Skip to main content
iScience logoLink to iScience
. 2024 Mar 25;27(5):109555. doi: 10.1016/j.isci.2024.109555

Unveiling the role of inorganic nanoparticles in Earth’s biochemical evolution through electron transfer dynamics

Xiao-Lan Huang 1,
PMCID: PMC11024932  PMID: 38638571

Summary

This article explores the intricate interplay between inorganic nanoparticles and Earth’s biochemical history, with a focus on their electron transfer properties. It reveals how iron oxide and sulfide nanoparticles, as examples of inorganic nanoparticles, exhibit oxidoreductase activity similar to proteins. Termed “life fossil oxidoreductases," these inorganic enzymes influence redox reactions, detoxification processes, and nutrient cycling in early Earth environments. By emphasizing the structural configuration of nanoparticles and their electron conformation, including oxygen defects and metal vacancies, especially electron hopping, the article provides a foundation for understanding inorganic enzyme mechanisms. This approach, rooted in physics, underscores that life’s origin and evolution are governed by electron transfer principles within the framework of chemical equilibrium. Today, these nanoparticles serve as vital biocatalysts in natural ecosystems, participating in critical reactions for ecosystem health. The research highlights their enduring impact on Earth’s history, shaping ecosystems and interacting with protein metal centers through shared electron transfer dynamics, offering insights into early life processes and adaptations.

Subject area: chemistry, inorganic chemistry, biochemistry

Graphical abstract

graphic file with name fx1.jpg


Chemistry; Inorganic chemistry; Biochemistry

Introduction

The oxidoreductases, key players in orchestrating electron transfer (ET) reactions across various organisms, are vital components in enzymatic activities.1,2,3 Their versatility is demonstrated through interactions with a wide array of substrates, both organic and inorganic, though they may generate reactive byproducts. Some oxidoreductases directly transfer electrons or utilize mediators such as cytochrome c to solid surfaces.4,5,6,7 This phenomenon, termed direct electron transfer (DET),4,5,6,7,8,9,10 was initially documented in 1977 by Eddows and Hill11 and Yeh and Kuwna.12 Their studies highlighted Cyc c’s reversible electrochemistry on gold and tin-doped indium oxide electrodes, observed through cyclic voltammetry. Notably, horseradish peroxidase (HRP)13 and laccase (Lc)14 also adhere to carbon electrodes, showcasing DET capacity. Impressively, more than 100 enzymes, with a predominant presence of oxidoreductases, are known to function under DET conditions.7,15,16,17,18,19,20,21 They play pivotal roles in metabolism, photosynthesis, and detoxification reactions. In the primordial conditions of the early Earth, oxidoreductases could have been crucial in primordial energy metabolism and the synthesis of biomolecules.

In 2007, Fe3O4 nanoparticles were found to exhibit peroxidase (POD)-like activity, giving rise to the field of nanozymes.22 Since then, nanozymes have gained significant attention in the fields of chemistry and biomedicine,23,24,25,26,27,28,29 finding applications in diverse areas, including medical, agriculture, and environmental protection. At the same time with another independent research, it was also observed that inorganic iron oxide nanoparticles can serve as inorganic phosphatase30,31,32 (Figure S1). This observation led to the proposal of the “inorganic enzyme hypothesis," suggesting that the metal architecture of inorganic nanoparticles bears similarities to proteins based on the experimental evidence.31,32,33

While the mechanisms of inorganic nanozyme continue to be debated,34,35,36,37 the notion that inorganic nanozymes should be considered inorganic enzymes has been gaining traction.38,39 This perspective underscores their role as biocatalysts and their potential contribution to the origin and development of life.

This review delves into inorganic nanomaterials as biocatalysts, focusing on their role as examples of inorganic oxidoreductases. It explores their involvement in ET processes, considering the relationship between the metal structure of nanoparticles and electron hopping, as well as the influence of the electron configuration of the metals involved. This sheds light on their potential as key players in biocatalysts and their relevance to the origin of life and its earliest evolution.

Inorganic nanoparticles with intrinsic oxidoreductase activity

In 2007, Gao et al. made a groundbreaking discovery regarding the POD activity of Fe3O4 nanoparticles.22 These nanoparticles exhibited catalytic activity in the oxidation of various substrates, showing kinetics following Michaelis–Menten curves. The apparent Michaelis-Menten constants (Km) of HRP and 30 nm Fe3O4 particles measured with H2O2 as substrate are 3.70 and 154 mM, respectively. In contrast, the respective Km values of HRP and 30 nm Fe3O4 measured with 3,3,5,5-tetramethylbenzidine (TMB) as substrate are 0.434 and 0.098 mM. The catalytic rates (kcat) of HRP and 30 nm Fe3O4 particles measured with H2O2 or TMB are 3.48 × 103 and 8.58 × 104 s−1 or 4.0 × 103 and 3.02 × 104 s−1, respectively.22 Thus, in terms of their catalytic efficiency (kcat/Km) the nanoparticles (H2O2: 560 mM−1 s−1; TMB: 3.1 × 105 mM−1 s−1) are at least as good as their biological counterpart (H2O2: 940 mM−1 s−1; TMB: 9.2 × 103 mM−1 s−1). Notably, the activity was related to the particle size. For nanoparticles with a diameter of 300 nm, the catalytic efficiency of a single nanoparticle was comparable to that of a single HRP molecule.22 Other inorganic iron oxide nanoparticles also displayed intrinsic peroxidase-like activity, although at a lower specific activity compared to HRP.40,41,42,43,44,45

Additionally, some iron oxide nanoparticles have exhibited multiple activities, including catalases (CAT), superoxide dismutase (SOD), and oxidases (OXD) activities associated with H2O2 and free radicals.43,45,46,47,48,49,50,51 Ferrihydrites, in particular, have been found to possess high intrinsic CAT-like activity and low intrinsic POD-like activity due to the abundance of hydroxyl groups in their nanocrystals49 (Figure S2). Actually, Ferrihydrites, functioning as inorganic CAT, have demonstrated in vivo their ability to produce O2 within the tumor microenvironment, leading to a substantial enhancement of the therapeutic effect of radiotherapy.49,52

Iron oxide nanoparticles were not the only inorganic nanomaterials to demonstrate oxidoreductase activity. Iron sulfide nanoparticles, including various phases such as FeS, Fe1−xS, and FeS2, were also found to catalyze peroxidase substrates, albeit with different kinetic parameters compared to HRP53,54,55,56 (Figure S3). For instance, Fe7S8 nanowires exhibited particularly high POD activity.53 The apparent Km value for Fe7S8 NWs with TMB was determined to be 0.548 mM, significantly lower than that of Fe3O4 and even five times lower than that of HRP, indicating that Fe7S8 NWs have a higher affinity for TMB.22,53 Moreover, the Km value for Fe7S8 NWs with H2O2 (0.895 mM) indicates a similar affinity for H2O2 compared to HRP and better H2O2 affinity than Fe3O4 nanoparticles.22,53 Furthermore, iron polysulfide nanoparticles (Fe1−xS and Fe3S4) displayed both POD and CAT activities, allowing them to break down H2O2 into free radicals and oxygen.57 These nanoparticles, especially Cys0.5-nFeS, demonstrated the highest activity.57

Density function theory (DFT) studies highlighted the importance of the architecture of inorganic nanomaterials in their catalytic activity (Figure S4). The mechanism for inorganic POD activity involves the chemisorption of H2O2 onto the nanomaterial surface, generating two hydroxyl adsorbates that undergo two reduction processes to remove the hydroxyl groups from the surface.58 The chemisorption energy and thus the POD-like activity are dependent on the metal architecture in the nanomaterials58 as will be discussed in more detail in the following section.

Importantly, it was found that even naturally occurring iron oxide nanoparticles in magnetosomes from magnetotactic bacteria (MTB) retained POD activity after removing the protein membrane,59,60 suggesting that the POD activity is an inherent property of the nanoparticles themselves. Biogenic iron oxide nanoparticles from Burkholderia sp YN01v (Fe3O4),61,62 Comamonas testosterone (Fe1.44O0.32(OH)3.86),63 and Acinetobacter strains (Fe0.96O0.88(OH)1.12)64 also exhibited intrinsic POD, SOD, and CAT activity (Figure S5A).

Ferritin, an iron storage protein present in various organisms, including archaea, bacteria, and eukaryotes, features an iron core that serves as a natural nano-structure with biogenetic iron minerals.65,66,67,68 Belarusian scientists found in 1999 that iron-containing crystallites in ferritins have POD activities,69 as biocatalysts, following Michaelis-Menten kinetics in the oxidation of TMB by hydrogen peroxide in acetate buffer solution (pH 4.2) in aqueous solution before the observation of Gao et al.22 Linear Lineweaver–Burk plots have been presented both for TMB (up to 3.0 mM) and H2O2 (up to 50 mM). The POD activity of the mineral core of ferritin was further confirmed with TMB, o-phenylenediamine (OPD), and N, N-diethyl-1,4-phenylenediamine (DPD) as substrates at 50°C in 201170 (Figure S5B). Their experiments also indicated that the POD activity of horse spleen ferritin was derived from its ferric nanocore but not from the ferritin protein.70 A recent study on various ferritins from different biological groups potentially spanning different evolutionary lineages has further shown that these biogenic iron cores, after the removal of proteins from ferritins, exhibit enzyme-like activity (POD, CAT, OXD, and SOD) due to their metal structure, rather than the organic compounds that make up ferritins, particularly the amino acid sequences in these proteins (Figure S5C).71 For Archaea, the study examined ferritins from Pyrococcus furiosus (WP_011011871.1, labeled as pfFn), Pyrococcus yayanosii (WP_013905435.1, labeled as pyFn), and DNA protection protein from Sulfolobus solfataricus (NP_343470.1, labeled as ssDps). Bacteria were represented by Bacterioferritin (EcBfr) from Escherichia coli str. K-12 substr. MG1655 (NP_417795.1), Non-heme ferritin (EcFTn) from Enterobacteriaceae (WP_000917208.1), and DNA binding protein (EcDps) from starved Escherichia coli str. K-12 substr. MG1655 (NP_415333.1). Eukaryotic ferritins were represented by Heavy chain (HFn) from Homo sapiens (NP_002023.2) and Light chain (LFn) from Homo sapiens (NP_000137.2). This classification into different groups highlights the diversity of the iron cores from ferritins and their enzyme-like activity across various biological domains.71

Photoreactivation, the responsiveness of certain enzymes and inorganic nanozymes to light, plays a crucial role in various biological and chemical processes. This phenomenon is particularly prominent in oxidoreductases, enzymes involved in ET during redox reactions.72 PODs and OXDs containing iron-containing heme groups, such as HRP isozymes, demonstrate sensitivity to ultraviolet (UV) and blue light.73 The exact mechanisms behind this light-induced transformation remain uncertain, but aromatic amino acids such as tryptophan and tyrosine are implicated in electron tunneling processes.

This light-induced influence extends to inorganic nanozymes, including magnetite within MTB,59,60 α-Fe2O3 nanoparticles,74 TiO2 nanoparticles,75 Cu2Se hollow nanocubes,76 NH−MoO3−x nanobelts,77 cobalt sulfide (Co9S8) nanodot,78 and CuO.79 For example, α-Fe2O3 nanoparticles exhibit enhanced substrate affinity and catalytic efficiency under light irradiation.74 Magnetosomes and naked magnetosomes also demonstrate heightened sensitivity to light, leading to improved substrate affinity.60 CuO nanorods show increased peroxidase-like activity under visible light, with a significant improvement in the generation of ⋅OH radicals79 (Figure S6). Composite nanoparticles such as TiO2 nanotubes coated with MoS2 nanoflowers (TiO2NTs@MoS2),80 plasmonic gold nanoparticle-modified Mn3O4 nanozyme (Mn3O4-Au),81 Fe3O4/Ag/Bi2MoO6 nanoparticle82 and quaternized chitosan (QCS)/silver (Ag)/cobalt phosphide (CoP) nanocomposites,83 all exhibiting photo responsive attributes. This characteristic allows for the modulation and enhancement of their catalytic activity using light energy.

Furthermore, the phenomenon of photoreactivation extends to near-infrared-II (NIR-II) light (1000–1700 nm). Cu2MoS4 (CMS) nanoparticles showcase both oxidase and peroxidase activities upon NIR-II light exposure.84 In the presence of NIR-II light, CMS NPs demonstrate OXD activity using L-ascorbic acid (AA) as a substrate. The Km and Vmax of CMS NPs for AA are determined as 12.06 μM and 0.11 μM s−1, respectively. Furthermore, CMS NPs display OXD activity with glutathione (GSH) and TMB as substrates.84 Under NIR-II light irradiation, GSH levels exhibit a decrease of more than 2-fold compared to reactions without light irradiation, indicating the enhanced OXD activity of CMS NPs. POD activity of Cu2MoS4 nanoparticles is also evident under NIR-II light. The Km of CMS NPs for TMB and H2O2 is determined as 1.36 mM and 25.46 mM, respectively. The Vmax of CMS NPs for TMB and H2O2 measures 27.29 × 10−8 M s−1 and 42.81 × 10−8 M s−1, respectively.84 Notably, the absorbance of the oxidized TMB product following treatment with CMS NPs and NIR-II light shows a 2-fold increase compared to CMS NPs alone, indicating the efficient acceleration of POD catalysis by NIR-II light.84 They demonstrate enhanced activity under NIR-II light, indicating the potential for utilizing this range of light in catalytic applications.

Cytochrome c (Cyt c) is a versatile protein found in a wide range of organisms, playing a crucial role in mediating ET during oxidative reactions.85 It belongs to the c-type Cytochrome family and is characterized by its distinctive CXXCH amino acid motif, which forms the binding site for the heme group embedded within the protein’s structure.86 Cyt c is primarily located in the inner membranes of mitochondria and chloroplasts, where it orchestrates electron transport chains essential for energy production.20,87 It shuttles electrons from oxidoreductases to terminal electron acceptor proteins or electrodes, facilitating processes such as ATP synthesis and chloroplast photosynthesis.88 The heme group (Haem) within Cyt c’s versatile active site enables its participation in various catalytic activities and ET reactions.89

In addition to its role in protein environments, Cyt c interacts with inorganic nanoparticles, exhibiting enzyme-like behaviors. For instance, when combined with WO3 nanoparticles, Cyt c demonstrates significant peroxidase-like activity, especially in the presence of water and hydrogen peroxide.90 Similarly, Cu2O nanoparticles mimic the functionality of the mitochondrial enzyme Cyt c oxidase,91 while CeVO4 nanoparticles serve as mitochondrial enzyme CcO, utilizing Cyt c as a biological electron donor in the four-electron reduction of molecular oxygen, thus integrating SOD-like functions.92 CdS nanorods accept electrons transferred from Cyt c as a bio-electron acceptor, catalyze the formation of superoxide anions, and exhibit NADH oxidase-like activity.93 Under physiological pH conditions, CdS nanorods, with the assistance of light, couple with dehydrogenase to recycle NADH.

Cyt c electron mediators are believed to have emerged around 3.5 billion years ago during the Archean era, a crucial epoch in Earth’s geological and environmental history. During this period, MTB demonstrated exceptional adaptations and biomineralization capabilities,94,95 utilizing specialized organelles known as magnetosomes to navigate using Earth’s magnetic field.96,97,98 These magnetosomes housed magnetic minerals such as magnetite (Fe3O4) and greigite (Fe3S4),94,95 with proteins containing the crucial CXXCH c-type Cytochrome-binding motifs integral to their structure.95,96,99,100,101 The Archean atmosphere contained significant concentrations of H2O2, presenting challenges for nascent life forms.102,103,104,105,106,107 Reactive oxygen species (ROS), including H2O2, could induce oxidative damage to cellular components. Defense mechanisms such as SOD, CAT, and POD evolved early to protect against ROS-induced damage across diverse domains of life.108,109,110,111,112

Laboratory investigations suggest that diverse iron oxide nanoparticles, precursors to magnetite, initially formed within MTB.94,113,114,115,116,117,118,119,120 These nanoparticles became encapsulated within lipid bilayer membranes, akin to structures found in mitochondria and chloroplasts.121 Iron oxide nanoparticles such as Fe3O4 and ferrihydrite could traverse lipid bilayers and enter cellular compartments without disrupting plasma membranes.122,123 The POD activity exhibited by magnetite within magnetosomes likely played a pivotal role in driving the formation of magnetosomes, as it conferred a survival advantage in environments rich in hydrogen peroxide.59,60,108,109,110,111,112 This “life fossil oxidoreductase” concept highlights the enduring role of magnetite nanoparticles within magnetosomes, acting as vital enzymes for ET in redox reactions.38 It suggests an adaptation to endure high hydrogen peroxide conditions and oxidative stress throughout Earth’s history.

Genetic analyses of magnetosome assembly reveal intricate processes that offer invaluable insights into the early evolutionary stages shaping life’s complexity.124,125 The convergence of Cyt c electron mediators and magnetosomes, coupled with the concept of the “life fossil oxidoreductase,”38 illuminates the innovative strategies adopted by early life during the Archean era to thrive amidst challenges such as hydrogen peroxide. This underscores the pioneering use of inorganic nanoparticles for survival and adaptation, ultimately contributing to the evolution and diversity of life forms we observe today.

In summary, inorganic nanoparticles, such as iron oxide, iron sulfide, and biogenic nanoparticles, exhibit significant oxidoreductase activities, including peroxidase, catalase, oxidase, and superoxide dismutase-like functions. Furthermore, certain nanoparticles demonstrate increased catalytic abilities in response to light, especially in the UV, blue, and NIR-II spectrum due to photoirradiation, along with the mediating effect of Cytochrome c for electron transfer. These groundbreaking concepts suggest that these nanoparticles serve as enduring biocatalysts crucial for redox reactions, particularly in early life and the emergence of life. These groundbreaking concepts suggest that these nanoparticles serve as enduring biocatalysts crucial for redox reactions, particularly in early life and the emergence of life. This adaptation allowed different life forms to thrive in environments with high levels of ROS, such as hydrogen peroxidase. This symbiotic relationship played a pivotal role in the various types of organisms' ability to thrive and adapt in environments characterized by heightened ROS levels. Together, these findings highlight the innovative utilization of inorganic nanoparticles for survival and adaptation on Earth, paving the way for the evolution and diversity of life forms we observe today.39

Metal architectures in inorganic nanoparticles

The metal architecture in iron oxides is primarily determined by the ferric-ferrous composition (Fe (III)/Fetotal) and the hydroxylation ratio (OH/Fetotal)126,127,128 (Figure S7A). A fundamental structural unit of ferrihydrite and other iron oxide nanoparticles is the Back-Figges δ-Keggin cluster (Fe13), comprising 13 iron and 40 oxygen atoms129,130 (Figure S7B). Within this structure, central tetrahedrally coordinated Fe is linked to 12 peripheral octahedrally coordinated Fe atoms, forming edge-sharing groups of three through oxo bridges. Iron oxide nanoparticles between 2 and 6 nm can be visualized as a three-dimensional arrangement of these clusters, with adjacent clusters connected by various combinations of edge, corner, and face-shared octahedra, forming oxo bridges within the cluster129 (Figure S7C). The distance between Fe atoms depends on the type of connections, with corner sharing having the longest distance (3.39–3.70 Å) and face sharing the shortest (2.88 Å)126,131 (Figure S7D).

Magnetite possesses an inverse cubic spinel structure, where metal ions are distributed between tetrahedral and octahedral sites. Tetrahedral sites house Fe (III) ions, while octahedral sites contain both Fe (III) and Fe (II) ions in equal measure. This leads to the chemical formula Fe(III)tetra[Fe(II)Fe(III)]octa(O2−)4, indicating the position of ferrous and ferric ions within the structure132 (Figure S7E). Similarly, maghemite displays a spinel crystal structure with all iron cations in the trivalent state, and the charge neutrality of the cell is maintained through the presence of cation vacancies. The maghemite structure can be approximated as a cubic unit cell with the composition, (Fe8III)A[Fe403III83]BO32,where ◻ represents a vacancy, A indicates tetrahedral, and B octahedral coordination sphere.126,133,134,135 In contrast, hematite possesses a primitive rhombohedral structure, with Fe (III) and oxygen atoms arranged differently, forming layers along the 3-fold axis in a hexagonal pattern.126,136,137,138In the crystal structure of trigonal hematite, the oxygen atoms are stacked in approximately closed-packed layers along the 3-fold axis. These layers are arranged in a hexagonal pattern.139

It is important to highlight that mixed-valent Fe minerals in the environment exhibit limited stability and undergo transformations through both abiotic and biotic pathways. Various factors such as oxygen levels, ROS, light, nitrate (NO3), different forms of iron (II) and iron (III), and phosphorus concentrations can influence the transformation of iron oxide minerals140,141 or the metal architecture of iron oxide nanomaterials126,128 (Figure S8A). These external influences induce changes in the oxidation state, reactivity, and properties of iron, affecting its behavior in biological and biogeochemical processing.126,142,143,144,145 For instance, the study by Usman et al. demonstrates the transformation of ferric (oxyhydr)oxides, such as ferrihydrite, lepidocrocite, and goethite, into magnetite in the presence of Fe (II) ions.146 The reactivity and transformation sequence were found to be dependent on several factors, including the initial mineralogy of the oxyhydroxides, aging time, and solution chemistry. The order of reactivity for the transformation into magnetite was observed to be ferrihydrite > lepidocrocite > goethite146 (Figure S8B).

Solar irradiation also contributes to the formation of iron oxide nanoparticles. Even in the absence of oxygen in the atmosphere, low levels of ferric ions (Fe III) would have been generated at the ocean surface due to photooxidation.147 Solar irradiation promotes the transformation of ferrihydrite into goethite compared to dark conditions.148,149 For example, enhanced pathways of sunlight-driven ferrihydrite transformation in the presence of dissolved oxygen were observed, demonstrating the significant role of sunlight in the conversion of ferrihydrite to goethite.149 Photoinduced ET processes at the ferrihydrite interface generate redox active species, including hole-electron pairs, reactive radicals, and Fe (II). The production of hydroxyl radicals occurs through water oxidation, reduction of dissolved oxygen, and photolysis of Fe(III)-hydroxyl complexes149 (Figure S8C). Under acidic conditions, superoxide radicals act as oxidants for Fe (II) reoxidation, promoting the transformation of ferrihydrite. It is worth noting that the presence of inorganic ions such as chloride, sulfate, and nitrate not only influences the hydrolysis and precipitation of Fe (III) but also impacts the generation of radicals through photoinduced charge transfer reactions.149

Magnetite (Fe3O4) nanoparticles can be converted into maghemite (γ-Fe2O3) not only by oxidation with oxygen but also through various ions and/or ETs across the solid–solution interface.150 This involves the elimination of an electron from a surface ferrous ion, leading to the formation of a ferric ion and a cationic vacancy in the octahedral lattice, maintaining electrical neutrality. Electron mobility within the solid allows for the renewal of the ferrous sites at the surface, resulting in maghemite formation within the particles. At the same time, cationic vacancies are created by the migration of ferric ions to the surface, preserving the electroneutrality of the particles. The oxidation conditions significantly influence this process.150 Even in the absence of oxygen, 10 nm-sized stoichiometric magnetite particles (Fe (II)/Fe (III) = 0.5) in aqueous solutions over a biologically and environmentally relevant pH range (4–7) are still not stable. Fe (II) is released into the solution due to the H+-promoted dissolution process, leading to the partial oxidation of magnetite to a magnetite-maghemite solid solution.151 Under harsh conditions, such as exposure to a 0.07 mol HNO3 solution at 60°C in an air atmosphere, the process is much slower, with the complete oxidation of ferrous iron ions being observed, but the crystal phase of maghemite has not yet formed over 24 h.152 Moreover, recent research has revealed that when magnetite 111 is prepared under oxidizing or reducing conditions, the metal architecture at the surface, i.e., iron versus oxygen fractional surface terminations, is also different. A larger fraction of Fe-termination was found in the magnetite preparation under reducing conditions.153

When iron oxide nanoparticles interact with bacteria, especially with Shewanella oneidensis MR-1, a strain known for its metal-reducing abilities, remarkable transformations occur. This bacterium induces a significant reduction in hematite, resulting in the generation of substantial amounts of Fe (II). This process also leads to a noteworthy alteration in the crystalline structure of hematite, transitioning from a hexagonal to a cubic system. The bacterially catalyzed the reductive dissolution of hematite gives rise to intermediate states and the emergence of unique chemical environments characterized by Fe(II)/Fe(III) complexes with monodentate and bidentate coordination patterns. This transformative process involves the cleavage of iron-oxygen bonds within the hematite structure, resulting in the development of microstructures that form complexes with iron. These microstructures play a pivotal role in facilitating the creation of biogenic magnetite. Furthermore, observations from electron paramagnetic resonance (EPR) spectra indicate a diminishing EPR intensity over time, suggesting changes in the composition and local structures of both hematite and magnetite throughout the transformation process. DFT calculations further support the earlier observation regarding the octahedral configuration that arises as a consequence of Fe (II) production.154

In another study, Fe3O4 nanoparticles were exposed to a solution containing E. coli for 1 hour. As a result, approximately 46–48% of γ-Fe2O3 was generated, while only 22% of γ-Fe2O3 was generated in the control group (pure water, pH 5, 30°C). This was determined by comparing the changes in Fe−K-edge XANES spectra. No structural modifications of maghemite nanoparticles were observed, indicating that maghemite remained highly stable after direct contact with bacteria.155

Metal ion architecture in nanoparticles significantly influences inorganic oxidoreductase activities, particularly the POD activity of iron oxide nanoparticles.49,156,157,158,159,160,161,162,163,164,165,166,167,168 Fe3O4 nanoparticles demonstrate higher POD activity compared to γ-Fe2O3 and α-Fe2O3 nanoparticles.169 Notably, cycling catalysts can transform Fe3O4 to γ-Fe2O3, reducing ET rates and subsequently lowering POD activity. Specific POD activities for Fe3O4, α-Fe2O3, and γ-Fe2O3 nanoparticles are 1.79, 0.45, and 0.03 U·mg−1, respectively169 (Figure S9). The specific activity of Fe3O4 nanoparticles decreases over time, as supported by spectroscopic data.169 Prolonged reaction times lead to the gradual oxidation of interior Fe2+, affecting both surface and internal Fe2+ contributions to POD-like activity. Fe2+ ions within Fe3O4 nanoparticles facilitate ET to the surface, enabling sustained catalytic reactions. However, excess oxidized Fe3+ions migrating from the lattice act as the rate-limiting step, gradually diminishing the catalytic activity of the regenerated nanoparticles.169

The interplay between minerals and microbiology highlights the pivotal role of metal architecture in nanoparticles, a phenomenon widespread in nature. This is exemplified in the interaction between proliferating fungi and iron-rich hematite, resulting in the generation of biogenic ferrihydrite nanoparticles with inherent POD activity.170 Alterations in the non-lattice oxygen within the iron architecture largely contribute to their catalytic behavior, expanding our understanding of metal architecture’s influence on enzyme-like activity. Fungi actively mediate the catalytic reactions, confirmed by advanced microscopy techniques and biomass analyses170 (Figure S10). Another example involves incubating magnetite nanoparticles with T. guizhouense, resulting in a significant enhancement of their POD activity.171 Detailed spectroscopy analysis reveals changes in the metal architecture, particularly affecting non-lattice oxygen. This showcases nature’s ability to generate nanoparticles with oxidoreductase activity through modifications in their metal architecture. These examples illustrate the dynamic nature of metal architectures in iron oxide nanoparticles during fungal interactions. The interplay between fungi and iron oxide minerals leads to structural transformations directly impacting catalytic properties. This insight illuminates the intricate relationship between microorganisms and nanomaterials, emphasizing the role of metal architecture in dictating catalytic behaviors within these complex biological and chemical processes.

The study related to PVC dichlorination residues and iron chips treated with subcritical water revealed the presence of Fe2O3 nanoparticles in wastewater, demonstrating their intrinsic peroxidase-like properties driven by the unique metal architecture of iron oxide nanoparticles.172 This highlights the dynamic nature of metal architectures in iron oxide nanomaterials. Both ferrous and ferric ions within these nanoparticles occupy octahedral and tetrahedral coordinated interstitial sites within a closely packed anionic lattice of O2−/OH. The distinctive properties of these nanoparticles arise from their electron configuration and oxygen coordination, including the presence of oxygen vacancies (OVs).126,135,171,173,174,175,176,177,178 This perspective extends to various iron sulfide nanoparticles. Much like iron oxide nanoparticles, diverse iron sulfide particles can undergo structural modifications due to chemical and microbial interactions.179,180,181,182,183,184,185,186,187,188 These interactions can lead to shifts in atomic and ionic arrangements within the nanoparticles, ultimately reshaping their metal architecture and influencing their properties. The metal architecture plays a crucial role in determining the functionalities and reactivity of these nanomaterials, making them remarkably adaptable, particularly in response to biological changes and catalytic behaviors, especially in the context of nanozymes. The interplay between iron oxide and iron sulfide nanomaterials with environmental factors and biological systems opens up new avenues for understanding the origins of life, the evolution of proteins, and their ecological functions. Examining these interactions sheds light on the fundamental processes that contributed to the emergence of life on our planet. Keep in mind that the story does not end with the elucidation of these static attributes. Nanoparticle structures are not rigid and unchanging entities; instead, they respond dynamically to a myriad of external influences, including both abiotic and biotic factors.

Electron conduction mechanisms in inorganic nanoparticles

In solid-state physics, energy bands are crucial for understanding a material’s electrical properties. The valence band (VB) is the highest energy level filled with electrons involved in bonding.189 In contrast, the conduction band (CB) allows electrons to move freely, acting as charge carriers for electric current.189 Electron conductivity, or electrical conductivity, depends on how easily electrons move through a material.190,191 Electron mobility, which measures this ease of movement, and carrier concentration, the abundance of charge carriers, influence conductivity. Semiconductors and insulators are defined by their band gap, the energy difference between the valence and conduction bands.189 Materials characterized by elevated electron conductivity, featuring heightened electron mobility and carrier concentration, exhibit superior capabilities for effective ET, thereby catalyzing ET reactions adeptly. These materials hold paramount significance in shepherding the ET steps integral to biological functions and technological applications alike. Conversely, materials endowed with diminished electron conductivity may confront hindrances in expediting ET processes. In such scenarios, the pace or efficiency of ET reactions could wane, impairing the overall functionality of implicated systems.

In the realm of inorganic nanoparticles, the band gap plays a pivotal role in ET processes, determining the energy required for electron migration between the nanoparticle and other molecules or surfaces. ET is a fundamental process influencing interactions between inorganic nanomaterials and various organic entities, including large organic compounds such as proteins and DNA,5,6,7,192 as well as microorganisms.4,193,194 Numerous factors influence ET efficiency, with the band gap closely linked to the nanoparticles' electron configuration and structural characteristics, including size, structure, surface effects, composition, doping, and quantum confinement. This process holds significant sway over various biological mechanisms, especially in the realm of energy transfer within oxidoreductases from a physical standpoint.

To exemplify, the electron configuration of Fe (II) (iron ion with a +2 charge) and Fe (III) (iron ion with a +3 charge) is [Ar] 3d6 and [Ar] 3d5, respectively. In contrast, a neutral oxygen atom (O) assumes an electron configuration housing 2 unpaired electrons within the 2p subshell. These electron configurations contribute to a multifaceted electronic landscape within nanoparticles, fostering electron delocalization and mobility. Within the crystalline lattice, the transfer of electrons unfolds amid Fe (II) and Fe (III) ions, both inhabiting diverse oxidation states. The fusion of Fe 3d electrons with oxygen affords leeway for electron delocalization and mobility within the nanoparticle’s structural matrix. In turn, this underpins a phenomenon christened electron hopping—where localized electrons traverse between Fe (II) and Fe (III) ions. The impetus propelling this electron hopping derives from the oxidation state disparity between ions and the band gap, a metric representing the energy prerequisite for electron transitions.

The occurrence of electron hopping is postulated by theoretical models pertaining to myriad iron oxide nanomaterials, encompassing green rust,195 magnetite,196 iron oxyhydroxides,197,198 and hematite.199,200 Additionally, experimental observations have confirmed electron hopping on the surfaces of iron (oxyhydr)oxide,201 α-Fe2O3,202,203 and γ-Fe2O3204 NPs. The electron small-polaron hopping rate, quantifying the velocity of electron hopping, emerges as a pivotal factor impacting the kinetics of numerous iron redox reactions. This rate hinges on the short-range structural topology201 or lattice expansion of iron oxide,202 holding precedence over nanoparticle dimensions, suspension pH, or the infusion of multiple electrons per nanoparticle.205 These nuances underscore the central role played by the short-range structural topology and lattice expansion of iron oxide in orchestrating efficient electron hopping, highlighting the elemental role of electron hopping stemming from the nanoparticles' iron and oxygen electron configuration.

The electrical conductivity of iron oxide nanoparticles is subject to an array of influences, including the crystal structure and the spatial disposition of iron ions. Notably, charge-ordering (CO) emerges as a pivotal phenomenon underpinning electrical conductivity properties. Charge ordering denotes the organized arrangement of distinct metal oxidation states within a material’s crystal lattice. In the context of iron oxide nanoparticles, charge-ordering surfaces in magnetite at temperatures below the Verwey transition temperature of 120 K.206 Within iron oxide nanoparticles, the Fe-Fe distance within octahedral chains wields significant sway over charge ordering and electron mobility. The gap separating iron ions within these chains reverberates through the extent of charge transfer and electron hopping across different oxidation states, thereby wielding sway over the overall electrical conductivity of the material.

Both α-Fe2O3 and γ-Fe2O3 emerge as n-type semiconductors,126 denoting their capacity to conduct electricity through the orchestrated movement of electrons. Hematite’s band gap measures 2.2 eV, whereas maghemite’s slightly undercuts it at 2.03 eV.126 On the converse, wüstite (FeO) upholds status as a p-type semiconductor,126 with electricity conduction transpiring through hole movement, characterized by a band gap of 2.3 eV.126 In the scheme of things, goethite, lepidocrocite, and δ-FeOOH host comparably modest electrical conductivities in contrast to hematite and maghemite. Goethite exhibits a band gap of 2.10 eV, lepidocrocite clocks in at 2.06 eV, and δ-FeOOH boasts a band gap of 1.94 eV.126 An intriguing case manifests in magnetite, renowned for its multifarious electrical properties, embodying both p-type and n-type semiconductors within its structural confines. It features a minute band gap measuring 0.1 eV.126

Intriguingly, magnetite’s electrical conductivity, boasting a Fe (III)/total Fe ratio of 2/3, stands several multiples higher compared to hematite, characterized by a Fe (III)/total Fe ratio of 1.207 Magnetite operates as an electron conductor bearing a distinct spin orientation, while donning the mantle of insulator or semiconductor for electrons aligned with the opposite spin orientation. This unique property categorizes magnetite as a half-metallic material.208 Rigorous experimental scrutiny and density-functional calculations validate the presence of spin-split band energies, substantiating its half-metallic nature.209,210,211,212

Furthermore, investigations into sphere-shaped magnetite nanoparticles unveil shifts in electron mobility aligned with temperature shifts across an expansive range (173–373 K).213 These nanoparticles' electrical conductivity intertwines with AC frequency and temperature interactions, as evidenced in conductivity measurements (Figure S11A). The dielectric properties and AC conductivity at 273 K provide insights into the behavior of grain boundaries and grain conductivity, revealing distinct characteristics related to long-range and short-range mobility.213 These findings highlight the coexistence of two conduction mechanisms—barrier hopping and non-overlapping small polaron tunneling—across a range of temperatures and frequencies213 (Figure S11B). At low temperatures, the dominant mechanism governing the conductivity in Fe3O4 nanoparticles is charge carrier hopping, where electrons move between Fe (III) and Fe (II) ions. This mechanism involves two distinct modes of hopping, resulting in a smooth transition in conductivity as the temperature increases. At high temperatures, the conductivity mechanism still involves electron hopping between Fe(III) and Fe(II) ions, but the specific mode of hopping may undergo changes. These transitions between different modes of conductivity reflect the complex relationship between temperature and the structural dynamics of the nanoparticles, affecting the mobility of electrons within the nanoparticles. In essence, the transport of electrons is governed by the tunneling of small polarons, electron hopping, and movement between Fe (III) and Fe (II) ions.213

Additionally, the electrical conductivity of magnetite evolves in response to pressure alterations. Empirical studies divulge that under pressure ranging from 0 to 20 GPa, the resistivity of Fe3O4 plummets by over an order of magnitude, reaching a nadir. However, beyond 20 GPa, the resistivity resurges, ultimately doubling by 48 GPa214 (Figure S11C). This propensity for resistivity change underscores the pronounced susceptibility of Fe3O4’s electrical conductivity to shifts in pressure. Notably, the electronic attributes of Fe2O3 and FeO also undergo transformation under pressure and structural modifications.215,216,217

Temperature and pressure stand out as pivotal factors sculpting iron architecture in nanomaterials, accentuating alterations in Fe-Fe distance. These shifts, in turn, reverberate through the atomic structure of iron oxide nanoparticles, impacting the CO phenomenon and ET processes. The minimal Fe-Fe distance within the crystalline structure looms large as a critical factor dictating the nature and temperature at which charge-ordering transpires218,219,220,221,222,223,224 (Figure S11D). The electrical conductivity properties of iron oxide nanoparticles are further molded by the transmission of electrons between Fe (III) and Fe (II) ions. Adjustments to the Fe-Fe distance can impede or facilitate electron migration between these ions, invariably precipitating changes in electrical conductivity. Appreciating the nexus between the Fe-Fe distance, CO, and ET in iron oxide nanoparticles assumes pivotal significance in unraveling their electrical conductivity. Notably, shifts in the Fe-Fe distance—stemming from temperature and pressure shifts—unearth invaluable insights into the mechanics of electron transport, such as electron hopping and tunneling of small polarons. These shifts in the Fe-Fe distance stand as an inherent aspect of the metals' atomic structure and electron configuration, engendering repercussions for charge ordering and ET.

Moving on to iron sulfides, greigite (Fe3S4) exhibits characteristics akin to the mixed-valence ferrimagnetism observed in Fe3O4225,226,227 (Figure S12A). Pyrite (FeS2) acts as a semiconductor with a discernible band gap of 0.90 eV, displaying both p-type and n-type conductivity228,229 (Figure S12B). The electronic properties of pyrite are further influenced by the presence of sulfur vacancies, imparting notable effects on its conductivity.230,231 Pertinently, electron hopping materializes between Fe (II) and Fe (III) on octahedral crystal surfaces and between conductive sulfur-deficient grain cores ensconced within nominally stoichiometric FeS2229,232,233,234 (Figure S12C).

Fe7S8, distinguished by its monoclinic structure, showcases heightened magnetic anisotropy energy compared to FeS, making it a subject of keen interest in magnetism studies.235 It is worth noting that some iron sulfides, such as troilite, have demonstrated superconductivity below 4.5 K236,237,238 (Figure S12D). Furthermore, iron sulfides, much like their oxide counterparts, exhibit dynamic responses to changes in pressure239 and temperature.240 The crystal structure of iron sulfide (FeS) adopts a tetragonal configuration, with iron atoms coordinated by four equidistant sulfur atoms241,242 (Figure S12E). On the other hand, pyrrhotite (Fe1-xS) takes on a hexagonal crystal structure reminiscent of the NiAs arrangement.235 Other variations of Fe-S minerals, not mentioned, also contribute to the rich landscape of electronic behavior in this class of compounds. Additionally, it is important to consider iron deficiency in some iron sulfides, which can occur when the mineral lacks a sufficient amount of iron atoms.235,239,240 This deficiency affects the electronic properties of the material and can lead to alterations in its conductivity and magnetic behavior. Recognizing the impact of iron deficiency on the properties of iron sulfides introduces an additional layer of complexity to their behavior and electronic characteristics.

In summary, exploring ET mechanisms in inorganic nanomaterials, including both iron oxides and sulfides, reveals a complex tapestry of properties. These properties encompass band gaps, conductivity, and structural dynamics, all intricately linked to the electron configuration and coordination with oxygen or sulfur, including the various defects and vacancies. Importantly, these nanoparticle structures introduce an additional layer of complexity to the already intricate interplay of factors, being dynamic and responsive to external influences.

Mechanisms of electron transfer in inorganic enzymes

In biological systems, ET is fundamental to reduction and oxidation reactions. Electrons play a central role, regardless of whether a biocatalyst or protein facilitates the process.243 Oxidoreductases, though not net electron producers themselves, excel at facilitating electron-transfer processes, making them vital in both biological and catalytic domains. Their essence lies in enabling electron flow.244,245

Oxidoreductases act as adept mediators, shuttling electrons either externally or within intricate intramolecular pathways—an event termed electron transport (ETp).246,247 This process primarily occurs through ionic channels within the biological milieu. The efficiency of electron transport within proteins correlates intimately with their electron-transfer proficiency and redox potential.244,245

To grasp the intricacies of the inorganic catalysis process, consider the essential role of electrons in biological processing. In many cases, the electron flux is low, and reaction rates are sluggish. In contrast to metal ions, which directly participate in these biochemical reactions, inorganic nanoparticles act as catalysts due to their unique structure and surface properties. Consider a scenario where a simple chemical equilibrium process involving electron transfer is established with a fixed equilibrium constant (Keq) for the biological reaction. While most of the electrons involved in the reaction are provided by the chemical constituents of the system, nanoparticles with electron hopping behavior can generate additional electrons, disrupting the chemical equilibrium governed by the fixed Keq. This disruption significantly enhances their role as catalysts, leading to increased electron flux and higher reaction rates. This phenomenon aligns with the second law of thermodynamics, which states that the disorder in a closed system always increases over time. For instance, Fe3O4 nanoparticles exhibit a distinctive POD activity by adsorbing H2O2 onto their surface, initiating redox reactions.49,58,169 This process differs from the conventional Fenton reaction, where the generation of hydroxyl radicals takes place in solution with metal ions.248,249 The pivotal step in hydroxyl radical generation with Fe3O4 nanoparticles involves redox reactions between H2O2 and the Fe (II) sites on the Fe3O4 surface.250,251 This orchestrated surge re-establishes equilibrium, injecting momentum into biological redox cascades, expediting metabolic processes. This symbolizes the essence of oxidoreductase functionality—a robust sustenance of ET flux, propelling the rhythmic cadence of reduction and oxidation. This interplay fosters a catalytic process fueled by the intrinsic electron hopping propensity of these nanoparticles distinct from conventional protein-based catalysis. as exemplified by HRP, wherein the genesis and retention of hydroxyl radicals transpire within the Fe-porphyrin ring throughout the ET sequence.252,253 The Fe (III) site nestled proximal to His170 and enveloped by a porphyrin architecture interfaces with Arg38 and His42, harmonizing to reorient H2O2.253,254

In addition to biological systems, a diverse array of inorganic nanomaterials, including transition metal oxides, sulfides, and selenides, exhibit remarkable electron transfer capabilities and hopping behavior. Some even demonstrate properties similar to those of superconductors, comparable to their inorganic counterparts such as iron oxides and sulfides, as we discussed earlier.

One fascinating category is that of inorganic enzymes, a special class of nanomaterials with protein-like catalytic activities.38 These biocatalysts, which may have predated all proteins or enzymes, perform functions akin to enzymes and display similar catalytic kinetics.38 For instance.

