Fe2O3/γ-Fe2O3Iron oresDrying, grinding, and calcination22.8–42.5NO = 500 ppm; NH3 = 500 ppm; O2 = 3 vol %; SO2 (when used) = 150 ppm; H2O (when used) = 5 vol%; N2 = 145 mL·min−1; GHSV = 20,000 h−1Selectivity catalytic reduction activity above 80% at 170–350 °C and N2 selectivity (above 90% up to 250 °C) at low temperatures[152]
Decomposition of hydrogen peroxideFerrihydriteAcid mine drainageVacuum drying270Cat = 1 g·L−1; H2O2 = 0.02 MWaste ferrihydrite shows the same catalytic activity for H2O2 decomposition as the commercial goethite-based catalyst[150]
Photo-Fenton degradation of ENR and LEVWüstiteNDND108.7Cat = 10 mg·L−1; ENR = LEV = 0.05 mmol·L−1; H2O2 = 1.0 mmol·L−1; pH = 6.5; T = 35 °C; t = 180 min100% and 55 % of antibiotic activity elimination in 180 min for ENR and LEV, respectively; complete antibiotic activity elimination for ENR in the next four recycling cycles[151]
ND = No data.
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

Jiménez-Vázquez, A.; Jaimes-López, R.; Morales-Bautista, C.M.; Pérez-Rodríguez, S.; Gochi-Ponce, Y.; Estudillo-Wong, L.A. Catalytic Applications of Natural Iron Oxides and Hydroxides: A Review. Catalysts 2025, 15, 236. https://doi.org/10.3390/catal15030236

AMA Style

Jiménez-Vázquez A, Jaimes-López R, Morales-Bautista CM, Pérez-Rodríguez S, Gochi-Ponce Y, Estudillo-Wong LA. Catalytic Applications of Natural Iron Oxides and Hydroxides: A Review. Catalysts. 2025; 15(3):236. https://doi.org/10.3390/catal15030236

Chicago/Turabian Style

Jiménez-Vázquez, Adriana, Raciel Jaimes-López, Carlos Mario Morales-Bautista, Samuel Pérez-Rodríguez, Yadira Gochi-Ponce, and Luis Alberto Estudillo-Wong. 2025. "Catalytic Applications of Natural Iron Oxides and Hydroxides: A Review" Catalysts 15, no. 3: 236. https://doi.org/10.3390/catal15030236

APA Style

Jiménez-Vázquez, A., Jaimes-López, R., Morales-Bautista, C. M., Pérez-Rodríguez, S., Gochi-Ponce, Y., & Estudillo-Wong, L. A. (2025). Catalytic Applications of Natural Iron Oxides and Hydroxides: A Review. Catalysts, 15(3), 236. https://doi.org/10.3390/catal15030236

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop