Next Article in Journal
Site-Specific Bioinspired Architecture—A Case Study of the Allen–Lambe House by Frank Lloyd Wright: The Pragmatic versus the Naturalistic, Intent versus Realization
Previous Article in Journal
A Highly Compact Zip Chain Arm with Origami-Inspired Folding Chain Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Preparation of Iron-Based Sulfides and Their Applications in Biomedical Fields

State Key Laboratory of Bioelectronics, Jiangsu Key Laboratory of Biomaterials and Devices, School of Biological Science and Medical Engineering, Southeast University, Nanjing 210009, China
*
Author to whom correspondence should be addressed.
Biomimetics 2023, 8(2), 177; https://doi.org/10.3390/biomimetics8020177
Submission received: 15 March 2023 / Revised: 14 April 2023 / Accepted: 21 April 2023 / Published: 24 April 2023

Abstract

:
Recently, iron-based sulfides, including iron sulfide minerals and biological iron sulfide clusters, have attracted widespread interest, owing to their excellent biocompatibility and multi-functionality in biomedical applications. As such, controlled synthesized iron sulfide nanomaterials with elaborate designs, enhanced functionality and unique electronic structures show numerous advantages. Furthermore, iron sulfide clusters produced through biological metabolism are thought to possess magnetic properties and play a crucial role in balancing the concentration of iron in cells, thereby affecting ferroptosis processes. The electrons in the Fenton reaction constantly transfer between Fe2+ and Fe3+, participating in the production and reaction process of reactive oxygen species (ROS). This mechanism is considered to confer advantages in various biomedical fields such as the antibacterial field, tumor treatment, biosensing and the treatment of neurodegenerative diseases. Thus, we aim to systematically introduce recent advances in common iron-based sulfides.

Graphical Abstract

1. Introduction

There is a significant trend in developing the next generation of materials that are suitable for practical clinical needs. Different from the new energy, optoelectronic device and electrochemical fields, the biomedical field not only requires materials with reasonable physical, chemical, electronic and optical properties, but also pays more attention to higher biological safety, to facilitate implantation or injection into the human body [1,2,3]. The size of materials used for injection will be strictly controlled to avoid long-term toxicity in the body due to low renal clearance, which will result in impaired liver and kidney function [4,5,6]. The diversity of the internal microenvironment and application scenarios therefore brings more challenges to nanomaterials. As a result, it is of great importance to develop and optimize the performance of materials to support multi-function platforms with excellent biocompatibility in the biomedical field.
Superparamagnetic iron oxide nanoparticles (SPIONs) were the first generation of nanomaterials to achieve accurate clinical applications. They are approved by the United States Food and Drug Administration (FDA) as a contrast agent for magnetic resonance imaging (MRI) due to their low toxicity, biocompatibility, and biodegradability, which implies that iron-based materials have potential clinical practicability [7,8,9,10,11]. Subsequently, iron-based sulfides came into view as a type of iron-based material, and gained a lot of interest as a potential object that can be developed and used in biomedical applications. Iron-based sulfides have an enormous variety of compositions involving Fe1-xS [12,13,14], FeS [15,16,17], FeS2 [18,19,20], Fe3S4 [21,22,23] and some ternary compounds [24,25]. It is worth noting that the iron atoms in iron-sulfide compounds can bind to sulfhydryl ligands on cysteine in vivo to form a ferromagnetic and antiferromagnetic iron sulfide cluster [26,27]. In addition, sulfur atoms have 16 protons and 3 layers of extrinsic electrons, giving iron-based sulfides a more complex electronic structure, which further facilitates higher electron and charge transfer efficiency than iron-based oxides [28]. This also makes them more complex with varied properties, because physical and chemical properties often depend on the electronic structure. In contrast, iron-based sulfides can break through the opposition between reliable biosafety and superior multi-functionality, providing new opportunities for efficient biomedical applications.
Although the properties of iron-based sulfides are outstanding, they are still in an initial stage in terms of the biomedical field, and a comprehensive presentation of their properties is absent. Herein, we present the structure and properties of iron-based sulfides, consisting of iron sulfide minerals in nature and biological iron sulfide clusters in organisms. In the aspect of performance, the outstanding magnetic properties and catalytic activities are mainly described, and the structure is found to have a significant influence on them. Next, the preparation methods are presented based on two aspects: molecular crystal synthesis and biosynthesis. In addition, several biomedical applications based on electron transport mechanisms are proposed. Finally, the challenges and opportunities faced by iron-based sulfides in the biomedical field are subsequently clarified to boost a deeper understanding in the future.

2. Structure and Properties of Iron-Based Sulfides

2.1. Structure of Iron-Based Sulfides

Iron is a widely distributed element in nature, being found in soil particles, rocks, water sediments and living organisms [29]. Under specific temperatures and pressures, a chemical reaction will occur between iron and oxygen or sulfur to form natural minerals (iron oxide minerals and iron sulfide minerals). In organisms, iron ions are prone to combine with the cysteine residues of proteins, which may be followed by the formation of iron sulfide clusters. Herein, the common natural iron sulfide minerals include pyrite, pyrrhotite, marcasite and mackinawite. Table 1 shows their crystal structures. The basic structural units of iron sulfide clusters in living organisms are also mentioned.
Pyrite, the most abundant iron-based sulfide with FeS2 as the main component, and its crystal structure are attributed to the cubic system, in which there are one iron atom and six nearby sulfur atoms combined in an octahedral configuration with identical Fe–S bond distances. Each S atom consists of three iron atoms, and another S atom in a tetrahedral structure [30]. The length of the Fe–S bond is 2.26 Å, and the length of the S–S bond is 2.15 Å. Overall, FeS2 can be viewed as consisting of Fe2+ and S22− in a cubic symmetry with a slight distortion [31]. The band gap energy of bulk pyrite is 0.95 eV, and the optical absorption coefficient (5 × 105 cm−1) is high [32]; thus, pyrite has been found to exhibit semiconducting properties.
Based on the fragile character of the S–S bond of pyrite, the sulfur atoms can be quickly released, resulting in the formation of another mineral named pyrrhotite [33]. The chemical formula of pyrrhotite is usually written as Fe1−xS (x ranges from 0 to 0.125), followed by a distinct Fe/S ratio, which results in a higher amount of S compared with FeS [34]. Thus far, four substances with near stoichiometric Fe/S ratios have been explored: Fe7S8, Fe9S10, Fe10S11 and Fe11S12. Among them, Fe7S8 and Fe9S10 are the two representative substances with various crystal structures, corresponding to monoclinic and hexagonal structures, respectively [35]. Consequently, the coordination modes between Fe and S atoms are vividly diverse.
The main chemical component of marcasite is the same as that of pyrite (FeS2), but its crystal structure varies. Marcasite is in an orthogonal system, which is distinguished by edge-sharing FeS6 units along the c-axis of the unit cell and corner-sharing edges in other directions [36]. Regardless of whether pyrite or marcasite is described as trigonally distorted FeS6 octahedra and tetrahedrally coordinated S atoms, the S atom is coordinated to three Fe atoms and one other S atom, and it is worth noting that the property of marcasite is considered less stable than that of pyrite. Due to the similar characteristics between pyrite and marcasite in some crystallization directions, they can converge into each other under certain conditions. Recently, studies based on theoretical investigations revealed that the band gap of marcasite is around 0.8–1.1 eV, which is similar to that of pyrite [37]. These findings demonstrate that pyrite and marcasite possess semiconductor properties that have potential applications in transistors and electronic devices.
Mackinawite is a layered iron sulfide mineral with a tetragonal structure, in which the Fe ion has +2 valence and the S ion has −2 valence [38,39]. It is a highly active precursor of other stable iron sulfide minerals (pyrite or marcasite). In the crystal structure, sulfur atoms and iron atoms exhibit tetrahedral coordination, connected by covalent bonds. Therefore, the structure of mackinawite is tight and stable. However, whether tetrahedral FeS is classified as a metal or a semiconductor phase has been a controversial issue. This dilemma was resolved when Huang’s team prepared tetrahedral FeS with excellent crystallization and stability under ordinary conditions, where the superconductivity was observed to be below 5 K for the first time in tetrahedral FeS, strongly suggesting that FeS behaves as a paramagnetic metal phase in the normal state [17]. In general, due to the abundant valence of the Fe atom and its lively properties, iron sulfide minerals exist in various forms and are often accompanied by the doping of other substances, which can be applied in multiple fields.
In addition to iron-based sulfides in mineral form occurring during geological processes, the iron element is also widely distributed in organisms in the form of iron sulfide clusters, as the oldest classes of bioinorganic cofactors. From the perspective of structure, the Fe and S atoms exist in various stoichiometries, and the coordinating ligands of the irons in the cluster are different, in which cysteine generally completes tetrahedral S coordination at each Fe site [40]. [2Fe-2S] [41] and [4Fe-4S] [42] are the most generic common stoichiometries of biological iron sulfide clusters, and some other patterns, such as [3Fe-4S] [43], [4Fe-3S] [44], [8Fe-7S] [45] and [4Fe-5S] [46], have also be explored with the development of research (Figure 1). Conversion reactions have existed in homogeneous or heterogeneous systems of initial or intermediate clusters. Iron sulfur clusters readily accept or provide a single electron, participating in complex chemical reactions in which two [2Fe-2S] can transform into one [4Fe-4S] ([2Fe-2S]2+, 1+ Biomimetics 08 00177 i001 [4Fe-4S]2+), and the process is reversible [47]. dimerization based on electron transfer exists universally in iron sulfide clusters of disparate structures, enriching the forms and types of iron sulfide clusters to satisfy the functional requirements of specific biological scenarios. On account of the conspicuous structural plasticity and versatile electronic characteristics, iron sulfide clusters play an important role in electron transfer, biological catalysis, ATP production, regulation of gene expression and synthesis of proteins [48].