  • (1)

    Mixed-valence vanadium pentoxide V2O5 demonstrates semiconducting traits255 due to electron hopping dynamics within V4+ and V5+ ions,256 exhibiting intrinsic POD, OXD, CAT, glucose oxidase (GOX), and glutathione peroxidase (GPx) activity.162,257,258,259,260

  • (2)

    Mn3O4 nanoparticles demonstrate the polaron hopping of electron holes between Mn4+ and Mn3+ within octahedral sites,261 resulting in intrinsic CAT, GPx, and SOD activities.164,262

  • (3)

    MnO2 nanoparticles display direct electron hops within Mn - Mn chains,263,264 resulting in POD, CAT, OXD, and SOD activities.265,266

  • (4)

    Co3O4 NPs exhibit semiconducting attributes marked by Co3+-Co2+ hopping,204,267,268 concurrently demonstrating intrinsic POD and CAT activities.159,269,270,271

  • (5)

    Titanium dioxide (TiO2) nanoparticles, characterized by lower band gaps and electron hopping,272,273,274 also exhibit intrinsic POD activity,275 particularly with the photoelectron effect.276

  • (6)

    Nanocrystalline ceria (CeO2), with high electron conductivity and hopping attributes,277,278 can convert Ce4+ to Ce3+ due to oxygen vacancies,279 functioning as oxidoreductases (POD, CAT, OXD, SOD, nucleases, and photolyases),161,165,280,281,282 even phosphatase.281,283,284,285

  • (7)

    Cu2O/CuO nanoparticles, also known for their superconductivity,286,287 demonstrate intrinsic POD activity.163,288,289

In comparison to inorganic oxide nanoparticles, inorganic iron sulfide usually has a high electron transfer rate coupled with low resistivity, and even superconductivity as previously described,236,237,238 leading to improved corresponding enzyme activity.53,54,55,56,57 Additionally, molybdenum disulfide (MoS2) possesses both semiconductor properties290 and electron hopping behavior,291 allowing it to naturally act as POD, CAT, and SOD.292,293 Other nanoparticles such as α-FeSe, manganese selenide (MnSe), and molybdenum selenide (MoSe2) also demonstrate intrinsic POD activity55,294,295,296 due to their lower band gaps and electron hopping,297,298 including the superconductivity of α-FeSe.236,299,300

Moreover, certain inorganic nanoparticles, similar to proteinaceous enzyme, also respond to light, influencing their enzyme activity, including the light-sensitive α-Fe2O3 nanoparticles,74 TiO2 NP75,276 and various combinations thereof.80,301,302 The activity can also be influenced by nanoparticle modification, as seen with MoO3 nanoparticles. After modification, NH-MoO3-x nanobelts exhibit sensitivity to light-induced electron dynamics, leading to changes in their band gap77 (Figure S13). Comparatively, MoO3 nanobelts, with a specific band gap of 2.592 eV, provide a reference point.77 NH-MoO3-x nanobelts, with a reduced band gap of 1.793 eV, facilitate electron exchange through defect energy levels.77 This interaction not only improves electron conductivity but also triggers light-induced POD, OXD, and CAT activities, emphasizing the generation of ROS activity, creating a captivating interplay between temperature, energy bands, and catalytic responses. When these inorganic nanomaterials encounter light beyond their band gaps, an intriguing phenomenon unfolds. Electrons make a direct leap from valence to conduction levels, generating electron-hole pairs. Guided by photons and influenced by temperature, these pairs release surplus energy at the band’s edge, amplifying enzymatic efficiency. This orchestrated dance of electrons, orchestrated by both light and temperature, leads to an exhilarating enhancement of catalytic potency.

The introduction of Cyc-c electron mediators, akin to their protein counterparts, seamlessly woven into this intricate dance, stands as a testament to the multifaceted elegance inherent in these nanomaterials—an orchestration of catalytic prowess directed by the skillful hands of light, structure, and the enigmatic Cyc-c electron mediators. This remarkable contribution extends far beyond the domain of enzyme activity within synthesized nanoparticles. For instance, cadmium sulfide coupled with Cyc c serves as a bifunctional NADH oxidase and Cyc c reductase, showcasing the versatile roles that Cyc-c mediators play in orchestrating ET dynamics.93 Furthermore, CeVO4 nanozymes, mimicking the functions of the mitochondrial enzyme CcO, utilize Cyt c as a biological electron donor in the four-electron reduction of molecular oxygen, thus integrating SOD-like functions92 (Figure S14). These instances underscore the vast potential of Cyc-c mediators in influencing catalytic transformations.

This harmonious interplay extends its influence into uncharted horizons, particularly within the realm of the origin of life and its earliest evolution. In the opulent symphony of inorganic oxidoreductases, electron transfer modes come to the forefront, as illustrated in Figure 1. Here, electron hopping in the inorganic nanoparticles plays an essential role, with the interplay of light and Cyc-c offering a visual testament to the intricate pathways that pave the way for catalytic prowess. This visual representation vividly captures the essence of Cyc-c electron mediation, weaving threads of energy and transformation into the very fabric of these nanomaterials' reactivity.

Figure 1.

Figure 1

Electron Transfer Modes of the Inorganic Oxidoreductase

(A) Electron Hopping in Inorganic Nanomaterials (Adopted from electron hopping behavior from iron oxide nanoparticles, reprinting from Katz J E et al.,201 copyright@2012, The American Association for the Advancement of Science), Multiples reports support this observation (green rust,195 magnetite,196 iron oxyhydroxides,197,198,201 hematite,199,200,202,203 γ-Fe2O3; 204 FeS2,229,230,231,232,233,234 Iron polysulfides (e.g., Fe1-xS),235,239,240 V2O5,256 Mn3O4,261 MnO2,263,264 Co3O4,204,267,268 TiO2,272,273,274 CeO2,277,278 α-FeSe,291 MnSe,297 and MoSe2298).

(B) Photoelectron Effect and Photothermal Effect. Multiple reports support this observation.59,60,74,75,76,77,78,79,80,81,82,83,84,276,301,302

(C) Electron mediation by Cytochrome c (Adopted from Direct Electron Transfer of Enzymes Facilitated by Cytochromes and reprinting from Ma and Ludwing87) Multiple reports support this observation.90,91,92,93,96,99,100,101

Like the echoes of a symphony that linger long after the final note, the role of Cyc-c electron mediators reverberates beyond the laboratory, resonating through the corridors of scientific insight. It enriches our understanding of the delicate interplay between inorganic nanozymes and the grand theater of biochemical processes. In this grand symphony, electron hopping, coupled with the influence of light, finds its place as an essential and harmonious melody within the composition of nanoscale catalysis. Moreover, Cyc-c emerges as a bridge between the inorganic and organic worlds, particularly significant in the context of the origin of life, linking the elemental processes of early Earth to the intricate biochemical ballets of living organisms.

The influence of metal centers is deeply ingrained in biochemical transformations, dating back to the earliest stages of life, as exemplified by the compelling case of MTB.94,95 Remarkably, these orchestrations transcend the boundaries of inorganic nanoparticles, finding harmonious alignment with life’s ancient redox processes. This alignment is evident in the formation of biogenic iron oxide nanoparticles through biomineralization,94,95 referred to as “life fossil oxidoreductases,”38 where the POD activity59,60 and Cyc-c mediators play a pivotal role in modulating electron transfer pathways within the magnetite of MTB,95,96,99,100,101 as described in Section inorganic nanoparticles with intrinsic oxidoreductase activity.

All iron cores in ferritins can also be recognized as another example of “life fossil oxidoreductases."38 The observation of superoxide-diminishing activity in various biogenic iron cores of ferritin further emphasizes the significance of the metal structure in early life.71 After the removal of the protein, the iron core from prokaryotic ferritin, specifically both pfFn and pyFn from archaea P. furiosus or P. yayanosii, demonstrates higher SOD activity than that of eukaryotic ferritins, such as HFn and LFn from H. sapiens71 (Figure S5C). The iron/phosphate ratio in the iron core, a characteristic mainly determined by the structures of ferritins, also affects their SOD catalytic capability.71 The presence of phosphate in the iron core transforms it from a single crystalline structure to an amorphous iron phosphate-like structure, leading to a decreased affinity for the hydrogen proton of the ferrihydrite-like core. For example, ssDps, the DNA protection protein from the archaea Sulfolobus solfataricus, exhibits this behavior.71 Additionally, eukaryotes also have a higher phosphorus content due to evolutionary processes.71 These findings underscore the significance of the composition and evolution of ferritin structures in the context of early life and have potential implications for understanding the role of inorganic materials in biochemical processes.

The significance of these findings extends beyond synthetic experiments, reaching profound depths. In the synesthetic choreography of electron movement, these mediators establish a dynamic bridge connecting light, structure, and catalytic activity. This bridge not only unites nanomaterials and biochemical systems but also transcends temporal boundaries, linking the intricate dance of electrons in contemporary catalysis with the primordial rhythms of life’s redox processes.

The intricate dance of inorganic biocatalysts and their biological counterparts, proteinaceous oxidoreductases, forms a cornerstone of this narrative. These inorganic biocatalysts often exhibit striking parallels with their cognate oxidoreductases (proteinaceous enzymes)13,14,15,16,17,18,21 and universally possess DET capabilities.7 Even enzymes engaged in the hydrolysis of phosphate esters, known as phosphatases, share profound associations with protein-mediated ET processes.303,304,305 These phosphatases mirror the attributes of vanadium haloperoxidases, a subclass of oxidoreductases, embodying the essence of ET paradigms.306,307,308,309

The characteristics of biocatalysis (inorganic enzyme) are deeply rooted in the intrinsic metallic architecture properties of nanoparticles, including defects such as oxygen or metal vacancies,310,311,312,313,314,315,316 which are fundamental aspects of their nature. In particular, inorganic iron oxide and sulfur nanoparticles, with their complex metal architectures, exhibit unique electrical attributes. The electron configurations of Fe(III) and Fe(II) ions are distinct: Fe(III) is 1s2 2s2 2p⁶ 3s2 3p⁶ 3d⁵, and Fe(II) is 1s2 2s2 2p⁶ 3s2 3p⁶ 3d⁶. Within the atomic realm, the d orbitals of iron split into two sets with differing energy levels, a phenomenon known as d-orbital splitting, influenced by the crystal field effect from ligand interactions within coordination complexes. This energy difference governs the coloration and reactivity of iron compounds. The electron configuration of the oxide ion (O2⁻) is 1s2 2s2 2p⁶, with the additional two electrons in the 2p orbital. Similarly, the electron configuration of sulfide ions (S2⁻) is 1s2 2s2 2p⁶ 3s2 3p⁶, with the additional two electrons also in the 2p orbital. This intricate interplay of electron arrangement unfolds through the coordination of iron’s electron configuration with oxygen or sulfur moieties, accentuated by vacancies, lending complexity to the ensemble. This interplay results in semiconductivity, characterized by small band gaps, facilitating efficient electron translocation. Additionally, it extends into superconductivity, akin to pyrrhotite. This intrinsic proclivity for spontaneous electron hopping manifests as a dance across octahedral surface locales, choreographed by Fe 3d electrons in partnership with oxygen or sulfur partners. In essence, the symphony of electron configurations, orbital interactions, and atomic vacancies composes a harmonious score, producing the electrical attributes of inorganic iron oxide and sulfur nanoparticles. This resonant interplay showcases the elegance of atomic structures and their emergent properties, shaping nanoscale reactivity and catalysis.

It is noted that electron hopping, where electrons move between localized sites within a material, is fundamental and observed in both inorganic nanoparticles and biomacromolecules such as proteins and DNA.317 In inorganic nanoparticles, this process occurs within the crystal lattice, influenced by factors such as particle size, shape, and surface chemistry. For example, smaller nanoparticles may exhibit enhanced electron hopping due to pronounced quantum effects. In biomacromolecules, electron hopping is vital in various biological processes.247,318 Proteins, with their complex structures, facilitate electron transfer through specific amino acid residues or cofactors. Metal ions or organic cofactors within proteins can act as electron carriers, allowing electrons to move along redox-active sites. Similarly, DNA, despite its role in genetic information storage, exhibits electron hopping behavior under certain conditions, influencing processes such as DNA repair and oxidative damage.

The mechanisms underlying electron hopping in both inorganic nanoparticles and biomacromolecules are intricately linked to their structures and characteristics. In inorganic nanoparticles, the arrangement of atoms and the presence of defects or impurities affect electron hopping. In biomacromolecules, the specific arrangement of amino acids or nucleotides, as well as the presence of metal ions or organic cofactors, modulate this process. Understanding these mechanisms is essential for elucidating the role of ET in biological systems and for designing functional materials with tailored electronic properties. Moreover, comprehending the connection between the metal center structure of proteins and the metal structure in inorganic nanoparticles is crucial for understanding the evolution and emergence of life.39 This dynamic relationship has profound implications for biological functions and technological applications, potentially shaping the course of evolution on Earth.

Bridge of inorganic nanoparticles in biomolecular evolution

In our journey back to the conditions of early Earth, a spotlight is cast upon the establishment of fundamental biomolecules that have left indelible marks in the annals of our planet’s history. Among these molecules, primitive amino acids,319,320,321 adenine,321,322 nucleobases,323 nucleotides,324,325 sugars,326,327 lipids,328,329 and thiodepsipeptides330 stand as the cornerstones of this biochemical narrative. In this intricate interplay, amino acids emerge as protagonists, serving as the elemental building blocks of proteins and playing a pivotal role as conduits for ET processes, particularly within the realm of oxidoreductases. These amino acids boast diverse functional groups, ranging from amines to carboxylic acids, each eagerly participating in the choreography of ET reactions.

Amidst this intricate dance, the stage is set for the introduction of key electron relays—tryptophan and tyrosine residues. These entities come forth as maestros of electron flow, orchestrating efficient electron hopping and rapid electron tunneling across substantial distances within the intricate folds of metalloenzymes.331,332,333,334,335,336 As the dance continues, an additional layer of complexity is woven into the narrative—the interplay of light activation. This interplay introduces a symphonic interlude, where photons harmonize with matter, enriching the catalytic activity of enzymes and contributing to the captivating intricacies of biochemical transformations.

Within this ensemble, certain proteins stand out as virtuosos, housing heme groups with iron at their core. Consider POD,253,337,338 CAT,339,340 Fe-SOD,341 and cytochrome P450 peroxygenase,342,343 as prime examples. Together, they collaborate to construct the iconic Fe-oxo-Fe bridge, a structural marvel crucial to the catalytic procession of metal centers. This structure assumes a pivotal role in the delicate ballet of electron acceptance and donation during redox reactions.344,345,346 Through their intricate coordination, they imbue the dance of electrons with grace and purpose, ushering in moments of transformation that define the very essence of life’s processes.

The journey of functional groups extends beyond nature’s enzymatic boundaries, venturing into the realm of nanomaterials. While the selection of amino acids by humans for nanomaterial modification showcases deliberate innovation, the natural selection of amino acids within proteins through evolution is a result of intricate biological processes. This diverse array of amino acids and their arrangement, i.e., proteins, which has been shaped over millions of years, represents one of the nature’s greatest products—driving the complexity and functionality of living organisms. Precise nanomaterial modification with amino acids or other functional groups illustrates human ingenuity, yielding heightened ET rates and catalytic efficiency.347,348,349,350,351,352,353,354,355,356,357,358,359,360,361,362,363,364,365,366,367 These adept adjustments amplify surface interactions, culminating in more efficient ET processes. A notable example of this symphony is evident in histidine-modified Fe3O4 nanozymes, where a significant increase in apparent affinity Km for the substrate H2O2 leads to a remarkable 20-fold enhancement in catalytic efficiency kcat/Km.347 Furthermore, research involving irony polysulfide nanoparticles (Fe1−xS and Fe3S4) unveils POD and CAT activities.57 Even when molybdenum trioxide nanoparticles reveal limited inherent electron conductivity,368 the introduction of triphenylphosphonium ions emerges as a masterful orchestrator, directing the activities of sulfite-oxidizing enzymes with precision and finesse.369 Similarly, the strategic utilization of near-Infrared-II (NIR-II) light unveils its transformative power, deftly lowering the bandgaps of NH−MoO3−x nanoparticles. This nuanced reduction in bandgaps acts as a catalyst, triggering the liberation of electrons and catalyzing the remarkable functions of POD, CAT, and OXD.77

The formation and role of the metal centers stand as pivotal chapters intricately intertwined with the broader fabric of metalloenzymes structure and function. Nestled within the oxidoreductase protein, the metal center is not a passive participant but rather a testament to the symphonic interplay between its structural stabilization by amino acid sequences and its catalytic prowess within the Fe-oxo-Fe architecture.38,39 This catalytic potential, however, does not arise in isolation; it is accentuated by a harmonious ensemble of hydrogen bonds meticulously woven among the protein’s amino acids. This ensemble acts as a stabilizing force, amplifying the catalytic capabilities of the metal center to unprecedented heights. These insights resonate deeply, highlighting the crucial role of functional groups, particularly amino acids, in further fortifying the stability and selection of metal centers in metalloenzymes. Their impact reverberates across the secondary and tertiary levels of protein structure, where they engage in a seamless interplay with the evolving frameworks of metalloenzymes. Like skilled artisans, these functional groups assume the role of stabilizers—a concept elegantly championed by Huang.38,39 Yet, their contributions extend beyond mere stability; they lay the very foundation upon which the early pathways of life were etched, igniting the emergence of catalytic processes that have intricately woven the tapestry of life as we know it today.

The evolution of metalloenzymes has often been spotlighted through the lens of primary and secondary protein structures.334,335,336,370,371,372,373 This focus heavily relies upon gene codes or DNA, casting somewhat limited light on tertiary/quaternary structures and the integral metal centers. Within these tertiary structures, an enigma takes root—an enigma of metal center formation. While many studies have delved into the intricacies of folds, a significant emphasis has been placed on the secondary structure,374,375 often relegating metal centers and tertiary structures376,377,378 to the shadows. This fact suggests that the metal center could be considered a primary determinant in the early stages of protein folding, potentially preceding the involvement of amino acid residues in stabilizing the protein’s structure.38,39 This hypothesis aligns with the intricate relationship between inorganic nanoparticles, metal centers, and protein folding, highlighting the significance of metal ions or inorganic nanoparticles in shaping the functional properties and evolution of metalloproteins.

The story of ET and its evolution remains partially obscured, despite acknowledgments of the role played by iron-sulfur cluster proteins as electron carriers and transfer agents. Amidst these uncertainties, the pivotal question lingers—did metal centers precede primary and secondary protein structures, or did they arise in tandem or even later in the evolutionary journey? Does the formation of these metal centers intertwine with DNA/RNA synthesis, or does it emerge as a distinct process38? The selection of metals within metalloenzymes appears to be intricately woven with the fabric of geochemical processes, rather than solely being dictated by genetic materials such as DNA or RNA.249,379,380,381,382,383,384,385

A pivotal study conducted by Eck and Dayhoff in 1966 heralded a paradigm-shifting concept, unveiling the remarkable potential of ferredoxin to harness photon energy even in the absence of a fully established genetic code.386 This extraordinary attribute emanated from its inorganic active sites—clusters of [FeS]—and its elegantly simple, repetitive amino acid sequences. Subsequent experiments and computational models lent credence to this audacious hypothesis, suggesting the birth of a protoferredoxin, characterized by [2Fe-2S] and [4Fe-4S] motifs. These motifs emerged through the photooxidation of ferrous ions under UV light, facilitated by the duplication of an iron-sulfur tripeptide motif.387,388 The significance of ferredoxins spans the spectrum of life, from ancient anaerobic bacteria to sophisticated plants and animals, making them invaluable subjects for studying the evolution of metalloproteins.389,390,391

Weiss et al.'s pioneering research brilliantly illuminated the prevalence of FeS clusters and radical reaction mechanisms within the last universal common ancestor (LUCA), the progenitor of all cellular life forms.392 This revelation provided a tantalizing glimpse into the early utilization of FeS clusters as trailblazers that facilitated ET and orchestrated redox reactions within the nascent cells of the primordial world. Simultaneously, explorations into the enigmatic realm of hydrothermal vents, often considered the cradles of life’s origin, unveiled remarkably similar observations.393,394,395 Notably, the metal centers of Fe-S cluster proteins and diverse iron sulfides exhibited intriguing parallels, showcasing the elegant design of nature396,397,398 (Figure S15). For instance, the [2Fe 2S] motifs adorning eukaryotic ferredoxins and Rieske proteins displayed a symphony of coordination between paired iron atoms and two equidistant sulfur atoms.399,400 In contrast, the quartet of iron atoms, harmoniously coordinated with four equidistant sulfur atoms in high-potential iron-sulfur proteins and iron regulatory proteins (IRPs), composed the resonant tune of [4Fe 4S].401,402,403,404 Rubredoxin, with its graceful choreography, highlighted a single iron atom partnered with four equidistant sulfur atoms, a [1Fe 4S] motif of eloquent simplicity. Moreover, these intricate dancers on the molecular stage were accompanied by their corresponding protein ligands—the harmonious rhythms of [3Fe 3S] in bacterial ferredoxins,405 and the intricate melodies of [6Fe 6S] in Desulfovibrio Vulgaris (Hildenborough).406

The catalytic activity of cubane-type Fe4S4 clusters in metalloproteins, such as biotin synthase,407 aconitase,408 and (E)-4-hydroxy-3-methylbut-2-enyl pyrophosphate reductase (IspH),409 as well as in synthetic M4S4 clusters, highlights their role in the emergence of life and the formation of organic compounds from inorganic precursors.410 Recent studies have revealed that Fe–S clusters with low-valent Fe1+ centers can adopt various electronic configurations crucial for their catalytic activity. For instance, when CO binds to a synthetic [Fe4S4]0 cluster, it triggers the generation of Fe1+ centers through intracluster electron transfer, facilitating redox reactions. Similarly, CO binding to an [Fe4S4]+ cluster induces electron delocalization, enabling the activation of C–O bonds without highly negative redox states.411 Metalloproteins containing Fe4S4 clusters can catalyze the reduction of CO and CO2 to hydrocarbons,412,413,414 which is significant in the context of Earth’s early life.

Pyruvate is a central metabolite in Archaea, Bacteria, and Eukarya, where iron-sulfur enzymes play a crucial role in connecting pyruvate to carbon fixation pathways and thioester biochemistry.415,416 The FeS/S/FeS2 system can catalyze the interconversion of hydroxyl acids and keto acids.417 Recent studies have shown that natural iron sulfide pyrrhotite acts as an oxidoreductase catalyst in the conversion of pyruvic acid to lactic acid418 and in CO2 reduction.419 However, these studies have not demonstrated the inorganic oxidoreductase activity of these nanoparticles with enzyme kinetics, contrasting with our approach in this review, which aligns with the mineral surface approach focusing on the iron-sulfur world420 and its relevance to evolutionary biochemistry.421,422

Of exceptional note, the symphony of iron sulfides, featuring a solitary iron atom coordinated with four equidistant sulfur atoms, resonated with the enthralling duet of superconductivity and high oxidoreductase activity. This symphonic revelation hinted at their potential roles within the biochemical orchestration of LUCA’s primordial milieu and the profound odyssey of the genetic code’s evolution. As we peer into this majestic tapestry woven by the interplay of matter and energy, we catch glimpses of the cosmic dance that has shaped the narrative of life itself—an unfolding tale of harmonious interactions and exquisite choreography that spans the epochs of time.

In the midst of this breathtaking panorama, whether inorganic or biological, emerges a luminous narrative—an ode to the equilibrium of chemical processes. The irresistible allure of electrons creates an invitation that transcends boundaries, beckoning toward an indomitable symphony that harmoniously merges the worlds of inorganic matter and organic life. It is a symphony that orchestrates a harmonious marriage, guiding a dance of reactivity and transformation that resonates through the annals of both chemistry and biology. Just as musical notes blend and harmonize to compose a masterpiece, the dance of electrons, photons, and Cyc-c mediators creates a resonant composition—a masterpiece of catalytic brilliance that elegantly unfurls on the grand stage of nature’s theater.

Recent advancements in synthesizing various inorganic nanoparticles, attributed to their intrinsic enzyme-like activity, have found wide applications in biomedicine.25,26,27,50,51,423,424 Fe2O3 nanocubes exhibit enhanced POD activity under visible light, enabling a photo-assisted colorimetric method for detecting glucose at low concentrations (0.1 mM detection limit).74 Ferrihydrite NPs, possessing CAT activity, enhance the effectiveness of radiotherapy,49 while magnetoferritin NPs target and visualize tumor tissues due to the iron oxide core catalyzing the oxidation of peroxidase substrates in the presence of hydrogen peroxide, producing a color reaction used to visualize tumor tissues.425 Additionally, dietary iron oxide NPs with CAT activity mitigate neurodegeneration in a Drosophila-Alzheimer’s disease model.426 The high POD and CAT activities of inorganic iron polysulfide nanoparticles are responsible for effectively inhibiting Pseudomonas aeruginosa and Staphylococcus aureus, including drug-resistant strains.57 These activities enable them to disrupt pathogenic biofilms, making them valuable for disinfecting implant devices such as ventilators and blood catheters. Additionally, their enzyme activity accelerates the healing of infected wounds, further extending their potential for preventing or treating biofilm-related infections.57 Conversely, remarkably active CeVO4, acting as CcO, catalyzes the four-electron reduction of dioxygen to water in the respiratory electron transport chain without releasing any partially reduced oxygen species (PROS) such as superoxide, peroxide, and hydroxyl radicals.92 These findings highlight the potential of iron oxide and other nanoparticles as inorganic enzymes for therapeutic and diagnostic purposes.28,51,427

The synthesis of inorganic iron oxide nanoparticles is an important technique in agriculture for regulating plant ROS levels in various environments.428,429,430 Unlike in biomedicine, where nanoparticles are typically injected into the body, plants can uptake these nanoparticles themselves.431 For example, inorganic iron oxide nanoparticles include those naturally occurring in soils432,433,434 and synthesized ones,435,436,437,438 which can be absorbed by roots,435,436,437,438,439,440,441,442 leaves,443 or even the surface of seeds.444 Once inside the plant, these nanoparticles can move to different parts of the plant, affecting various physiological processes.445 One notable effect of these nanoparticles is their ability to regulate ROS levels in plants, effectively acting as inorganic enzymes to manage oxidative stress,446,447,448 including various abiotic stresses such as drought446 and salt stress,449 as well as combating virus443 and fungi diseases,450 and reducing heavy metal contamination.451,452,453,454 Additionally, inorganic iron oxide nanoparticles have been found to increase the production of photosynthetic pigments in plants and electron transfer, enhancing their photosynthesis and leading to improved growth and development,436,439,441,442,447,451 and can even improve nitrogen fixation in soybeans.455 These nanoparticles can also influence plants at the molecular level, impacting gene expression and metabolic pathways,430,431,436,453,456 which can further affect their growth patterns. These applications demonstrate the diverse potential of inorganic nanoparticles as inorganic enzymes in addressing environmental challenges and contributing to sustainable agricultural practices.

It is worth noting that inorganic oxidoreductase is still functional in our current Earth environment. Recent results indicate that Fe2O3 nanoparticles obtained from PVC dichlorination residues and iron chips treated with subcritical water exhibit inherent peroxidase-like properties.172 Biogenic iron oxide nanoparticles, derived from the interaction of fungi with iron oxide nanoparticles,170,171 as well as magnetite nanoparticles enclosed within magnetosomes in magnetotactic bacteria,59,60 also demonstrate oxidoreductase activity. Furthermore, biogenic iron oxide nanoparticles from bacteria contribute to this narrative.61,62,63,64,71 Pyrrhotite, a natural iron sulfide mineral, has also been observed to act as an oxidoreductase in the conversion of pyruvic acid to lactic acid.418 It is anticipated that all iron oxide nanoparticles with the same metal structure on the current Earth will continue to function as biocatalysts, a realization yet to be fully acknowledged.

In comparison to naturally occurring iron oxide nanoparticles, the number of synthesized nanoparticles remains limited.457 Natural iron oxide nanoparticles are widely distributed in diverse environments, including soils, water, rocks, and living organisms458 (Figure S16). Extensive research and documentation on these nanoparticles have been conducted. Examples of such environments include high pH hydrothermal vents, ice sheets, fly ash, and street dust, and magnetosomes from MTB, as well as other biogenic iron minerals. These nanoparticles form through various mechanisms, resulting in different sizes, shapes, and structures. Similarly, iron sulfide nanoparticles can be found in hydrothermal vent plumes459,460 and many marine sediments.461,462,463 They have also been identified in the early Earth, predating the existence of life itself.98,142,464,465,466,467 Some of these nanoparticles are believed to be directly associated with the origins of life and the evolution of living organisms.394,395,420,468,469,470,471,472,473,474

It is important to note that not all nanoparticles exhibit biocatalytic properties. The activity of inorganic enzymes is directly linked to their metal structure, which can be influenced by both biological and environmental conditions. Therefore, further research is needed to elucidate the nuanced relationship between the structure of inorganic enzymes and their catalytic activity in different environmental contexts. Additionally, the role of inorganic enzymes in ecosystems, particularly in the pre-protein world, remains an understudied area that holds great potential for uncovering fundamental aspects of biochemical evolution and the origin of life. Future studies focusing on the function of natural inorganic nanoparticles or inorganic nanocolloids are necessary to deepen our understanding of their significance in biological and environmental systems, including addressing the challenge of ROS in water eutrophication and toxicity, air pollution, and global climate change. This underscores the importance of studying natural nanoparticles and their functions, which are currently underexplored.

Furthermore, the applications of inorganic nanoparticles, especially those synthesized, are not limited to current biomedicine, agriculture, and environmental assessment. They can also be further extended to address various challenges in our lives related to global climate change, water quality, and sustainable development. Research in these areas could lead to innovative solutions that leverage the unique properties of inorganic nanoparticles to tackle pressing global issues.

Conclusion

In summary, the research illuminates the extraordinary oxidoreductase activity displayed by inorganic iron oxide and sulfide nanoparticles. This activity is driven by their distinctive metal architecture and electron conductivity. These characteristics set the stage for electron hopping, a phenomenon vital for disrupting chemical equilibrium and initiating reactions. Additionally, the involvement of light activation and Cyc c electron mediators further amplifies their catalytic potential.

These inorganic nanoparticles, akin to fascinating “life fossil oxidoreductases," provide a unique window into the origins and development of life on Earth. They represent one of the earliest classes of oxidoreductases known as inorganic enzymes, predating the emergence of more complex biological molecules. They represent one of the earliest-known classes of oxidoreductases, predating the emergence of more complex biological molecules. In ancient environments characterized by hydrogen peroxide and free radicals, these nanoparticles played a pivotal role in fundamental biochemical processes, including redox reactions, detoxification, and nutrient cycling.

Today, inorganic iron oxide and sulfide nanoparticles continue to serve as essential biocatalysts in natural ecosystems. Found in soils, sediments, and aquatic environments, they mediate crucial reactions that contribute to the overall health and functioning of ecosystems. Their enduring impact on Earth’s history underscores their significance, shaping the delicate symphony of our planet’s ecosystems.

The electron transfer property stands as a vital bridge between inorganic nanoparticles and proteins, particularly in the context of metal centers and their evolution. This property is a fundamental aspect that unites their catalytic capabilities. The coordination of metal centers within proteins is resonant with the metal architecture of inorganic nanoparticles, and both contribute significantly to the electron transfer dynamics essential for various biochemical processes.

As we delve deeper into the world of inorganic enzymes, exploring their electron transfer capabilities, electron hopping phenomena, and their responsiveness to light and Cyc c mediation, we gain invaluable insights into the fundamental principles that governed early life processes and adaptations. These nanoparticles continue to be pivotal players in the ongoing dance of life on our planet, leaving an indelible mark on Earth’s intricate web of existence.

Moving forward, while current studies are more focused on the biomedical applications of synthesized nanoparticles, it is important to broaden the scope to include the exploration of natural nanoparticles and their role in ecosystem function. Understanding the characteristics of enzyme activity exhibited by both natural and synthesized nanoparticles is crucial for comprehensively assessing their potential impact. It is important to emphasize that the activity of nanoparticles is inherent to their nanoparticle physics and is not a product of deliberate design by scientists. This activity has existed even before the protein world and should be researched and applied in a broader context, encompassing environmental challenges and sustainable development.

The scope of this field should not be limited to synthesized nanoparticles with biomedical applications. Understanding the role of natural inorganic nanoparticles as inorganic enzymes in our current ecosystem is crucial for addressing environmental challenges, including their impact on plant growth and water purification mechanisms (e.g., eutrophication and chemical detoxification), as well as their interactions with ROS in the atmosphere and air pollution contributing to global climate change. Such studies can broaden our understanding of the significance of these nanoparticles in biogeochemistry and their potential applications in diverse fields, including climate science, atmospheric chemistry, soil science, agriculture, and environmental management.

Furthermore, studying the characteristics and activities of nanoparticles provides an opportunity to gain insights into the origin of life and evolution. By understanding how nanoparticles interacted with early Earth environments and potentially played a role in the emergence of life, we can gain a deeper understanding of the processes that led to the development of life on our planet. This broader perspective can provide valuable insights into the fundamental principles of biology and evolution.

This study highlights the pivotal role of inorganic enzymes in bridging ancient and modern biology, illustrating their significance from ecosystem to evolution in both chemistry and biology. The hypothesis on inorganic enzymes stemmed from the author’s observation of phosphate ester behavior in artificial seawater in 2006,30 triggered by Graham Cairns-Smith’s book “Seven Clues to the Origin of Life.”475 I dedicate this work to my wife, Wei Sun, and my son, Jack Jixiang Huang, for their unwavering support and sacrifices throughout my academic journey, especially during the COVID-19 pandemic when I worked independently from home. My inspiration was ignited by J.D. Bernal’s seminal work “The Physical Basis of Life”476 and by the persistent questioning of Dr. Michael J. Russell regarding the energy for these intricate systems. I also acknowledge the guidance and encouragement of Drs. Gerhard Schenk and the late R.J.P. Williams, as well as the support of Drs. Robert Atlas, Jia-Zhong Zhang, and Peter B. Ortneras, for their instrumental contributions to the development of my research.

Acknowledgments

X-L.H. is an independent researcher.

The groundwork for this project began in June 2022, following my exploration into 'Inorganic Nanozymes: Artificial or Inorganic Enzymes.' By December 2022, I had finalized the initial draft independently. Starting in October 2022, I consulted closely with scientists including Dr. Michael J. Russell (Jet Propulsion Laboratory), Dr. Harry B. Gray (California Institute of Technology), and Dr. Gerhard Schenk and colleagues: Dr. Marcelo Monteiro Pedroso, Dr. Jeffrey Harmer, Dr. Lawrence R. Gahan, and Dr. Gordon Southam (The University of Queensland). Despite this collaboration, I independently completed the final version at Stony Brook University. It's important to note that all statements, findings, conclusions, and recommendations in this work are solely attributed to me and may not align with the perspectives of Stony Brook University or any affiliated agency.

In this review, I emphasize a central idea: that the core essence of life is rooted in the principles of physics intricately woven with chemical processes. This revelation dawned on me during a poignant journey to China, where I paid tribute to my mother. She was not only a retired professor but also a resilient soul who, like millions of others, suffered due to the Covid pandemic and eventually succumbed to it. My parents, the late Drs. Shi-Xiong Huang (1927-2018) and Jing-Xiong Ji (1930-2023), were pioneers in the field of public health in Anhui and China. They stood as unwavering pillars of support throughout my academic journey, providing invaluable encouragement and guidance over the years. Their steadfast belief in my endeavors has been an unending source of strength.

Author contributions

Conceptualization, X-L H.; methodology, X-L H.; investigation, X-L H.; visualization, X-L H.; writing – original draft, X-L H.; writing – review and editing, X-L H.

Declaration of interests

There are no conflicts to declare.

Footnotes

Supplemental information can be found online at https://doi.org/10.1016/j.isci.2024.109555.

Supplemental information

Document S1. Figure S1–S16
mmc1.pdf (3.3MB, pdf)

