Next Article in Journal
The Uplift and Denudation History of the Jianfeng Pluton on Hainan Island, China
Previous Article in Journal
The Role of Fluid Chemistry in the Diagenetic Transformation of Detrital Clay Minerals: Experimental Insights from Modern Estuarine Sediments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Utilization of Natural Mineral Materials in Environmental Remediation: Processes and Applications

1
Sinomine Rock & Mineral Analysis Tianjin Co., Ltd., Fang 5th Road, Binhai New Area, Tianjin 300450, China
2
School of Environmental Science and Engineering, China-Singapore Joint Center for Sustainable Water Management, Tianjin University, Tianjin 300350, China
3
Hunan Port Shipping and Water Resources Group Co., Ltd., No. 196, Xinyao Road (South), Tianxin District, Changsha 410004, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(3), 318; https://doi.org/10.3390/min15030318
Submission received: 31 December 2024 / Revised: 28 February 2025 / Accepted: 14 March 2025 / Published: 19 March 2025
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)

Abstract

:
The discharge of wastewater containing persistent organic pollutants presents significant ecological and health challenges due to their toxicity and resilience. Recent advances in advanced oxidation processes (AOPs) and other remediation mechanisms, notably utilizing natural mineral materials (NMMs), offer promising solutions to these challenges. NMMs, with their cost-effectiveness, accessibility, eco-friendly nature, non-toxicity, and unique structural properties, have shown significant promise in environmental remediation and could effectively replace conventional catalysts in related applications. These minerals enable the activation of oxidants, generating reactive oxygen species crucial for the degradation of pollutants. This article reviews the mechanisms of NMMs in various AOPs, including photocatalysis, Fenton-like reactions, and persulfate-activation-based processes, and discusses the potential of these materials in enhancing pollutant degradation efficiency, with a focus on the activation of persulfates and the photo-induced redox processes. The synergy between photocatalytic properties and catalytic activation provided by NMMs offers a robust approach to managing water pollution without the drawbacks of secondary waste production, thus supporting sustainable remediation efforts.

1. Introduction

The uncontrolled release of wastewater has become a pressing issue for ecosystems and public health, primarily due to its adverse impacts on water quality and the availability of potable water. The presence of organic pollutants, which are both resilient and toxic, further complicates efforts in environmental remediation [1]. Recent advancements have introduced various methods to address these environmental contaminants, such as absorption, membrane separation, biodegradation, and especially advanced oxidation processes (AOPs) [2]. These AOPs are highly regarded for their capacity to efficiently degrade or mineralize harmful organics via reactive oxygen species (ROS), such as hydroxyl (•OH; redox potential: 2.8 V vs. NHE), sulfate ( S O 4 , 2.5–3.1 V), and superoxide radicals (•O2). Furthermore, AOPs can generate ROS through activation processes involving peroxymonosulfate, persulfate, and hydrogen peroxide, catalyzed by ultrasound [3], microwave radiation [4], thermal methods [5], and ultraviolet radiation [6]. AOPs can efficiently degrade recalcitrant organic pollutants and broad-spectrum treatment capabilities in environmental applications, leveraging highly reactive oxidative radicals to enable rapid mineralization. Additionally, their mild reaction conditions and versatile applicability across diverse systems allow for the effective treatment of complex environmental matrices while minimizing toxic byproduct formation, demonstrating strong environmental compatibility and a promising potential for practical use.
Natural mineral materials (NMMs) and their synthesized derivatives have garnered significant attention in advanced oxidation processes (AOPs) due to their cost-effectiveness, abundance, non-toxicity, and physicochemical stability. Geological data reveal significant global reserves of key transition metals: iron (164.3 billion tons proven, 582.8 billion tons estimated), manganese (2.034 billion tons proven, 8.073 billion tons estimated), and copper (821.63 million tons proven, 1.92 billion tons estimated) [7]. Enriched with transition metals, these materials function as efficient heterogeneous catalysts for peroxide activation, effectively degrading organic pollutants with minimal metal ion leaching [8]. Certain minerals—such as hematite [9], sphalerite [10], and anatase [11,12]—also demonstrate intrinsic photocatalytic properties, enabling contaminant removal under light irradiation. However, the long-term stability of these catalysts under the conditions of a fluctuating pH and coexisting anions remains insufficiently explored, and the mechanisms behind metal leaching during repeated catalytic cycles warrant a more in-depth investigation.
This review examines the potential of NMMs in various AOPs—including photocatalysis, Fenton reactions [13], Fenton-like processes [14], and persulfate activation [15]—focusing on their capacity to generate reactive species for decomposing persistent organic pollutants. It summarizes recent progress in employing mineral-based photocatalysts for water decontamination, highlighting how mineral composition and structure influence photocatalytic performance. The discussion also addresses current research gaps, stresses the sustainability and ecological safety of these materials, and proposes directions for future development.

2. Mechanism of Environmental Purification Driven by NMMs

2.1. Photocatalytic Redox Reaction

There are abundant natural semiconductor minerals in the key zones of the Earth’s surface; notably, many kinds of metal oxide minerals and metal sulfide minerals exhibit typical semiconductor characteristics. Research has demonstrated that various oxides and sulfides in natural metallic minerals—such as hematite (Fe2O3), anatase (TiO2), pyrite (FeS2), sphalerite (ZnS), and greenockite (CdS)—exhibit photocatalytic activity. Table 1 summarizes the characteristic parameters, including band gap energy and maximal wavelength, of representative metal oxide and sulfide semiconductor minerals. These findings reveal the critical conditions for generating photoinduced charge carriers in such minerals, providing a significant theoretical foundation for the study of photocatalytic mineral materials [16,17].
Under light irradiation, the electrons in the valance band of semiconductor minerals will transition to the conduction band, and thus a hole will remain in the valance band. Then, the oxidative holes and reductive electrons will trigger a series of redox reactions. These reactions can influence the transformation of environmental pollutants and trigger the environmental purification process, such as the oxidative degradation of organic pollutants and the reduction in high-valance-state toxic metals. In fact, oxygen and water often participate in photocatalytic reactions and induce the production of various reactive oxygen species (ROS), such as •OH, •O2, and H2O2 (Table 2). The photo-generated holes and produced ROS are responsible for the degradation of pollutants via oxidation, while the photo-generated electrons are responsible for the reduction in high-valance-state toxic metals, such as Cr(VI).
The photocatalytic redox efficiency of a semiconductor is mainly determined by the light response range and the separation efficiency of the photo-generated electron–holes. Generally, ion doping and substitution phenomena will inevitably occur in naturally formed semiconductor minerals due to the complex actual environment, while impurity ions play a positive role in widening the light response range and promoting the separation of the photo-generated electron–holes of semiconductor minerals. This facilitates the photocatalytic redox process for environmental purification.

2.2. Fenton-like Process and Persulfate-Based Advanced Oxidation Process

Fenton reactions have found widespread application in environmental remediation due to their effective oxidative capabilities [18]. These reactions primarily rely on the generation of highly ROS, such as •OH, to facilitate the degradation of organic pollutants, the reduction and transformation of heavy metals, and the inactivation of microorganisms. In the classic Fenton process, Fe2+ reacts with hydrogen peroxide (H2O2) to produce •OH, with the reaction being sustained by the iron redox cycle. Specifically, Fe2+ reacts with H2O2 to generate •OH and Fe3+ (Equation (1)), and the resulting Fe3+ can be reduced back to Fe2+ (Equation (2)), thus maintaining the reaction cycle. This process allows the Fenton reaction to efficiently degrade a wide range of organic pollutants [19,20,21,22]. However, the practical application of the Fenton reaction is limited by its narrow effective pH range (pH > 3) and the generation of excessive iron hydroxide sludge [23]. Notably, transition metal catalysts are not limited to iron-based systems. Recent studies indicate that the Cu(I)/Cu(II) redox couple exhibits faster electron transfer kinetics compared to the Fe(II)/Fe(III) counterpart, facilitating efficient •OH generation under near-neutral pH conditions—the catalytic cycle of copper-based Fenton-like reactions, as described in Equations (3) and (4) [24]. Moreover, copper-based materials demonstrate excellent catalytic activity within a pH range of 3 to 7, addressing a key limitation of traditional iron-based Fenton systems [25]. The Cu+/H2O2 system, with a reaction rate constant k of 1 × 10⁴ M−1s−1, significantly outperforms the Fe2+/H2O2 system (k = 63–76 M−1s−1), highlighting the considerable potential of copper-based materials in Fenton-like reactions [26].
F e 2 + + H 2 O 2 F e 3 + + O H + O H
F e 3 + + H 2 O 2 F e 2 + + H O 2 + H +
C u + + H 2 O 2 C u 2 + + O H + O H
C u 2 + + H 2 O 2 C u + + H O 2 + H +
To mitigate the drawbacks of homogeneous Fenton reactions, heterogeneous catalysts have been developed to activate H2O2 without iron sludge formation while extending the operational pH range [27]. Concurrently, persulfates (PDS) and peroxymonosulfates (PMS) have emerged as alternative oxidants for wastewater treatment. Although persulfates can directly degrade pollutants, their slow reactivity necessitates activation to generate sulfate radicals (SO4•−) or •OH [28]. Activation methods include thermal energy, which cleaves the O-O bond in persulfate to produce SO4 (Equations (5) and (6)) [29], and ultrasound (US), where cavitation-induced localized heating and pressure collapse persulfate molecules into radicals (Equation (7)) [30]. Transition metal ions (Fe2+, Co2+) further activate persulfates via redox reactions, generating SO4 (Equations (8) and (9)) [31,32]. These radicals degrade pollutants into intermediates, ultimately mineralizing them into harmless products like CO2 and H2O (Equations (10) and (11)) [33]. Beyond persulfates, transition metals also activate peroxides such as peracetic acid (PAA). Iron-mediated systems drive PAA activation: (1) Fe2+ transfers electrons to PAA, yielding acetoxy radicals (CH3C(O)O•, Eh0= 1.60 V Equation (12)) [34]; and (2) Fe2+ induces the homolytic cleavage of PAA to produce •OH (Equation (13)) [35]. Such multi-path radical generation underscores the versatility of transition metal catalysts in advanced oxidation processes, enabling efficient pollutant degradation across diverse aquatic environments.
H S O 5 h e a t S O 4 + O H
S 2 O 8 2 h e a t 2 S O 4
H S O 5 u l t r a s o n i c S O 4 + O H
H S O 5 + M n + S O 4 + M n + 1 + O H
S 2 O 8 2 + M n + S O 4 + M n + 1 + S O 4 2
p o l l u t a n t s + S O 4 d e g r a d e d   p r o d u c t s
degraded   products + S O 4 C O 2 + H 2 O
F e 2 + + C H 3 C ( O ) OOH   F e 3 + + C H 3 C ( O ) O + O H
F e 2 + + C H 3 C ( O ) OOH   F e 3 + + C H 3 C ( O ) O + O H

2.3. Photocatalytic-Based Persulfate Activation

Traditional persulfate activation methods, though effective, are often constrained by operational parameters such as temperature and pH, resulting in high costs and limited practical efficiency [36]. Light activation addresses these limitations by reducing the energy barrier for persulfate decomposition while providing supplementary energy, thereby enhancing the generation of reactive radicals (SO4 and •OH) critical for pollutant degradation (Equations (14)–(20)) [37]. Semiconductor photocatalysts optimize this process by improving electron transfer kinetics to boost radical production. Concurrently, redox-active natural mineral materials (NMMs), including Fe-, Cu-, and Mn-based minerals, demonstrate versatile activation capabilities under both illuminated and dark conditions. Their large specific surface areas enhance persulfate adsorption, enabling localized decomposition and sustained radical generation through surface-bound metal redox cycling [38]. The synergy between photocatalytic and mineral-mediated activation broadens the environmental adaptability of persulfate systems, offering a robust and efficient strategy for pollutant remediation.
Photocatalyst + h ν e - + h +
e + H S O 5 S O 4 + O H
h + + H 2 O O H + H +
e + O 2 O 2
F e 3 + + e F e 2 +
F e 2 + + H S O 5 F e 3 + + S O 4 + O H
S O 4 - / h + / O 2 - / OH + Pollutants degraded   products

2.4. Other Remediation Mechanisms

NMMs exhibit versatile remediation mechanisms beyond conventional photocatalytic and oxidative pathways. In ozonation systems, transition metal sites on mineral surfaces catalyze O3 decomposition via redox cycling, generating •OH and •O2 radicals, while synergistically adsorbing pollutants to enhance degradation efficiency [39]. Piezocatalysis leverages piezoelectric materials that generate charges under mechanical vibration (energy conversion efficiency up to 78%). These charges drive electrochemical redox reactions to produce reactive species, enabling efficient degradation of organic pollutants in dye wastewater. The mechanism integrates the piezoelectric effect with charge transfer in solution, directly converting mechanical energy into reactive intermediates for catalytic processes [40]. For reductive remediation, sulfide-rich minerals or structural Fe2+ drive contaminant detoxification through direct electron transfer or aqueous electron (eaq) generation, achieving critical pathways such as dehalogenation (C-Cl bond cleavage) and Cr(VI) reduction [41]. Notably, minerals can orchestrate oxidation–reduction synergy by simultaneously activating oxidants and reductants, enabling parallel organic pollutant degradation and heavy metal immobilization. This multifunctionality hinges on the spatiotemporal control of reactive species diffusion and interfacial partitioning, positioning natural minerals as adaptive, eco-compatible solutions for complex co-contamination scenarios.

3. NMMs for Sustainable Pollutants Degradation

3.1. Metal Oxide Materials

3.1.1. Iron Oxide Minerals

Iron (Fe) is the most abundant and widely distributed transition metal element in the Earth’s crust and exists in the form of stable iron oxides and hydroxides, such as hematite, goethite, and magnetite. The outstanding oxidation–reduction and photochemical activities of iron oxides (hydroxides) have a non-negligible impact and regulatory effect on environmental pollutants [42].
Hematite (hexagonal system) is one of the most common iron oxide minerals found in strongly weathered soils, such as red soil and brick red soil, because its stable mineral structure, which is widely distributed in the sediments and surface soil, is 70% iron. This mineral has attracted significant attention due to its abundant occurrence, economic potential, non-toxic nature, and exceptional reactivity to visible light. Its narrow band gap, ranging from 2.0 to 2.2 eV, allows it to absorb visible-light wavelengths up to 600 nm, thereby promoting chemical reactions through the generation of photoinduced charge carriers [43,44]. Hematite shows great potential in environmental remediation due to its advantages, such as high specific surface area, fast electron transport, and high redox activity. Many studies have reported on the treatment of heavy metal and organic pollutants by hematite. Liu et al. investigated the effect of hematite on the fate and transport of Cr under sunlight irradiation, with the reaction mechanism illustrated in Figure 1a [45]. It was found that hematite can induce the photocatalytic redox of Cr, thus enhancing the immobilization of Cr, and the acidic condition is more conducive to the immobilization of Cr. The photocatalytic oxidation of Mn2+ by hematite was also reported, where the oxidized Mn2+(aq) will further form tunneled Mn (III/IV) oxides, thus remediating Mn2+(aq)-contaminated water [46]. Oxalate, which widely exists in the environment, has been proven to be a favorable factor for the photocatalytic performance of hematite. A solar/hematite/oxalate system was constructed to drive roxarsone degradation and induce the simultaneous As(V) immobilization. It was found that within 6 h, 85.1% of roxarsone can be converted to inorganic As(V) by the oxidation effect of •OH produced in the solar/hematite/oxalate system, and the reaction mechanism is shown in Figure 1b. The production process of •OH can be attributed to the following steps: Firstly, oxalate chelated with Fe(III) on the surface of hematite to form Fe(III)-oxalate. Then, Fe(III)-oxalate could be reduced to Fe(II)-oxalate under light irradiation, while •CO2 with reducibility was produced simultaneously. The produced •CO2 could further induce the activation of oxygen molecules; thus, the ROS, including •OH, could be produced with the assistance of Fe(II) [47]. Moreover, the similar oxalate/hematite can also induce the degradation of dissolved organic matter by the oxidative ROS produced under light irradiation (Figure 1c) [48].
As an iron-rich mineral, hematite can also serve as a catalyst for the activation of oxidants in the advanced oxidation process. Liu et al. modified natural hematite using a reducing agent (NaBH4), and this modified hematite was applied to the activation of peracetic acid for the efficient degradation of cefazolin. The high content of Fe(II) in the modified hematite mainly contributed to the activation of peracetic acid, where the O2, CH3C(O)OO•, and •OH produced in this process were identified as the activated species for cefazolin degradation [49]. Additionally, the photo-Fenton process using natural hematite and siderite as heterogeneous catalysts has also been reported, and it exhibited an outstanding effect in 4-chlorophenol removal [50].
Goethite (α-FeOOH) is also a commonly distributed natural iron hydroxide mineral with a similar structure to gibbsite, of which the anions are arranged in a hexagonal, close-packed structure. Goethite is a thermodynamically stable mineral, making it the termination phase of the transformation process of many iron oxides. The abundant structure of hydroxyl groups and surface function groups endow goethite with good surface adsorption performance and reactivity; thus, it can influence the transformation and fate of many environmental contaminants. Another paper reported the synergistic promotion effect of goethite and humic acid on the photodegradation of bifenthrin under light irradiation. The improved photocatalytic activity of goethite for bifenthrin degradation can be attributed to its extended light adsorption and improved electron–hole separation induced by humic acid, as shown in Figure 2. In this system, both the oxidative species, •OH and 1O2, and reducing species, such as hydrogen atoms (H•), contributed to the degradation of bifenthrin [51]. The element substitution phenomenon often occurs in goethite, and Al, Mn, Cr, and Ni can replace the lattice Fe in goethite, which can change the band structure of goethite and further influence its photocatalytic performance. In a typical example, the influence of Mn doping on the band structure has been investigated by both theoretical calculation and experimental confirmation. It was reported that Mn-substitution could cause the reduction in the band gap of goethite when the Mn-substitution is lower than 3–4 mol%. Accordingly, the photocatalytic performance of goethite for methylene blue degradation increased due to its improved response [52].
In addition to hematite and goethite, other iron oxides (hydroxides) such as magnetite, ferrihydrite, ilmenite, and lepidocrocite were also investigated for their environmental remediation effect [53]. Wang et al. reported a natural-magnetite-based PMS activation system, where hydroxylamine was introduced into this system to synergistically enhance its degradation effect toward organic pollutants. The results showed that the degradation efficiency of rhodamine B (rhB) was increased by 61.82% after the addition of hydroxylamine, which was caused by the promotion effect of hydroxylamine on Fe2+/Fe3+ conversion in the magnetite/peroxymonosulfate system [54]. Compared to the monometallic Fe, the cooperation of multiple transition metal elements in natural minerals was confirmed to be a favorable factor for Fe(III)/Fe(II) circulation and ROS generation in sulfate-radical-advanced oxidation processes. The natural Ti- and V-enriched magnetite was applied as the catalyst in the UV-assisted PMS activation system, which achieved a 100% degradation efficiency of BPS within 20 min. It was suggested that the photo-generated electron–holes of TiO2 and the subsequent electron transfer by V(III)/V(IV)/V(V) redox cycles had effectively accelerated the Fe(III)/Fe(II) circulation [55]. Similarly, the other bimetallic natural mineral–ilmenite showed excellent bacterial inactivation performance in the persulfate-mediated catalytic and photocatalytic system [56]. Ferrihydrite is a widely studied natural iron oxide with the characteristics of low crystallinity and nanoscale particle size, showing high reactivity with environmental pollutants. In addition to the well-known good adsorption performance of ferrihydrite, its semiconductor properties have also been noticed. Wang et al. clarified the effective photocatalytic reduction in U(VI) by ferrihydrite under both anaerobic and aerobic conditions [57]. These catalysts have two critical advantages: (1) sustained oxidant activation via surface-bound Fe3+/Fe2+ redox cycles, which effectively minimizes iron ion leaching and prevents hydroxide precipitation and (2) broad pH tolerance, allowing them to maintain a high catalytic activity from acidic to neutral conditions. Their structural integrity ensures exceptional reusability, with most systems retaining their efficiency after multiple cycles. This synergy of stability, pH adaptability, and recyclability positions heterogeneous Fenton catalysts as sustainable alternatives to conventional homogeneous systems.

3.1.2. Titanium Oxide Mineral

Titanium dioxide occurs in phase structure forms of anatase, rutile, and brookite, which are typical semiconductor minerals. Among them, rutile is the most stable structure thermodynamically, making it the most frequently occurring titanium dioxide mineral in nature. Some trace impurity ions, such as V and Fe, can sometimes be doped into the structure of natural rutile, which will affect its band structure and photocatalytic performance. Li et al. determined the band gap of natural rutile (2.7 eV) using synchrotron-based O K-edge X-ray absorption and emission spectra, finding it to be narrower than that of synthetic rutile (3.0 eV), where the doping V and Fe ions were considered the main leading factor for the narrowed band gap of rutile from the DFT calculation results. This narrow band gap endows rutile with visible-light response properties, which can more efficiently drive photocatalytic redox in the Earth’s surface environment; the energy band and structure of natural rutile are shown in Figure 3a,b [58]. The photocatalytic performance of V-doping natural rutile for the degradation of organic pollutants, such as methyl orange and halohydrocarbons, was examined. It was found that the natural rutile bearing V5+ and Fe3+ substitution exhibited good methyl orange degradation performance with the presence of H2O2 under visible-light irradiation, and 60.59% of methyl orange was removed within 1 h, which is comparable to the commercial nano-TiO2 (P25). The increased light absorption and promoted electron–hole separation efficiency induced by the electron trapping effect of high-valance metal ions (V5+ and Fe3+) were both reasonable for the good photocatalytic performance of natural rutile. Lu et al. conducted heating, quenching, and electron irradiation treatments of natural rutile to change its surface characteristics. Quenching improved the exposure of adsorbed water and V on the surface of rutile, leading to its high photoactivity for halohydrocarbon degradation (trichloroethylene and tetrachloroethylene). In addition to photocatalytic activity, ion substitution can also influence the Fenton catalytic activity of natural rutile. Rutile with different levels of V and Fe doping was obtained by the hydrogen annealing of natural rutile, thus elucidating the influence of doping ions on the Fenton catalytic activity of rutile. Hydrogen annealing led to the formation of surface Fe(II) and bulk V(III) in rutile, benefiting the Fenton catalytic reaction and organic pollutant degradation [59].
Outside of engineering applications, the photocatalytic effect of titanium dioxide also drives the oxidation and transformation process of many substances in the natural environment, such as the fixation of NOx, the oxidation of Mn2+, and manganese oxide formation [60]. Nitrogen oxides (NOx = NO + NO2) in the atmosphere are considered a threat to human health, and the photocatalytic oxidation of NOx to nitrate (NO3) is an efficient way to reduce NOx pollution. It was reported that the small amount (≤5%) of titanium dioxide (anatase and rutile) naturally occurring in soils could lead to NOx fixation by oxidizing NOx to NO3 under light irradiation. It provides an abiotic renoxification route for N-fixation and NOx pollution alleviation with the leading role of titanium dioxide in soils, and the mechanism of the redox reaction of manganese (Mn) is very important for environmental evolution. Jung et al. found that the TiO2 minerals could catalyze the rapid oxidation of Mn2+ (aq) in a neutral environment under light irradiation, and the reaction rate was even higher than those of the reported biotic/abiotic processes. This finding provided new insight into the formation of manganese oxide minerals in the natural environment. Moreover, the main minerals in the red sediment from the Danxia landform was determined to be iron oxide (hematite) and titanium oxide (anatase). It was confirmed that the combination of anatase and hematite played a significant role in the photocatalytic reaction, where the photocurrent produced by hematite was improved 23-fold at 0.8 V (vs. Ag/AgCl) after coupling with anatase; the reaction process is shown in Figure 3c [61].
Figure 3. Schematic diagram for the structure of natural rutile: (a) band gap and (b) cell [58], Copyright © 2018, Elsevier; (c) reaction mechanism of titanium dioxide-catalyzed dissolution of Mn2+ and formation of tunnel-structured manganese oxide [61], Copyright © 2018, MDPI.
Figure 3. Schematic diagram for the structure of natural rutile: (a) band gap and (b) cell [58], Copyright © 2018, Elsevier; (c) reaction mechanism of titanium dioxide-catalyzed dissolution of Mn2+ and formation of tunnel-structured manganese oxide [61], Copyright © 2018, MDPI.
Minerals 15 00318 g003

3.1.3. Manganese Oxide Mineral

Manganese oxides, naturally abundant and diverse, serve as crucial environmental mineral materials due to their structural versatility and redox properties. The Mn in the manganese oxides generally exists in the form of Mn2+, Mn3+, and Mn4+; they can undergo mutual transformation under specific conditions; and the variable valance states of Mn endow manganese oxides with outstanding redox activity. Their crystal structures originate from the arrangement of [MnO6] octahedral units, which interconnect via shared edges or corners to form chain-like frameworks. These chains further assemble into porous architectures with rectangular or square channels, enhancing their environmental purification performance through heavy metal adsorption and organic pollutant degradation [62]. Birnessite, a prominent manganese oxide, effectively oxidizes contaminants such as Co(II), Cr(III), V(IV), Pu(IV), and Se(IV) to higher oxidation states [63,64,65,66]. Li et al. demonstrated that the birnessite-mediated chemical oxidation of As(III) involves Mn(IV) reduction to Mn(II), although the resulting MnOOH byproduct inhibits further As(III) oxidation. By contrast, photocatalytic oxidation using birnessite achieved superior As(III) removal efficiency, as the process minimized MnOOH interference, highlighting its regulatory role in arsenic speciation and environmental mobility (Figure 4) [67]. Additionally, manganese oxides function as heterogeneous catalysts in advanced oxidation processes. For instance, hydrothermally aged natural manganese oxides exhibited enhanced catalytic ozonation performance in 4-chlorophenol degradation, attributed to redox cycling between Mn3+/Mn2+ and Fe3+/Fe2+, which amplified the generation of reactive oxygen species (ROS) [68]. Similarly, natural manganese ores activated PMS to efficiently degrade phenol, TBBPA, and RhB, underscoring their broad applicability in pollutant remediation [69]. The structural adaptability and redox plasticity of manganese oxides position them as multifunctional materials for environmental applications. However, challenges persist in optimizing their stability, recyclability, and selectivity under real-world conditions.
To elucidate the enhancement mechanism of natural minerals in pollutant degradation, we selected typical metal oxide materials for performance comparison (Table 3). Notably, the introduction of mineral components significantly improved degradation efficiency while also achieving cost savings and mitigating the risk of metal leaching contamination.

3.2. Metal Sulfide Materials

3.2.1. Pyrite

The degradation effects of natural metal sulfide minerals on pollutants are summarized in Table 4. As the most abundant and most widely distributed metal sulfide mineral on the Earth, pyrite (FeS2) has become a research hotspot due to its excellent catalytic performance, which was considered to have great advantages in environmental remediation, cost-effectiveness of pyrite as catalyst for aqueous pollutant solutions is confirmed by numerous researchers. In typical applications, pyrite is often used as a heterogeneous catalyst to drive the persulfate-based advanced oxidation reaction due to the high content of Fe2+. Zhu et al. constructed a pyrite/persulfate system for the removal of cyclohexanoic acid in real petroleum wastewater. Under optimal conditions (2 g/L of pyrite, 4 mM PS), more than 90% of cyclohexanoic acid can be removed in a wide pH range. It was revealed that the Fe2+ serves as the active reaction sites for persulfate activation, and SO4•−, Fe(IV), and •OH contribute to cyclohexanoic acid degradation. Moreover, the interaction between Fe3+ and S played a positive role in promoting the cycle of Fe3+/Fe2+, endowing the pyrite with outstanding recycling performance. The reaction process is shown in Figure 5a [74]. Interestingly, the pyrite/persulfate system was also applied for the innovative removal of Microcystis aeruginosa [75]. Additionally, the pyrite-driven persulfate activation showed good performance for the treatment of the metal ion–organic complex; for example, the V(IV)–citrate complex in groundwater can be efficiently degraded with the natural pyrite/PMS system assisted with alkali precipitation, providing an efficient way for actual complex wastewater treatment [76].
In addition to the persulfate-based advanced oxidation process, pyrite/H2O2 Fenton-like processes have also been widely reported. The functional mechanism of pyrite/H2O2 systems for naphthalene degradation was investigated. The results indicated that the homogeneous reaction mainly contributed to hydrogen peroxide disintegration and naphthalene degradation in the pyrite/H2O2 system. Moreover, the sulfur species from pyrite was confirmed as an important factor for promoting the Fe2+/Fe3+ cycle, which further facilitated the continuous production of reactive oxygen species, thus enhancing the degradation of naphthalene [84]. Wang et al. proposed an ingenious pre-reaction scheme to improve the reaction activity of the pyrite/H2O2 system, where the contaminant solution was added after a period of pre-reaction between pyrite and H2O2. As illustrated in Figure 5b, the pre-reaction of pyrite and H2O2 caused the in situ formation of H+ and thus promoted the Fe2+ release, which further boosted the generation of ROS and led to the efficient degradation of dyes [85]. In addition, Dai et al. demonstrated that FeS2 effectively activates PAA to degrade tetracycline (TC), with sulfur speciation critically influencing ROS generation and Fe(II) regeneration [86]. DFT calculations revealed that sulfur species facilitate electron transfer between CH3COO• radicals and FeS2, driving dual roles in regulating ROS and sustaining Fe(II)/Fe(III) cycling. This theoretical insight aligns with experimental observations showing sulfur-mediated enhancement in SO4•− production and PAA activation efficiency. Furthermore, the integration of electrospun hydrophilic FNC fiber films with the FeS2/PAA system maintained robust TC degradation in real wastewater, showcasing its practical applicability.
The modification of pyrite via ball milling is considered efficient for improving the catalytic activity of pyrite. It was found that the production of S vacancies (VS) and the improved reducibility of S(-II) on the pyrite induced by the ball-milling process played multiple roles in promoting the catalytic activation of the persulfate, with the reaction mechanism shown in Figure 5c. It accelerated Fe(III) reduction and Fe(II) dissolution in the pyrite/persulfate system and thus promoted the production of reactive species (RS) for monochlorobenzene degradation. Interestingly, it was also found that the ball-milling treatment improved the contribution of FeIV for monochlorobenzene degradation, indicating that the RS generation pathway was changed [87]. Sun et al. emphasized the dominant function of the size effect on improvements in the catalytic performance of pyrite in the pyrite/persulfate system induced by the mechanical treatment of pyrite. Different grinding treatment times result in different pyrite particle sizes. The fitting results showed that the decomposition rate was almost linearly related to the particle size of pyrite [88]. In recent work, by optimizing the ball-milling conditions, the catalytic efficiency of modified pyrite was increased by 60-fold compared to natural pyrite, and the ball-milled pyrite Fenton system was able to degrade dyes and a series of antibiotics within 30 min [89].
Figure 5. (a) The efficient removal of naphthenic acids from real petroleum wastewater by the natural-pyrite-activated persulfate system [74], Copyright © 2023, Elsevier; (b) the reaction scheme in pyrite/H2O2 system assisted by pre-reaction [85], Copyright © 2021, Elsevier; and (c) the reaction mechanism of the ball-milled pyrite/PDS system [87], Copyright © 2024, Elsevier.
Figure 5. (a) The efficient removal of naphthenic acids from real petroleum wastewater by the natural-pyrite-activated persulfate system [74], Copyright © 2023, Elsevier; (b) the reaction scheme in pyrite/H2O2 system assisted by pre-reaction [85], Copyright © 2021, Elsevier; and (c) the reaction mechanism of the ball-milled pyrite/PDS system [87], Copyright © 2024, Elsevier.
Minerals 15 00318 g005

3.2.2. Chalcopyrite

Chalcopyrite (tetragonal system) is one of the most widely distributed sulfur-containing minerals, with Cu, Fe, and S mass fractions of 35.78, 32.73, and 28.31%, respectively. It accounts for more than 70% of the total copper content in the Earth’s crust. The effective electron transfer properties of chalcopyrite, due to its complex metal oxidation states, such as Cu(I), Cu(II), Fe(II), and Fe(III), enable it to exhibit excellent catalytic performance in the degradation of organic pollutants through peroxide activation. Wang et al. used natural chalcopyrite (NCP) to activate peroxymonosulfate (PMS) for the remediation of groundwater contaminated with aged leachate [90]. As illustrated in Figure 6a, the exceptional performance of NCP-activated PMS is primarily attributed to the promotion of Fe3+/Fe2+ and Cu2+/Cu+ cycling facilitated by S2−. Moreover, SO4 and •OH were identified as the dominant reactive oxygen species (ROS) in the NCP/PMS system under varying initial pH conditions [91].
Compared with single metal sulfides such as FeS and CuS, Cu/Fe bimetallic sulfides exhibit high catalytic activities due to the synergistic effect of Cu-Fe active sites. Nie et al. compared the PMS activation of BPA removal between CuFe2O4 and CuFeS2. According to reports, bimetallic sulfides are much more effective than bimetallic oxides due to the role of S2− in improving Fe3+/Fe2+and Cu2+/Cu+ cycling [80]. In addition, chalcopyrite (CuFeS2) is a copper iron bimetallic sulfide mineral and a promising heterogeneous Fenton catalyst. The synergistic effect exhibited by Cu and Fe in minerals stimulates the activation of H2O2, resulting in the production of more active hydroxyl radicals (•OH). Yang et al. investigated the use of thermally activated natural chalcopyrite for the Fenton-like degradation of organic dye Rhodamine B (RhB). This study showed that thermal activation at 300 °C changed the valence of surface elements of natural chalcopyrite, rather than altering its main chemical phase. Heat-activated chalcopyrite significantly improved the Fenton-like degradation of RhB. The Cu+ and Fe2+ species on the surface of chalcopyrite can activate H2O2 to produce • OH through electron transfer reaction, accompanied by Cu2+/Fe3+, and the reaction process is shown in Figure 6b [81].

3.2.3. Sphalerite

Sphalerite is a widely distributed natural semiconductor mineral with the theoretical chemical composition of ZnS, which is a wide-band-gap semiconductor (Eg = 3.68 eV) [90]. However, natural sphalerite generally contains various impurities, such as Fe2+, Cd2 +, and Cu2+, resulting in the narrowing of its band gap [92,93]. Therefore, natural sphalerite usually shows good photocatalytic performance under sunlight or visible-light irradiation. Lu et al. detected the chemical composition of natural sphalerite by electron microprobe analysis (EMPA) (as shown in Table 5). In addition to Zn, several other metal elements, including Fe, Cu, Cd, Ag, and Mn, were detected, and the Fe content was highest (13.74–18.65%). The results showed that the band gap (Eg) of natural sphalerite (2.8 eV) was significantly narrower than that of pure ZnS (3.4 eV). It was proposed that the Fe 3d and S 3p orbitals both contributed to the valence band of natural sphalerite and thus led to a negative shift in the valence band, thereby decreasing the band gap of natural sphalerite. Moreover, the trace non-isoelectronic substitution of Ag+ for Zn 2+ also produced a negative charge, which could serve as the capture sites for holes. As a result, the natural sphalerite exhibited a higher photocatalytic MB degradation activity than pure ZnS under visible-light irradiation [94]. Based on the photocatalytic effect of natural sphalerite, it could also be applied for the treatment of halohydrocarbons and oily wastewater. The hydrometallurgical process, especially solvent extraction technology, exhibits good application prospects in metal recovery from spent lithium-ion batteries. However, the solvent extraction process will inevitably produce oily wastewater, causing environmental issues. It was reported that natural sphalerite could achieve an efficient degradation of oily wastewater (from solvent extraction) under visible-light irradiation with the assistance of H2O2, and 86.20% total organic carbon can be removed under optimal conditions. The natural doped Fe in sphalerite could introduce an impurity level, thus narrowing the band gap of sphalerite. Meanwhile, Fe could also serve as the capture site for electrons to promote the separation of electron–holes [95,96]. Moreover, the applications of natural sphalerite in photocatalytic sterilization have also been reported [97,98].

3.3. Silicates

Silicates are widely used as catalyst support in water treatment due to their abundance, low cost, and stability. Their high surface area, developed pore structure, and strong cation exchange capacity enhance adsorption, catalysis, and reusability. Studies have focused on integrating catalysts with clay minerals to improve performance and provide additional active sites for peroxide activation in water treatment.
Kaolinite, a classic 2D layered mineral, consists of a silicon–oxygen tetrahedron layer and an aluminum–oxygen octahedron layer, with distinct basal and edge planes. Its abundant surface-bound and structural hydroxyl groups, adsorption sites, and micro-scale properties make kaolinite a promising catalyst for activating PMS. Li et al. reported that natural kaolinite’s weak photochemical activity can be enhanced by constructing Kaol/NiFe-LDO heterostructures. As shown in Figure 7a, the interfacial chemical bonds and Z-scheme band structure enable efficient PMS activation, generating sulfate radicals via photo-generated hole oxidation. Compared to natural kaolinite, the Kaol/NiFe-LDO heterostructure achieved a significantly improved degradation of RhB and orange II, emphasizing its potential in environmental applications [99]. Li et al. explored the use of natural negatively charged kaolinite with abundant hydroxyl groups to activate PMS for atrazine degradation. The results showed that hydroxyl and sulfate radicals are the key reactive species, and factors such as catalyst loading, pH, PMS dosage, and inorganic ions significantly affect the degradation efficiency. Notably, the natural kaolinite used for PMS activation possessed excellent stability, and the structure is shown in Figure 7b [100]. Zhang et al. utilized the natural layered clay mineral kaolinite as a carrier, combining it with g-C3N4 to load single iron atoms (FeSA-NGK), resulting in the fabrication of a novel composite material. This material demonstrated outstanding pollutant degradation performance under the synergistic action of PMS and visible light. The introduction of kaolinite not only increased the loading amount of single Fe atoms by 14.2% but also enhanced the concentration of N vacancies, optimizing the electronic structure and thereby significantly enhancing the catalytic activity of the material [101].
Halloysite nanotubes, a type of clay mineral, possess a distinctive nanotube morphology, predominantly composed of Al2Si2O5(OH)4·nH2O, where n = 4 for a 1.0 nm wall-packing spacing (row halloysite) and 2 for a 0.72 nm spacing (dried sample). These nanotubes exhibit diameters between 0.1 and 0.4 µm and a length that does not exceed 0.5 µm [102]. Characterized by their expansive specific surface area, distinctive hollow cylindrical form, remarkable stability, exceptional biocompatibility, and low biotoxicity, these HNTs offer remarkable adaptability across a spectrum of uses such as adsorption, energy storage, and catalytic processes [102,103]. Zhang et al. demonstrated that halloysite nanotubes (HNTs), modified with nanomanganese cobaltate (MCO@HNTs), effectively activate PMS for antibiotic degradation, as shown in Figure 3c. The system degrades ornidazole (ONZ) over a wide pH range (6.08–11.00) with minimal ion leaching and strong resistance to natural water interference. Singlet oxygen (1O2) was identified as the main ROS driving ONZ degradation [104]. Another study developed a novel catalyst using nano-CoFe2O4-decorated halloysite nanotubes (CoFe2O4/HNT), which efficiently degrades the antibiotic ornidazole (ONZ) through the activation of PMS. HNT served as a carrier to enhance the dispersion of CoFe2O4 and, due to its positively charged inner wall, facilitated the loading of anionic macromolecules, which is essential for enhancing catalytic efficiency. The CoFe2O4/HNT plus PMS system nearly completely degraded ONZ within an hour, demonstrating good pH tolerance and resistance to anion interference [105]. Zhao et al. utilized the Coulomb interaction to deposit 1,10-phenanthroline-coordinated Co2+ onto the surface of HNTs, and then incorporated atomically dispersed cobalt sites into nitrogen-doped hollow carbon nanotubes to fabricate the CoNC/NHCNTs-900 material. This material activated PMS effectively, leading to the efficient degradation of tetracycline (TC) pollutants with an unprecedented mineralization rate of 89.07%. The use of HNTs as a template prevented the aggregation of cobalt species and enhanced the material’s specific surface area, which facilitated the exposure of active sites and the transfer of reactants [106].
Montmorillonite is a monoclinic crystalline layered structure aluminosilicate clay mineral with the chemical formula (Na,Ca)0.33(Al,Mg)2Si4O10(OH)2·nH2O. It has received widespread attention due to its large specific surface area, excellent cation exchange capacity, and significant water absorption and swelling properties. Figure 7e shows the adsorption mechanism of cationic and anionic dyes on montmorillonite clay minerals [107]. The adsorption efficiency of montmorillonite can be significantly enhanced through appropriate modification. Its surface is covered with abundant hydroxyl functional groups, which can activate peroxides and act as active sites. Gao et al. enhanced the strong interaction between CoO and montmorillonite through a molten salt induction method, preparing an efficient CoO/Mt catalyst for the activation of PMS to degrade tetracycline (TC). In this work, the introduction of montmorillonite not only provided abundant surface hydroxyl groups to enhance the electronic structure of CoO nanoparticles but also significantly improved the stability and activity of the catalyst, and the reaction process is shown in Figure 7d. Treating montmorillonite with molten NaPO3 mitigated its dehydroxylation process, thereby more effectively modulating the electronic structure of CoO. This resulted in the CoO/Mt catalyst exhibiting excellent performance in the activation of PMS, achieving a 99% TC degradation efficiency within 7 min and maintaining a 96% removal rate after eight cycles, while also significantly reducing the leaching of Co ions [108]. Shi et al. developed a nanoscale zero-valent iron/montmorillonite (NZVI/Mt) composite material for the efficient removal of hexavalent chromium (Cr(VI)) from aqueous solutions. The NZVI, synthesized through a liquid-phase reduction method, when combined with natural montmorillonite, not only enhances the dispersion of NZVI to prevent aggregation but also leverages the adsorptive properties of montmorillonite to further increase the composite’s removal efficiency for Cr(VI). Under optimal conditions, the NZVI/Mt composite achieved a removal efficiency of up to 92.41% for Cr(VI) and demonstrated a maximum adsorption capacity of 67.75 mg/g [109]. Furthermore, Rehman et al. synthesized a cost-effective paired mineral–carbon material (PMC) by synergistically integrating montmorillonite, goethite, and rice-husk-derived carbon. The hierarchical porous structure (187 m2/g) with abundant defects enabled superior PMS activation, achieving rapid diethyl phthalate (DEP) degradation (k1 = 0.923 min−1, pH 6.0). DFT calculations revealed targeted radical (SO4/•OH/1O2) attack on DEP’s electron-rich sites (f0: 0.0837–0.1027). Beyond performance, the study emphasized circular economy benefits: agricultural waste (rice husk) was valorized into functional catalysts, while the PMC/PMS system significantly reduced DEP phytotoxicity, restoring the healthy growth of seedlings (biomass recovery > 90%). Eco-friendly synthesis by avoiding energy-intensive calcination steps and a radical-dominated metal-free pathway minimizes secondary pollution risks. Economically, the production cost of PMC on the lab-scale was estimated at USD 8.08/kg, 20% lower than that of commercial activated carbon (USD 10/kg), with scalable processes leveraging low-cost minerals and biomass. This work exemplifies a sustainable paradigm integrating “waste-to-resource” conversion, high-efficiency decontamination, and ecological risk mitigation, offering a competitive alternative for industrial wastewater treatment [110].
Figure 7. (a) The proposed mechanism for the improved photocatalytic ability of NFK HCs [99], Copyright © 2024, Elsevier; (b) the highly efficient activation of peroxymonosulfate by natural negatively charged kaolinite with abundant hydroxyl groups for the degradation of atrazine [100], Copyright © 2019, Elsevier; (c) nanomanganese cobaltate-decorated halloysite nanotubes for the degradation of ornidazole via peroxymonosulfate activation [104], Copyright © 2023, Elsevier; (d) Co3O4 and montmorillonite for promoted peroxymonosulfate activation toward tetracycline degradation [108], Copyright © 2024, Elsevier; (e) the mechanism of adsorption of cationic and anionic dyes onto smectite clay minerals [109], Copyright © 2022, Elsevier.
Figure 7. (a) The proposed mechanism for the improved photocatalytic ability of NFK HCs [99], Copyright © 2024, Elsevier; (b) the highly efficient activation of peroxymonosulfate by natural negatively charged kaolinite with abundant hydroxyl groups for the degradation of atrazine [100], Copyright © 2019, Elsevier; (c) nanomanganese cobaltate-decorated halloysite nanotubes for the degradation of ornidazole via peroxymonosulfate activation [104], Copyright © 2023, Elsevier; (d) Co3O4 and montmorillonite for promoted peroxymonosulfate activation toward tetracycline degradation [108], Copyright © 2024, Elsevier; (e) the mechanism of adsorption of cationic and anionic dyes onto smectite clay minerals [109], Copyright © 2022, Elsevier.
Minerals 15 00318 g007

4. Conclusions and Outlook

The application of natural minerals in environmental remediation has demonstrated substantial promise, particularly as heterogeneous catalysts for pollutant degradation. These minerals, enriched with metallic elements and endowed with distinct physicochemical properties, exhibit inherent multi-metallic compositions, defect sites, and porous structures that serve as naturally optimized platforms for regulating radical/non-radical pathways. While current research predominantly focuses on pyrite, chalcopyrite, and magnetite, emerging candidates, such as ilmenite, apatite, and tourmaline, have untapped potential due to their unique redox-active metal combinations and crystallographic defect densities. Existing studies emphasize the critical role of surface grain boundaries and redox-active metals in mediating electron transfer, yet the atomic-scale reaction kinetics and their structure–activity relationships with mineral-specific features remain poorly elucidated. Key challenges impeding technological translation include (1) poorly understood dynamic adsorption–catalysis coupling mechanisms at mineral–pollutant–oxidant ternary interfaces, particularly in municipal wastewater degradation scenarios where adsorption-dominated mineral carriers may compete with catalytic pathways; (2) the lack of a balanced assessment of energy efficiency from a lifecycle perspective; and (3) the irreversible deactivation of active sites in complex environmental matrices.
Future advancements require the synergistic integration of computational materials science and advanced characterization. For instance, DFT simulations could decide how defect engineering in minerals such as chalcopyrite redistributes electron clouds to suppress electron–hole recombination. Meanwhile, mineral-based heterojunction composites can be developed through band-gap engineering for electron transfer channel modulation. By artificially synthesizing crystal structures that mimic natural minerals (e.g., double-layered hydroxide [111,112,113], nanostructured composites [114,115]), researchers can overcome the inherent limitations of natural counterparts, achieving the precise regulation of active sites and stability [116]. This synthetic strategy and the utilization of unmodified natural minerals are complementary pillars in mineral-mediated catalysis; while synthetic approaches enable tailored control over electronic and structural properties, natural minerals inherently minimize environmental disruption due to their geochemical compatibility. To synergistically advance both pathways, future research on all-natural mineral composites must integrate multidimensional evaluation frameworks that concurrently assess synthetic analogs in terms of mining feasibility, regeneration potential, and ecotoxicological impacts. Such dual-focused efforts will bridge lab-scale innovations to engineered catalytic systems, optimizing the performance of these systems without compromising sustainability.

Author Contributions

Conceptualization, D.X. and L.G.; methodology, D.X. and Y.Y.; resources, D.X. and L.G.; writing—original draft preparation, D.X.; writing—review and editing, D.X. and Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

Di Xu is an employee of Sinomine Rock & Mineral Analysis Tianjin Co., Ltd., and Lingqun Gan is an employee of Hunan Port Shipping and Water Resources Group Co., Ltd. The paper reflects the views of the scientists and not the company.

References

  1. Lee, H.; Sam, K.; Coulon, F.; De Gisi, S.; Notarnicola, M.; Labianca, C. Recent Developments and Prospects of Sustainable Remediation Treatments for Major Contaminants in Soil: A Review. Sci. Total Environ. 2024, 912, 168769. [Google Scholar] [CrossRef] [PubMed]
  2. He, L.; Carrera, P.; Wang, C.; Duc, V.N.; Heynderickx, P.M.; Feng, R.; Guo, S.; Wu, D. Assessment of Hydrogen Peroxide, Persulfate, and Peroxymonosulfate as Oxidizing Agents in Electrochemical Oxidation of Pyridine. Chem. Eng. J. 2025, 503, 158125. [Google Scholar] [CrossRef]
  3. Babu, S.G.; Ashokkumar, M.; Neppolian, B. The Role of Ultrasound on Advanced Oxidation Processes. Top. Curr. Chem. 2016, 374, 75. [Google Scholar] [CrossRef]
  4. Verma, P.; Samanta, S.K. Microwave-Enhanced Advanced Oxidation Processes for the Degradation of Dyes in Water. Environ. Chem. Lett. 2018, 16, 969–1007. [Google Scholar] [CrossRef]
  5. Hu, K.; Zhou, P.; Yang, Y.; Hall, T.; Nie, G.; Yao, Y.; Duan, X.; Wang, S. Degradation of Microplastics by a Thermal Fenton Reaction. ACS EST Eng. 2022, 2, 110–120. [Google Scholar] [CrossRef]
  6. Chuang, Y.-H.; Chen, S.; Chinn, C.J.; Mitch, W.A. Comparing the UV/Monochloramine and UV/Free Chlorine Advanced Oxidation Processes (AOPs) to the UV/Hydrogen Peroxide AOP under Scenarios Relevant to Potable Reuse. Environ. Sci. Technol. 2017, 51, 13859–13868. [Google Scholar] [CrossRef]
  7. Liang, C.; Yin, S.; Huang, P.; Yang, S.; Wang, Z.; Zheng, S.; Li, C.; Sun, Z. The Critical Role of Minerals in Persulfate-Based Advanced Oxidation Process: Catalytic Properties, Mechanism, and Prospects. Chem. Eng. J. 2024, 482, 148969. [Google Scholar] [CrossRef]
  8. Elmaci, G.; Frey, C.E.; Kurz, P.; Zümreoğlu-Karan, B. Water Oxidation Catalysis by Birnessite@iron Oxide Core–Shell Nanocomposites. Inorg. Chem. 2015, 54, 2734–2741. [Google Scholar] [CrossRef]
  9. Huang, X.; Chen, Y.; Walter, E.; Zong, M.; Wang, Y.; Zhang, X.; Qafoku, O.; Wang, Z.; Rosso, K.M. Facet-Specific Photocatalytic Degradation of Organics by Heterogeneous Fenton Chemistry on Hematite Nanoparticles. Environ. Sci. Technol. 2019, 53, 10197–10207. [Google Scholar] [CrossRef]
  10. Zhang, J.; Wageh, S.; Al-Ghamdi, A.; Yu, J. New Understanding on the Different Photocatalytic Activity of Wurtzite and Zinc-Blende CdS. Appl. Catal. B Environ. 2016, 192, 101–107. [Google Scholar] [CrossRef]
  11. Ma, J.; Zhang, L.; Fan, Z.; Sun, S.; Feng, Z.; Li, W.; Ding, H. Construction of R-TiO2/n-TiO2 Heterophase Photocatalysts for Efficient Degradation of Organic Pollutants. J. Alloys Compd. 2023, 968, 172127. [Google Scholar] [CrossRef]
  12. Zhao, Y.; Wang, Y.; Xie, W.; Li, Z.; Zhou, Y.; Qin, R.; Wang, L.; Zhou, J.; Ren, G. One-Step-Modified Biochar by Natural Anatase for Eco-Friendly Cr(VI) Removal. Sustainability 2024, 16, 8056. [Google Scholar] [CrossRef]
  13. Gong, C.; Zhai, J.; Wang, X.; Zhu, W.; Yang, D.; Luo, Y.; Gao, X. Synergistic Improving Photo-Fenton and Photo-Catalytic Degradation of Carbamazepine over FeS2/Fe2O3/Organic Acid with H2O2 in-Situ Generation. Chemosphere 2022, 307, 136199. [Google Scholar] [CrossRef]
  14. Liu, C.; Lin, S.; Liu, Y.; Li, M.; Shen, W.; Jiang, N.; Li, F.; Tian, J. Disclosing the Influence Mechanism of Facet-Dependent Pyrite Photo-Activation and Photo-Dissolution Processes on the Reduction of Cr(VI). Environ. Pollut. 2024, 359, 124578. [Google Scholar] [CrossRef] [PubMed]
  15. Zheng, J.; Fan, C.; Li, X.; Yang, Q.; Wang, D.; Duan, A.; Pan, S. Tourmaline/ZnAL-LDH Nanocomposite Based Photocatalytic System for Efficient Degradation of Mixed Pollutant Wastewater. Sep. Purif. Technol. 2024, 345, 127306. [Google Scholar] [CrossRef]
  16. Lu, A.H. Photocatalytic Properties of Inorganic Mineral Natural Self-Purification. Acta Petrol. Mineral. 2003, 22, 323–331. [Google Scholar]
  17. Toledo, A.G.R.; Bevilaqua, D.; Panda, S.; Akcil, A. Hydrometallurgical Processing of Sulfide Minerals from the Perspective of Semiconductor Electrochemistry: A Review. Miner. Eng. 2023, 204, 108409. [Google Scholar] [CrossRef]
  18. Babuponnusami, A.; Muthukumar, K. A Review on Fenton and Improvements to the Fenton Process for Wastewater Treatment. J. Environ. Chem. Eng. 2014, 2, 557–572. [Google Scholar] [CrossRef]
  19. Neyens, E.; Baeyens, J. A Review of Classic Fenton’s Peroxidation as an Advanced Oxidation Technique. J. Hazard. Mater. 2003, 98, 33–50. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Zhu, R.; Xi, Y.; Zhu, J.; Zhu, G.; He, H. Strategies for Enhancing the Heterogeneous Fenton Catalytic Reactivity: A Review. Appl. Catal. B Environ. 2019, 255, 117739. [Google Scholar] [CrossRef]
  21. Wang, L.; Luo, D.; Hamdaoui, O.; Vasseghian, Y.; Momotko, M.; Boczkaj, G.; Kyzas, G.Z.; Wang, C. Bibliometric Analysis and Literature Review of Ultrasound-Assisted Degradation of Organic Pollutants. Sci. Total Environ. 2023, 876, 162551. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, J.; Tang, J. Fe-Based Fenton-like Catalysts for Water Treatment: Preparation, Characterization and Modification. Chemosphere 2021, 276, 130177. [Google Scholar] [CrossRef]
  23. Liu, Y.; Wang, J. Multivalent Metal Catalysts in Fenton/Fenton-like Oxidation System: A Critical Review. Chem. Eng. J. 2023, 466, 143147. [Google Scholar] [CrossRef]
  24. Pan, S.; Cao, B.; Yuan, D.; Jiao, T.; Zhang, Q.; Tang, S. Complexes of Cupric Ion and Tartaric Acid Enhanced Calcium Peroxide Fenton-like Reaction for Metronidazole Degradation. Chin. Chem. Lett. 2024, 35, 109185. [Google Scholar] [CrossRef]
  25. Lee, H.; Seong, J.; Lee, K.-M.; Kim, H.-H.; Choi, J.; Kim, J.-H.; Lee, C. Chloride-Enhanced Oxidation of Organic Contaminants by Cu(II)-Catalyzed Fenton-like Reaction at Neutral pH. J. Hazard. Mater. 2018, 344, 1174–1180. [Google Scholar] [CrossRef] [PubMed]
  26. Luo, X.; Hu, H.; Pan, Z.; Pei, F.; Qian, H.; Miao, K.; Guo, S.; Wang, W.; Feng, G. Efficient and Stable Catalysis of Hollow Cu9S5 Nanospheres in the Fenton-like Degradation of Organic Dyes. J. Hazard. Mater. 2020, 396, 122735. [Google Scholar] [CrossRef]
  27. Li, T.; Wang, X.; Chen, Y.; Liang, J.; Zhou, L. Producing OH, SO4 and O2 in Heterogeneous Fenton Reaction Induced by Fe3O4-Modified Schwertmannite. Chem. Eng. J. 2020, 393, 124735. [Google Scholar] [CrossRef]
  28. Chen, F.; Huang, X.-T.; Bai, C.-W.; Zhang, Z.-Q.; Duan, P.-J.; Sun, Y.-J.; Chen, X.-J.; Zhang, B.-B.; Zhang, Y.-S. Advancements in Heterogeneous Activation of Persulfates: Exploring Mechanisms, Challenges in Organic Wastewater Treatment, and Innovative Solutions. Chem. Eng. J. 2024, 481, 148789. [Google Scholar] [CrossRef]
  29. Ji, Y.; Dong, C.; Kong, D.; Lu, J.; Zhou, Q. Heat-Activated Persulfate Oxidation of Atrazine: Implications for Remediation of Groundwater Contaminated by Herbicides. Chem. Eng. J. 2015, 263, 45–54. [Google Scholar] [CrossRef]
  30. Kang, D.W.; Kim, J.H.; Lim, J.H.; Kim, Y.; Kang, M.; Shin, J.; Son, S.; Yun, H.; Kim, H.; Park, S.; et al. Promoted Type I and II ROS Generation by a Covalent Organic Framework through Sonosensitization and PMS Activation. ACS Catal. 2022, 12, 9621–9628. [Google Scholar] [CrossRef]
  31. Anipsitakis, G.P.; Dionysiou, D.D. Radical Generation by the Interaction of Transition Metals with Common Oxidants. Environ. Sci. Technol. 2004, 38, 3705–3712. [Google Scholar] [CrossRef] [PubMed]
  32. Kohantorabi, M.; Moussavi, G.; Giannakis, S. A Review of the Innovations in Metal- and Carbon-Based Catalysts Explored for Heterogeneous Peroxymonosulfate (PMS) Activation, with Focus on Radical vs. Non-Radical Degradation Pathways of Organic Contaminants. Chem. Eng. J. 2021, 411, 127957. [Google Scholar] [CrossRef]
  33. Lee, J.; von Gunten, U.; Kim, J.-H. Persulfate-Based Advanced Oxidation: Critical Assessment of Opportunities and Roadblocks. Environ. Sci. Technol. 2020, 54, 3064–3081. [Google Scholar] [CrossRef]
  34. Kim, J.; Du, P.; Liu, W.; Luo, C.; Zhao, H.; Huang, C.-H. Cobalt/Peracetic Acid: Advanced Oxidation of Aromatic Organic Compounds by Acetylperoxyl Radicals. Environ. Sci. Technol. 2020, 54, 5268–5278. [Google Scholar] [CrossRef]
  35. Xing, D.; Shao, S.; Yang, Y.; Zhou, Z.; Jing, G.; Zhao, X. Mechanistic Insights into the Efficient Activation of Peracetic Acid by Pyrite for the Tetracycline Abatement. Water Res. 2022, 222, 118930. [Google Scholar] [CrossRef] [PubMed]
  36. Tian, D.; Zhou, H.; Zhang, H.; Zhou, P.; You, J.; Yao, G.; Pan, Z.; Liu, Y.; Lai, B. Heterogeneous Photocatalyst-Driven Persulfate Activation Process under Visible Light Irradiation: From Basic Catalyst Design Principles to Novel Enhancement Strategies. Chem. Eng. J. 2022, 428, 131166. [Google Scholar] [CrossRef]
  37. Sundaram, I.M.; Kalimuthu, S.; Ponniah, G. Highly Active ZnO Modified g-C3N4 Nanocomposite for Dye Degradation under UV and Visible Light with Enhanced Stability and Antimicrobial Activity. Compos. Commun. 2017, 5, 64–71. [Google Scholar] [CrossRef]
  38. Huang, W.; Xiao, S.; Zhong, H.; Yan, M.; Yang, X. Activation of Persulfates by Carbonaceous Materials: A Review. Chem. Eng. J. 2021, 418, 129297. [Google Scholar] [CrossRef]
  39. Wang, D.; Xu, H.; Ma, J.; Lu, X.; Qi, J.; Song, S. Strong Promoted Catalytic Ozonation of Atrazine at Low Temperature Using Tourmaline as Catalyst: Influencing Factors, Reaction Mechanisms and Pathways. Chem. Eng. J. 2018, 354, 113–125. [Google Scholar] [CrossRef]
  40. Wang, X.; Jia, Y.; Wang, Y.; Xu, X.; Qin, L.; Wu, Z. Natural Piezoelectric Tourmaline Mineral for Piezocatalytic Decomposition of Organic Dyes under Vibration. J. Am. Ceram. Soc. 2024, 107, 1682–1690. [Google Scholar] [CrossRef]
  41. Abazari, R.; Heshmatpour, F.; Balalaie, S. Pt/Pd/Fe Trimetallic Nanoparticle Produced via Reverse Micelle Technique: Synthesis, Characterization, and Its Use as an Efficient Catalyst for Reductive Hydrodehalogenation of Aryl and Aliphatic Halides under Mild Conditions. ACS Catal. 2013, 3, 139–149. [Google Scholar] [CrossRef]
  42. Belaidi, S.; Mammeri, L.; Mechakra, H.; Remache, W.; Benhamouda, K.; Larouk, S.; Kribeche, M.A.; Sehili, T. UV and Solar Light Induced Natural Iron Oxide Activation: Characterization and Photocatalytic Degradation of Organic Compounds. Int. J. Chem. React. Eng. 2019, 17, 20180027. [Google Scholar] [CrossRef]
  43. Huang, H.H.; Lu, M.C.; Chen, J.-N. Catalytic Decomposition of Hydrogen Peroxide and 2-Chlorophenol with Iron Oxides. Water Res. 2001, 35, 2291–2299. [Google Scholar] [CrossRef] [PubMed]
  44. Garrido-Ramírez, E.G.; Theng, B.K.G.; Mora, M.L. Clays and Oxide Minerals as Catalysts and Nanocatalysts in Fenton-like Reactions—A Review. Appl. Clay Sci. 2010, 47, 182–192. [Google Scholar] [CrossRef]
  45. Liu, J.; Zhu, R.; Chen, Q.; Zhou, H.; Liang, X.; Ma, L.; Parker, S.C. The Significant Effect of Photo-Catalyzed Redox Reactions on the Immobilization of Chromium by Hematite. Chem. Geol. 2019, 524, 228–236. [Google Scholar] [CrossRef]
  46. Choi, J.; Choi, W.; Hwang, H.; Tang, Y.; Jung, H. Natural Sunlight-Driven Oxidation of Mn2+(Aq) and Heterogeneous Formation of Mn Oxides on Hematite. Chemosphere 2024, 348, 140734. [Google Scholar] [CrossRef]
  47. Chen, N.; Wan, Y.; Zhan, G.; Wang, X.; Li, M.; Zhang, L. Simulated Solar Light Driven Roxarsone Degradation and Arsenic Immobilization with Hematite and Oxalate. Chem. Eng. J. 2020, 384, 123254. [Google Scholar] [CrossRef]
  48. Huang, X.; Zhao, Q.; Young, R.P.; Zhang, X.; Walter, E.D.; Chen, Y.; Nakouzi, E.; Taylor, S.D.; Loring, J.S.; Wang, Z.; et al. Photo-Production of Reactive Oxygen Species and Degradation of Dissolved Organic Matter by Hematite Nanoplates Functionalized by Adsorbed Oxalate. Environ. Sci. Nano 2020, 7, 2278–2292. [Google Scholar] [CrossRef]
  49. Liu, X.; Li, Z.; Jin, L.; Wang, H.; Huang, Y.; Huang, D.; Liu, X. Peracetic Acid Activation by Modified Hematite for Water Purification: Performance, Degradation Pathways, and Mechanism. Langmuir 2024, 40, 15301–15309. [Google Scholar] [CrossRef]
  50. Bel Hadjltaief, H.; Sdiri, A.; Gálvez, M.E.; Zidi, H.; Da Costa, P.; Ben Zina, M. Natural Hematite and Siderite as Heterogeneous Catalysts for an Effective Degradation of 4-Chlorophenol via Photo-Fenton Process. ChemEngineering 2018, 2, 29. [Google Scholar] [CrossRef]
  51. Dai, M.; Dong, X.; Yang, Y.; Wu, Y.; Chen, L.; Jiang, C.; Guo, Z.; Yang, T. Mechanistic Insight into the Impact of Interaction between Goethite and Humic Acid on the Photooxidation and Photoreduction of Bifenthrin. Environ. Res. 2024, 252, 118779. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, X.; Liu, H.; Zong, M.; Chen, M.; He, H.; Wang, R.; Lu, X. Mn Substitution and Distribution in Goethite and Influences on Its Photocatalytic Properties: A Combined Study Using First-Principles Calculations and Photocatalytic Experiments. Am. Mineral. 2023, 108, 968–977. [Google Scholar] [CrossRef]
  53. Hu, S.; Li, H.; Wang, P.; Liu, C.; Shi, Z.; Li, F.; Liu, T. Interfacial Photoreactions of Cr(VI) and Oxalate on Lepidocrocite Surface under Oxic and Acidic Conditions: Reaction Mechanism and Potential Implications for Contaminant Degradation in Surface Waters. Chem. Geol. 2021, 583, 120481. [Google Scholar] [CrossRef]
  54. Wang, C.; Liu, H.; Li, X.; Cao, Y.; Jia, K. Degradation of Organic Pollutants in Natural Magnetite-Activated Peroxymonosulfate System Enhanced by Hydroxylamine: Catalytic Performance, Reaction Pathway, and Enhancement Mechanism. J. Water Process Eng. 2024, 66, 106000. [Google Scholar] [CrossRef]
  55. Liu, H.; Li, C.; Zhang, T.; Xu, Z.; Li, Y.; Li, B.; Tian, S. UV Facilitated Synergistic Effects of Polymetals in Ore Catalyst on Peroxymonosulfate Activation: Implication for the Degradation of Bisphenol S. Chem. Eng. J. 2022, 431, 133989. [Google Scholar] [CrossRef]
  56. Xia, D.; He, H.; Liu, H.; Wang, Y.; Zhang, Q.; Li, Y.; Lu, A.; He, C.; Wong, P.K. Persulfate-Mediated Catalytic and Photocatalytic Bacterial Inactivation by Magnetic Natural Ilmenite. Appl. Catal. B Environ. 2018, 238, 70–81. [Google Scholar] [CrossRef]
  57. Wang, Y.; Wang, J.; Ding, Z.; Wang, W.; Song, J.; Li, P.; Liang, J.; Fan, Q. Light Promotes the Immobilization of U(VI) by Ferrihydrite. Molecules 2022, 27, 1859. [Google Scholar] [CrossRef] [PubMed]
  58. Li, Y.; Xu, X.; Li, Y.; Ding, C.; Wu, J.; Lu, A.; Ding, H.; Qin, S.; Wang, C. Absolute Band Structure Determination on Naturally Occurring Rutile with Complex Chemistry: Implications for Mineral Photocatalysis on Both Earth and Mars. Appl. Surf. Sci. 2018, 439, 660–671. [Google Scholar] [CrossRef]
  59. Lu, A.; Li, Y.; Lv, M.; Wang, C.; Yang, L.; Liu, J.; Wang, Y.; Wong, K.-H.; Wong, P.-K. Photocatalytic Oxidation of Methyl Orange by Natural V-Bearing Rutile under Visible Light. Sol. Energy Mater. Sol. Cells 2007, 91, 1849–1855. [Google Scholar] [CrossRef]
  60. Jung, H.; Snyder, C.; Xu, W.; Wen, K.; Zhu, M.; Li, Y.; Lu, A.; Tang, Y. Photocatalytic Oxidation of Dissolved Mn2+ by TiO2 and the Formation of Tunnel Structured Manganese Oxides. ACS Earth Space Chem. 2021, 5, 2105–2114. [Google Scholar] [CrossRef]
  61. Xiao, Y.; Li, Y.; Ding, H.; Li, Y.; Lu, A. The Fine Characterization and Potential Photocatalytic Effect of Semiconducting Metal Minerals in Danxia Landforms. Minerals 2018, 8, 554. [Google Scholar] [CrossRef]
  62. Zhu, S.; Ho, S.-H.; Jin, C.; Duan, X.; Wang, S. Nanostructured Manganese Oxides: Natural/Artificial Formation and Their Induced Catalysis for Wastewater Remediation. Environ. Sci. Nano 2020, 7, 368–396. [Google Scholar] [CrossRef]
  63. Crowther, D.L.; Dillard, J.G.; Murray, J.W. The Mechanisms of Co(II) Oxidation on Synthetic Birnessite. Geochim. Cosmochim. Acta 1983, 47, 1399–1403. [Google Scholar] [CrossRef]
  64. Scott, M.J.; Morgan, J.J. Reactions at Oxide Surfaces. 2. Oxidation of Se(IV) by Synthetic Birnessite. Environ. Sci. Technol. 1996, 30, 1990–1996. [Google Scholar] [CrossRef]
  65. Rajapaksha, A.U.; Vithanage, M.; Ok, Y.S.; Oze, C. Cr(VI) Formation Related to Cr(III)-Muscovite and Birnessite Interactions in Ultramafic Environments. Environ. Sci. Technol. 2013, 47, 9722–9729. [Google Scholar] [CrossRef]
  66. Abernathy, M.J.; Schaefer, M.V.; Vessey, C.J.; Liu, H.; Ying, S.C. Oxidation of V(IV) by Birnessite: Kinetics and Surface Complexation. Environ. Sci. Technol. 2021, 55, 11703–11712. [Google Scholar] [CrossRef] [PubMed]
  67. Li, P.; Wang, Y.; Wang, J.; Wang, W.; Ding, Z.; Liang, J.; Fan, Q. The Photocatalytic Oxidation of As(III) on Birnessite. npj Clean Water 2024, 7, 19. [Google Scholar] [CrossRef]
  68. Ghanbari, S.; Fatehizadeh, A.; Ebrahimi, A.; Bina, B.; Taheri, E.; Iqbal, H.M.N. Hydrothermally Improved Natural Manganese-Containing Catalytic Materials to Degrade 4-Chlorophenol. Environ. Res. 2023, 226, 115641. [Google Scholar] [CrossRef]
  69. Yao, Z.; Chen, R.; Han, N.; Sun, H.; Wong, N.H.; Ernawati, L.; Wang, S.; Sunarso, J.; Liu, S. Natural Manganese Ores for Efficient Removal of Organic Pollutants via Catalytic Peroxymonosulfate-Based Advanced Oxidation Processes. Asia-Pac. J. Chem. Eng. 2023, 18, e2907. [Google Scholar] [CrossRef]
  70. Zhong, C.; Jiang, Y.; Liu, Q.; Sun, X.; Yu, J. Natural Siderite Derivatives Activated Peroxydisulfate toward Oxidation of Organic Contaminant: A Green Soil Remediation Strategy. J. Environ. Sci. 2023, 127, 615–627. [Google Scholar] [CrossRef]
  71. Ding, J.; Shen, L.; Yan, R.; Lu, S.; Zhang, Y.; Zhang, X.; Zhang, H. Heterogeneously Activation of H2O2 and Persulfate with Goethite for Bisphenol a Degradation: A Mechanistic Study. Chemosphere 2020, 261, 127715. [Google Scholar] [CrossRef]
  72. Tian, L.; Wang, L.; Wei, S.; Zhang, L.; Dong, D.; Guo, Z. Enhanced Degradation of Enoxacin Using Ferrihydrite-Catalyzed Heterogeneous Photo-Fenton Process. Environ. Res. 2024, 251, 118650. [Google Scholar] [CrossRef] [PubMed]
  73. Yuan, J.; Ding, Z.; Gong, X.; Yu, A.; Wen, S.; Bai, S. Degradation of Residual Xanthate in Flotation Tailing Wastewater by Activated H2O2 with Natural Ilmenite: A Heterogeneous Fenton-like Reaction Catalyzed at Natural pH. Sep. Purif. Technol. 2024, 332, 125819. [Google Scholar] [CrossRef]