2.2. Properties of Iron-Based Sulfides

Iron, as a very active element in group VIII transition metals, has various valences such as 0, +2 and +3, and possesses a face-centered cubic unit cell or hexagonal lattice structure proximately to metal elements (Pt, Pd, Mo, W, Ni, Co, etc.) owning traditional catalytic activity. Moreover, its d-atomic orbital is not complete, leading to an instinctively catalytic activity of iron-based sulfides [49]. One prominent catalytic property is based on the Fenton reaction [50], where reductive Fe2+ catalyzes H2O2 to produce potent oxidizing hydroxyl radicals (·OH). This performance of iron-based sulfides has been extensively applied in wastewater treatment [51], cancer catalytic therapy [52] and antibacterial therapy [53]. On top of that, hematite is a typical magnetic mineral in nature, and the S element is one of the oxygen group elements, so some iron-based sulfides have magnetic properties, including pyrrhotite (Fe1−xS) and melnikovite (Fe3S4). Interestingly, biological iron sulfide clusters also have magnetic properties in vivo [54]. These results indicate that iron sulfide clusters can not only participate in energy metabolism and the enzymatic reactions of cells but can also be developed as magnetic receptor proteins to gift organisms with magnetic sensitivity. The Fenton reaction was first presented by Henry J. Fenton, subsequently drawing significant interest in the field of wastewater treatment [55,56]. During the process of the reaction, potent oxidative hydroxyl radicals occur in the presence of Fe2+ and H2O2, which play an essential role in degrading water pollution. Chemodynamic therapy (CDT) is an efficient strategy for cancer treatment inspired by the Fenton reaction to generate highly cytotoxic ·OH [57]. In cancer cells, H2O2 is an overproduced non-radical reactive oxygen species, which can easily diffuse across biological membranes. If iron exists, H2O2 will be converted to •OH locally according to the Fenton reaction and further damage the biological system including lipids, DNA and amino acids in proteins [58]. Meanwhile, the accumulation of •OH will drastically exacerbate oxidative stress in cells and subsequently induce cell death. Hence, significant attention has been paid to Fenton catalysts such as iron minerals, including FeS2 [59], FeS [60] and Fe1-xS [13]. However, the process of the Fenton reaction is complicated and constituted of a chain of reactions. As shown in Reaction 1, •OH is firstly generated based on H2O2 and Fe2+. Then, the produced Fe3+ is reduced by H2O2 and the reproduced Fe2+ (Reaction (2)), involving a circular process. The continuous operation of this cycle is rigorous and requires large amounts of H2O2 and an optimal pH range (pH = 3.0–4.0). Nevertheless, if the pH is above 3.0, Fe3+ can easily convert into inactive Fe(OH)3 precipitation and significantly block further reactions. Therefore, the heterogeneous Fenton reaction has been proposed, in which •OH is continuously generated, avoiding the precipitation of Fe(OH)3. Besides the reaction between Fe ions and H2O2 (Reactions (1) and (2)), the heterogeneous Fenton reaction concerns the electron exchange between surface Fe and H2O2 (Reactions (3) and (4)). It is worth noting that the efficiency of the Fenton reaction will be decreased in the presence of Fe3+ or Fe(Ⅲ) reduction, because of its lower rate constant (0.001–0.01 M−1s−1) [61]. Therefore, it is a reliable method to excite the reaction system by reducing and consuming Fe3+ [62]. Additionally, the activity of the Fenton reaction is related to the temperature, time, pH, concentration, surface morphology and stability of substrates. Most reactions are limited by a narrow pH range (pH = 3.0–5.0) and accompanied by limited kinetics, poor permeability, the short half-life of hydroxyl radicals and other deficiencies [63]. To overcome these deficiencies, materials need to be systematically designed, including regulating their morphology and electron structure at the micro level, to provide more active sites to effectively enhance the Fenton catalytic performance.
Fe2+ + H2O2 → Fe3+ +·OH + OH
Fe3+ + H2O2 → Fe2+ + HO2· + H+
≡Fe(Ⅲ) + H2O2 → ≡Fe(Ⅱ) + HO2·+ H+
≡Fe(Ⅱ) + H2O2 → ≡Fe(Ⅲ) + HO· + OH
The magnetic properties of iron-based sulfides have been intensely investigated for numerous applications, such as magnetic resonance imaging contrast agents, magnetothermal therapy, magnetically responsive photodetectors and magnetically responsive biosensors. It is well known that iron oxide nanoparticles have excellent magnetic properties. Parts of iron-based sulfides have also been found to have magnetic properties, and the size, morphology and composition of the materials will affect the performance of the magnetic properties. Moreover, crystallinity is positively correlated with magnetic properties [7]. Clinically, MRI contrast agents can be divided into longitudinal relaxation contrast agents (T1) and transverse relaxation contrast agents (T2). In Table 2, the magnetic properties of various iron-based sulfides are presented. As a whole, the transverse relaxation rate (R2) of iron-based sulfides is much better than the longitudinal relaxation rate (R1), so they are mainly used in T2-weighted magnetic resonance imaging. Liu’s group developed strong superparamagnetic FeS nanoplates using a simple one-step method, in which the transverse relaxation rate of magnetic resonance (R2) was 209.8 mM−1 S−1, much higher than most of the T2 contrast agents that have been used clinically (105.93 mM−1 S−1 for Fe2O3 nanoparticles, 72 mM−1 S−1 for ferumoxsil, 151 mM−1 S−1 for ferrixan, 98.3 mM−1 S−1 for ferumoxide) [64]. This suggests that iron-based sulfides may be a promising class of magnetic nanomaterials with potential prospects for future clinical translation. Subsequently, different types of iron-based sulfides such as Fe1−xS, FeS2, Fe3S4 and some heterostructures containing iron sulfides have been developed to satisfy the requirements of the specific application (Figure 2).
In addition, as a type of iron sulfide cluster binding protein, each protein monomer binds an iron sulfide cluster in the form of a ferro-disulfide, which has been found to have potential magnetic properties and proved to be related to the magnetic sensing ability of some species, becoming a breakthrough in the field of biological magnetic sensing. Xie’s group proposed a protein-based biocompass model, which indicated the existence of an iron sulfide protein called the magnetoreceptor (MagR), which is assembled via linear polymerization to form a rod-like protein complex (magnetosensor). Like a small magnetic bar, it has a north and south pole. It was observed under an electron microscope that the MagR protein complex could sense the weak magnetic field of the earth, and it had an obvious intrinsic magnetic moment [56,65,66,67]. Another interesting finding is that transfection of clMagR/clCry4 stimulated the formation of iron oxide nanoparticles in the presence of Fe3+, and further impacted MRI T2 contrast properties [68]. Based on this, further research needs to be carried out to reveal the mysterious mechanism of animal magnetic induction and biological navigation, which will be a promising boost in the development of a new generation of bionic navigation and magnetic induction devices.

3. Synthetic Methods of Iron-Based Sulfides

Molecular crystal synthesis and biosynthesis are both important methods for the synthesis of iron-based sulfides, but there are significant differences in the objects and principles. Molecular crystal synthesis refers to the study of how to prepare and construct molecular structures with specific structures and properties, focusing on the exploration of reaction principles and reaction conditions. Biosynthesis focuses on the synthesis of molecules in living organisms, involving energy transfer, electron transfer and enzymatic kinetics. The synthesis of iron-based sulfides will be elaborated from these two dimensions.

3.1. Molecular Crystal Synthesis

The typical molecular crystal synthesis of iron-based sulfides includes the chemical vapor deposition method, hydrothermal synthesis, chemical coprecipitation and high-temperature pyrolysis. Among the above methods, common iron sources that can be used in the laboratory are FeCl2·4H2O, Fe(NO3)3·9H2O, FeSO4, etc., and usual sulfur sources include sodium thiosulphate (Na2S2O3), sulfur powder (S), sodium sulfide (Na2S) and thiourea (NH2CSNH2). The time, the temperature during the reaction and the proportions of the iron and sulfur sources have a significant influence on the composition and structure of the final crystals.
The chemical vapor deposition method is a standard method for preparing two-dimensional ultra-thin materials. Thus far, different types of two-dimensional iron-based sulfides have been attempted. For example, pure semiconductor-phase FeS2 films can be directly grown on the CoS2 substrate at high temperatures, and di-tert-butyl disulfide (TBDS) has been used as a novel sulfur precursor [75]. A single layer of high-quality tetragonal FeS films was prepared via in situ topological preparation by Shigekawa [76]. In addition, the direction of growth can be controlled by introducing a template. Regular hexagonal FeS2 crystals were grown at 600 °C based on the template of graphene [77]. Doping and phase regulation can be carried out using the CVD method to obtain two-dimensional materials with excellent electrical, optical, magnetic and mechanical properties. Although the growth mechanism based on chemical reactions is complex, many studies have shown that the reaction temperature, gas flow rate and molten salt all have an influence on the product and its performance.
Hydrothermal synthesis is a widely used method applied in laboratory and industrial production due to its advantages of a simple operation and high yield. The hydrothermal synthesis of FeS2 microspheres [78], metastable FeS2 [19], FeS2 quantum dots [79], FeS nanosheets [80], FeS nanodots [81], Fe3S4 microspheres [82], Fe3S4 nanoflowers [83] and other iron-based sulfides have been reported thus far. Melonie et al. synthesized highly crystalline FeS nanosheets via surfactant-assisted hydrothermal synthesis, and showed great control over the size, shape and thickness of the final product by changing the amount of the iron source and surfactant [84]. From previous studies, it can be summarized that surfactants such as cetyltrimethylammonium bromide (CTAB), cetyltrimethylammonium chloride (CTAC) and polyvinylpyrrolidone (PVP) will spontaneously form spherical or rod-like micelles attached to the surface of the crystal nucleus to restrict the growth of materials and affect the morphology and particle size of the product, when they disperse in the solvent [85]. If the template is introduced during the hydrothermal reaction, then the crystal nucleus would attach to the template and facilitate the growth of composite or heterogeneous structures. Inspired by this mechanism, many materials with composite structures have been created, such as rGO/FeS composite [86], FeS2@graphene [87], rGO/FeS2 hybrid microspheres [87] and Fe3S4 NPs@rGO [88].
The procedure of the solvothermal method is similar to that of hydrothermal synthesis, the most significant difference being the choice of solvent, and that the solvothermal method prefers to utilize organic solvents (N’N-dimethylformamide, methanol, ethylenediamine, oleylamine, polyethylene glycol, etc.) as the solvent. Under high temperatures and pressures, the nucleation and recrystallization processes occur. The structural and chemical defects of crystal growth can be improved by adjusting the factors that affect the crystal growth rate (temperature, pressure, time, solvent and gas atmosphere), which can adjust the morphology, size, pore size and function of the crystal [89]. FeS nanosheet [90], rhombic FeS2 [91], octahedral FeS2 [92], FeS2 nanoparticle [93] and magnetic Fe3S4 [94] have been created using this method and applied in some areas. However, this method is not green, and the biocompatibility of the product is demanding up to the standard in the biomedical field, so further modification steps need to be performed before application.