References

  • 1.Williams R.J. The natural selection of the chemical elements. Cell. Mol. Life Sci. 1997;53:816–829. doi: 10.1007/s000180050102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Falkowski P.G., Fenchel T., Delong E.F. The microbial engines that drive earth's biogeochemical cycles. Science. 2008;320:1034–1039. doi: 10.1126/science.1153213. [DOI] [PubMed] [Google Scholar]
  • 3.Kim J.D., Senn S., Harel A., Jelen B.I., Falkowski P.G. Discovering the electronic circuit diagram of life: Structural relationships among transition metal binding sites in oxidoreductases. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2013;368 doi: 10.1098/rstb.2012.0257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kracke F., Vassilev I., Krömer J.O. Microbial electron transport and energy conservation - The foundation for optimizing bioelectrochemical systems. Front. Microbiol. 2015;6:575. doi: 10.3389/fmicb.2015.00575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Milton R.D., Minteer S.D. Direct enzymatic bioelectrocatalysis: differentiating between myth and reality. J. R. Soc. Interface. 2017;14 doi: 10.1098/rsif.2017.0253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Chen H., Simoska O., Lim K., Grattieri M., Yuan M., Dong F., Lee Y.S., Beaver K., Weliwatte S., Gaffney E.M., Minteer S.D. Fundamentals, Applications, and Future Directions of Bioelectrocatalysis. Chem. Rev. 2020;120:12903–12993. doi: 10.1021/acs.chemrev.0c00472. [DOI] [PubMed] [Google Scholar]
  • 7.Ratautas D., Dagys M. Nanocatalysts Containing Direct Electron Transfer-Capable Oxidoreductases: Recent Advances and Applications. Catalysts. 2019;10:9. doi: 10.3390/catal10010009. [DOI] [Google Scholar]
  • 8.Ferapontova E.E., Shleev S., Ruzgas T., Stoica L., Christenson A., Tkac J., Yaropolov A.I., Gorton L. In: Perspectives in Bioanalysis. Paleček E., Scheller F., Wang J., editors. Elsevier; 2005. Direct Electrochemistry of Proteins and Enzymes; pp. 517–598. [DOI] [Google Scholar]
  • 9.Dominguez-Benetton X., Srikanth I., Satyawali Y., Vanbroekhoven K., Pant D. Enzymatic Electrosynthesis: An Overview on the Progress in Enzyme- Electrodes for the Production of Electricity, Fuels and Chemicals. J. Microb. Biochem. Technol. 2013;2013:1–20. doi: 10.4172/1948-5948.S6-007. [DOI] [Google Scholar]
  • 10.Kano K., Shirai O., Kitazumi Y., Sakai K., Xia H.-Q. In: Enzymatic Bioelectrocatalysis. Kano K., Shirai O., Kitazumi Y., Sakai K., Xia H.-Q., editors. Springer Singapore; 2021. Fundamentals of DET-Type Bioelectrocatalysis; pp. 57–78. [DOI] [Google Scholar]
  • 11.Eddowes M.J., Hill H.A.O. Novel method for the investigation of the electrochemistry of metalloproteins: cytochrome c. J. Chem. Soc., Chem. Commun. 1977:771b–772b. doi: 10.1039/C3977000771B. [DOI] [Google Scholar]
  • 12.Peter Y., Theodore K. Reversible ecectrode reaction of cytochrome c. Chem. Lett. 1977;6:1145–1148. doi: 10.1246/cl.1977.1145. [DOI] [Google Scholar]
  • 13.Yaropolov A.I., Malovik V., Varfolomeyev S.D., Berezin I.V. Electroreduction of hydrogen peroxide on an electrode with immobilized peroxidase. Dokl. Akad. Nauk SSSR. 1979;249:1399. [Google Scholar]
  • 14.Tarasevich M.R., Yaropolov A.I., Bogdanovskaya V.A., Varfolomeev S.D. Electrocatalysis of a cathodic oxygen reductionby laccase. J. Electroanal. Chem. 1979;104:393–403. doi: 10.1016/S0022-0728(79)81047-5. [DOI] [Google Scholar]
  • 15.Gorton L., Lindgren A., Larsson T., Munteanu F.D., Ruzgas T., Gazaryan I. Direct electron transfer between heme-containing enzymes and electrodes as basis for third generation biosensors. Anal. Chim. Acta X. 1999;400:91–108. doi: 10.1016/S0003-2670(99)00610-8. [DOI] [Google Scholar]
  • 16.Ferapontova E.E., Ruzgas T., Gorton L. Direct Electron Transfer of Heme- and Molybdopterin Cofactor-Containing Chicken Liver Sulfite Oxidase on Alkanethiol-Modified Gold Electrodes. Anal. Chem. 2003;75:4841–4850. doi: 10.1021/ac0341923. [DOI] [PubMed] [Google Scholar]
  • 17.Shleev S., Tkac J., Christenson A., Ruzgas T., Yaropolov A.I., Whittaker J.W., Gorton L. Direct electron transfer between copper-containing proteins and electrodes. Biosens. Bioelectron. 2005;20:2517–2554. doi: 10.1016/j.bios.2004.10.003. [DOI] [PubMed] [Google Scholar]
  • 18.Liu Y., Yuan R., Chai Y., Tang D., Dai J., Zhong X. Direct electrochemistry of horseradish peroxidase immobilized on gold colloid/cysteine/nafion-modified platinum disk electrode. Sens. Actuators B Chem. 2006;115:109–115. doi: 10.1016/j.snb.2005.08.048. [DOI] [Google Scholar]
  • 19.Léger C., Bertrand P. Direct Electrochemistry of Redox Enzymes as a Tool for Mechanistic Studies. Chem. Rev. 2008;108:2379–2438. doi: 10.1021/cr0680742. [DOI] [PubMed] [Google Scholar]
  • 20.Liu J., Chakraborty S., Hosseinzadeh P., Yu Y., Tian S., Petrik I., Bhagi A., Lu Y. Metalloproteins Containing Cytochrome, Iron–Sulfur, or Copper Redox Centers. Chem. Rev. 2014;114:4366–4469. doi: 10.1021/cr400479b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bollella P., Gorton L., Antiochia R. Direct Electron Transfer of Dehydrogenases for Development of 3rd Generation Biosensors and Enzymatic Fuel Cells. Sensors. 2018;18:1319. doi: 10.3390/s18051319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Gao L., Zhuang J., Nie L., Zhang J., Zhang Y., Gu N., Wang T., Feng J., Yang D., Perrett S., Yan X. Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007;2:577–583. doi: 10.1038/nnano.2007.260. [DOI] [PubMed] [Google Scholar]
  • 23.Huang Y., Ren J., Qu X. Nanozymes: Classification, Catalytic Mechanisms, Activity Regulation, and Applications. Chem. Rev. 2019;119:4357–4412. doi: 10.1021/acs.chemrev.8b00672. [DOI] [PubMed] [Google Scholar]
  • 24.Liang M., Yan X. Nanozymes: From New Concepts, Mechanisms, and Standards to Applications. Acc. Chem. Res. 2019;52:2190–2200. doi: 10.1021/acs.accounts.9b00140. [DOI] [PubMed] [Google Scholar]
  • 25.Singh S. Nanomaterials Exhibiting Enzyme-Like Properties (Nanozymes): Current Advances and Future Perspectives. Front. Chem. 2019;7:46. doi: 10.3389/fchem.2019.00046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Yang W., Yang X., Zhu L., Chu H., Li X., Xu W. Nanozymes: Activity origin, catalytic mechanism, and biological application. Coord. Chem. Rev. 2021;448 doi: 10.1016/j.ccr.2021.214170. [DOI] [Google Scholar]
  • 27.Hong C., Meng X., He J., Fan K., Yan X. Nanozyme: A promising tool from clinical diagnosis and environmental monitoring to wastewater treatment. Particuology. 2022;71:90–107. doi: 10.1016/j.partic.2022.02.001. [DOI] [Google Scholar]
  • 28.Wei H., Li G., Li J. 1st Edition. Springer; 2023. Biomedical Nanozymes, from Diagnostics to Therapeutics. [DOI] [Google Scholar]
  • 29.Yang L., Dong S., Gai S., Yang D., Ding H., Feng L., Yang G., Rehman Z., Yang P. Deep Insight of Design, Mechanism, and Cancer Theranostic Strategy of Nanozymes. Nano-Micro Lett. 2023;16:28. doi: 10.1007/s40820-023-01224-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Huang X.L., Zhang J.Z. University of Miami; 2007. Sediment-water Exchange of Dissolved Organic Phosphorus in Florida Bay.https://cimas.rsmas.miami.edu/_assets/pdf/annual-reports/2008_annual_report.pdf [Google Scholar]
  • 31.Huang X.L., Zhang J.Z. Hydrolysis of glucose-6-phosphate in aged, acid-forced hydrolysed nanomolar inorganic iron solutions - An inorganic biocatalyst? RSC Adv. 2012;2:199–208. doi: 10.1039/c1ra00353d. [DOI] [Google Scholar]
  • 32.Huang X.L. Hydrolysis of Phosphate Esters Catalyzed by Inorganic Iron Oxide Nanoparticles Acting as Biocatalysts. Astrobiology. 2018;18:294–310. doi: 10.1089/ast.2016.1628. [DOI] [PubMed] [Google Scholar]
  • 33.Huang X.L. In: Nanocatalysts. Sinha I., Shukla M., editors. Interchopen; 2019. Iron Oxide Nanoparticles: An Inorganic Phosphatase; pp. 97–124. [DOI] [Google Scholar]
  • 34.Scott S., Zhao H., Dey A., Gunnoe T.B. Nano-Apples and Orange-Zymes. ACS Catal. 2020;10:14315–14317. doi: 10.1021/acscatal.0c05047. [DOI] [Google Scholar]
  • 35.Wei H., Gao L., Fan K., Liu J., He J., Qu X., Dong S., Wang E., Yan X. Nanozymes: A clear definition with fuzzy edges. Nano Today. 2021;40 doi: 10.1016/j.nantod.2021.101269. [DOI] [Google Scholar]
  • 36.Zandieh M., Liu J. Nanozyme Catalytic Turnover and Self-Limited Reactions. ACS Nano. 2021;15:15645–15655. doi: 10.1021/acsnano.1c07520. [DOI] [PubMed] [Google Scholar]
  • 37.Robert A., Meunier B. How to Define a Nanozyme. ACS Nano. 2022;16:6956–6959. doi: 10.1021/acsnano.2c02966. [DOI] [PubMed] [Google Scholar]
  • 38.Huang X.-L. What are inorganic nanozymes? Artificial or inorganic enzymes. New J. Chem. 2022;46:15273–15291. doi: 10.1039/D2NJ02088B. [DOI] [Google Scholar]
  • 39.Huang X.-L., Harmer J.R., Schenk G., Southam G. Inorganic Fe-O and Fe-S Oxidoreductases: Paradigms for Prebiotic Chemistry and the Evolution of Enzymatic Activity in Biology. Front. Chem. 2024;12 doi: 10.3389/fchem.2024.1349020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Zhang X.Q., Gong S.W., Zhang Y., Yang T., Wang C.Y., Gu N. Prussian blue modified iron oxide magnetic nanoparticles and their high peroxidase-like activity. J. Mater. Chem. 2010;20:5110–5116. doi: 10.1039/c0jm00174k. [DOI] [Google Scholar]
  • 41.Peng C., Jiang B., Liu Q., Guo Z., Xu Z., Huang Q., Xu H., Tai R., Fan C. Graphene-templated formation of two-dimensional lepidocrocite nanostructures for high-efficiency catalytic degradation of phenols. Energy Environ. Sci. 2011;4:2035–2040. doi: 10.1039/c0ee00495b. [DOI] [Google Scholar]
  • 42.Chaudhari K.N., Chaudhari N.K., Yu J.S. Peroxidase mimic activity of hematite iron oxides (α-Fe2O3) with different nanostructures. Catal. Sci. Technol. 2012;2:119–124. doi: 10.1039/c1cy00124h. [DOI] [Google Scholar]
  • 43.Chen Z., Yin J.J., Zhou Y.T., Zhang Y., Song L., Song M., Hu S., Gu N. Dual enzyme-like activities of iron oxide nanoparticles and their implication for diminishing cytotoxicity. ACS Nano. 2012;6:4001–4012. doi: 10.1021/nn300291r. [DOI] [PubMed] [Google Scholar]
  • 44.Liu Q., Zhang L., Li H., Jia Q., Jiang Y., Yang Y., Zhu R. One-pot synthesis of porphyrin functionalized γ-Fe2O3 nanocomposites as peroxidase mimics for H2O2 and glucose detection. Mater. Sci. Eng. C. 2015;55:193–200. doi: 10.1016/j.msec.2015.05.028. [DOI] [PubMed] [Google Scholar]
  • 45.Gao L., Fan K., Yan X. Iron Oxide Nanozyme: A Multifunctional Enzyme Mimetic for Biomedical Applications. Theranostics. 2017;7:3207–3227. doi: 10.7150/thno.19738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Qin T., Ma R., Yin Y., Miao X., Chen S., Fan K., Xi J., Liu Q., Gu Y., Yin Y., et al. Catalytic inactivation of influenza virus by iron oxide nanozyme. Theranostics. 2019;9:6920–6935. doi: 10.7150/thno.35826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Gu Y., Huang Y., Qiu Z., Xu Z., Li D., Chen L., Jiang J., Gao L. Vitamin B(2) functionalized iron oxide nanozymes for mouth ulcer healing. Sci. China Life Sci. 2020;63:68–79. doi: 10.1007/s11427-019-9590-6. [DOI] [PubMed] [Google Scholar]
  • 48.Guo S., Guo L. Unraveling the Multi-Enzyme-Like Activities of Iron Oxide Nanozyme via a First-Principles Microkinetic Study. J. Phys. Chem. C. 2019;123:30318–30334. doi: 10.1021/acs.jpcc.9b07802. [DOI] [Google Scholar]
  • 49.Zhang R., Chen L., Liang Q., Xi J., Zhao H., Jin Y., Gao X., Yan X., Gao L., Fan K. Unveiling the active sites on ferrihydrite with apparent catalase-like activity for potentiating radiotherapy. Nano Today. 2021;41 doi: 10.1016/j.nantod.2021.101317. [DOI] [Google Scholar]
  • 50.Gao L. In: Iron Oxide Nanoparticles. Huang X.-L., editor. 2022. Enzyme-Like Property (Nanozyme) of Iron Oxide Nanoparticles.https://www.intechopen.com [DOI] [Google Scholar]
  • 51.Gao L., Yan X. Nanozymes: Biomedical Applications of Enzymatic Fe(3)O(4) Nanoparticles from In Vitro to In Vivo. Adv. Exp. Med. Biol. 2019;1174:291–312. doi: 10.1007/978-981-13-9791-2_9. [DOI] [PubMed] [Google Scholar]
  • 52.Zu Y., Wang Z., Yao H., Yan L. Oxygen-generating biocatalytic nanomaterials for tumor hypoxia relief in cancer radiotherapy. J. Mater. Chem. B. 2023;11:3071–3088. doi: 10.1039/d2tb02751h. [DOI] [PubMed] [Google Scholar]
  • 53.Yao W.T., Zhu H.Z., Li W.G., Yao H.B., Wu Y.C., Yu S.H. Intrinsic peroxidase catalytic activity of Fe7S8 nanowires templated from [Fe16S20]/diethylenetriamine hybrid nanowires. ChemPlusChem. 2013;78:723–727. doi: 10.1002/cplu.201300075. [DOI] [PubMed] [Google Scholar]
  • 54.Dai Z., Liu S., Bao J., Ju H. Nanostructured FeS as a mimic peroxidase for biocatalysis and biosensing. Chemistry. 2009;15:4321–4326. doi: 10.1002/chem.200802158. [DOI] [PubMed] [Google Scholar]
  • 55.Dutta A.K., Maji S.K., Srivastava D.N., Mondal A., Biswas P., Paul P., Adhikary B. Synthesis of FeS and FeSe nanoparticles from a single source precursor: a study of their photocatalytic activity, peroxidase-like behavior, and electrochemical sensing of H2O2. ACS Appl. Mater. Interfaces. 2012;4:1919–1927. doi: 10.1021/am300408r. [DOI] [PubMed] [Google Scholar]
  • 56.Ding C., Yan Y., Xiang D., Zhang C., Xian Y. Magnetic Fe3S4 nanoparticles with peroxidase-like activity, and their use in a photometric enzymatic glucose assay. Microchim. Acta. 2016;183:625–631. doi: 10.1007/s00604-015-1690-6. [DOI] [Google Scholar]
  • 57.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. doi: 10.1038/s41467-018-06164-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Shen X., Wang Z., Gao X., Zhao Y. Density Functional Theory-Based Method to Predict the Activities of Nanomaterials as Peroxidase Mimics. ACS Catal. 2020;10:12657–12665. doi: 10.1021/acscatal.0c03426. [DOI] [Google Scholar]
  • 59.Hu L., Song T., Ma Q., Chen C., Pan W., Xie C., Nie L., Yang W., Hafeli U., Schutt W., Zborowski M. Bacterial magnetic nanoparticles as peroxidase mimetics and application in immunoassay. AIP Conf. Proc. 2010;1311:369–374. doi: 10.1063/1.3530039. [DOI] [Google Scholar]
  • 60.Li K., Chen C., Chen C., Wang Y., Wei Z., Pan W., Song T. Magnetosomes extracted from Magnetospirillum magneticum strain AMB-1 showed enhanced peroxidase-like activity under visible-light irradiation. Enzyme Microb. Technol. 2015;72:72–78. doi: 10.1016/j.enzmictec.2015.02.009. [DOI] [PubMed] [Google Scholar]
  • 61.Pan Y., Li N., Mu J., Zhou R., Xu Y., Cui D., Wang Y., Zhao M. Biogenic magnetic nanoparticles from Burkholderia sp. YN01 exhibiting intrinsic peroxidase-like activity and their applications. Appl. Microbiol. Biotechnol. 2015;99:703–715. doi: 10.1007/s00253-014-5938-6. [DOI] [PubMed] [Google Scholar]
  • 62.Pan Y., Wang Y., Fan X., Wang W., Yang X., Cui D., Zhao M. Bacterial intracellular nanoparticles exhibiting antioxidant properties and the significance of their formation in ROS detoxification. Environ. Microbiol. Rep. 2019;11:140–146. doi: 10.1111/1758-2229.12733. [DOI] [PubMed] [Google Scholar]
  • 63.Ahmed A., Abagana A., Cui D., Zhao M. De novo iron oxide hydroxide, ferrihydrite produced by comamonas testosteroni exhibiting intrinsic peroxidase-like activity and their analytical applications. BioMed Res. Int. 2019;2019 doi: 10.1155/2019/7127869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Abagana A.Y., Zhao M., Alshahrani M.Y., Rehman K.U., Andleeb S., Wang J., Bukhari S.M. Hydrogen iron oxide from an Acinetobacter strain exhibiting intrinsic peroxidase-like activity and its catalytic mechanism and applications. Biomass Convers. Biorefin. 2022;14:3453–3462. doi: 10.1007/s13399-022-02370-y. [DOI] [Google Scholar]
  • 65.Andrews S.C. The Ferritin-like superfamily: Evolution of the biological iron storeman from a rubrerythrin-like ancestor. Biochim. Biophys. Acta. 2010;1800:691–705. doi: 10.1016/j.bbagen.2010.05.010. [DOI] [PubMed] [Google Scholar]
  • 66.Le Brun N.E., Crow A., Murphy M.E.P., Mauk A.G., Moore G.R. Iron core mineralisation in prokaryotic ferritins. Biochim. Biophys. Acta. 2010;1800:732–744. doi: 10.1016/j.bbagen.2010.04.002. [DOI] [PubMed] [Google Scholar]
  • 67.Narayanan S., Firlar E., Rasul M.G., Foroozan T., Farajpour N., Covnot L., Shahbazian-Yassar R., Shokuhfar T. On the structure and chemistry of iron oxide cores in human heart and human spleen ferritins using graphene liquid cell electron microscopy. Nanoscale. 2019;11:16868–16878. doi: 10.1039/c9nr01541h. [DOI] [PubMed] [Google Scholar]
  • 68.Plays M., Müller S., Rodriguez R. Chemistry and biology of ferritin. Metallomics. 2021;13 doi: 10.1093/mtomcs/mfab021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Arapova G.S., Eryomin A.N., Metelitza D.I. Catalytic peroxidase-like activity of an iron-containing crystallite isolated from ferritin in aqueous solution and reversed micelles of aerosol OT in heptane. Russ. J. Bioorganic Chem. 1999;25:369–376. [Google Scholar]
  • 70.Tang Z., Wu H., Zhang Y., Li Z., Lin Y. Enzyme-mimic activity of ferric nano-core residing in ferritin and its biosensing applications. Anal. Chem. 2011;83:8611–8616. doi: 10.1021/ac202049q. [DOI] [PubMed] [Google Scholar]
  • 71.Ma L., Zheng J.-J., Zhou N., Zhang R., Fang L., Yang Y., Gao X., Chen C., Yan X., Fan K. A natural biogenic nanozyme for scavenging superoxide radicals. Nat. Commun. 2024;15:233. doi: 10.1038/s41467-023-44463-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Reed N.L., Yoon T.P. Oxidase reactions in photoredox catalysis. Chem. Soc. Rev. 2021;50:2954–2967. doi: 10.1039/d0cs00797h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Neves-Petersen M.T., Klitgaard S., Carvalho A.S.L., Petersen S.B., Aires de Barros M.R., Pinho e Melo E. Photophysics and photochemistry of horseradish peroxidase A2 upon ultraviolet illumination. Biophys. J. 2007;92:2016–2027. doi: 10.1529/biophysj.106.095455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Zhu M., Dai Y., Wu Y., Liu K., Qi X., Sun Y. Bandgap control of α-Fe(2)O(3) nanozymes and their superior visible light promoted peroxidase-like catalytic activity. Nanotechnology. 2018;29 doi: 10.1088/1361-6528/aaddc2. [DOI] [PubMed] [Google Scholar]
  • 75.Lin S., Tan W., Han P., Li X., Li J., Nie Z., Li K. A photonanozyme with light-empowered specific peroxidase-mimicking activity. Nano Res. 2022;15:9073–9081. doi: 10.1007/s12274-022-4538-5. [DOI] [Google Scholar]
  • 76.Wang X., Zhong X., Lei H., Geng Y., Zhao Q., Gong F., Yang Z., Dong Z., Liu Z., Cheng L. Hollow Cu2Se Nanozymes for Tumor Photothermal-Catalytic Therapy. Chem. Mater. 2019;31:6174–6186. doi: 10.1021/acs.chemmater.9b01958. [DOI] [Google Scholar]
  • 77.Zhou Z., Wang Y., Peng F., Meng F., Zha J., Ma L., Du Y., Peng N., Ma L., Zhang Q., et al. Intercalation-Activated Layered MoO3 Nanobelts as Biodegradable Nanozymes for Tumor-Specific Photo-Enhanced Catalytic Therapy. Angew. Chem. Int. Ed. 2022;61 doi: 10.1002/anie.202115939. [DOI] [PubMed] [Google Scholar]
  • 78.Lin S., Wang Y., Chen Z., Li L., Zeng J., Dong Q., Wang Y., Chai Z. Biomineralized Enzyme-Like Cobalt Sulfide Nanodots for Synergetic Phototherapy with Tumor Multimodal Imaging Navigation. ACS Sustain. Chem. Eng. 2018;6:12061–12069. doi: 10.1021/acssuschemeng.8b02386. [DOI] [Google Scholar]