  74. Zhu, S.; Li, Z.; Yu, M.; Wang, Q.; Chen, C.; Ma, J. Efficient Removal of Naphthenic Acids from Real Petroleum Wastewater by Natural Pyrite Activated Persulfate System. J. Environ. Manag. 2023, 348, 119239. [Google Scholar] [CrossRef]
  75. Zheng, X.; Niu, X.; Zhang, D.; Ye, X.; Ma, J.; Lv, M.; Lin, Z. Removal of Microcystis Aeruginosa by Natural Pyrite-Activated Persulfate: Performance and the Significance of Iron Species. Chem. Eng. J. 2022, 428, 132565. [Google Scholar] [CrossRef]
  76. Han, Y.; Sun, S.; Zhang, B.; Du, J.; Duan, X. Activation of Peroxymonosulfate by Natural Pyrite for Efficient Degradation of V(IV)-Citrate Complex in Groundwater. J. Colloid Interface Sci. 2022, 617, 683–693. [Google Scholar] [CrossRef]
  77. Mashayekh-Salehi, A.; Akbarmojeni, K.; Roudbari, A.; van der Hoek, J.P.; Nabizadeh, R.; Dehghani, M.H.; Yaghmaeian, K. Use of Mine Waste for H2O2-Assisted Heterogeneous Fenton-like Degradation of Tetracycline by Natural Pyrite Nanoparticles: Catalyst Characterization, Degradation Mechanism, Operational Parameters and Cytotoxicity Assessment. J. Clean. Prod. 2021, 291, 125235. [Google Scholar] [CrossRef]
  78. Peng, S.; Feng, Y.; Liu, Y.; Wu, D. Applicability Study on the Degradation of Acetaminophen via an H2O2/PDS-Based Advanced Oxidation Process Using Pyrite. Chemosphere 2018, 212, 438–446. [Google Scholar] [CrossRef]
  79. Li, C.; Yuan, D.; Yang, K.; Wang, H.; Wang, Z.; Zhang, Q.; Tang, S. Reutilization of Pyrite Tailings in Peracetic Acid-Based Advanced Oxidation Process for Water Purification. Sep. Purif. Technol. 2025, 354, 129155. [Google Scholar] [CrossRef]
  80. Nie, W.; Mao, Q.; Ding, Y.; Hu, Y.; Tang, H. Highly Efficient Catalysis of Chalcopyrite with Surface Bonded Ferrous Species for Activation of Peroxymonosulfate toward Degradation of Bisphenol a: A Mechanism Study. J. Hazard. Mater. 2019, 364, 59–68. [Google Scholar] [CrossRef]
  81. Yang, J.; Jia, K.; Lu, S.; Cao, Y.; Boczkaj, G.; Wang, C. Thermally Activated Natural Chalcopyrite for Fenton-like Degradation of Rhodamine B: Catalyst Characterization, Performance Evaluation, and Catalytic Mechanism. J. Environ. Chem. Eng. 2024, 12, 111469. [Google Scholar] [CrossRef]
  82. Yang, K.; Zhai, Z.; Liu, H.; Zhao, T.; Yuan, D.; Jiao, T.; Zhang, Q.; Tang, S. Peracetic Acid Activation by Natural Chalcopyrite for Metronidazole Degradation: Unveiling the Effects of Cu-Fe Bimetallic Sites and Sulfur Species. Sep. Purif. Technol. 2023, 305, 122500. [Google Scholar] [CrossRef]
  83. Yuan, T.; Wang, X.; Zhao, X.; Liu, T.; Zhang, H.; Lv, Y.; Wang, L. Efficient Degradation of Minocycline by Natural Bornite-Activated Hydrogen Peroxide and Persulfate: Kinetics and Mechanisms. Environ. Sci. Pollut. Res. 2021, 28, 69314–69328. [Google Scholar] [CrossRef]
  84. Yang, R.; Zeng, G.; Zhou, Z.; Xu, Z.; Lyu, S. Naphthalene Degradation Dominated by Homogeneous Reaction in Fenton-like Process Catalyzed by Pyrite: Mechanism and Application. Sep. Purif. Technol. 2023, 310, 123150. [Google Scholar] [CrossRef]
  85. Wang, C.; Sun, R.; Huang, R.; Cao, Y. A Novel Strategy for Enhancing Heterogeneous Fenton Degradation of Dye Wastewater Using Natural Pyrite: Kinetics and Mechanism. Chemosphere 2021, 272, 129883. [Google Scholar] [CrossRef]
  86. Dai, H.; Huang, Z.; Wang, C.; Hu, Y.; Xu, X.; Zhao, F.; Pi, Y.; Qian, J.; Hu, F.; Peng, X. Activation of PAA by Pyrite for the Degradation of Tetracycline with Adjusted pH Conditions: The Overlooked Role of Sulfur species. Sep. Purif. Technol. 2025, 361, 131225. [Google Scholar] [CrossRef]
  87. Qiu, R.; Zhang, P.; Zhang, Z.; Wang, C.; Wang, Q.; Rončević, S.D.; Sun, H. The Interface Mechanism of Ball-Milled Natural Pyrite Activating Persulfate to Degrade Monochlorobenzene in Soil: Intrinsic Synergism of S and Fe Species. Sep. Purif. Technol. 2024, 341, 126946. [Google Scholar] [CrossRef]
  88. Sun, Y.; Zhang, W.; Xiang, W.; Liu, X.; Wu, X.; Zhou, T. Catalytic Activation of Persulfate by Mechanically-Treated Natural Pyrite: Revisit on the Dominant Role of Sulfur Vacancies and Size Effect. Chem. Eng. J. 2022, 450, 138269. [Google Scholar] [CrossRef]
  89. Wu, C.; Su, L.; Zhang, B.; Wang, X.; Wang, M.; Zhang, J.; Wang, Q. Ball Milling of Pyrite in Air to Significantly Promote the Fenton Reaction: A Mechanistic Study. J. Hazard. Mater. 2024, 480, 136081. [Google Scholar] [CrossRef]
  90. Khosravi, A.A.; Kundu, M.; Jatwa, L.; Deshpande, S.K.; Bhagwat, U.A.; Sastry, M.; Kulkarni, S.K. Green Luminescence from Copper Doped Zinc Sulphide Quantum Particles. Appl. Phys. Lett. 1995, 67, 2702–2704. [Google Scholar] [CrossRef]
  91. Wang, H.; Liao, B.; Hu, M.; Ai, Y.; Wen, L.; Yang, S.; Ye, Z.; Qin, J.; Liu, G. Heterogeneous Activation of Peroxymonosulfate by Natural Chalcopyrite for Efficient Remediation of Groundwater Polluted by Aged Landfill Leachate. Appl. Catal. B Environ. 2022, 300, 120744. [Google Scholar] [CrossRef]
  92. Zhou, B.X.; Zhou, Z.G.; Huang, G.F.; Xiong, D.N.; Yan, Q.; Huang, W.Q.; Wang, Q.L. Mass Production of ZnxCd1−xS Nanoparticles with Enhanced Visible Light Photocatalytic Activity. Mater. Lett. 2015, 158, 432–435. [Google Scholar] [CrossRef]
  93. Lee, G.-J.; Anandan, S.; Masten, S.J.; Wu, J.J. Photocatalytic Hydrogen Evolution from Water Splitting Using Cu Doped ZnS Microspheres under Visible Light Irradiation. Renew. Energy 2016, 89, 18–26. [Google Scholar] [CrossRef]
  94. Li, Y.; Lu, A.; Wang, C.; Wu, X. Characterization of Natural Sphalerite as a Novel Visible Light-Driven Photocatalyst. Sol. Energy Mater. Sol. Cells 2008, 92, 953–959. [Google Scholar] [CrossRef]
  95. Yang, X.; Li, Y.; Lu, A.; Yan, Y.; Wang, C.; Wong, P.-K. Photocatalytic Reduction of Carbon Tetrachloride by Natural Sphalerite under Visible Light Irradiation. Sol. Energy Mater. Sol. Cells 2011, 95, 1915–1921. [Google Scholar] [CrossRef]
  96. Wu, M.; Song, S.; Wang, T.; Sun, W.; Xu, S.; Yang, Y. Natural Sphalerite Photocatalyst for Treatment of Oily Wastewater Produced by Solvent Extraction from Spent Lithium-Ion Battery Recycling. Appl. Catal. B Environ. 2022, 313, 121460. [Google Scholar] [CrossRef]
  97. Wang, W.; Liu, Y.; Li, G.; Liu, Z.; Wong, P.K.; An, T. Mechanism Insights into Bacterial Sporulation at Natural Sphalerite Interface with and without Light Irradiation: The Suppressing Role in Bacterial Sporulation by Photocatalysis. Environ. Int. 2022, 168, 107460. [Google Scholar] [CrossRef] [PubMed]
  98. Chen, Y.; Lu, A.; Li, Y.; Zhang, L.; Yip, H.Y.; Zhao, H.; An, T.; Wong, P.-K. Naturally Occurring Sphalerite as a Novel Cost-Effective Photocatalyst for Bacterial Disinfection under Visible Light. Environ. Sci. Technol. 2011, 45, 5689–5695. [Google Scholar] [CrossRef]
  99. Li, C.; Jing, H.; Lv, B.; Zhou, D.; Fu, S.; Tang, X.; Wang, Z.; Wu, W.; Jiang, D. Interfacial Chemical Bonds Enhanced Photocatalytic Performance of Kaolinite/NiFe Layered Double Oxide Heterostructure for Photocatalytic Activated Peroxymonosulfate. Appl. Clay Sci. 2024, 260, 107520. [Google Scholar] [CrossRef]
  100. Li, C.; Huang, Y.; Dong, X.; Sun, Z.; Duan, X.; Ren, B.; Zheng, S.; Dionysiou, D.D. Highly Efficient Activation of Peroxymonosulfate by Natural Negatively-Charged Kaolinite with Abundant Hydroxyl Groups for the Degradation of Atrazine. Appl. Catal. B Environ. 2019, 247, 10–23. [Google Scholar] [CrossRef]
  101. Zhang, X.; Li, C.; Wang, X.; Yang, S.; Tan, Y.; Yuan, F.; Zheng, S.; Dionysiou, D.D.; Sun, Z. Defect Engineering Modulated Iron Single Atoms with Assist of Layered Clay for Enhanced Advanced Oxidation Processes. Small 2022, 18, 2204793. [Google Scholar] [CrossRef] [PubMed]
  102. Li, Y.; Yuan, X.; Jiang, L.; Dai, H.; Zhao, Y.; Guan, X.; Bai, J.; Wang, H. Manipulation of the Halloysite Clay Nanotube Lumen for Environmental Remediation: A Review. Environ. Sci. Nano 2022, 9, 841–866. [Google Scholar] [CrossRef]
  103. Massaro, M.; Colletti, C.G.; Lazzara, G.; Milioto, S.; Noto, R.; Riela, S. Halloysite Nanotubes as Support for Metal-Based Catalysts. J. Mater. Chem. A Mater. Energy Sustain. 2017, 5, 13276–13293. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Li, Y.; Bi, H.; Zhou, S.; Chen, J.; Zhang, S.; Huang, Y.; Chang, F.; Zhang, H.; Wågberg, T.; et al. Nanomanganese Cobaltate-Decorated Halloysite Nanotubes for the Complete Degradation of Ornidazole via Peroxymonosulfate Activation. J. Colloid Interface Sci. 2023, 630, 855–866. [Google Scholar] [CrossRef]
  105. Zhang, L.; Dong, G.; Yang, Y.; Niu, Y.; Gao, W.; Li, Z. Enhanced Ornidazole Degradation via Peroxymonosulfate Activation Using Nano CoFe2O4-Decorated Halloysite Nanotubes: A High-Efficiency and Stable Catalyst Approach. Adv. Sustain. Syst. 2024, 9, 2400677. [Google Scholar] [CrossRef]
  106. Zhao, X.; Duan, L.; Chen, M.; Yang, P.; Liu, Q.; Liu, Y.; Zhang, H.; He, Z.; Hu, G.; Zhou, Y. Regulating the Microenvironment of Atomically Dispersed Cobalt to Achieve the Record-Breaking Mineralization and 100. Chem. Eng. J. 2024, 485, 149928. [Google Scholar] [CrossRef]