3.2. Biosynthesis

Recently, more attention has been focused on biosynthesis owing to its advantages of being inexpensive, environmentally friendly and energy-efficient. It is an old strategy to prepare complex and pharmacologically active natural products, replacing plant separation and extraction with a lower extraction efficiency. Notably, it is a method that has the potential to synthesize nanomaterials by employing living organisms as biofactories, having special advantages compared with traditional chemical methods [95]. Biosynthesized nanoparticles are endowed with excellent biocompatibility, water solubility and an ultrasmall size, which are desirable in the biomedical field and applicable in bioimaging, biosensing, drug delivery and treatment [96]. FeS nanoparticles are representative iron-based sulfides that can be predominantly biosynthesized through sulfate-reducing bacteria (SRB) in anoxic environments [97]. Nonetheless, little is known about the mechanism of this process, especially regarding the electron transfer capacity and enzyme kinetics in cells. In previous studies, evidence can be found that the production of FeS is related to the metabolic process of SRB, which reduces sulfate (SO42−) as an electron acceptor to form S2− [98]. However, which biological groups, signaling pathways and protein expressions are involved during the reaction remain unclear. Therefore, in-depth exploration of biological mechanisms is urgently needed, and will contribute to the green controllable synthesis of related materials on morphological and structural scales.
Iron sulfide clusters are types of iron-based sulfides widely existing in organisms, which have important biological functions. Iron sulfide clusters play an important role in cellular processes such as electron transport, redox reactions and signal transduction. The biosynthetic pathway of iron sulfide clusters is not fully defined, but some studies have shown that the biosynthetic process involves the interaction and regulation of various enzymes, proteins and cofactors. In organisms, iron ions are easily oxidized, so special transporters must be used to transport iron ions to the synthetic iron sulfur cluster. For example, iron sulfur cluster assembly protein A (IscA) can collect free iron in cells and transfer it to the iron sulfur cluster assembly enzyme (IscU) for assembly. Finally, the assembly of the iron sulfur cluster is promoted and completed on the IscU scaffold protein, because the scaffold protein has high structural stability and iron sulfide cluster binding ability. In conclusion, the biosynthesis of iron sulfide clusters is a complex process that has not been described explicitly here. Further study will help to better understand the formation and function of iron sulfide clusters in organisms and provide a theoretical basis for the design and synthesis of new iron sulfide cluster structures.

4. Biomedical Applications Based on Electron Transport Mechanism

The Fenton reaction uses H2O2 as a chain propagation medium to mediate the conversion between Fe2+ and Fe3+ through an electron transfer reaction, resulting in the generation of ·OH. Based on this mechanism, it can be used for antibacterial agents, biosensor construction and chemodynamic therapy. In recent years, ferroptosis, a new, highly iron-dependent programmed cell death mechanism that differs from apoptosis or necrosis, has also been widely explored [99,100,101]. When iron-based nanoparticles enter cells, lipid peroxidation and elevated reactive oxygen species (ROS) levels will occur, followed by the ferroptosis of cells. Ferroptosis can therefore be seen as the result of damage to the electron transfer chain and can be used in the efficient treatment of tumors [102] and neurodegeneration [103]. Based on this, this section mainly summarizes biomedical applications based on electron transport mechanisms, including in the antibacterial field, tumor treatment, biosensing and degenerative neurological diseases.

4.1. Antibacterial Field

The abuse of antibiotics leads to an increase in bacterial resistance, which is the dilemma faced by traditional antibacterial agents. Therefore, eliminating bacterial infection based on nanotechnology is a new solution [104]. Some inorganic metal elements naturally have good sterilization and inhibition of microbial reproduction, such as silver, copper and zinc. In addition, iron has also been found to have potential antibacterial ability. For example, iron oxide nanoparticles have peroxidase activity, which can catalyze hydrogen peroxide and regulate the level of reactive oxygen species (ROS) in the body, to achieve efficient sterilization and removal of bacterial biofilm [105]. Gao’s group focused on the iron–sulfur protein system, a class of proteins that affect the body’s redox homeostasis by regulating the amount of sulfur, by converting organic sulfur compounds of garlic into inorganic nano-iron sulfides, in which the antibacterial activity improved more than 500 times (Figure 3) [106]. Differently, the antibacterial mechanism of iron sulfides mainly includes the following two aspects, dissimilar to iron oxides: (1) Nano-iron-based sulfides have enzyme-like activity, can catalyze H2O2 to produce free radicals ·OH, and can promote lipid peroxidation of the cell membrane. (2) Hydrogen sulfide (H2S) gas will be produced under weakly acidic conditions by iron-based sulfides. Bacteria can be killed by the release of hydrogen polysulfide, and this small inorganic molecule can quickly permeate through the bacteria and significantly increase the level of ROS and lipid oxidation, eventually leading to bacterial death. Together, the bacterial death induced by iron sulfides has typical iron death characteristics. Along with the significant inhibition of the activity of the complex in the bacterial respiratory chain, glutathione (GSH) will be oxidized into oxidized glutathione (GSSG), resulting in GSH depletion and lipid peroxidation of the cell membrane [107]. However, not all iron sulfides have antibacterial activity, among which Fe3S4 and Fe7S8 have good antibacterial activity and antibacterial selectivity to gram-negative bacteria, especially to Gardnerella, with a minimum inhibitory concentration of 7.8 μg/mL, with the antibacterial effect better than that of metronidazole [108]. A disadvantage of iron-based sulfides is that they are unstable in aqueous systems, so iron-based sulfides such as FeS can also be encapsulated in acrylamide hydrogels, and the porous structure will not only capture and enrich bacteria, but also effectively retain active antibacterial substances, further enhancing antibacterial properties and wound healing [109]. The unique antibacterial properties and mechanisms of iron-based sulfides make them a potential novel class of non-antibiotic drugs for the prevention and treatment of bacterial vaginitis, dental caries and wound infection.

4.2. Tumor Treatment

There has been a rapid revolution in the growing treatment of cancer, where some new treatments with low side effects and high specificity have been developed, including photothermal therapy [110], chemodynamic therapy [111], gas therapy [112], starvation therapy [113], etc. At the same time, cancer therapy is also expanding to multi-modal treatment. Inorganic nano-platforms are endowed with an intelligent response to special microenvironments, and they are easy to combine with exogenous physical tools (such as light, ultrasound, X-rays, magnetic fields, heat and electricity) to orthotopically activate the immune and other functions, leading to the implementation of more accurate and efficient cancer treatment. In the previous studies, iron-based sulfides have been explored that not only possess chemodynamic properties based on the Fenton reaction but also exhibit higher absorbance in the near infrared region (NIR), and this indicates iron-based sulfides can convert more absorbed NIR light into heart. This is why iron-based sulfides have potential applications in tumor treatment, including chemotherapy and photothermal therapy. For example, Zhang’s group presented an albumin-constrained strategy to synthesize tiny and highly dispersible ferrous sulfide (termed FeS@BSA) quantum dots at ambient conditions. FeS@BSA quantum dots have specific and strong absorption in the near infrared region, the temperature of FeS@BSA solution (1.2 mg/mL) rapidly increased up to 56.8 °C after being irradiated for 5 min by a 660 nm laser (2 W/cm2), which indicated the excellent photothermal conversion effect of FeS@BSA quantum dots [73]. The photothermal therapeutic effects of other iron-based sulfides on tumors are shown in Table 3. Later, Hou’s group reported a strategy based on carbonic anhydrase inhibitor (CAI)-modified ferrous sulfide nanoparticles (FeS-PEG-CAI NPs) that eliminate tumors by inducing acidosis to destroy the metabolic comorbidities within the tumor site, achieving PAI and MRI dual-mode guided chemotherapy/photothermal/gas therapy (Figure 4) [72]. If the size of FeS is controlled within the ultrasmall range, it will possess renal clearance properties, enhancing the potential for clinical transformation [73]. FeS was also found to be a drug-carrying nano-platform to load the anti-cancer drug Dox for multi-modal breast cancer treatment [114]. Yan’s group developed self-cascade pyrite nanozymes according to FeS2, contributing to apoptosis–ferroptosis synergistic tumor therapy. On top of that, they explained that when FeS2 interacts with the substrate H2O2, the ligand binding covalent bond is shorter than Fe3O4 nanoparticles. The surface exhibits many furl-like structures, conferring FeS2 greater binding competence with the substrate, resulting in the catalytic efficiency of H2O2 (Kcat/Km) being 4144 times higher than that of traditional Fe3O4 nanoparticles, and 3086 times higher than that of natural horseradish peroxidase [115].
In general, the advantages of iron-based sulfides are based on the following aspects: (1) They can consume the endogenous H2O2 of tumor cells through the Fenton reaction to produce ROS with high cytotoxicity, which will further induce the apoptosis of tumor cells. (2) Iron-based sulfides produce and release H2S under weakly acidic conditions, which has a specific inhibitory effect on the catalase activity of cancer cells, and then increases the content of H2O2 in tumor cells, helping to enhance the CDT performance simultaneously. (3) Due to the strong optical absorption of iron sulfides in the near-infrared region, they can be used in the photothermal therapy of tumors based on the photothermal effect. (4) In particular, some iron-based sulfides can be seen as the initiators of ferroptosis that cause the inactivation of intracellular glutathione peroxidase 4 (GPX4) directly or indirectly, which leads to the collapse of the intracellular antioxidant system and, ultimately, cell death. At present, the difficulty is to overcome some universal problems such as the poor pharmacokinetics, low accumulation level of the target site and low biocompatibility, striving for the clinical transformation of iron-based sulfides.

4.3. Bio-Sensing

Hydrogen peroxide (H2O2) and glutathione (GSH) are important molecules in life processes, which can be generally used as key biomarkers and prognostic indicators in biomedical diagnosis. Iron-based sulfides have peroxidase catalytic activity due to the electron transfer between Fe2+ and H2O2, which can be used to detect H2O2, GSH and other bioactive substances [116]. Therefore, they have the potential to be used to construct biosensors with high sensitivity, a low detection limit and a wide detection range. FeS2 nanoparticles have been shown to have high peroxidase catalytic activity, which can efficiently catalyze H2O2 to produce ·OH. When 3,3′,5,5′-tetramethylbenzidine (TMB) is used as the indicator of ·OH, the reaction system will change from colorless to blue, and there is a prominent absorption peak (652 nm) in the UV–Vis spectrum. When GSH is added, the blue color fades, and the absorption peak disappears (Figure 5). The detection limits of H2O2 and GSH are as low as 0.91 μM and 0.15 μM based on this colorimetric biosensor [117]. Huang et al. constructed FeS2/SiO2 mesoporous hollow spheres with higher activity on this basis. The Mi constant (Km) is 18 times smaller than that of FeS2 nanoparticles, and the catalytic efficiency is 16 times that of FeS2 nanoparticles. The response to H2O2 and GSH at room temperature takes only 1 min, which provides a material base for the construction of biosensors with high sensitivity and a rapid response [118].

4.4. Neurodegenerative Diseases

Neurodegenerative diseases, including Alzheimer’s disease, Parkinson’s disease and multiple sclerosis, are common in the elderly. Previously, it has been noted that iron disorders are linked to various neurodegenerative diseases, such as Parkinson’s disease and multiple sclerosis. To be specific, these brain regions are typically occupied by iron-rich microglia cells, which can induce an iron overload followed by a significant change in the transcriptional state of microglia, which will lead to the threat of ferroptosis [119]. Iatrogenic iron in metal implants is a potential risk factor for developing neurological diseases, along with a higher incidence of Parkinson’s disease. In the face of an attack of excess iron, glial cells act as neuroprotectors to accumulate more extracellular iron by upregulating divalentmetaltransporter 1 (DMT1), while neurons limit iron uptake by increasing DMT1 degradation [120]. However, over time, the microglia become overwhelmed and die, releasing stored iron ions and causing neurons to die together, which will further induce neurodegenerative disease [103]. These findings suggest that ferroptosis may be an essential mechanism of neuronal loss in neurodegenerative diseases; therefore, inhibiting neuronal iron death may interfere with the progression of neurodegenerative diseases. Ferroptosis suppressor protein 1 (FSP1), calcium-independent phospholipase iPLA2β, ubiquinol, glutathione peroxidase 4 (GPX4), NFS1 cysteine desulfurase, frataxin (FXN) and iron sulfur cluster assembly enzyme (ISCU) are all critical regulators of ferroptosis in the extrinsic or intrinsic pathway [121,122,123,124,125,126,127]. Among them, NFS1, FXN and ISCU play an essential role in producing iron sulfide clusters. Specifically, NFS1 is a rate-limiting enzyme in the biosynthesis of iron sulfide clusters, mainly providing sulfide and catalyzing the formation of iron sulfide clusters [128]. FXN serves as a transporter, delivering iron to the ISCU scaffold protein, where the final assembly is completed [129,130,131]. These results provide evidence for a way to treat neurodegenerative diseases by adjusting ferroptosis and controlling the synthesis of iron sulfide clusters, including intracellular iron uptake, iron utilization and conversion, and enzyme catalysis.

5. Conclusions

Iron-based sulfides have a variety of structures with multi-functionality, and good application potential in the biomedical field, including in the antibacterial field, tumor treatment, bioimaging, biosensing and the treatment of neurodegenerative diseases. In addition to their excellent optical, electrical, magnetic and thermal properties, the primary reasons why they can be widely used in the biomedical field are their biosafety and universal applicability. Although iron-based sulfides are effective materials in electrocatalysis and water pollution treatment, their research in the biomedical field is still in the early stages. Over the past decade, considerable progress has been achieved regarding their catalytic properties, magnetic properties and ferroptosis regulation. However, controlled synthesis is still in its infancy, and there is still a long way to go to resolve the dilemma regarding the biological mechanisms of iron sulfide clusters. In addition, if the clinical application prospects of iron-based sulfides are to be realized, it is also necessary to explore the suitable dosages for different diseases. Whereas numerous studies have shown that the iron balance is crucial for health, iron overload has been revealed to be linked to the pathogenesis of diseases such as Parkinson’s and Alzheimer’s. Therefore, the biological metabolic pathways and potential side effects of iron-based sulfides urgently need to be researched. It is highly likely that, in the future, iron-based sulfides will be able to serve as clinical drug candidates and show potential in novel biomedical fields such as self-assembling microrobots [132], tissue engineering [133], and biomimetic protein [134].

Author Contributions

The manuscript was written and edited with contributions from all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (No. 2021YFA1201403) and China Science and Technology Innovation 2030-Major Project (No. 2022ZD0211704).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the National Key Research and Development Program of China (No. 2021YFA1201403) and China Science and Technology Innovation 2030-Major Project (No. 2022ZD0211704). The authors gratefully acknowledge their financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qi, C.; Lin, J.; Fu, L.H.; Huang, P. Calcium-based biomaterials for diagnosis, treatment, and theranostics. Chem. Soc. Rev. 2018, 47, 357–403. [Google Scholar] [CrossRef] [PubMed]
  2. Fenton, O.S.; Olafson, K.N.; Pillai, P.S.; Mitchell, M.J.; Langer, R. Advances in biomaterials for drug delivery. Adv. Mater. 2018, 30, e1705328. [Google Scholar] [CrossRef] [PubMed]
  3. Guttenplan, A.P.M.; Tahmasebi Birgani, Z.; Giselbrecht, S.; Truckenmuller, R.K.; Habibovic, P. Chips for biomaterials and biomaterials for chips: Recent advances at the interface between microfabrication and biomaterials research. Adv. Healthc. Mater. 2021, 10, e2100371. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, J.; Li, Y.; Nie, G. Multifunctional biomolecule nanostructures for cancer therapy. Nat. Rev. Mater. 2021, 6, 766–783. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, C.; Fan, W.; Cheng, W.X.; Gu, Y.; Chen, Y.; Zhou, W.; Yu, X.F.; Chen, M.; Zhu, M.; Fan, K.; et al. Red emissive carbon dot superoxide dismutase nanozyme for bioimaging and ameliorating acute lung injury. Adv. Funct. 2023, 2213856. [Google Scholar] [CrossRef]
  6. Colby, A.H.; Kirsch, J.; Patwa, A.N.; Liu, R.; Hollister, B.; McCulloch, W.; Burdette, J.E.; Pearce, C.J.; Oberliels, N.H.; Colson, Y.L.; et al. Radiolabeled biodistribution of expansile nanoparticles: Intraperitoneal administration results in tumor specific accumulation. ACS Nano 2023, 17, 2212–2221. [Google Scholar] [CrossRef]
  7. Gu, N.; Zhang, Z.; Li, Y. Adaptive iron-based magnetic nanomaterials of high performance for biomedical applications. Nano Res. 2021, 15, 1–17. [Google Scholar] [CrossRef]
  8. Ye, D.; Li, M.; Xie, Y.; Chen, B.; Han, Y.; Liu, S.; Wei, Q.H.; Gu, N. Optical imaging and high-accuracy quantification of intracellular iron contents. Small 2021, 17, e2005474. [Google Scholar] [CrossRef]
  9. Li, M.; Li, J.; Chen, J.; Liu, Y.; Cheng, X.; Yang, F.; Gu, N. Platelet membrane biomimetic magnetic nanocarriers for targeted delivery and in situ generation of nitric oxide in early ischemic stroke. ACS Nano 2020, 14, 2024–2035. [Google Scholar] [CrossRef]
  10. Li, Y.; Chen, Z.; Gu, N. In vitro biological effects of magnetic nanoparticles. Chin. Sci. Bull. 2012, 57, 3972–3978. [Google Scholar] [CrossRef]
  11. Dong, H.; Du, W.; Dong, J.; Che, R.; Kong, F.; Cheng, W.; Ma, M.; Gu, N.; Zhang, Y. Depletable peroxidase-like activity of Fe3O4 nanozymes accompanied with separate migration of electrons and iron ions. Nat. Commun. 2022, 13, 5365. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, Y.; Xu, R.; Wang, Z.; Hao, X.; Zhang, C.; Zhao, H.; Li, W.; Wang, S.; Dong, Y.; Huang, Z.; et al. Carbon-free crystal-like Fe1-xS as an anode for potassium-ion batteries. ACS Appl. Mater. Interfaces 2021, 13, 55218–55226. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Z.; Luo, Y.; Hu, Y.; Liang, K.; He, G.; Chen, Q.; Wang, Q.; Chen, H. Photothermo-promoted nanocatalysis combined with H2S-mediated respiration inhibition for efficient cancer therapy. Adv. Funct. 2020, 31, 2007991. [Google Scholar] [CrossRef]
  14. Xiao, Y.; Hwang, J.-Y.; Belharouak, I.; Sun, Y.-K. Na storage capability investigation of a carbon nanotube-encapsulated Fe1–xS composite. ACS Energy Lett. 2017, 2, 364–372. [Google Scholar] [CrossRef]
  15. Sun, Y.; Danish, M.; Ali, M.; Shan, A.; Li, M.; Lyu, Y.; Qiu, Z.; Sui, Q.; Zang, X.; Lyu, S. Trichloroethene degradation by nanoscale CaO2 activated with Fe(II)/FeS: The role of FeS and the synergistic activation mechanism of Fe(II)/FeS. Chem. Eng. J. 2020, 394, 124830. [Google Scholar] [CrossRef]
  16. Brumovsky, M.; Filip, J.; Malina, O.; Oborna, J.; Sracek, O.; Reichenauer, T.G.; Andryskova, P.; Zboril, R. Core-shell Fe/FeS nanoparticles with controlled shell thickness for enhanced trichloroethylene removal. ACS Appl. Mater. Interfaces 2020, 12, 35424–35434. [Google Scholar] [CrossRef]
  17. Lai, X.; Zhang, H.; Wang, Y.; Wang, X.; Zhang, X.; Lin, J.; Huang, F. Observation of superconductivity in tetragonal FeS. J. Am. Chem. Soc. 2015, 137, 10148–10151. [Google Scholar] [CrossRef]
  18. Wu, L.; Dzade, N.Y.; Gao, L.; Scanlon, D.O.; Ozturk, Z.; Hollingsworth, N.; Weckhuysen, B.M.; Hensen, E.J.; de Leeuw, N.H.; Hofmann, J.P. Enhanced photoresponse of FeS2 films: The role of marcasite-pyrite phase junctions. Adv. Mater. 2016, 28, 9602–9607. [Google Scholar] [CrossRef]
  19. Kitchaev, D.A.; Ceder, G. Evaluating structure selection in the hydrothermal growth of FeS2 pyrite and marcasite. Nat. Commun. 2016, 7, 13799. [Google Scholar] [CrossRef]
  20. Sun, W.; Liu, S.; Li, Y.; Wang, D.; Guo, Q.; Hong, X.; Xie, K.; Ma, Z.; Zheng, C.; Xiong, S. Monodispersed FeS2 electrocatalyst anchored to nitrogen-doped carbon host for lithium-sulfur batteries. Adv. Funct. 2022, 32, 2205471. [Google Scholar] [CrossRef]
  21. Preiner, M.; Igarashi, K.; Muchowska, K.B.; Yu, M.; Varma, S.J.; Kleinermanns, K.; Nobu, M.K.; Kamagata, Y.; Tuysuz, H.; Moran, J.; et al. A hydrogen-dependent geochemical analogue of primordial carbon and energy metabolism. Nat. Ecol. Evol. 2020, 4, 534–542. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, Q.; Gao, J.; Cao, C.; Yin, G.; Jiang, Z.; Ge, M.; Xiao, X.; Lee, W.-K.; Wang, J. Insights into enhanced sodium ion storage mechanism in Fe3S4: The coupling of surface chemistry, microstructural regulation and 3D electronic transport. Nano Energy 2019, 62, 384–392. [Google Scholar] [CrossRef]
  23. Wang, R.; Wu, X.; Tian, Z.; Hu, T.; Cai, C.; Wu, G.; Jiang, G.; Liu, B. Sustained release of hydrogen sulfide from anisotropic ferrofluid hydrogel for the repair of spinal cord injury. Bioact. Mater. 2023, 23, 118–128. [Google Scholar] [CrossRef]
  24. Sun, Y.; Ding, S.; Xia, B.; Duan, J.; Antonietti, M.; Chen, S. Biomimetic FeMo(Se, Te) as joint electron pool promoting nitrogen electrofixation. Angew Chem. Int. Ed. Engl. 2022, 61, e202115198. [Google Scholar]
  25. Pang, K.; Xu, X.; Wei, Y.; Ying, T.; Li, W.; Yang, J.; Li, X.; Jiang, Y.; Zhang, G.; Tian, W. Integrating ferromagnetism and ferroelectricity in an iron chalcogenide monolayer: A first-principles study. Nanoscale 2022, 14, 14231–14239. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, C.W.; McCabe, J.W.; Russell, D.H.; Barondeau, D.P. Molecular mechanism of ISC iron-sulfur cluster biogenesis revealed by high-resolution native mass spectrometry. J. Am. Chem. Soc. 2020, 142, 6018–6029. [Google Scholar] [CrossRef]
  27. Zhang, B.; Bandyopadhyay, S.; Shakamuri, P.; Naik, S.G.; Huynh, B.H.; Couturier, J.; Rouhier, N.; Johnson, M.K. Monothiol glutaredoxins can bind linear [Fe3S4]+ and [Fe4S4]2+ clusters in addition to [Fe2S2]2+ clusters: Spectroscopic characterization and functional implications. J. Am. Chem. Soc. 2013, 135, 15153–15164. [Google Scholar] [CrossRef]
  28. Wang, Y.; Ren, B.; Zhen Ou, J.; Xu, K.; Yang, C.; Li, Y.; Zhang, H. Engineering two-dimensional metal oxides and chalcogenides for enhanced electro- and photocatalysis. Sci. Bull. 2021, 66, 1228–1252. [Google Scholar] [CrossRef]
  29. Turcheniuk, K.; Singhal, V.; Yushin, G. Ten years left to redesign lithium-ion batteries. Nature 2018, 559, 467–470. [Google Scholar] [CrossRef]
  30. Moslemi, H.; Gharabaghi, M. A review on electrochemical behavior of pyrite in the froth flotation process. J. Ind. Eng. Chem. 2017, 47, 1–18. [Google Scholar] [CrossRef]
  31. Choi, H.; Seo, J.Y.; Uhm, Y.R.; Sun, G.M.; Kim, C.S. Crystalline structure and magnetic properties of pyrite FeS. AIP Adv. 2021, 11, 015131. [Google Scholar] [CrossRef]
  32. Chandrawat, G.S.; Singh, J.; Tripathi, J.; Sharma, A.; Gupta, M.; Sathe, V.; Tripathi, S. Synthesis and structural characterization of FeS2 nanoparticles using rietveld refinement. AIP Conf. Proc. 2019, 2100, 020023. [Google Scholar]
  33. Yu, W.J.; Liu, C.; Zhang, L.; Hou, P.X.; Li, F.; Zhang, B.; Cheng, H.M. Synthesis and electrochemical lithium storage behavior of carbon nanotubes filled with iron sulfide nanoparticles. Adv. Sci. 2016, 3, 1600113. [Google Scholar] [CrossRef] [PubMed]
  34. Özdeniz, A.H.; Kelebek, S. A study of self-heating characteristics of a pyrrhotite-rich sulphide ore stockpile. Int. J. Min. Sci. Technol. 2013, 23, 381–386. [Google Scholar] [CrossRef]
  35. Tang, X.; Chen, Y. A review of flotation and selective separation of pyrrhotite: A perspective from crystal structures. Int. J. Min. Sci. Technol. 2022, 32, 847–863. [Google Scholar] [CrossRef]
  36. Schmøkel, M.S.; Bjerg, L.; Cenedese, S.; Jørgensen, M.R.V.; Chen, Y.-S.; Overgaard, J.; Iversen, B.B. Atomic properties and chemical bonding in the pyrite and marcasite polymorphs of FeS2: A combined experimental and theoretical electron density study. Chem. Sci. 2014, 5, 1408–1421. [Google Scholar] [CrossRef]
  37. Dzade, N.Y.; de Leeuw, N.H. Periodic DFT+U investigation of the bulk and surface properties of marcasite FeS. Phys. Chem. Chem. Phys. 2017, 19, 27478–27488. [Google Scholar] [CrossRef]
  38. Bone, S.E.; Bargar, J.R.; Sposito, G. Mackinawite (FeS) reduces mercury(II) under sulfidic conditions. Environ. Sci. Technol. 2014, 48, 10681–10689. [Google Scholar] [CrossRef]
  39. Dzade, N.Y.; Roldan, A.; de Leeuw, N.H. A DFT-D2 study of the adsorption and dissociation of water on cleanand oxygen-covered {001} and {011} surfaces of mackinawite (FeS). J. Phys. Chem. C 2016, 120, 21441–21450. [Google Scholar] [CrossRef]