  • 79.Karim M.N., Singh M., Weerathunge P., Bian P., Zheng R., Dekiwadia C., Ahmed T., Walia S., Della Gaspera E., Singh S., et al. Visible-Light-Triggered Reactive-Oxygen-Species-Mediated Antibacterial Activity of Peroxidase-Mimic CuO Nanorods. ACS Appl. Nano Mater. 2018;1:1694–1704. doi: 10.1021/acsanm.8b00153. [DOI] [Google Scholar]
  • 80.Lin Y., Liu X., Liu Z., Xu Y. Visible-Light-Driven Photocatalysis-Enhanced Nanozyme of TiO2 Nanotubes@MoS2 Nanoflowers for Efficient Wound Healing Infected with Multidrug-Resistant Bacteria. Small. 2021;17 doi: 10.1002/smll.202103348. [DOI] [PubMed] [Google Scholar]
  • 81.Han X., Liu R., Zhang H., Zhou Q., Feng W., Hu K. Enhanced Peroxidase-mimicking Activity of Plasmonic Gold-modified Mn3O4 Nanocomposites through Photoexcited Hot Electron Transfer. Chem. Asian J. 2021;16:1603–1607. doi: 10.1002/asia.202100337. [DOI] [PubMed] [Google Scholar]
  • 82.Cao C., Zou H., Yang N., Li H., Cai Y., Song X., Shao J., Chen P., Mou X., Wang W., Dong X. Fe3O4/Ag/Bi2MoO6 Photoactivatable Nanozyme for Self-Replenishing and Sustainable Cascaded Nanocatalytic Cancer Therapy. Adv. Mater. 2021;33 doi: 10.1002/adma.202106996. [DOI] [PubMed] [Google Scholar]
  • 83.Han H., Xu X., Kan H., Tang Y., Liu C., Wen H., Wu L., Jiang Y., Wang Z., Liu J., Wang F. Synergistic photodynamic/photothermal bacterial inactivation over heterogeneous quaternized chitosan/silver/cobalt phosphide nanocomposites. J. Colloid Interface Sci. 2022;616:304–315. doi: 10.1016/j.jcis.2022.02.068. [DOI] [PubMed] [Google Scholar]
  • 84.Shan J., Yang K., Xiu W., Qiu Q., Dai S., Yuwen L., Weng L., Teng Z., Wang L. Cu2MoS4 Nanozyme with NIR-II Light Enhanced Catalytic Activity for Efficient Eradication of Multidrug-Resistant Bacteria. Small. 2020;16 doi: 10.1002/smll.202001099. [DOI] [PubMed] [Google Scholar]
  • 85.Ambler R.P. Sequence variability in bacterial cytochromes c. Biochim. Biophys. Acta. 1991;1058:42–47. doi: 10.1016/S0005-2728(05)80266-X. [DOI] [PubMed] [Google Scholar]
  • 86.Mavridou D.A.I., Ferguson S.J., Stevens J.M. Cytochrome c assembly. IUBMB Life. 2013;65:209–216. doi: 10.1002/iub.1123. [DOI] [PubMed] [Google Scholar]
  • 87.Ma S., Ludwig R. Direct Electron Transfer of Enzymes Facilitated by Cytochromes. Chemelectrochem. 2019;6:958–975. doi: 10.1002/celc.201801256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Kalpage H.A., Wan J., Morse P.T., Zurek M.P., Turner A.A., Khobeir A., Yazdi N., Hakim L., Liu J., Vaishnav A., et al. Cytochrome c phosphorylation: Control of mitochondrial electron transport chain flux and apoptosis. Int. J. Biochem. Cell Biol. 2020;121 doi: 10.1016/j.biocel.2020.105704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Malström B.G. Cytochrome c oxidase Structure and catalytic activity. Biochim. Biophys. Acta Rev. Bioenerg. 1979;549:281–303. doi: 10.1016/0304-4173(79)90003-X. [DOI] [PubMed] [Google Scholar]
  • 90.Deng Z., Gong Y., Luo Y., Tian Y. WO3 nanostructures facilitate electron transfer of enzyme: application to detection of H2O2 with high selectivity. Biosens. Bioelectron. 2009;24:2465–2469. doi: 10.1016/j.bios.2008.12.037. [DOI] [PubMed] [Google Scholar]
  • 91.Chen M., Wang Z., Shu J., Jiang X., Wang W., Shi Z.-H., Lin Y.-W. Mimicking a Natural Enzyme System: Cytochrome c Oxidase-Like Activity of Cu2O Nanoparticles by Receiving Electrons from Cytochrome c. Inorg. Chem. 2017;56:9400–9403. doi: 10.1021/acs.inorgchem.7b01393. [DOI] [PubMed] [Google Scholar]
  • 92.Singh N., Mugesh G. CeVO4 Nanozymes Catalyze the Reduction of Dioxygen to Water without Releasing Partially Reduced Oxygen Species. Angew. Chem. Int. Ed. 2019;58:7797–7801. doi: 10.1002/anie.201903427. [DOI] [PubMed] [Google Scholar]
  • 93.Wang H., Chen J., Dong Q., Jia X., Li D., Wang J., Wang E. Cadmium sulfide as bifunctional mimics of NADH oxidase and cytochrome c reductase takes effect at physiological pH. Nano Res. 2022;15:5256–5261. doi: 10.1007/s12274-022-4150-8. [DOI] [Google Scholar]
  • 94.Uebe R., Schüler D. Magnetosome biogenesis in magnetotactic bacteria. Nat. Rev. Microbiol. 2016;14:621–637. doi: 10.1038/nrmicro.2016.99. [DOI] [PubMed] [Google Scholar]
  • 95.Strbak O., Hnilicova P., Gombos J., Lokajova A., Kopcansky P. Magnetotactic Bacteria: From Evolution to Biomineralization and Biomedical Applications. Minerals. 2022;12:1403. doi: 10.3390/min12111403. [DOI] [Google Scholar]
  • 96.Lower B.H., Bazylinski D.A. The Bacterial Magnetosome: A Unique Prokaryotic Organelle. J. Mol. Microbiol. Biotechnol. 2013;23:63–80. doi: 10.1159/000346543. [DOI] [PubMed] [Google Scholar]
  • 97.Lin W., Paterson G.A., Zhu Q., Wang Y., Kopylova E., Li Y., Knight R., Bazylinski D.A., Zhu R., Kirschvink J.L., Pan Y. Origin of microbial biomineralization and magnetotaxis during the Archean. Proc. Natl. Acad. Sci. USA. 2017;114:2171–2176. doi: 10.1073/pnas.1614654114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Goswami P., He K., Li J., Pan Y., Roberts A.P., Lin W. Magnetotactic bacteria and magnetofossils: ecology, evolution and environmental implications. NPJ Biofilms Microbiomes. 2022;8:43. doi: 10.1038/s41522-022-00304-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Komeili A., Li Z., Newman D.K., Jensen G.J. Magnetosomes are cell membrane invaginations organized by the actin-like protein MamK. Science. 2006;311:242–245. doi: 10.1126/science.1123231. [DOI] [PubMed] [Google Scholar]
  • 100.Siponen M.I., Adryanczyk G., Ginet N., Arnoux P., Pignol D. Magnetochrome: a c-type cytochrome domain specific to magnetotatic bacteria. Biochem. Soc. Trans. 2012;40:1319–1323. doi: 10.1042/bst20120104. [DOI] [PubMed] [Google Scholar]
  • 101.Taoka A., Eguchi Y., Shimoshige R., Fukumori Y. Recent advances in studies on magnetosome-associated proteins composing the bacterial geomagnetic sensor organelle. Microbiol. Immunol. 2023;67:228–238. doi: 10.1111/1348-0421.13062. [DOI] [PubMed] [Google Scholar]
  • 102.Dröge W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002;82:47–95. doi: 10.1152/physrev.00018.2001. [DOI] [PubMed] [Google Scholar]
  • 103.Halliwell B. Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiol. 2006;141:312–322. doi: 10.1104/pp.106.077073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Ślesak I., Libik M., Karpinska B., Karpinski S., Miszalski Z. The role of hydrogen peroxide in regulation of plant metabolism and cellular signalling in response to environmental stresses. Acta Biochim. Pol. 2007;54:39–50. doi: 10.18388/abp.2007_3267. [DOI] [PubMed] [Google Scholar]
  • 105.Sies H. Hydrogen peroxide as a central redox signaling molecule in physiological oxidative stress: Oxidative eustress. Redox Biol. 2017;11:613–619. doi: 10.1016/j.redox.2016.12.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Taverne Y.J., Merkus D., Bogers A.J., Halliwell B., Duncker D.J., Lyons T.W. Reactive Oxygen Species: Radical Factors in the Evolution of Animal Life: A molecular timescale from Earth's earliest history to the rise of complex life. Bioessays. 2018;40 doi: 10.1002/bies.201700158. [DOI] [PubMed] [Google Scholar]
  • 107.Taverne Y.J., Caron A., Diamond C., Fournier G., Lyons T.W. In: Oxidative Stress. Sies H., editor. Academic Press; 2020. Chapter 5 - Oxidative stress and the early coevolution of life and biospheric oxygen; pp. 67–85. [DOI] [Google Scholar]
  • 108.Runnegar B. Precambrian oxygen levels estimated from the biochemistry and physiology of early eukaryotes. Glob. Planet. Change. 1991;5:97–111. doi: 10.1016/0921-8181(91)90131-F. [DOI] [Google Scholar]
  • 109.Castresana J., Lübben M., Saraste M., Higgins D.G. Evolution of cytochrome oxidase, an enzyme older than atmospheric oxygen. EMBO J. 1994;13:2516–2525. doi: 10.1002/j.1460-2075.1994.tb06541.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Castresana J., Saraste M. Evolution of energetic metabolism: the respiration-early hypothesis. Trends Biochem. Sci. 1995;20:443–448. doi: 10.1016/S0968-0004(00)89098-2. [DOI] [PubMed] [Google Scholar]
  • 111.Lenton T.M. Evolution on Planet Earth: The Impact of the Physical Environment. Elsevier Ltd; 2003. The coupled evolution of life and atmospheric oxygen; pp. 35–53. [DOI] [Google Scholar]
  • 112.Neubeck A., Freund F. Sulfur Chemistry May Have Paved the Way for Evolution of Antioxidants. Astrobiology. 2020;20:670–675. doi: 10.1089/ast.2019.2156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Frankel R.B., Papaefthymiou G.C., Blakemore R.P., O'Brien W. Fe3O4 precipitation in magnetotactic bacteria. BBA Mol. Cell Res. 1983;763:147–159. doi: 10.1016/0167-4889(83)90038-1. [DOI] [Google Scholar]
  • 114.Faivre D., Böttger L.H., Matzanke B.F., Schüler D. Intracellular magnetite biomineralization in bacteria proceeds by a distinct pathway involving membrane-bound ferritin and an iron(II) species. Angew. Chem. Int. Ed. 2007;46:8495–8499. doi: 10.1002/anie.200700927. [DOI] [PubMed] [Google Scholar]
  • 115.Staniland S., Ward B., Harrison A., Van Der Laan G., Telling N. Rapid magnetosome formation shown by real-time x-ray magnetic circular dichroism. Proc. Natl. Acad. Sci. USA. 2007;104:19524–19528. doi: 10.1073/pnas.0704879104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Fdez-Gubieda M.L., Muela A., Alonso J., García-Prieto A., Olivi L., Fernández-Pacheco R., Barandiarán J.M. Magnetite biomineralization in magnetospirillum gryphiswaldense: Time-resolved magnetic and structural studies. ACS Nano. 2013;7:3297–3305. doi: 10.1021/nn3059983. [DOI] [PubMed] [Google Scholar]
  • 117.Baumgartner J., Morin G., Menguy N., Perez Gonzalez T., Widdrat M., Cosmidis J., Faivre D. Magnetotactic bacteria form magnetite from a phosphate-rich ferric hydroxide via nanometric ferric (oxyhydr)oxide intermediates. Proc. Natl. Acad. Sci. USA. 2013;110:14883–14888. doi: 10.1073/pnas.1307119110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Baumgartner J., Menguy N., Gonzalez T.P., Morin G., Widdrat M., Faivre D. Elongated magnetite nanoparticle formation from a solid ferrous precursor in a magnetotactic bacterium. J. R. Soc. Interface. 2016;13 doi: 10.1098/rsif.2016.0665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Le Nagard L., Zhu X., Yuan H., Benzerara K., Bazylinski D.A., Fradin C., Besson A., Swaraj S., Stanescu S., Belkhou R., Hitchcock A.P. Magnetite magnetosome biomineralization in Magnetospirillum magneticum strain AMB-1: A time course study. Chem. Geol. 2019;530 doi: 10.1016/j.chemgeo.2019.119348. [DOI] [Google Scholar]
  • 120.Wen T., Zhang Y., Geng Y., Liu J., Basit A., Tian J., Li Y., Li J., Ju J., Jiang W. Epsilon-Fe2O3 is a novel intermediate for magnetite biosynthesis in magnetotactic bacteria. Biomater. Res. 2019;23:13. doi: 10.1186/s40824-019-0162-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Gorby Y.A., Beveridge T.J., Blakemore R.P. Characterization of the bacterial magnetosome membrane. J. Bacteriol. 1988;170:834–841. doi: 10.1128/jb.170.2.834-841.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Zanella D., Bossi E., Gornati R., Bastos C., Faria N., Bernardini G. Iron oxide nanoparticles can cross plasma membranes. Sci. Rep. 2017;7 doi: 10.1038/s41598-017-11535-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Chilom C.G., Zorilă B., Bacalum M., Bălăşoiu M., Yaroslavtsev R., Stolyar S.V., Tyutyunnicov S. Ferrihydrite nanoparticles interaction with model lipid membranes. Chem. Phys. Lipids. 2020;226 doi: 10.1016/j.chemphyslip.2019.104851. [DOI] [PubMed] [Google Scholar]
  • 124.Liu P., Zheng Y., Zhang R., Bai J., Zhu K., Benzerara K., Menguy N., Zhao X., Roberts A.P., Pan Y., Li J. Key gene networks that control magnetosome biomineralization in magnetotactic bacteria. Natl. Sci. Rev. 2023;10 doi: 10.1093/nsr/nwac238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Murat D., Quinlan A., Vali H., Komeili A. Comprehensive genetic dissection of the magnetosome gene island reveals the step-wise assembly of a prokaryotic organelle. Proc. Natl. Acad. Sci. USA. 2010;107:5593–5598. doi: 10.1073/pnas.0914439107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Cornell R.M., Schwertmann U. Second Edition. Wiley-VCH Verlag GmbH & Co. KGaA; 2003. The Iron Oxides: Structure, Properties, Reactions, Occurrence and Uses. [DOI] [Google Scholar]
  • 127.Jolivet J.P., Tronc E., Chanéac C. Iron oxides: From molecular clusters to solid. A nice example of chemical versatility. Chem. Commun. (Camb) 2006;338:488–497. doi: 10.1016/j.crte.2006.04.014. [DOI] [Google Scholar]
  • 128.Jolivet J.P. Oxford University Press; 2019. Metal Oxide Nanostructures Chemistry: Synthesis from Aqueous Solutions. [DOI] [Google Scholar]
  • 129.Michel F.M., Ehm L., Antao S.M., Lee P.L., Chupas P.J., Liu G., Strongin D.R., Schoonen M.A.A., Phillips B.L., Parise J.B. The structure of ferrihydrite, a nanocrystalline material. Science. 2007;316:1726–1729. doi: 10.1126/science.1142525. [DOI] [PubMed] [Google Scholar]
  • 130.Michel F.M., Barrón V., Torrent J., Morales M.P., Serna C.J., Boily J.F., Liu Q., Ambrosini A., Cismasu A.C., Brown G.E., Jr. Ordered ferrimagnetic form of ferrihydrite reveals links among structure, composition, and magnetism. Proc. Natl. Acad. Sci. USA. 2010;107:2787–2792. doi: 10.1073/pnas.0910170107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Manceau A., Combes J.M. Structure of Mn and Fe oxides and oxyhydroxides: A topological approach by EXAFS. Phys. Chem. Miner. 1988;15:283–295. doi: 10.1007/BF00307518. [DOI] [Google Scholar]
  • 132.Girardet T., Venturini P., Martinez H., Dupin J.-C., Cleymand F., Fleutot S. Spinel Magnetic Iron Oxide Nanoparticles: Properties, Synthesis and Washing Methods. Appl. Sci. 2022;12:8127. doi: 10.3390/app12168127. [DOI] [Google Scholar]
  • 133.Greaves C. A powder neutron diffraction investigation of vacancy ordering and covalence in γ-Fe2O3. J. Solid State Chem. 1983;49:325–333. doi: 10.1016/S0022-4596(83)80010-3. [DOI] [Google Scholar]
  • 134.Morales M.P., Veintemillas-Verdaguer S., Montero M.I., Serna C.J., Roig A., Casas L., Martínez B., Sandiumenge F. Surface and internal spin canting in γ-Fe2O3 nanoparticles. Chem. Mater. 1999;11:3058–3064. doi: 10.1021/cm991018f. [DOI] [Google Scholar]
  • 135.Grau-Crespo R., Al-Baitai A.Y., Saadoune I., De Leeuw N.H. Vacancy ordering and electronic structure of γ-Fe2O3 (maghemite): a theoretical investigation. J. Phys. Condens. Matter. 2010;22 doi: 10.1088/0953-8984/22/25/255401. [DOI] [PubMed] [Google Scholar]
  • 136.Hill A.H., Jiao F., Bruce P.G., Harrison A., Kockelmann W., Ritter C. Neutron Diffraction Study of Mesoporous and Bulk Hematite, α-Fe2O3. Chem. Mater. 2008;20:4891–4899. doi: 10.1021/cm800009s. [DOI] [Google Scholar]
  • 137.Dzade N., Roldan A., de Leeuw N. A Density Functional Theory Study of the Adsorption of Benzene on Hematite (α-Fe2O3) Surfaces. Minerals. 2014;4:89–115. doi: 10.3390/min4010089. [DOI] [Google Scholar]
  • 138.Zhang H., Sun W., Xie X., He J., Zhang C. Insights into the Fracture Nature of Hematite from First Principles DFT Calculations. ACS Omega. 2023;8:8248–8255. doi: 10.1021/acsomega.2c06101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Nolze G., Winkelmann A. Exploring structural similarities between crystal phases using EBSD pattern comparison. Cryst. Res. Technol. 2014;49:490–501. doi: 10.1002/crat.201400091. [DOI] [Google Scholar]
  • 140.Kappler A., Bryce C., Mansor M., Lueder U., Byrne J.M., Swanner E.D. An evolving view on biogeochemical cycling of iron. Nat. Rev. Microbiol. 2021;19:360–374. doi: 10.1038/s41579-020-00502-7. [DOI] [PubMed] [Google Scholar]
  • 141.Usman M., Byrne J.M., Chaudhary A., Orsetti S., Hanna K., Ruby C., Kappler A., Haderlein S.B. Magnetite and Green Rust: Synthesis, Properties, and Environmental Applications of Mixed-Valent Iron Minerals. Chem. Rev. 2018;118:3251–3304. doi: 10.1021/acs.chemrev.7b00224. [DOI] [PubMed] [Google Scholar]
  • 142.Raiswell R., Canfield D.E. The iron biogeochemical cycle past and present. Geochem. Perspect. 2012;1:1–220. doi: 10.7185/geochempersp.1.1. [DOI] [Google Scholar]
  • 143.Fitzsimmons J.N., Conway T.M. Novel Insights into Marine Iron Biogeochemistry from Iron Isotopes. Ann. Rev. Mar. Sci. 2023;15:383–406. doi: 10.1146/annurev-marine-032822-103431. [DOI] [PubMed] [Google Scholar]
  • 144.Yu C., Xie S., Song Z., Xia S., Åström M.E. Biogeochemical cycling of iron (hydr-)oxides and its impact on organic carbon turnover in coastal wetlands: A global synthesis and perspective. Earth Sci. Rev. 2021;218 doi: 10.1016/j.earscirev.2021.103658. [DOI] [Google Scholar]
  • 145.Günther K. In: Diet for Iron Deficiency: Metabolism - Bioavailability - Diagnostics. Günther K., editor. Springer Berlin Heidelberg; 2023. Biochemistry of Iron; pp. 7–27. [DOI] [Google Scholar]
  • 146.Usman M., Abdelmoula M., Hanna K., Grégoire B., Faure P., Ruby C. FeII induced mineralogical transformations of ferric oxyhydroxides into magnetite of variable stoichiometry and morphology. J. Solid State Chem. 2012;194:328–335. doi: 10.1016/j.jssc.2012.05.022. [DOI] [Google Scholar]
  • 147.Braterman P.S., Cairns-Smith A.G., Sloper R.W. Photo-oxidation of hydrated Fe2+-significance for banded iron formations. Nature. 1983;303:163–164. doi: 10.1038/303163a0. [DOI] [Google Scholar]
  • 148.Shu Z., Liu L., Tan W., Suib S.L., Qiu G., Yang X., Zheng L., Liu F. Solar Irradiation Induced Transformation of Ferrihydrite in the Presence of Aqueous Fe2+ Environ. Sci. Technol. 2019;53:8854–8861. doi: 10.1021/acs.est.9b02750. [DOI] [PubMed] [Google Scholar]
  • 149.Shu Z., Pan Z., Wang X., He H., Yan S., Zhu X., Song W., Wang Z. Sunlight-Induced Interfacial Electron Transfer of Ferrihydrite under Oxic Conditions: Mineral Transformation and Redox Active Species Production. Environ. Sci. Technol. 2022;56:14188–14197. doi: 10.1021/acs.est.2c04594. [DOI] [PubMed] [Google Scholar]
  • 150.Jolivet J.-P., Tronc E. Interfacial electron transfer in colloidal spinel iron oxide. Conversion of Fe3O4-γFe2O3 in aqueous medium. J. Colloid Interface Sci. 1988;125:688–701. doi: 10.1016/0021-9797(88)90036-7. [DOI] [Google Scholar]
  • 151.Jungcharoen P., Pédrot M., Heberling F., Hanna K., Choueikani F., Catrouillet C., Dia A., Marsac R. Prediction of nanomagnetite stoichiometry (Fe(ii)/Fe(iii)) under contrasting pH and redox conditions. Environ. Sci. Nano. 2022;9:2363–2371. doi: 10.1039/D2EN00112H. [DOI] [Google Scholar]
  • 152.Schwaminger S.P., Bauer D., Fraga-García P., Wagner F.E., Berensmeier S. Oxidation of magnetite nanoparticles: impact on surface and crystal properties. CrystEngComm. 2017;19:246–255. doi: 10.1039/C6CE02421A. [DOI] [Google Scholar]
  • 153.Creutzburg M., Sellschopp K., Gleißner R., Arndt B., Vonbun-Feldbauer G.B., Vonk V., Noei H., Stierle A. Surface structure of magnetite (111) under oxidizing and reducing conditions. J. Phys. Condens. Matter. 2022;34 doi: 10.1088/1361-648X/ac4d5a. [DOI] [PubMed] [Google Scholar]
  • 154.Luo H.-W., Zhang X., Chen J.-J., Yu H.-Q., Sheng G.-P. Probing the biotransformation of hematite nanoparticles and magnetite formation mediated by Shewanella oneidensis MR-1 at the molecular scale. Environ. Sci. Nano. 2017;4:2395–2404. doi: 10.1039/C7EN00767A. [DOI] [Google Scholar]