  107. Al Kausor, M.; Sen Gupta, S.; Bhattacharyya, K.G.; Chakrabortty, D. Montmorillonite and Modified Montmorillonite as Adsorbents for Removal of Water Soluble Organic Dyes: A Review on Current Status of the Art. Inorg. Chem. Commun. 2022, 143, 109686. [Google Scholar] [CrossRef]
  108. Gao, C.; Zhao, G.; Zuo, X.; Yang, H. Molten Salt Induced Strong Interaction between Co3O4 and Montmorillonite for the Promoted Peroxymonosulfate Activation toward Tetracycline Degradation. Appl. Clay Sci. 2024, 255, 107414. [Google Scholar] [CrossRef]
  109. Shi, Y.; Wang, X.; Zhong, S.; Chen, W.; Feng, C.; Yang, S. Nano Zero-Valent Iron/Montmorillonite Composite for the Removal of Cr(VI) from Aqueous Solutions: Characterization, Performance, and Mechanistic Insights. Appl. Clay Sci. 2024, 253, 107345. [Google Scholar] [CrossRef]
  110. Rehman, S.; Sun, H.; Hanif, M.; Peng, T.; Tang, X.; Satti, O.T.; Bilal, M. Environmentally Friendly Nitrogen Doped Paired Mineral-Carbon for Catalytic Degradation of Diethyl Phthalate and Crop Damage Mitigation. Chem. Eng. J. 2025, 508, 160321. [Google Scholar] [CrossRef]
  111. Chen, G.; Nengzi, L.; Li, B.; Gao, Y.; Zhu, G.; Cheng, X. Octadecylamine Degradation through Catalytic Activation of Peroxymonosulfate by FeMn Layered Double Hydroxide. Sci. Total Environ. 2019, 695, 133963. [Google Scholar] [CrossRef] [PubMed]
  112. Hou, L.; Li, X.; Yang, Q.; Chen, F.; Wang, S.; Ma, Y.; Wu, Y.; Zhu, X.; Huang, X.; Wang, D. Heterogeneous Activation of Peroxymonosulfate Using Mn-Fe Layered Double Hydroxide: Performance and Mechanism for Organic Pollutant Degradation. Sci. Total Environ. 2019, 663, 453–464. [Google Scholar] [CrossRef] [PubMed]
  113. Taoufik, N.; Sadiq, M.; Abdennouri, M.; Qourzal, S.; Khataee, A.; Sillanpää, M.; Barka, N. Recent Advances in the Synthesis and Environmental Catalytic Applications of Layered Double Hydroxides-Based Materials for Degradation of Emerging Pollutants through Advanced Oxidation Processes. Mater. Res. Bull. 2022, 154, 111924. [Google Scholar] [CrossRef]
  114. Li, D.; Yao, Z.; Lin, J.; Tian, W.; Zhang, H.; Duan, X.; Wang, S. Nano-Sized FeVO4·1.1H2O and FeVO4 for Peroxymonosulfate Activation towards Enhanced Photocatalytic Activity. J. Environ. Chem. Eng. 2022, 10, 107199. [Google Scholar] [CrossRef]
  115. Hareendran, A.; Dais, E.; Shinoy, D.; Srikripa, S.; Shibu, G.M.; Kurian, M. Nitrogen- and Sulfur-Doped Zinc Ferrite Nanoparticles as Efficient Heterogeneous Catalysts in Advanced Oxidation Processes. J. Phys. Chem. Solids 2022, 161, 110398. [Google Scholar] [CrossRef]
  116. Mao, L.B.; Meng, Y.F.; Meng, X.S.; Yang, B.; Yang, Y.L.; Lu, Y.J.; Yang, Z.Y.; Shang, L.M.; Yu, S.-H. Matrix-Directed Mineralization for Bulk Structural Materials. J. Am. Chem. Soc. 2022, 144, 18175–18194. [Google Scholar] [CrossRef]
Figure 1. (a) Photo-catalyzed redox reactions on the immobilization of chromium by hematite [45], Copyright © 2019, Elsevier; (b) simulated solar-light-driven roxarsone degradation and arsenic immobilization with hematite and oxalate [47], Copyright © 2020, Elsevier; (c) the possible molecular structures of outer- and inner-sphere complexes of iron–oxalate on the hematite (001) surface with Fe-terminations, and a conceptual schematic representation of iron photochemical cycling coupling with reactive oxygen species generation [48], Copyright © 2014, Royal Society of Chemistry.
Figure 1. (a) Photo-catalyzed redox reactions on the immobilization of chromium by hematite [45], Copyright © 2019, Elsevier; (b) simulated solar-light-driven roxarsone degradation and arsenic immobilization with hematite and oxalate [47], Copyright © 2020, Elsevier; (c) the possible molecular structures of outer- and inner-sphere complexes of iron–oxalate on the hematite (001) surface with Fe-terminations, and a conceptual schematic representation of iron photochemical cycling coupling with reactive oxygen species generation [48], Copyright © 2014, Royal Society of Chemistry.
Minerals 15 00318 g001
Figure 2. Schematic illustration of the effect of HA on the photocatalytic process of Gt and the generation mechanisms of oxidized and reducing species in UV-Gt and UV-Gt-HA systems [51], Copyright © 2024, Elsevier.
Figure 2. Schematic illustration of the effect of HA on the photocatalytic process of Gt and the generation mechanisms of oxidized and reducing species in UV-Gt and UV-Gt-HA systems [51], Copyright © 2024, Elsevier.
Minerals 15 00318 g002
Figure 4. Schematic diagram of As(III) oxidation by birnessite [67], Copyright © 2024, Springer Nature.
Figure 4. Schematic diagram of As(III) oxidation by birnessite [67], Copyright © 2024, Springer Nature.
Minerals 15 00318 g004
Figure 6. (a) The reaction mechanism of the ball-milled pyrite/PDS system [87], Copyright © 2022, Elsevier. (b) The mechanism of RhB degradation in the thermally activated chalcopyrite/H2O2 system [81], Copyright © 2024, Elsevier.
Figure 6. (a) The reaction mechanism of the ball-milled pyrite/PDS system [87], Copyright © 2022, Elsevier. (b) The mechanism of RhB degradation in the thermally activated chalcopyrite/H2O2 system [81], Copyright © 2024, Elsevier.
Minerals 15 00318 g006
Table 1. Photocatalytic characteristic parameters of NMMS.
Table 1. Photocatalytic characteristic parameters of NMMS.
Mineral MaterialsBand Gap Eg/eVWavelength/nm
Hematite (Fe2O3)2.20565
Goethite (FeOOH)2.60478
Anatase (TiO2)3.20388
Pyrolusite (MnO2)0.254972
Ilmenite (FeTiO3)2.80444
Pyrite (FeS2)0.951309
Alabandite (MnS)3.00414
Molybdenite (MoS)1.171062
Greenokite (CdS)2.40518
Chalcopyrite (CuFeS2)0.353552
Table 2. Production mechanism of reactive oxygen species in the photocatalytic process.
Table 2. Production mechanism of reactive oxygen species in the photocatalytic process.
ReactionE (V) vs. NHE (pH = 0)
H2O + h+ → OH + H+2.38
2H2O + 2h+ → H2O2 + 2H+1.763
H2O2 + h+ → •O2 + 2H+1.72
O2 + e → •O2−0.33
O2 + H2O + e → H2O2 + OH−0.134
H2O2 + e → •OH + OH0.93
Table 3. The degradation of pollutants by metal oxide materials.
Table 3. The degradation of pollutants by metal oxide materials.
MineralTarget PollutantOxidant ConcentrationCatalyst DosageTimepHDegradation EfficiencyRef.
HematiteRoxarsoneOxalate, 2.0 mM0.5 g L−16 h4.585.10%[46]
HematiteCefazolinPAA, 0.4 mM0.3 g L−180 min798.10%[49]
Siderite2-chlorophenolPDS, 0.5 mM0.05 g⋅L−1180 min8.182.50%[70]
GoethiteBisphenol AH2O2, 1 mM0.5 g L−1240 min3.587.60%[71]
MagnetiteRhodamine BPMS, 1 mM0.2 g L−160 min794.78%[54]
Natural manganese-containing minerals4-ChlorophenolO3, 0.6 mg min−11.0 g L−115 min784.11%[68]
Ferrihydrite Enoxacin H2O2, 10 mM0.4 g L−1120 min389.70%[72]
Ilmenite Sodium butyl xanthateH2O2, 2 mM4.0 g L−180 min8.792.52%[73]
Table 4. The degradation of pollutants by metal sulfide materials.
Table 4. The degradation of pollutants by metal sulfide materials.
MineralsTarget PollutantsOxidant
Concentration
Catalyst DosageTimepHDegradation
Efficiency
Ref.
PyriteCyclohexanoic acid PS, 4.0 mM2.0 g L−1120 min3–1185.10%[74]
PyriteTetracyclineH2O2, 5 mM1.0 g L−160 min4.185%[77]
PyriteV(IV)-citratePMS, 5 mM8.0 g L−1180 minN.A.99.40%[76]
PyriteAcetaminophen PDS, 5 mM2.0 g L−1300 min4100%[78]
Pyrite sulfamethoxazolePAA, 460 μM0.3 g L−150 min5.893.67%[79]
ChalcopyriteBisphenol APMS, 0.3 mM0.1 g L−120 min699.70%[80]
ChalcopyriteRhodamine BH2O2, 43 mM0.75 g L−150 min5.197.20%[81]
ChalcopyritemetronidazolePAA, 460 μM4.0 g L−130 min383.92%[82]
BorniteTetracyclinePDS, 11.1 mM 3.5 g⋅L−1180 min4.587.50%[83]
Table 5. XPS binding energies and the corresponding atom ratio for comparisons between natural and pure sphalerite [94].
Table 5. XPS binding energies and the corresponding atom ratio for comparisons between natural and pure sphalerite [94].
CatalystBinding Energy (eV) Atom Ratio
Zn 2pS 2pO 1sFe 2pCu 2pZn/SFe/SCu/S
Natural sphalerite1022.15161.68, 162.86, 169.14532.11712.48932.260.6940.3290.033
Pure sphalerite1021.58161.7531.64--1.06--
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

Xu, D.; Yang, Y.; Gan, L. Utilization of Natural Mineral Materials in Environmental Remediation: Processes and Applications. Minerals 2025, 15, 318. https://doi.org/10.3390/min15030318

AMA Style

Xu D, Yang Y, Gan L. Utilization of Natural Mineral Materials in Environmental Remediation: Processes and Applications. Minerals. 2025; 15(3):318. https://doi.org/10.3390/min15030318

Chicago/Turabian Style

Xu, Di, Yongkui Yang, and Lingqun Gan. 2025. "Utilization of Natural Mineral Materials in Environmental Remediation: Processes and Applications" Minerals 15, no. 3: 318. https://doi.org/10.3390/min15030318

APA Style

Xu, D., Yang, Y., & Gan, L. (2025). Utilization of Natural Mineral Materials in Environmental Remediation: Processes and Applications. Minerals, 15(3), 318. https://doi.org/10.3390/min15030318

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