  40. Honarmand Ebrahimi, K.; Ciofi-Baffoni, S.; Hagedoorn, P.L.; Nicolet, Y.; Le Brun, N.E.; Hagen, W.R.; Armstrong, F.A. Iron-sulfur clusters as inhibitors and catalysts of viral replication. Nat. Chem. 2022, 14, 253–266. [Google Scholar] [CrossRef]
  41. Gourdoupis, S.; Nasta, V.; Calderone, V.; Ciofi-Baffoni, S.; Banci, L. IBA57 recruits ISCA2 to form a [2Fe-2S] cluster-mediated complex. J. Am. Chem. Soc. 2018, 140, 14401–14412. [Google Scholar] [CrossRef] [PubMed]
  42. Lu, Z.; Imlay, J.A. A conserved motif liganding the [4Fe-4S] cluster in [4Fe-4S] fumarases prevents irreversible inactivation of the enzyme during hydrogen peroxide stress. Redox Biol. 2019, 26, 101296. [Google Scholar] [CrossRef]
  43. Zhang, B.; Arcinas, A.J.; Radle, M.I.; Silakov, A.; Booker, S.J.; Krebs, C. First step in catalysis of the radical S-sdenosylmethionine methylthiotransferase miaB yields an intermediate with a [3Fe-4S]0-like auxiliary cluster. J. Am. Chem. Soc. 2020, 142, 1911–1924. [Google Scholar] [CrossRef] [PubMed]
  44. DeRosha, D.E.; Chilkuri, V.G.; Van Stappen, C.; Bill, E.; Mercado, B.Q.; DeBeer, S.; Neese, F.; Holland, P.L. Planar three-coordinate iron sulfide in a synthetic [4Fe-3S] cluster with biomimetic reactivity. Nat. Chem. 2019, 11, 1019–1025. [Google Scholar] [CrossRef]
  45. Chica, B.; Ruzicka, J.; Pellows, L.M.; Kallas, H.; Kisgeropoulos, E.; Vansuch, G.E.; Mulder, D.W.; Brown, K.A.; Svedruzic, D.; Peters, J.W.; et al. Dissecting electronic-structural transitions in the nitrogenase MoFe protein P-cluster during reduction. J. Am. Chem. Soc. 2022, 144, 5708–5712. [Google Scholar] [CrossRef] [PubMed]
  46. Caserta, G.; Zuccarello, L.; Barbosa, C.; Silveira, C.M.; Moe, E.; Katz, S.; Hildebrandt, P.; Zebger, I.; Todorovic, S. Unusual structures and unknown roles of FeS clusters in metalloenzymes seen from a resonance Raman spectroscopic perspective. Coordin. Chem. Rev. 2022, 452, 214287. [Google Scholar] [CrossRef]
  47. Holm, R.H.; Lo, W. Structural conversions of synthetic and protein-bound iron-sulfur clusters. Chem. Rev. 2016, 116, 13685–13713. [Google Scholar] [CrossRef] [PubMed]
  48. Srour, B.; Gervason, S.; Hoock, M.H.; Monfort, B.; Want, K.; Larkem, D.; Trabelsi, N.; Landrot, G.; Zitolo, A.; Fonda, E.; et al. Iron insertion at the assembly site of the ISCU scaffold protein is a conserved process initiating Fe-S cluster biosynthesis. J. Am. Chem. Soc. 2022, 144, 17496–17515. [Google Scholar] [CrossRef]
  49. Shen, J.; Xu, X.; Liu, J.; Wang, Z.; Zuo, S.; Liu, Z.; Zhang, D.; Liu, J.; Zhu, M. Unraveling the catalytic activity of Fe-based compounds toward Li2Sx in Li-S chemical system from d-p bands. Adv. Energy Mater. 2021, 11, 2100673. [Google Scholar] [CrossRef]
  50. Wang, J.; Dai, X.; Wang, H.; Liu, H.; Rabeah, J.; Bruckner, A.; Shi, F.; Gong, M.; Yang, X. Dihydroxyacetone valorization with high atom efficiency via controlling radical oxidation pathways over natural mineral-inspired catalyst. Nat. Commun. 2021, 12, 6840. [Google Scholar] [CrossRef]
  51. Yan, Q.; Zhang, J.; Xing, M. Cocatalytic Fenton reaction for pollutant control. Cell Rep. Phys. Sci. 2020, 1, 100149. [Google Scholar] [CrossRef]
  52. Wang, Y.; Gao, F.; Li, X.; Niu, G.; Yang, Y.; Li, H.; Jiang, Y. Tumor microenvironment-responsive fenton nanocatalysts for intensified anticancer treatment. J. Nanobiotechnol. 2022, 20, 69. [Google Scholar] [CrossRef] [PubMed]
  53. Li, F.; Huang, K.; Chang, H.; Liang, Y.; Zhao, J.; Yang, S.; Liu, F. A polydopamine coated nanoscale FeS theranostic platform for the elimination of drug-resistant bacteria via photothermal-enhanced Fenton reaction. Acta Biomater. 2022, 150, 380–390. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, J.; Jarocha, L.E.; Zollitsch, T.; Konowalczyk, M.; Henbest, K.B.; Richert, S.; Golesworthy, M.J.; Schmidt, J.; Dejean, V.; Sowood, D.J.C.; et al. Magnetic sensitivity of cryptochrome 4 from a migratory songbird. Nature 2021, 594, 535–540. [Google Scholar] [CrossRef]
  55. Meyerstein, D. Re-examining Fenton and Fenton-like reactions. Nat. Rev. Chem. 2021, 5, 595–597. [Google Scholar] [CrossRef]
  56. Thomas, N.; Dionysiou, D.D.; Pillai, S.C. Heterogeneous Fenton catalysts: A review of recent advances. J. Hazard. Mater. 2021, 404, 124082. [Google Scholar] [CrossRef]
  57. Chen, X.; Zhang, H.; Zhang, M.; Zhao, P.; Song, R.; Gong, T.; Liu, Y.; He, X.; Zhao, K.; Bu, W. Amorphous Fe-Based Nanoagents for Self-Enhanced Chemodynamic Therapy by Re-Establishing Tumor Acidosis. Adv. Funct. Mater. 2019, 30, 1908365. [Google Scholar] [CrossRef]
  58. Ranji-Burachaloo, H.; Gurr, P.A.; Dunstan, D.E.; Qiao, G.G. Cancer Treatment through Nanoparticle-Facilitated Fenton Reaction. ACS Nano 2018, 12, 11819–11837. [Google Scholar] [CrossRef]
  59. Xiao, S.; Lu, Y.; Feng, M.; Dong, M.; Cao, Z.; Zhang, X.; Chen, Y.; Liu, J. Multifunctional FeS2 theranostic nanoparticles for photothermal-enhanced chemodynamic/photodynamic cancer therapy and photoacoustic imaging. Chem. Eng. J. 2020, 396, 125294. [Google Scholar] [CrossRef]
  60. Cai, Y.; Fan, J.; Liu, Z. Enhanced degradation of tetracycline over FeS-based Fenton-like process: Autocatalytic decomposition of H2O2 and reduction of Fe(III). J. Hazard. 2022, 436, 129092. [Google Scholar] [CrossRef]
  61. Tang, Z.; Zhao, P.; Wang, H.; Liu, Y.; Bu, W. Biomedicine meets Fenton chemistry. Chem. Rev. 2021, 121, 1981–2019. [Google Scholar] [CrossRef]
  62. Yu, W.; Wen, Q.; Yang, J.; Xiao, K.; Zhu, Y.; Tao, S.; Liang, S.; Hu, S.; Wu, Q.; Hou, H.; et al. Novel insights into extracellular polymeric substance degradation, hydrophilic/hydrophobic characteristics, and dewaterability of waste activated sludge pretreated by hydroxylamine enhanced Fenton oxidation. ACS EST Eng. 2020, 1, 385–392. [Google Scholar] [CrossRef]
  63. 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 2019, 255, 117739. [Google Scholar] [CrossRef]
  64. Yang, K.; Yang, G.; Chen, L.; Cheng, L.; Wang, L.; Ge, C.; Liu, Z. FeS nanoplates as a multifunctional nano-theranostic for magnetic resonance imaging guided photothermal therapy. Biomaterials 2015, 38, 1–9. [Google Scholar] [CrossRef] [PubMed]
  65. Qin, S.; Yin, H.; Yang, C.; Dou, Y.; Liu, Z.; Zhang, P.; Yu, H.; Huang, Y.; Feng, J.; Hao, J.; et al. A magnetic protein biocompass. Nat. Mater. 2016, 15, 217–226. [Google Scholar] [CrossRef]
  66. Xue, L.; Hu, T.; Guo, Z.; Yang, C.; Wang, Z.; Qin, S.; Yang, P.; Xie, C.; Xu, J.; Li, N.; et al. A novel biomimetic magnetosensor based on magneto-optically involved conformational variation of MagR/Cry4 complex. Adv. Electron. Mater. 2020, 6, 1901168. [Google Scholar] [CrossRef]
  67. Zhou, Y.; Tong, T.; Wei, M.; Zhang, P.; Fei, F.; Zhou, X.; Guo, Z.; Zhang, J.; Xu, H.; Zhang, L.; et al. Towards magnetism in pigeon MagR: Iron- and iron-sulfur binding work indispensably and synergistically. Zool. Res. 2023, 44, 142–152. [Google Scholar] [CrossRef]
  68. Li, N.; Xue, L.; Mai, X.; Wang, P.; Zhu, C.; Han, X.; Xie, Y.; Wang, B.; Ge, Y.; Zhang, Y.; et al. Transfection of clMagR/clCry4 imparts MR-T2 imaging contrast properties to living organisms (E. coli) in the presence of Fe3+ by endogenous formation of iron oxide nanoparticles. Front. Mol. Biosci. 2023, 10, 1119356. [Google Scholar] [CrossRef]
  69. Jin, Q.; Liu, J.; Zhu, W.; Dong, Z.; Liu, Z.; Cheng, L. Albumin-assisted synthesis of ultrasmall FeS2 nanodots for imaging-guided photothermal enhanced photodynamic therapy. ACS Appl. Mater. Interfaces 2018, 10, 332–340. [Google Scholar] [CrossRef]
  70. Meng, Z.; Wei, F.; Ma, W.; Yu, N.; Wei, P.; Wang, Z.; Tang, Y.; Chen, Z.; Wang, H.; Zhu, M. Design and synthesis of “All-in-One” multifunctional FeS2 nanoparticles for magnetic resonance and near-infrared imaging guided photothermal therapy of tumors. Adv. Funct. Mater. 2016, 26, 8231–8242. [Google Scholar] [CrossRef]
  71. Tang, Z.; Zhang, H.; Liu, Y.; Ni, D.; Zhang, H.; Zhang, J.; Yao, Z.; He, M.; Shi, J.; Bu, W. Antiferromagnetic pyrite as the tumor microenvironment-mediated nanoplatform for self-enhanced tumor imaging and therapy. Adv. Mater. 2017, 29, 1701683. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, J.; Sun, Z.; Wang, S.; Zhao, C.; Xu, J.; Gao, S.; Yang, M.; Sheng, F.; Gao, S.; Hou, Y. Biodegradable ferrous sulfide-based nanocomposites for tumor theranostics through specific intratumoral acidosis-induced metabolic symbiosis disruption. J. Am. Chem. Soc. 2022, 144, 19884–19895. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, W.; Xiang, C.; Xu, Y.; Chen, S.; Zeng, W.; Liu, K.; Jin, X.; Zhou, X.; Zhang, B. Albumin-constrained large-scale synthesis of renal clearable ferrous sulfide quantum dots for T1-Weighted MR imaging and phototheranostics of tumors. Biomaterials 2020, 255, 120186. [Google Scholar] [CrossRef] [PubMed]
  74. Zhou, R.; Xu, H.; Qu, J.; Ohulchanskyy, T.Y. Hemoglobin nanocrystals for drugs free, synergistic theranostics of colon tumor. Small 2023, 19, e2205165. [Google Scholar] [CrossRef] [PubMed]
  75. Samad, L.; Cabán-Acevedo, M.; Shearer, M.J.; Park, K.; Hamers, R.J.; Jin, S. Direct chemical vapor deposition synthesis of phase-pure iron pyrite (FeS2) thin films. Chem. Mater. 2015, 27, 3108–3114. [Google Scholar] [CrossRef]
  76. Shigekawa, K.; Nakayama, K.; Kuno, M.; Phan, G.N.; Owada, K.; Sugawara, K.; Takahashi, T.; Sato, T. Dichotomy of superconductivity between monolayer FeS and FeSe. Proc. Natl. Acad. Sci. USA 2019, 116, 24470–24474. [Google Scholar] [CrossRef]
  77. Mutlu, Z.; Debnath, B.; Su, S.; Li, C.; Ozkan, M.; Bozhilov, K.N.; Lake, R.K.; Ozkan, C.S. Chemical vapor deposition and phase stability of pyrite on SiO. J. Mater. Chem. C 2018, 6, 4753–4759. [Google Scholar] [CrossRef]
  78. Wu, X.; Zhao, H.; Xu, J.; Wang, Y.; Dai, S.; Xu, T.; Liu, S.; Zhang, S.; Wang, X.; Li, X. Rational synthesis of marcacite FeS2 hollow microspheres for high-rate and long-life sodium ion battery anode. J. Alloys Compd. 2020, 825, 154173. [Google Scholar] [CrossRef]