  • 155.Auffan M., Achouak W., Rose J., Roncato M.A., Chanéac C., Waite D.T., Masion A., Woicik J.C., Wiesner M.R., Bottero J.Y. Relation between the redox state of iron-based nanoparticles and their cytotoxicity toward Escherichia coli. Environ. Sci. Technol. 2008;42:6730–6735. doi: 10.1021/es800086f. [DOI] [PubMed] [Google Scholar]
  • 156.Liu S., Lu F., Xing R., Zhu J.J. Structural effects of Fe3O4 nanocrystals on peroxidase-like activity. Chem. Eur J. 2011;17:620–625. doi: 10.1002/chem.201001789. [DOI] [PubMed] [Google Scholar]
  • 157.Puvvada N., Panigrahi P.K., Mandal D., Pathak A. Shape dependent peroxidase mimetic activity towards oxidation of pyrogallol by H2O2. RSC Adv. 2012;2:3270–3273. doi: 10.1039/C2RA01081J. [DOI] [Google Scholar]
  • 158.Cheng X.-L., Jiang J.-S., Jiang D.-M., Zhao Z.-J. Synthesis of Rhombic Dodecahedral Fe3O4 Nanocrystals with Exposed High-Energy {110} Facets and Their Peroxidase-like Activity and Lithium Storage Properties. J. Phys. Chem. C. 2014;118:12588–12598. doi: 10.1021/jp412661e. [DOI] [Google Scholar]
  • 159.Mu J., Zhang L., Zhao G., Wang Y. The crystal plane effect on the peroxidase-like catalytic properties of Co3O4 nanomaterials. Phys. Chem. Chem. Phys. 2014;16:15709–15716. doi: 10.1039/c4cp01326c. [DOI] [PubMed] [Google Scholar]
  • 160.Peng Y., Wang Z., Liu W., Zhang H., Zuo W., Tang H., Chen F., Wang B. Size- and shape-dependent peroxidase-like catalytic activity of MnFe2O4 Nanoparticles and their applications in highly efficient colorimetric detection of target cancer cells. Dalton Trans. 2015;44:12871–12877. doi: 10.1039/c5dt01585e. [DOI] [PubMed] [Google Scholar]
  • 161.Yang Y., Mao Z., Huang W., Liu L., Li J., Li J., Wu Q. Redox enzyme-mimicking activities of CeO(2) nanostructures: Intrinsic influence of exposed facets. Sci. Rep. 2016;6 doi: 10.1038/srep35344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Ghosh S., Roy P., Karmodak N., Jemmis E.D., Mugesh G. Nanoisozymes: Crystal-Facet-Dependent Enzyme-Mimetic Activity of V2O5 Nanomaterials. Angew. Chem. Int. Ed. 2018;57:4510–4515. doi: 10.1002/anie.201800681. [DOI] [PubMed] [Google Scholar]
  • 163.Jiang Y., Xia T., Shen L., Ma J., Ma H., Sun T., Lv F., Zhu N. Facet-Dependent Cu2O Electrocatalysis for Wearable Enzyme-Free Smart Sensing. ACS Catal. 2021;11:2949–2955. doi: 10.1021/acscatal.0c04797. [DOI] [Google Scholar]
  • 164.Xu Z., Qu A., Wang W., Lu M., Shi B., Chen C., Hao C., Xu L., Sun M., Xu C., Kuang H. Facet-Dependent Biodegradable Mn3O4 Nanoparticles for Ameliorating Parkinson's Disease. Adv. Healthc. Mater. 2021;10 doi: 10.1002/adhm.202101316. [DOI] [PubMed] [Google Scholar]
  • 165.Chen M., Zhou X., Xiong C., Yuan T., Wang W., Zhao Y., Xue Z., Guo W., Wang Q., Wang H., et al. Facet Engineering of Nanoceria for Enzyme-Mimetic Catalysis. ACS Appl. Mater. Interfaces. 2022;14:21989–21995. doi: 10.1021/acsami.2c04320. [DOI] [PubMed] [Google Scholar]
  • 166.Singh K., Gujju R., Bandaru S., Misra S., Babu K.S., Puvvada N. Facet-Dependent Bactericidal Activity of Ag3PO4 Nanostructures against Gram-Positive/Negative Bacteria. ACS Omega. 2022;7:16616–16628. doi: 10.1021/acsomega.2c00864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Zhang R., Zhao H., Fan K. Nanozymes: Design, Synthesis, and Applications. American Chemical Society; 2022. Structure-Activity Mechanism of Iron Oxide Nanozymes; pp. 1–35. [DOI] [Google Scholar]
  • 168.Tian T., Ai L., Liu X., Li L., Li J., Jiang J. Synthesis of Hierarchical FeWO4 Architectures with {100}-Faceted Nanosheet Assemblies as a Robust Biomimetic Catalyst. Ind. Eng. Chem. Res. 2015;54:1171–1178. doi: 10.1021/ie504114v. [DOI] [Google Scholar]
  • 169.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. doi: 10.1038/s41467-022-33098-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Yu G.H., Chi Z.L., Kappler A., Sun F.S., Liu C.Q., Teng H.H., Gadd G.M. Fungal Nanophase Particles Catalyze Iron Transformation for Oxidative Stress Removal and Iron Acquisition. Curr. Biol. 2020;30:2943–2950.e4. doi: 10.1016/j.cub.2020.05.058. [DOI] [PubMed] [Google Scholar]
  • 171.Chi Z.L., Zhao X.Y., Chen Y.L., Hao J.L., Yu G.H., Goodman B.A., Gadd G.M. Intrinsic enzyme-like activity of magnetite particles is enhanced by cultivation with Trichoderma guizhouense. Environ. Microbiol. 2021;23:893–907. doi: 10.1111/1462-2920.15193. [DOI] [PubMed] [Google Scholar]
  • 172.Qi Y., Sun Y., Song D., Wang Y., Xiu F. PVC dechlorination residues as new peroxidase-mimicking nanozyme and chemiluminescence sensing probe with high activity for glucose and ascorbic acid detection. Talanta. 2023;253 doi: 10.1016/j.talanta.2022.124039. [DOI] [Google Scholar]
  • 173.Yu X., Huo C.-F., Li Y.-W., Wang J., Jiao H. Fe3O4 surface electronic structures and stability from GGA+U. Surf. Sci. 2012;606:872–879. doi: 10.1016/j.susc.2012.02.003. [DOI] [Google Scholar]
  • 174.Noh J., Osman O.I., Aziz S.G., Winget P., Brédas J.L. Magnetite Fe3O4 (111) Surfaces: Impact of Defects on Structure, Stability, and Electronic Properties. Chem. Mater. 2015;27:5856–5867. doi: 10.1021/acs.chemmater.5b02885. [DOI] [Google Scholar]
  • 175.Erlebach A., Kurland H.-D., Grabow J., Müller F.A., Sierka M. Structure evolution of nanoparticulate Fe2O3. Nanoscale. 2015;7:2960–2969. doi: 10.1039/C4NR06989G. [DOI] [PubMed] [Google Scholar]
  • 176.Huang X., Ramadugu S.K., Mason S.E. Surface-Specific DFT + U Approach Applied to α-Fe2O3(0001) J. Phys. Chem. C. 2016;120:4919–4930. doi: 10.1021/acs.jpcc.5b12144. [DOI] [Google Scholar]
  • 177.Jian W., Wang S.-P., Zhang H.-X., Bai F.-Q. Disentangling the role of oxygen vacancies on the surface of Fe3O4 and γ-Fe2O3. Inorg. Chem. Front. 2019;6:2660–2666. doi: 10.1039/C9QI00351G. [DOI] [Google Scholar]
  • 178.Paidi V.K., Shepit M., Freeland J.W., Brewe D.L., Roberts C.A., van Lierop J. Intervening Oxygen Enabled Magnetic Moment Modulation in Spinel Nanostructures. J. Phys. Chem. C. 2021;125:26688–26697. doi: 10.1021/acs.jpcc.1c06494. [DOI] [Google Scholar]
  • 179.Davison W. The solubility of iron sulphides in synthetic and natural waters at ambient temperature. Aquat. Sci. 1991;53:309–329. doi: 10.1007/BF00877139. [DOI] [Google Scholar]
  • 180.Krupp R.E. Phase relations and phase transformations between the low-temperature iron sulfides mackinawite, greigite, and smythite. Eur. J. Mineral. 1994;6:265–278. doi: 10.1127/EJM/6/2/0265. [DOI] [Google Scholar]
  • 181.Benning L.G., Wilkin R.T., Barnes H.L. Reaction pathways in the Fe–S system below 100°C. Chem. Geol. 2000;167:25–51. doi: 10.1016/S0009-2541(99)00198-9. [DOI] [Google Scholar]
  • 182.Jørgensen B.B., Nelson D.C., Amend J.P., Edwards K.J., Lyons T.W. Sulfur Biogeochemistry - Past and Present. Geological Society of America; 2004. Sulfide oxidation in marine sediments: Geochemistry meets microbiology. [DOI] [Google Scholar]
  • 183.Csákberényi-Malasics D., Rodriguez-Blanco J.D., Kis V.K., Rečnik A., Benning L.G., Pósfai M. Structural properties and transformations of precipitated FeS. Chem. Geol. 2012;294–295:249–258. doi: 10.1016/j.chemgeo.2011.12.009. [DOI] [Google Scholar]
  • 184.Lan Y., Elwood Madden A.S., Butler E.C. Transformation of mackinawite to greigite by trichloroethylene and tetrachloroethylene. Environ. Sci. Process. Impacts. 2016;18:1266–1273. doi: 10.1039/C6EM00461J. [DOI] [PubMed] [Google Scholar]
  • 185.Wasmund K., Mußmann M., Loy A. The life sulfuric: microbial ecology of sulfur cycling in marine sediments. Environ. Microbiol. Rep. 2017;9:323–344. doi: 10.1111/1758-2229.12538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Li S.H., Chen Y.-H., Lee J.-J., Sheu H.-S. Phase transition of iron sulphide minerals under hydrothermal conditions and magnetic investigations. Phys. Chem. Miner. 2018;45:27–38. doi: 10.1007/s00269-017-0898-x. [DOI] [Google Scholar]
  • 187.Sano Y., Kyono A., Yoneda Y., Isaka N., Takagi S., Yamamoto G. Structure changes of nanocrystalline mackinawite under hydrothermal conditions. J. Mineral. Petrol. Sci. 2020;115:261–275. doi: 10.2465/jmps.190903. [DOI] [Google Scholar]
  • 188.Bolney R., Grosch M., Winkler M., van Slageren J., Weigand W., Robl C. Mackinawite formation from elemental iron and sulfur. RSC Adv. 2021;11:32464–32475. doi: 10.1039/D1RA03705F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Cox P.A. Oxford University Press; 1987. The Electronic Structure and Chemistry of Solids. [Google Scholar]
  • 190.Atkins P.W., Paula J.d. 9th Edition. Oxford University Presss; 2010. Physical Chemistry. [Google Scholar]
  • 191.Sze S.M., Lee M.-K. 3rd Edition. Wiley; 2012. Semiconductor Devices: Physics and Technology. [Google Scholar]
  • 192.Pumera M., Sánchez S., Ichinose I., Tang J. Electrochemical nanobiosensors. Sens. Actuators B Chem. 2007;123:1195–1205. doi: 10.1016/j.snb.2006.11.016. [DOI] [Google Scholar]
  • 193.Shi L., Dong H., Reguera G., Beyenal H., Lu A., Liu J., Yu H.-Q., Fredrickson J.K. Extracellular electron transfer mechanisms between microorganisms and minerals. Nat. Rev. Microbiol. 2016;14:651–662. doi: 10.1038/nrmicro.2016.93. [DOI] [PubMed] [Google Scholar]
  • 194.Lovley D.R. Syntrophy Goes Electric: Direct Interspecies Electron Transfer. Ann Rev Microbiol. 2017;71:643–664. doi: 10.1146/annurev-micro-030117-020420. [DOI] [PubMed] [Google Scholar]
  • 195.Wander M.C.F., Rosso K.M., Schoonen M.A.A. Structure and Charge Hopping Dynamics in Green Rust. J. Phys. Chem. C. 2007;111:11414–11423. doi: 10.1021/jp072762n. [DOI] [Google Scholar]
  • 196.Skomurski F.N., Kerisit S., Rosso K.M. Structure, charge distribution, and electron hopping dynamics in magnetite (Fe3O4) (100) surfaces from first principles. Geochim. Cosmochim. Acta. 2010;74:4234–4248. doi: 10.1016/j.gca.2010.04.063. [DOI] [Google Scholar]
  • 197.Alexandrov V., Rosso K.M. Electron transport in pure and substituted iron oxyhydroxides by small-polaron migration. J. Chem. Phys. 2014;140 doi: 10.1063/1.4882065. [DOI] [PubMed] [Google Scholar]
  • 198.Zarzycki P., Smith D.M., Rosso K.M. Proton Dynamics on Goethite Nanoparticles and Coupling to Electron Transport. J. Chem. Theory Comput. 2015;11:1715–1724. doi: 10.1021/ct500891a. [DOI] [PubMed] [Google Scholar]
  • 199.Iordanova N., Dupuis M., Rosso K.M. Charge transport in metal oxides: a theoretical study of hematite alpha-Fe2O3. J. Chem. Phys. 2005;122 doi: 10.1063/1.1869492. [DOI] [PubMed] [Google Scholar]
  • 200.Kerisit S., Rosso K.M. Computer simulation of electron transfer at hematite surfaces. Geochim. Cosmochim. Acta. 2006;70:1888–1903. doi: 10.1016/j.gca.2005.12.021. [DOI] [Google Scholar]
  • 201.Katz J.E., Zhang X., Attenkofer K., Chapman K.W., Frandsen C., Zarzycki P., Rosso K.M., Falcone R.W., Waychunas G.A., Gilbert B. Electron small polarons and their mobility in iron (oxyhydr)oxide nanoparticles. Science. 2012;337:1200–1203. doi: 10.1126/science.1223598. [DOI] [PubMed] [Google Scholar]
  • 202.Husek J., Cirri A., Biswas S., Baker L.R. Surface electron dynamics in hematite (α-Fe2O3): correlation between ultrafast surface electron trapping and small polaron formation. Chem. Sci. 2017;8:8170–8178. doi: 10.1039/C7SC02826A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Carneiro L.M., Cushing S.K., Liu C., Su Y., Yang P., Alivisatos A.P., Leone S.R. Excitation-wavelength-dependent small polaron trapping of photoexcited carriers in α-Fe2O3. Nat. Mater. 2017;16:819–825. doi: 10.1038/nmat4936. [DOI] [PubMed] [Google Scholar]
  • 204.Ibrahim E.M.M., Abdel-Rahman L.H., Abu-Dief A.M., Elshafaie A., Hamdan S.K., Ahmed A.M. Electric, thermoelectric and magnetic characterization of γ-Fe2O3 and Co3O4 nanoparticles synthesized by facile thermal decomposition of metal-Schiff base complexes. Mater. Res. Bull. 2018;99:103–108. doi: 10.1016/j.materresbull.2017.11.002. [DOI] [Google Scholar]
  • 205.Soltis J.A., Schwartzberg A.M., Zarzycki P., Penn R.L., Rosso K.M., Gilbert B. Electron Mobility and Trapping in Ferrihydrite Nanoparticles. ACS Earth Space Chem. 2017;1:216–226. doi: 10.1021/acsearthspacechem.7b00030. [DOI] [Google Scholar]
  • 206.Verwey E.J.W. Electronic Conduction of Magnetite (Fe3O4) and its Transition Point at Low Temperatures. Nature. 1939;144:327–328. doi: 10.1038/144327b0. [DOI] [Google Scholar]
  • 207.Tannhauser D.S. Conductivity in iron oxides. J. Phys. Chem. Solids. 1962;23:25–34. doi: 10.1016/0022-3697(62)90053-7. [DOI] [Google Scholar]
  • 208.de Groot R.A., Mueller F.M., Engen P.G.v., Buschow K.H.J. New Class of Materials: Half-Metallic Ferromagnets. Phys. Rev. Lett. 1983;50:2024–2027. doi: 10.1103/PhysRevLett.50.2024. [DOI] [Google Scholar]
  • 209.Coey J.M.D., Berkowitz A.E., Balcells L., Putris F.F., Parker F.T. Magnetoresistance of magnetite. Appl. Phys. Lett. 1998;72:734–736. doi: 10.1063/1.120859. [DOI] [Google Scholar]
  • 210.Soeya S., Hayakawa J., Takahashi H., Ito K., Yamamoto C., Kida A., Asano H., Matsui M. Development of half-metallic ultrathin Fe3O4 films for spin-transport devices. Appl. Phys. Lett. 2002;80:823–825. doi: 10.1063/1.1446995. [DOI] [Google Scholar]
  • 211.Dedkov Y.S., Rüdiger U., Güntherodt G. Evidence for the half-metallic ferromagnetic state of Fe3O4 by spin-resolved photoelectron spectroscopy. Phys. Rev. B. 2002;65 doi: 10.1103/PhysRevB.65.064417. [DOI] [Google Scholar]
  • 212.Mounkachi O., Lamouri R., Hamedoun M., Ez-Zahraouy H., Salmani E., Benyoussef A. Effect of Defects Disorder on the Half-Metallicity, Magnetic Properties, and Gap States of Fe3O4: a First-Principles Study. J. Supercond. Nov. Magn. 2017;30:3221–3224. doi: 10.1007/s10948-017-4108-3. [DOI] [Google Scholar]
  • 213.Radoń A., Łukowiec D., Kremzer M., Mikuła J., Włodarczyk P. Electrical Conduction Mechanism and Dielectric Properties of Spherical Shaped Fe₃O₄ Nanoparticles Synthesized by Co-Precipitation Method. Materials. 2018;11:735. doi: 10.3390/ma11050735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Morris E.R., Williams Q. Electrical resistivity of Fe3O4 to 48 GPa: Compression-induced changes in electron hopping at mantle pressures. J. Geophys. Res. 1997;102:18139–18148. doi: 10.1029/97JB00024. [DOI] [Google Scholar]
  • 215.Pasternak M.P., Rozenberg G.K., Machavariani G.Y., Naaman O., Taylor R.D., Jeanloz R. Breakdown of the Mott-Hubbard State in Fe2O3: A First-Order Insulator-Metal Transition with Collapse of Magnetism at 50 GPa. Phys. Rev. Lett. 1999;82:4663–4666. doi: 10.1103/PhysRevLett.82.4663. [DOI] [Google Scholar]
  • 216.Ohta K., Hirose K., Shimizu K., Ohishi Y. High-pressure experimental evidence for metal FeO with normal NiAs-type structure. Phys. Rev. B. 2010;82 doi: 10.1103/PhysRevB.82.174120. [DOI] [Google Scholar]
  • 217.Ohta K., Cohen R.E., Hirose K., Haule K., Shimizu K., Ohishi Y. Experimental and Theoretical Evidence for Pressure-Induced Metallization in FeO with Rocksalt-Type Structure. Phys. Rev. Lett. 2012;108 doi: 10.1103/PhysRevLett.108.026403. [DOI] [PubMed] [Google Scholar]
  • 218.Todo S., Takeshita N., Kanehara T., Mori T., Môri N. Metallization of magnetite (Fe3O4) under high pressure. J. Appl. Phys. 2001;89:7347–7349. doi: 10.1063/1.1359460. [DOI] [Google Scholar]
  • 219.Senn M.S., Wright J.P., Attfield J.P. Charge order and three-site distortions in the Verwey structure of magnetite. Nature. 2011;481:173–176. doi: 10.1038/nature10704. [DOI] [PubMed] [Google Scholar]
  • 220.Ovsyannikov S.V., Bykov M., Bykova E., Kozlenko D.P., Tsirlin A.A., Karkin A.E., Shchennikov V.V., Kichanov S.E., Gou H., Abakumov A.M., et al. Charge-ordering transition in iron oxide Fe4O5 involving competing dimer and trimer formation. Nat. Chem. 2016;8:501–508. doi: 10.1038/nchem.2478. [DOI] [PubMed] [Google Scholar]
  • 221.Hong K.H., Arevalo-Lopez A.M., Cumby J., Ritter C., Attfield J.P. Long range electronic phase separation in CaFe3O5. Nat. Commun. 2018;9:2975. doi: 10.1038/s41467-018-05363-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Ovsyannikov S.V., Bykov M., Bykova E., Glazyrin K., Manna R.S., Tsirlin A.A., Cerantola V., Kupenko I., Kurnosov A.V., Kantor I., et al. Pressure tuning of charge ordering in iron oxide. Nat. Commun. 2018;9:4142. doi: 10.1038/s41467-018-06457-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Cassidy S.J., Orlandi F., Manuel P., Clarke S.J. Single phase charge ordered stoichiometric CaFe3O5 with commensurate and incommensurate trimeron ordering. Nat. Commun. 2019;10:5475. doi: 10.1038/s41467-019-13450-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Ovsyannikov S.V., Bykov M., Medvedev S.A., Naumov P.G., Jesche A., Tsirlin A.A., Bykova E., Chuvashova I., Karkin A.E., Dyadkin V., et al. A Room-Temperature Verwey-type Transition in Iron Oxide, Fe5O6. Angew. Chem. Int. Ed. 2020;59:5632–5636. doi: 10.1002/anie.201914988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Chang L., Roberts A.P., Tang Y., Rainford B.D., Muxworthy A.R., Chen Q. Fundamental magnetic parameters from pure synthetic greigite (Fe3S4) J. Geophys. Res. 2008;113 doi: 10.1029/2007JB005502. [DOI] [Google Scholar]
  • 226.Roberts A.P., Chang L., Rowan C.J., Horng C.-S., Florindo F. Magnetic properties of sedimentary greigite (Fe3S4): An update. Rev. Geophys. 2011;49 doi: 10.1029/2010RG000336. [DOI] [Google Scholar]
  • 227.Wu M., Tse J.S., Pan Y. Electronic structures of greigite (Fe3S4): A hybrid functional study and prediction for a Verwey transition. Sci. Rep. 2016;6 doi: 10.1038/srep21637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Abraitis P.K., Pattrick R.A.D., Vaughan D.J. Variations in the compositional, textural and electrical properties of natural pyrite: a review. Int. J. Miner. Process. 2004;74:41–59. doi: 10.1016/j.minpro.2003.09.002. [DOI] [Google Scholar]
  • 229.Sorenson S.A., Patrow J.G., Dawlaty J.M. Electronic Dynamics in Natural Iron Pyrite Studied by Broadband Transient Reflection Spectroscopy. J. Phys. Chem. C. 2016;120:7736–7747. doi: 10.1021/acs.jpcc.5b11036. [DOI] [Google Scholar]
  • 230.Khalid S., Ahmed E., Khan Y., Nawaz S., Ramzan M., Khalid N.R., Ahmed W. In: Jackson M.J., Ahmed W., editors. Volume II. Springer International Publishing; 2018. Iron Pyrite (FeS2): Sustainable Photovoltaic Material; pp. 281–318. (Micro and Nanomanufacturing). [DOI] [Google Scholar]
  • 231.Rahman M.Z., Edvinsson T. What Is Limiting Pyrite Solar Cell Performance? Joule. 2019;3:2290–2293. doi: 10.1016/j.joule.2019.06.015. [DOI] [Google Scholar]
  • 232.Cabán-Acevedo M., Faber M.S., Tan Y., Hamers R.J., Jin S. Synthesis and Properties of Semiconducting Iron Pyrite (FeS2) Nanowires. Nano Lett. 2012;12:1977–1982. doi: 10.1021/nl2045364. [DOI] [PubMed] [Google Scholar]
  • 233.Cabán-Acevedo M., Liang D., Chew K.S., DeGrave J.P., Kaiser N.S., Jin S. Synthesis, Characterization, and Variable Range Hopping Transport of Pyrite (FeS2) Nanorods, Nanobelts, and Nanoplates. ACS Nano. 2013;7:1731–1739. doi: 10.1021/nn305833u. [DOI] [PubMed] [Google Scholar]
  • 234.Zhang X., Manno M., Baruth A., Johnson M., Aydil E.S., Leighton C. Crossover From Nanoscopic Intergranular Hopping to Conventional Charge Transport in Pyrite Thin Films. ACS Nano. 2013;7:2781–2789. doi: 10.1021/nn4003264. [DOI] [PubMed] [Google Scholar]