  79. Gao, W.; Razavi, R.; Fakhri, A. Preparation and development of FeS2 quantum dots on SiO2 nanostructures immobilized in biopolymers and synthetic polymers as nanoparticles and nanofibers catalyst for antibiotic degradation. Inter. J. Biol. Macromol. 2018, 114, 357–362. [Google Scholar] [CrossRef]
  80. Chen, K.; Cao, K.; Xing, C.; Hu, Y.; Liu, J.; He, Y.; Wang, J.; Li, A.; Qin, H. In-situ TEM study of the lithiation and delithiation of FeS nanosheets. J. Alloys Compd. 2016, 688, 946–952. [Google Scholar] [CrossRef]
  81. Zhu, C.; Wen, Y.; van Aken, P.A.; Maier, J.; Yu, Y. High lithium storage performance of FeS nanodots in porous graphitic carbon nanowires. Adv. Funct. Mater. 2015, 25, 2335–2342. [Google Scholar] [CrossRef]
  82. Li, J.; Zheng, J.; Wu, C.; Zhang, H.; Jin, T.; Wang, F.; Li, Q.; Shangguan, E. Facile synthesis of Fe3S4 microspheres as advanced anode materials for alkaline iron-based rechargeable batteries. J. Alloys Compd. 2021, 874, 159873. [Google Scholar] [CrossRef]
  83. Pan, F.; Liu, Z.; Deng, B.; Dong, Y.; Zhu, X.; Huang, C.; Shi, Z.; Lu, W. Magnetic Fe3S4 LTMCs micro-flowers@wax gourd aerogel-derived carbon hybrids as efficient and sustainable electromagnetic absorber. Carbon 2021, 179, 554–565. [Google Scholar] [CrossRef]
  84. Thomas, M.P.; Ullah, A.; Pham, R.H.; Djieutedjeu, H.; Selegue, J.P.; Guiton, B.S. Morphology control in the hydrothermal synthesis of FeS nanoplatelets. Cryst. Growth Des. 2020, 20, 5728–5735. [Google Scholar] [CrossRef]
  85. Zhao, T.; Elzatahry, A.; Li, X.; Zhao, D. Single-micelle-directed synthesis of mesoporous materials. Nat. Rev. Mater. 2019, 4, 775–791. [Google Scholar] [CrossRef]
  86. Zhao, C.; Shao, X.; Zhu, Z.; Zhao, C.; Qian, X. One-pot hydrothermal synthesis of RGO/FeS composite on Fe foil for high performance supercapacitors. Electrochim. Acta 2017, 246, 497–506. [Google Scholar] [CrossRef]
  87. Balakrishnan, B.; Balasingam, S.K.; Sivalingam Nallathambi, K.; Ramadoss, A.; Kundu, M.; Bak, J.S.; Cho, I.H.; Kandasamy, P.; Jun, Y.; Kim, H.-J. Facile synthesis of pristine FeS2 microflowers and hybrid rGO-FeS2 microsphere electrode materials for high performance symmetric capacitors. J. Ind. Eng. Chem. 2019, 71, 191–200. [Google Scholar] [CrossRef]
  88. Guo, S.P.; Li, J.C.; Xiao, J.R.; Xue, H.G. Fe3S4 Nanoparticles wrapped in an rGO matrix for promising energy storage: Outstanding cyclic and rate performance. ACS Appl. Mater. Interfaces 2017, 9, 37694–37701. [Google Scholar] [CrossRef]
  89. Huo, Y.; Xiu, S.J.; Meng, L.Y.; Quan, B. Solvothermal synthesis and applications of micro/nano carbons: A review. Chem. Eng. J. 2023, 451, 138572. [Google Scholar] [CrossRef]
  90. Cao, Z.; Song, H.; Cao, B.; Ma, J.; Chen, X.; Zhou, J.; Ma, Z. Sheet-on-sheet chrysanthemum-like C/FeS microspheres synthesized by one-step solvothermal method for high-performance sodium-ion batteries. J. Power Sources 2017, 364, 208–214. [Google Scholar] [CrossRef]
  91. Voronina, N.; Yashiro, H.; Myung, S.-T. Marcasite iron sulfide as a high-capacity electrode material for sodium storage. J. Mater. Chem. A 2018, 6, 17111–17119. [Google Scholar] [CrossRef]
  92. Venkateshalu, S.; Kumar, P.; Kollu, P.; Jeong, S.; Grace, A. Solvothermal synthesis and electrochemical properties of phase pure pyrite FeS2 for supercapacitor applications. Electrochim. Acta 2018, 290, 378–389. [Google Scholar] [CrossRef]
  93. Dewald, G.F.; Liaqat, Z.; Lange, M.A.; Tremel, W.; Zeier, W.G. Influence of iron sulfide nanoparticle sizes in solid-state batteries. Angew Chem. Int. Ed. Engl. 2021, 60, 17952–17956. [Google Scholar] [CrossRef]
  94. Liu, W.; Jin, L.; Xu, J.; Liu, J.; Li, Y.; Zhou, P.; Wang, C.; Dahlgren, R.A.; Wang, X. Insight into pH dependent Cr(VI) removal with magnetic Fe3S. Chem. Eng. J. 2019, 359, 564–571. [Google Scholar] [CrossRef]
  95. Tian, L.J.; Min, Y.; Li, W.W.; Chen, J.J.; Zhou, N.Q.; Zhu, T.T.; Li, D.B.; Ma, J.Y.; An, P.F.; Zheng, L.R.; et al. Substrate metabolism-driven assembly of high-quality CdSxSe1-x quantum dots in escherichia coli: Molecular mechanisms and bioimaging application. ACS Nano 2019, 13, 5841–5851. [Google Scholar] [CrossRef] [PubMed]
  96. He, S.; Hu, W.; Liu, Y.; Xie, Y.; Zhou, H.; Wang, X.; Chen, J.; Zhang, Y. Mechanism of efficient remediation of U(VI) using biogenic CMC-FeS complex produced by sulfate-reducing bacteria. J. Hazard Mater. 2021, 420, 126645. [Google Scholar] [CrossRef]
  97. Yang, H.; Gong, L.; Wang, H.; Dong, C.; Wang, J.; Qi, K.; Liu, H.; Guo, X.; Xia, B.Y. Preparation of nickel-iron hydroxides by microorganism corrosion for efficient oxygen evolution. Nat. Commun. 2020, 11, 5075. [Google Scholar] [CrossRef]
  98. Deng, X.; Dohmae, N.; Kaksonen, A.H.; Okamoto, A. Biogenic iron sulfide nanoparticles to enable extracellular electron uptake in sulfate-reducing bacteria. Angew Chem. Int. Ed. Engl. 2020, 59, 5995–5999. [Google Scholar] [CrossRef]
  99. Zheng, J.; Conrad, M. The metabolic underpinnings of ferroptosis. Cell Metab. 2020, 32, 920–937. [Google Scholar] [CrossRef]
  100. Liu, J.; Song, X.; Kuang, F.; Zhang, Q.; Xie, Y.; Kang, R.; Kroemer, G.; Tang, D. NUPR1 is a critical repressor of ferroptosis. Nat. Commun. 2021, 12, 647. [Google Scholar] [CrossRef]
  101. Shi, Z.; Naowarojna, N.; Pan, Z.; Zou, Y. Multifaceted mechanisms mediating cystine starvation-induced ferroptosis. Nat. Commun. 2021, 12, 4792. [Google Scholar] [CrossRef]
  102. Shen, Z.; Song, J.; Yung, B.C.; Zhou, Z.; Wu, A.; Chen, X. Emerging strategies of cancer therapy based on ferroptosis. Adv. Mater. 2018, 30, e1704007. [Google Scholar] [CrossRef]
  103. Ryan, S.K.; Zelic, M.; Han, Y.; Teeple, E.; Chen, L.; Sadeghi, M.; Shankara, S.; Guo, L.; Li, C.; Pontarelli, F.; et al. Microglia ferroptosis is regulated by SEC24B and contributes to neurodegeneration. Nat. Neurosci. 2023, 26, 12–26. [Google Scholar] [CrossRef] [PubMed]
  104. Gupta, A.; Mumtaz, S.; Li, C.H.; Hussain, I.; Rotello, V.M. Combatting antibiotic-resistant bacteria using nanomaterials. Chem. Soc. Rev. 2019, 48, 415–427. [Google Scholar] [CrossRef] [PubMed]
  105. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583. [Google Scholar] [CrossRef]
  106. Xu, Z.; Qiu, Z.; Liu, Q.; Huang, Y.; Li, D.; Shen, X.; Fan, K.; Xi, J.; Gu, Y.; Tang, Y.; et al. Converting organosulfur compounds to inorganic polysulfides against resistant bacterial infections. Nat. Commun. 2018, 9, 3713. [Google Scholar] [CrossRef] [PubMed]
  107. Shen, X.; Ma, R.; Huang, Y.; Chen, L.; Xu, Z.; Li, D.; Meng, X.; Fan, K.; Xi, J.; Yan, X.; et al. Nano-decocted ferrous polysulfide coordinates ferroptosis-like death in bacteria for anti-infection therapy. Nano Today 2020, 35, 100981. [Google Scholar] [CrossRef]
  108. Fang, L.; Ma, R.; Gao, X.J.; Chen, L.; Liu, Y.; Huo, Y.; Wei, T.; Wang, X.; Wang, Q.; Wang, H.; et al. Metastable iron sulfides gram-dependently counteract resistant gardnerella vaginalis for bacterial vaginosis treatment. Adv. Sci. 2022, 9, e2104341. [Google Scholar] [CrossRef] [PubMed]
  109. Shen, B.; Li, W.; Wang, Y.; Cheng, S.; Wang, X.; Zhu, L.; Zhang, Y.; Gao, L.; Jiang, L. Rapid capture and killing of bacteria by lyophilized nFeS-Hydrogel for improved healing of infected wounds. Biomater. Adv. 2023, 144, 213207. [Google Scholar] [CrossRef]
  110. Chen, Q.; Hu, Q.; Dukhovlinova, E.; Chen, G.; Ahn, S.; Wang, C.; Ogunnaike, E.A.; Ligler, F.S.; Dotti, G.; Gu, Z. Photothermal therapy promotes tumor infiltration and antitumor activity of CAR T cells. Adv. Mater. 2019, 31, e1900192. [Google Scholar] [CrossRef]
  111. Qin, X.; Wu, C.; Niu, D.; Qin, L.; Wang, X.; Wang, Q.; Li, Y. Peroxisome inspired hybrid enzyme nanogels for chemodynamic and photodynamic therapy. Nat. Commun. 2021, 12, 5243. [Google Scholar] [CrossRef] [PubMed]
  112. Zafonte, R.D.; Wang, L.; Arbelaez, C.A.; Dennison, R.; Teng, Y.D. Medical gas therapy for tissue, organ, and CNS protection: A systematic review of effects, mechanisms, and challenges. Adv. Sci. 2022, 9, e2104136. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, K.; Fang, Y.; He, Y.; Yin, H.; Guan, X.; Pu, Y.; Zhou, B.; Yue, W.; Ren, W.; Du, D.; et al. Extravascular gelation shrinkage-derived internal stress enables tumor starvation therapy with suppressed metastasis and recurrence. Nat. Commun. 2019, 10, 5380. [Google Scholar] [CrossRef] [PubMed]
  114. Ning, S.; Zheng, Y.; Qiao, K.; Li, G.; Bai, Q.; Xu, S. Laser-triggered combination therapy by iron sulfide-doxorubicin@functionalized nanozymes for breast cancer therapy. J. Nanobiotechnol. 2021, 19, 344. [Google Scholar] [CrossRef]
  115. Meng, X.; Li, D.; Chen, L.; He, H.; Wang, Q.; Hong, C.; He, J.; Gao, X.; Yang, Y.; Jiang, B.; et al. High-performance self-cascade pyrite nanozymes for apoptosis-ferroptosis synergistic tumor therapy. ACS Nano 2021, 15, 5735–5751. [Google Scholar] [CrossRef]
  116. Duan, Y.; Li, Q.; He, P.; Li, Y.; Song, J.; Wang, J.; Liu, J.; Zhou, J.; Chen, F.; Huang, Z.; et al. Ultrathin FeS nanosheets with high chemodynamic activity for sensitive colorimetric detection of H2O2 and glutathione. Chin. Chem. Lett. 2022, 33, 3217–3220. [Google Scholar] [CrossRef]
  117. Song, C.; Ding, W.; Zhao, W.; Liu, H.; Wang, J.; Yao, Y.; Yao, C. High peroxidase-like activity realized by facile synthesis of FeS2 nanoparticles for sensitive colorimetric detection of H2O2 and glutathione. Biosens. Bioelectron. 2020, 151, 111983. [Google Scholar] [CrossRef]
  118. Huang, X.; Xia, F.; Nan, Z. Fabrication of FeS2/SiO2 double mesoporous hollow spheres as an artificial peroxidase and rapid determination of H2O2 and glutathione. ACS Appl. Mater. Interfaces 2020, 12, 46539–46548. [Google Scholar] [CrossRef]
  119. Tian, R.; Abarientos, A.; Hong, J.; Hashemi, S.H.; Yan, R.; Drager, N.; Leng, K.; Nalls, M.A.; Singleton, A.B.; Xu, K.; et al. Genome-wide CRISPRi/a screens in human neurons link lysosomal failure to ferroptosis. Nat. Neurosci. 2021, 24, 1020–1034. [Google Scholar] [CrossRef]
  120. Xia, M.; Liang, S.; Li, S.; Ji, M.; Chen, B.; Zhang, M.; Dong, C.; Chen, B.; Gong, W.; Wen, G.; et al. Iatrogenic iron promotes neurodegeneration and activates self-protection of neural cells against exogenous iron attacks. Function 2021, 2, zqab003. [Google Scholar] [CrossRef]
  121. Doll, S.; Freitas, F.P.; Shah, R.; Aldrovandi, M.; da Silva, M.C.; Ingold, I.; Goya Grocin, A.; Xavier da Silva, T.N.; Panzilius, E.; Scheel, C.H.; et al. FSP1 is a glutathione-independent ferroptosis suppressor. Nature 2019, 575, 693–698. [Google Scholar] [CrossRef]
  122. Chen, D.; Chu, B.; Yang, X.; Liu, Z.; Jin, Y.; Kon, N.; Rabadan, R.; Jiang, X.; Stockwell, B.R.; Gu, W. iPLA2beta-mediated lipid detoxification controls p53-driven ferroptosis independent of GPX. Nat. Commun. 2021, 12, 3644. [Google Scholar]
  123. Stockwell, B.R. A powerful cell-protection system prevents cell death by ferroptosis. Nature 2019, 575, 597–598. [Google Scholar] [CrossRef] [PubMed]
  124. Liu, H.R.; Forouhar, F.; Seibt, T.; Saneto, R.; Wigby, K.; Friedman, J.; Xia, X.; Shchepinov, M.S.; Ramesh, S.K.; Conrad, M.; et al. Characterization of a patient-derived variant of GPX4 for precision therapy. Nat. Chem. Biol. 2022, 18, 91. [Google Scholar] [CrossRef] [PubMed]
  125. Lin, J.F.; Hu, P.S.; Wang, Y.Y.; Tan, Y.T.; Yu, K.; Liao, K.; Wu, Q.N.; Li, T.; Meng, Q.; Lin, J.Z.; et al. Phosphorylated NFS1 weakens oxaliplatin-based chemosensitivity of colorectal cancer by preventing PANoptosis. Signal Transduct. Target. Ther. 2022, 7, 54. [Google Scholar] [CrossRef] [PubMed]
  126. Du, J.; Zhou, Y.; Li, Y.; Xia, J.; Chen, Y.; Chen, S.; Wang, X.; Sun, W.; Wang, T.; Ren, X.; et al. Identification of Frataxin as a regulator of ferroptosis. Redox Biol. 2020, 32, 101483. [Google Scholar] [CrossRef] [PubMed]
  127. Du, J.; Wang, T.; Li, Y.; Zhou, Y.; Wang, X.; Yu, X.; Ren, X.; An, Y.; Wu, Y.; Sun, W.; et al. DHA inhibits proliferation and induces ferroptosis of leukemia cells through autophagy dependent degradation of ferritin. Free Radic. Biol. Med. 2019, 131, 356–369. [Google Scholar] [CrossRef]
  128. Gerber, J.; Muhlenhoff, U.; Lill, R. An interaction between frataxin and Isu1/Nfs1 that is crucial for Fe/S cluster synthesis on Isu. EMBO Rep. 2003, 4, 906–911. [Google Scholar] [CrossRef]
  129. Gervason, S.; Larkem, D.; Mansour, A.B.; Botzanowski, T.; Muller, C.S.; Pecqueur, L.; Le Pavec, G.; Delaunay-Moisan, A.; Brun, O.; Agramunt, J.; et al. Physiologically relevant reconstitution of iron-sulfur cluster biosynthesis uncovers persulfide-processing functions of ferredoxin-2 and frataxin. Nat. Commun. 2019, 10, 3566. [Google Scholar] [CrossRef]
  130. Fox, N.G.; Yu, X.; Feng, X.; Bailey, H.J.; Martelli, A.; Nabhan, J.F.; Strain-Damerell, C.; Bulawa, C.; Yue, W.W.; Han, S. Structure of the human frataxin-bound iron-sulfur cluster assembly complex provides insight into its activation mechanism. Nat. Commun. 2019, 10, 2210. [Google Scholar] [CrossRef]
  131. Tong, W.H.; Rouault, T.A. Functions of mitochondrial ISCU and cytosolic ISCU in mammalian iron-sulfur cluster biogenesis and iron homeostasis. Cell Metab. 2006, 3, 199–210. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, Y.Z.; Nitta, T.; Hiratsuka, Y.; Morishima, K. In situ integrated microrobots driven by artificial muscles built from biomolecular motors. Sci. Robot. 2022, 7, eaba8212. [Google Scholar] [CrossRef] [PubMed]
  133. Yu, R.; Zhang, H.; Guo, B. Conductive biomaterials as bioactive wound dressing for wound healing and skin tissue engineering. Nanomicro Lett. 2021, 14, 1. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, C.; Wang, X.; Zhang, W.; Ma, D.; Li, F.; Jia, R.; Shi, M.; Wang, Y.; Ma, G.; Wei, W. Shielding ferritin with a biomineralized shell enables efficient modulation of tumor microenvironment and targeted delivery of diverse therapeutic agents. Adv. Mater. 2022, 34, e2107150. [Google Scholar] [CrossRef]
Figure 1. Common forms of [FeS] clusters. These clusters have various functional roles in different proteins. (a) [2Fe-2S] cluster; (b) [3Fe-4S] cluster; (c) [4Fe-3S] cluster; (d) [4Fe-4S] cluster; (e) [4Fe-4S]RS cluster; (f) siroheme-[4Fe-4S] cluster; (g) [4Fe-5S] cluster; (h) non-cubane [4Fe-4S] cluster and (i) [4Fe-4S]-Sn-[4Fe-4S] clusters (n = 1, 5) [46]. Copyright 2022 Elsevier.
Figure 1. Common forms of [FeS] clusters. These clusters have various functional roles in different proteins. (a) [2Fe-2S] cluster; (b) [3Fe-4S] cluster; (c) [4Fe-3S] cluster; (d) [4Fe-4S] cluster; (e) [4Fe-4S]RS cluster; (f) siroheme-[4Fe-4S] cluster; (g) [4Fe-5S] cluster; (h) non-cubane [4Fe-4S] cluster and (i) [4Fe-4S]-Sn-[4Fe-4S] clusters (n = 1, 5) [46]. Copyright 2022 Elsevier.
Biomimetics 08 00177 g001
Figure 2. T2-weighted MRI performance of FeS2@BSA-Ce6 nanodots. (a) MR images of FeS2@BSA nanodots and IONPs; (b) relative relaxation rate R2 of FeS2@BSA nanodots, FeS2@BSA-Ce6 nanocomplex and iron oxide nanoparticles (IONPs); (c) MR images and (d) quantified MR signals of 4T1 tumor-bearing nude mice before and 8 h after iv injection of FeS2@BSA-Ce6 nanodots [69]. Means ± SD are shown (n = 3) (*** p < 0.001) Copyright 2017 American Chemical Society.
Figure 2. T2-weighted MRI performance of FeS2@BSA-Ce6 nanodots. (a) MR images of FeS2@BSA nanodots and IONPs; (b) relative relaxation rate R2 of FeS2@BSA nanodots, FeS2@BSA-Ce6 nanocomplex and iron oxide nanoparticles (IONPs); (c) MR images and (d) quantified MR signals of 4T1 tumor-bearing nude mice before and 8 h after iv injection of FeS2@BSA-Ce6 nanodots [69]. Means ± SD are shown (n = 3) (*** p < 0.001) Copyright 2017 American Chemical Society.
Biomimetics 08 00177 g002
Figure 3. Antibacterial activity of nFeS. (a) Antibacterial activity of nFeS converted from different organosulfur sources against S. mutans UA159; (b) antibacterial (S. mutans UA159) activity of organosulfur compounds; (c) dependence of antibacterial (S. mutans UA159) efficacy of Cys-nFeS on the amount of cysteine input into the solvothermal synthesis; (dh) antibacterial activity of nFeS on P. aeruginosa, E. coli, S. enteritidis, S. aureus and S. aureus (MDR), respectively; (i,j) ROS level and lipid peroxidation of bacteria treated with Cys-nFeS; (k) genomic DNA degradation of bacteria treated with Cys-nFeS; (l) SEM image of bacteria treated with Cys-nFeS [106]. Means ± SD are shown (n = 3) (** p < 0.01, *** p < 0.001, **** p < 0.0001) Copyright 2018 Nature Publishing Group.
Figure 3. Antibacterial activity of nFeS. (a) Antibacterial activity of nFeS converted from different organosulfur sources against S. mutans UA159; (b) antibacterial (S. mutans UA159) activity of organosulfur compounds; (c) dependence of antibacterial (S. mutans UA159) efficacy of Cys-nFeS on the amount of cysteine input into the solvothermal synthesis; (dh) antibacterial activity of nFeS on P. aeruginosa, E. coli, S. enteritidis, S. aureus and S. aureus (MDR), respectively; (i,j) ROS level and lipid peroxidation of bacteria treated with Cys-nFeS; (k) genomic DNA degradation of bacteria treated with Cys-nFeS; (l) SEM image of bacteria treated with Cys-nFeS [106]. Means ± SD are shown (n = 3) (** p < 0.01, *** p < 0.001, **** p < 0.0001) Copyright 2018 Nature Publishing Group.
Biomimetics 08 00177 g003
Figure 4. (a) Degradation behavior and (b) tumor therapy mechanism of FeS-PEG-CAI NPs [72]. Copyright 2022 American Chemical Society.
Figure 4. (a) Degradation behavior and (b) tumor therapy mechanism of FeS-PEG-CAI NPs [72]. Copyright 2022 American Chemical Society.
Biomimetics 08 00177 g004
Figure 5. (a) Schematic illustration of colorimetric detection with H2O2 and GSH based on FeS2 NPs as the sensing platforms; (b) UV–Vis absorption spectra of the reaction system with different reagents; linear calibration plots for the detection of (c) H2O2 and (d) GSH. Inset: photographs of corresponding samples [117]. Copyright 2020 Elsevier.
Figure 5. (a) Schematic illustration of colorimetric detection with H2O2 and GSH based on FeS2 NPs as the sensing platforms; (b) UV–Vis absorption spectra of the reaction system with different reagents; linear calibration plots for the detection of (c) H2O2 and (d) GSH. Inset: photographs of corresponding samples [117]. Copyright 2020 Elsevier.
Biomimetics 08 00177 g005
Table 1. The crystal structure of common iron sulfide minerals.
Table 1. The crystal structure of common iron sulfide minerals.
MineralsMain ComponentCrystallographic SystemReference
PyriteFeS2Cubic system[30]
PyrrhotiteFe1-xSHexagonal or monoclinic system[35]
MarcasiteFeS2Orthogonal system[36]
MackinawiteFeSTetragonal system[38,39]
Table 2. Magnetic properties of various iron-based sulfides.
Table 2. Magnetic properties of various iron-based sulfides.
NameR1 (mM−1 S−1)R2 (mM−1 S−1)MorphologyReference
FeS2/85.36Nanodots[69]
/31.836Nanoparticles[70]
118.14Nanocrystals[71]
FeS/209.8Nanoplates[64]
/40.159Nanoparticles[72]
5.35/Quantum dots[73]
Fe1-xS/36.09Nanocrystals[74]
Table 3. Effect of photothermal therapy mediated by various iron-based sulfides.
Table 3. Effect of photothermal therapy mediated by various iron-based sulfides.
NameLaserPhotothermal Conversion EfficiencyWeight Extinction Coefficient (L g−1 cm−1)Tumor TypeReference
BSO-FeS2808 nm49.5%/4T1 cells[59]
FeS2@BSA-Ce6808 nm//4T1 cells[69]
FeS2-350915 nm33.1%/7721 cells[70]
FeS2-PEG808 nm28.6%/4T1 cells[71]
FeS@BSA660 nm30.04%/4T1 cells[73]
FeS-PEG808 nm/15.5 4T1 cells[64]
FeS-PEG-CAI1064 nm56.51%/4T1 cells[72]
Fe1-xS-PVP808 nm24%/PAN-02 cells[13]
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

Duan, Y.; Sun, J. Preparation of Iron-Based Sulfides and Their Applications in Biomedical Fields. Biomimetics 2023, 8, 177. https://doi.org/10.3390/biomimetics8020177

AMA Style

Duan Y, Sun J. Preparation of Iron-Based Sulfides and Their Applications in Biomedical Fields. Biomimetics. 2023; 8(2):177. https://doi.org/10.3390/biomimetics8020177

Chicago/Turabian Style

Duan, Yefan, and Jianfei Sun. 2023. "Preparation of Iron-Based Sulfides and Their Applications in Biomedical Fields" Biomimetics 8, no. 2: 177. https://doi.org/10.3390/biomimetics8020177

Article Metrics

Back to TopTop