  • 235.Živković A., King H.E., Wolthers M., de Leeuw N.H. Magnetic structure and exchange interactions in pyrrhotite end member minerals: hexagonal FeS and monoclinic Fe7S8. J. Phys. Condens. Matter. 2021;33 doi: 10.1088/1361-648X/ac1cb2. [DOI] [PubMed] [Google Scholar]
  • 236.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. doi: 10.1021/jacs.5b06687. [DOI] [PubMed] [Google Scholar]
  • 237.Guo Z., Sun F., Han B., Lin K., Zhou L., Yuan W. Iron vacancy in tetragonal Fe1−xS crystals and its effect on the structure and superconductivity. Phys. Chem. Chem. Phys. 2017;19:9000–9006. doi: 10.1039/C7CP00068E. [DOI] [PubMed] [Google Scholar]
  • 238.Kuhn S.J., Kidder M.K., Parker D.S., dela Cruz C., McGuire M.A., Chance W.M., Li L., Debeer-Schmitt L., Ermentrout J., Littrell K.C., et al. Structure and property correlations in FeS. Phys. C Supercond. 2017;534:29–36. doi: 10.1016/j.physc.2016.12.006. [DOI] [Google Scholar]
  • 239.Takele S., Hearne G.R. Electrical transport, magnetism, and spin-state configurations of high-pressure phases of FeS. Phys. Rev. B. 1999;60:4401–4403. doi: 10.1103/PhysRevB.60.4401. [DOI] [Google Scholar]
  • 240.Roberts D.M., Landin A.R., Ritter T.G., Eaves J.D., Stoldt C.R. Nanocrystalline Iron Monosulfides Near Stoichiometry. Sci. Rep. 2018;8:6591. doi: 10.1038/s41598-018-24739-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Sakkopoulos S., Vitoratos E., Argyreas T. Energy-band diagram for pyrrhotite. J. Phys. Chem. Solids. 1984;45:923–928. doi: 10.1016/0022-3697(84)90135-5. [DOI] [Google Scholar]
  • 242.Sagnotti L. In: Encyclopedia of Geomagnetism and Paleomagnetism. Gubbins D., Herrero-Bervera E., editors. Springer; 2007. Iron Sulfides; pp. 454–459. [DOI] [Google Scholar]
  • 243.Xu Y., Zhou Z., Deng N., Fu K., Zhu C., Hong Q., Shen Y., Liu S., Zhang Y. Molecular insights of nanozymes from design to catalytic mechanism. Sci. China Chem. 2023;66:1318–1335. doi: 10.1007/s11426-022-1529-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Wierzbinski E., Venkatramani R., Davis K.L., Bezer S., Kong J., Xing Y., Borguet E., Achim C., Beratan D.N., Waldeck D.H. The Single-Molecule Conductance and Electrochemical Electron-Transfer Rate Are Related by a Power Law. ACS Nano. 2013;7:5391–5401. doi: 10.1021/nn401321k. [DOI] [PubMed] [Google Scholar]
  • 245.Venkatramani R., Wierzbinski E., Waldeck D.H., Beratan D.N. Breaking the simple proportionality between molecular conductances and charge transfer rates. Faraday Discuss. 2014;174:57–78. doi: 10.1039/C4FD00106K. [DOI] [PubMed] [Google Scholar]
  • 246.Winkler J.R., Gray H.B. Electron flow through metalloproteins. Chem. Rev. 2014;114:3369–3380. doi: 10.1021/cr4004715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Bostick C.D., Mukhopadhyay S., Pecht I., Sheves M., Cahen D., Lederman D. Protein bioelectronics: a review of what we do and do not know. Rep. Prog. Phys. 2018;81 doi: 10.1088/1361-6633/aa85f2. [DOI] [PubMed] [Google Scholar]
  • 248.Wink D.A., Nims R.W., Saavedra J.E., Utermahlen W.E., Jr., Ford P.C. The Fenton oxidation mechanism: reactivities of biologically relevant substrates with two oxidizing intermediates differ from those predicted for the hydroxyl radical. Proc. Natl. Acad. Sci. USA. 1994;91:6604–6608. doi: 10.1073/pnas.91.14.6604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Valdez C.E., Smith Q.A., Nechay M.R., Alexandrova A.N. Mysteries of metals in metalloenzymes. Acc. Chem. Res. 2014;47:3110–3117. doi: 10.1021/ar500227u. [DOI] [PubMed] [Google Scholar]
  • 250.Yuan B., Chou H.-L., Peng Y.-K. Disclosing the Origin of Transition Metal Oxides as Peroxidase (and Catalase) Mimetics. ACS Appl. Mater. Interfaces. 2021;14:22728–22736. doi: 10.1021/acsami.1c13429. [DOI] [PubMed] [Google Scholar]
  • 251.Qiu Y., Yuan B., Mi H., Lee J.-H., Chou S.-W., Peng Y.-K. An Atomic Insight into the Confusion on the Activity of Fe3SO4 Nanoparticles as Peroxidase Mimetics and Their Comparison with Horseradish Peroxidase. J. Phys. Chem. Lett. 2022;13:8872–8878. doi: 10.1021/acs.jpclett.2c02331. [DOI] [PubMed] [Google Scholar]
  • 252.Rodríguez-López J.N., Lowe D.J., Hernández-Ruiz J., Hiner A.N., García-Cánovas F., Thorneley R.N. Mechanism of reaction of hydrogen peroxide with horseradish peroxidase: identification of intermediates in the catalytic cycle. J. Am. Chem. Soc. 2001;123:11838–11847. doi: 10.1021/ja011853+. [DOI] [PubMed] [Google Scholar]
  • 253.Veitch N.C. Horseradish peroxidase: A modern view of a classic enzyme. Phytochemistry. 2004;65:249–259. doi: 10.1016/j.phytochem.2003.10.022. [DOI] [PubMed] [Google Scholar]
  • 254.Poulos T.L., Kraut J. The stereochemistry of peroxidase catalysis. J. Biol. Chem. 1980;255:8199–8205. doi: 10.1016/S0021-9258(19)70630-9. [DOI] [PubMed] [Google Scholar]
  • 255.Sanchez C., Babonneau F., Morineau R., Livage J., Bullot J. Semiconducting properties of V2O5 gels. Philos. Mag. B. 1983;47:279–290. doi: 10.1080/13642812.1983.9728310. [DOI] [Google Scholar]
  • 256.Sanchez C., Morineau R., Livage J. Electrical conductivity of amorphous V2O5. Phys. Stat. Sol. 1983;76:661–666. doi: 10.1002/pssa.2210760232. [DOI] [Google Scholar]
  • 257.André R., Natálio F., Humanes M., Leppin J., Heinze K., Wever R., Schröder H., Müller W.E.G., Tremel W. V2O5 nanowires with an intrinsic peroxidase-like activity. Adv. Funct. Mater. 2011;21:501–509. doi: 10.1002/adfm.201001302. [DOI] [Google Scholar]
  • 258.Vernekar A.A., Sinha D., Srivastava S., Paramasivam P.U., D'Silva P., Mugesh G. An antioxidant nanozyme that uncovers the cytoprotective potential of vanadia nanowires. Nat. Commun. 2014;5:5301. doi: 10.1038/ncomms6301. [DOI] [PubMed] [Google Scholar]
  • 259.Ding Y., Ren G., Wang G., Lu M., Liu J., Li K., Lin Y. V2O5 Nanobelts Mimick Tandem Enzymes To Achieve Nonenzymatic Online Monitoring of Glucose in Living Rat Brain. Anal. Chem. 2020;92:4583–4591. doi: 10.1021/acs.analchem.9b05872. [DOI] [PubMed] [Google Scholar]
  • 260.Zeng X., Wang H., Ma Y., Xu X., Lu X., Hu Y., Xie J., Wang X., Wang Y., Guo X., et al. Vanadium Oxide Nanozymes with Multiple Enzyme-Mimic Activities for Tumor Catalytic Therapy. ACS Appl. Mater. Interfaces. 2023;15:13941–13955. doi: 10.1021/acsami.2c20878. [DOI] [PubMed] [Google Scholar]
  • 261.DORRIS S.E., MASON T.O. Electrical Properties and Cation Valencies in Mn3O4. J. Am. Ceram. Soc. 1988;71:379–385. doi: 10.1111/j.1151-2916.1988.tb05057.x. [DOI] [Google Scholar]
  • 262.Singh N., Geethika M., Eswarappa S.M., Mugesh G. Manganese-Based Nanozymes: Multienzyme Redox Activity and Effect on the Nitric Oxide Produced by Endothelial Nitric Oxide Synthase. Chemistry. 2018;24:8393–8403. doi: 10.1002/chem.201800770. [DOI] [PubMed] [Google Scholar]
  • 263.Devaraj S., Munichandraiah N. Effect of Crystallographic Structure of MnO2 on Its Electrochemical Capacitance Properties. J. Phys. Chem. C. 2008;112:4406–4417. doi: 10.1021/jp7108785. [DOI] [Google Scholar]
  • 264.Farooq M.U., Muhammad Z., Khalid S., Fatima K., Zou B. Magnetic coupling in 3D-hierarchical MnO2 microsphere. J. Mater. Sci. Mater. Electron. 2019;30:2802–2808. doi: 10.1007/s10854-018-0556-1. [DOI] [Google Scholar]
  • 265.Huang Y., Liu Z., Liu C., Ju E., Zhang Y., Ren J., Qu X. Self-Assembly of Multi-nanozymes to Mimic an Intracellular Antioxidant Defense System. Angew. Chem. Int. Ed. 2016;55:6646–6650. doi: 10.1002/anie.201600868. [DOI] [PubMed] [Google Scholar]
  • 266.Tang M., Zhang Z., Sun T., Li B., Wu Z. Manganese-Based Nanozymes: Preparation, Catalytic Mechanisms, and Biomedical Applications. Adv. Healthc. Mater. 2022;11 doi: 10.1002/adhm.202201733. [DOI] [PubMed] [Google Scholar]
  • 267.Cheng C.-S., Serizawa M., Sakata H., Hirayama T. Electrical conductivity of Co3O4 films prepared by chemical vapour deposition. Mater. Chem. Phys. 1998;53:225–230. doi: 10.1016/S0254-0584(98)00044-3. [DOI] [Google Scholar]
  • 268.Pham H.H., Cheng M.-J., Frei H., Wang L.-W. Surface Proton Hopping and Fast-Kinetics Pathway of Water Oxidation on Co3O4 (001) Surface. ACS Catal. 2016;6:5610–5617. doi: 10.1021/acscatal.6b00713. [DOI] [Google Scholar]
  • 269.Mu J., Wang Y., Zhao M., Zhang L. Intrinsic peroxidase-like activity and catalase-like activity of Co3O4 nanoparticles. Chem. Commun. 2012;48:2540–2542. doi: 10.1039/c2cc17013b. [DOI] [PubMed] [Google Scholar]
  • 270.Wang Q., Chen J., Zhang H., Wu W., Zhang Z., Dong S. Porous Co3O4 nanoplates with pH-switchable peroxidase- and catalase-like activity. Nanoscale. 2018;10:19140–19146. doi: 10.1039/C8NR06162A. [DOI] [PubMed] [Google Scholar]
  • 271.Li W., Wang J., Zhu J., Zheng Y.-Q. Co3O4 nanocrystals as an efficient catalase mimic for the colorimetric detection of glutathione. J. Mater. Chem. B. 2018;6:6858–6864. doi: 10.1039/C8TB01948G. [DOI] [PubMed] [Google Scholar]
  • 272.Setvin M., Franchini C., Hao X., Schmid M., Janotti A., Kaltak M., Van De Walle C.G., Kresse G., Diebold U. Direct view at excess electrons in TiO2 rutile and anatase. Phys. Rev. Lett. 2014;113 doi: 10.1103/PhysRevLett.113.086402. [DOI] [PubMed] [Google Scholar]
  • 273.Xu Z.F., Tong C.J., Si R.T., Teobaldi G., Liu L.M. Nonadiabatic Dynamics of Polaron Hopping and Coupling with Water on Reduced TiO2. J. Phys. Chem. Lett. 2022;13:857–863. doi: 10.1021/acs.jpclett.1c04231. [DOI] [PubMed] [Google Scholar]
  • 274.Fattakhova-Rohlfing D., Zaleska A., Bein T. Three-dimensional titanium dioxide nanomaterials. Chem. Rev. 2014;114:9487–9558. doi: 10.1021/cr500201c. [DOI] [PubMed] [Google Scholar]
  • 275.Zhang L., Han L., Hu P., Wang L., Dong S. TiO2 nanotube arrays: Intrinsic peroxidase mimetics. Chem. Commun. 2013;49:10480–10482. doi: 10.1039/c3cc46163g. [DOI] [PubMed] [Google Scholar]
  • 276.Zhu Y., Cao H., Tang L., Yang X., Li C. Immobilization of horseradish peroxidase in three-dimensional macroporous TiO2 matrices for biosensor applications. Electrochim. Acta. 2009;54:2823–2827. doi: 10.1016/j.electacta.2008.11.025. [DOI] [Google Scholar]
  • 277.Tuller H.L., Nowick A.S. Small polaron electron transport in reduced CeO2 single crystals. J. Phys. Chem. Solids. 1977;38:859–867. doi: 10.1016/0022-3697(77)90124-X. [DOI] [Google Scholar]
  • 278.Kim S., Maier J. On the Conductivity Mechanism of Nanocrystalline Ceria. J. Electrochem. Soc. 2002;149 doi: 10.1149/1.1507597. [DOI] [Google Scholar]
  • 279.Esch F., Fabris S., Zhou L., Montini T., Africh C., Fornasiero P., Comelli G., Rosei R. Electron Localization Determines Defect Formation on Ceria Substrates. Science. 2005;309:752–755. doi: 10.1126/science.1111568. [DOI] [PubMed] [Google Scholar]
  • 280.Montini T., Melchionna M., Monai M., Fornasiero P. Fundamentals and Catalytic Applications of CeO2-Based Materials. Chem. Rev. 2016;116:5987–6041. doi: 10.1021/acs.chemrev.5b00603. [DOI] [PubMed] [Google Scholar]
  • 281.Xiao G., Li H., Zhao Y., Wei H., Li J., Su H. Nanoceria-Based Artificial Nanozymes: Review of Materials and Applications. ACS Appl. Nano Mater. 2022;5:14147–14170. doi: 10.1021/acsanm.2c03009. [DOI] [Google Scholar]
  • 282.Ma Y., Tian Z., Zhai W., Qu Y. Insights on catalytic mechanism of CeO2 as multiple nanozymes. Nano Res. 2022;15:10328–10342. doi: 10.1007/s12274-022-4666-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Tan Z., Chen Y.-C., Zhang J., Chou J.-P., Hu A., Peng Y.-K. Nanoisozymes: The Origin behind Pristine CeO2 as Enzyme Mimetics. Chem. Eur J. 2020;26:10598–10606. doi: 10.1002/chem.202001597. [DOI] [PubMed] [Google Scholar]
  • 284.Dhall A., Burns A., Dowding J., Das S., Seal S., Self W. Characterizing the phosphatase mimetic activity of cerium oxide nanoparticles and distinguishing its active site from that for catalase mimetic activity using anionic inhibitors. Environ. Sci. Nano. 2017;4:1742–1749. doi: 10.1039/C7EN00394C. [DOI] [Google Scholar]
  • 285.Tan Z., Wu T.-S., Soo Y.-L., Peng Y.-K. Unravelling the true active site for CeO2-catalyzed dephosphorylation. Appl. Catal. B. 2020;264 doi: 10.1016/j.apcatb.2019.118508. [DOI] [Google Scholar]
  • 286.Ito T., Takagi H., Ishibashi S., Ido T., Uchida S. Normal-state conductivity between CuO2 planes in copper oxide superconductors. Nature. 1991;350:596–598. doi: 10.1038/350596a0. [DOI] [Google Scholar]
  • 287.Sidorov N.S., Palnichenko A.V., Khasanov S.S. Superconductivity in the metallic-oxidized copper interface. Phys. C: Supercond. Appl. 2011;471:247–249. doi: 10.1016/j.physc.2011.02.006. [DOI] [Google Scholar]
  • 288.Liu T., Xiao B., Xiang F., Tan J., Chen Z., Zhang X., Wu C., Mao Z., Luo G., Chen X., Deng J. Ultrasmall copper-based nanoparticles for reactive oxygen species scavenging and alleviation of inflammation related diseases. Nat. Commun. 2020;11:2788. doi: 10.1038/s41467-020-16544-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Zhu Y., Zhang Z., Song X., Bu Y. A facile strategy for synthesis of porous Cu2O nanospheres and application as nanozymes in colorimetric biosensing. J. Mater. Chem. B. 2021;9:3533–3543. doi: 10.1039/d0tb03005h. [DOI] [PubMed] [Google Scholar]
  • 290.Radisavljevic B., Radenovic A., Brivio J., Giacometti V., Kis A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011;6:147–150. doi: 10.1038/nnano.2010.279. [DOI] [PubMed] [Google Scholar]
  • 291.Qiu H., Xu T., Wang Z., Ren W., Nan H., Ni Z., Chen Q., Yuan S., Miao F., Song F., et al. Hopping transport through defect-induced localized states in molybdenum disulphide. Nat. Commun. 2013;4:2642. doi: 10.1038/ncomms3642. [DOI] [PubMed] [Google Scholar]
  • 292.Chen T., Zou H., Wu X., Liu C., Situ B., Zheng L., Yang G. Nanozymatic Antioxidant System Based on MoS2 Nanosheets. ACS Appl. Mater. Interfaces. 2018;10:12453–12462. doi: 10.1021/acsami.8b01245. [DOI] [PubMed] [Google Scholar]
  • 293.Yu Y., Lu L., Yang Q., Zupanic A., Xu Q., Jiang L. Using MoS2 Nanomaterials to Generate or Remove Reactive Oxygen Species: A Review. ACS Appl. Nano Mater. 2021;4:7523–7537. doi: 10.1021/acsanm.1c00751. [DOI] [Google Scholar]
  • 294.Dutta A.K., Maji S.K., Mondal A., Karmakar B., Biswas P., Adhikary B. Iron selenide thin film: Peroxidase-like behavior, glucose detection and amperometric sensing of hydrogen peroxide. Sens. Actuators B Chem. 2012;173:724–731. doi: 10.1016/j.snb.2012.07.070. [DOI] [Google Scholar]
  • 295.Qiao F., Chen L., Li X., Li L., Ai S. Peroxidase-like activity of manganese selenide nanoparticles and its analytical application for visual detection of hydrogen peroxide and glucose. Sens. Actuators B Chem. 2014;193:255–262. doi: 10.1016/j.snb.2013.11.108. [DOI] [Google Scholar]
  • 296.Wu X., Chen T., Wang J., Yang G. Few-layered MoSe2 nanosheets as an efficient peroxidase nanozyme for highly sensitive colorimetric detection of H2O2 and xanthine. J. Mater. Chem. B. 2018;6:105–111. doi: 10.1039/c7tb02434g. [DOI] [PubMed] [Google Scholar]
  • 297.Liu K., Ma X., Xu S., Li Y., Zhao M. Tunable sliding ferroelectricity and magnetoelectric coupling in two-dimensional multiferroic MnSe materials. npj Comput. Mater. 2023;9:16. doi: 10.1038/s41524-023-00972-2. [DOI] [Google Scholar]
  • 298.Suri D., Patel R.S. Electron and thermal transport via variable range hopping in MoSe2 single crystals. Appl. Phys. Lett. 2017;110 doi: 10.1063/1.4984953. [DOI] [Google Scholar]
  • 299.Hsu F.-C., Luo J.-Y., Yeh K.-W., Chen T.-K., Huang T.-W., Wu P.M., Lee Y.-C., Huang Y.-L., Chu Y.-Y., Yan D.-C., Wu M.-K. Superconductivity in the PbO-type structure α-FeSe. Proc. Natl. Acad. Sci. USA. 2008;105:14262–14264. doi: 10.1073/pnas.0807325105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Liu X., Liu D., Zhang W., He J., Zhao L., He S., Mou D., Li F., Tang C., Li Z., et al. Dichotomy of the electronic structure and superconductivity between single-layer and double-layer FeSe/SrTiO3 films. Nat. Commun. 2014;5:5047. doi: 10.1038/ncomms6047. [DOI] [PubMed] [Google Scholar]
  • 301.Xia Y., Yin L. Core–shell structured α-Fe2O3@TiO2 nanocomposites with improved photocatalytic activity in the visible light region. Phys. Chem. Chem. Phys. 2013;15:18627–18634. doi: 10.1039/C3CP53178C. [DOI] [PubMed] [Google Scholar]
  • 302.Boruah P.K., Das M.R. Dual responsive magnetic Fe3O4-TiO2/graphene nanocomposite as an artificial nanozyme for the colorimetric detection and photodegradation of pesticide in an aqueous medium. J. Hazard Mater. 2020;385 doi: 10.1016/j.jhazmat.2019.121516. [DOI] [PubMed] [Google Scholar]
  • 303.Lindqvist Y., Johansson E., Kaija H., Vihko P., Schneider G. Three-dimensional structure of a mammalian purple acid phosphatase at 2.2 Å resolution with a μ-(hydr)oxo bridged di-iron center. J. Mol. Biol. 1999;291:135–147. doi: 10.1006/jmbi.1999.2962. [DOI] [PubMed] [Google Scholar]
  • 304.Schenk G., Mitić N., Hanson G.R., Comba P. Purple acid phosphatase: A journey into the function and mechanism of a colorful enzyme. Coord. Chem. Rev. 2013;257:473–482. doi: 10.1016/j.ccr.2012.03.020. [DOI] [Google Scholar]
  • 305.Mitić N., Smith S.J., Neves A., Guddat L.W., Gahan L.R., Schenk G. The catalytic mechanisms of binuclear metallohydrolases. Chem. Rev. 2006;106:3338–3363. doi: 10.1021/cr050318f. [DOI] [PubMed] [Google Scholar]
  • 306.Hemrika W., Renirie R., Dekker H.L., Barnett P., Wever R. From phosphatases to vanadium peroxidases: A similar architecture of the active site. Proc. Natl. Acad. Sci. USA. 1997;94:2145–2149. doi: 10.1073/pnas.94.6.2145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Plass W., Bangesh M., Nica S., Buchholz A. In: ACS Symposium Series. Kustin K., Pessoa J.C., Crans D.C., editors. 2007. Model studies of vanadium-dependent haloperoxidation: Structural and functional lessons synthetic and computational models of supramolecular interactions and the formation of peroxo species. [Google Scholar]
  • 308.Leblanc C., Vilter H., Fournier J.B., Delage L., Potin P., Rebuffet E., Michel G., Solari P.L., Feiters M.C., Czjzek M. Vanadium haloperoxidases: From the discovery 30 years ago to X-ray crystallographic and V K-edge absorption spectroscopic studies. Coord. Chem. Rev. 2015;301–302:134–146. doi: 10.1016/j.ccr.2015.02.013. [DOI] [Google Scholar]
  • 309.Langeslay R.R., Kaphan D.M., Marshall C.L., Stair P.C., Sattelberger A.P., Delferro M. Catalytic Applications of Vanadium: A Mechanistic Perspective. Chem. Rev. 2019;119:2128–2191. doi: 10.1021/acs.chemrev.8b00245. [DOI] [PubMed] [Google Scholar]
  • 310.Tompkins F.C. Superficial Chemistry and Solid Imperfections. Nature. 1960;186:3–6. doi: 10.1038/186003a0. [DOI] [Google Scholar]
  • 311.Wang C.M., Baer D.R., Thomas L.E., Amonette J.E., Antony J., Qiang Y., Duscher G. Void formation during early stages of passivation: Initial oxidation of iron nanoparticles at room temperature. J. Appl. Phys. 2005;094308 doi: 10.1063/1.2130890. [DOI] [Google Scholar]
  • 312.Ganduglia-Pirovano M.V., Hofmann A., Sauer J. Oxygen vacancies in transition metal and rare earth oxides: Current state of understanding and remaining challenges. Surf. Sci. Rep. 2007;62:219–270. doi: 10.1016/j.surfrep.2007.03.002. [DOI] [Google Scholar]
  • 313.Cabot A., Puntes V.F., Shevchenko E., Yin Y., Balcells L., Marcus M.A., Hughes S.M., Alivisatos A.P. Vacancy coalescence during oxidation of iron nanoparticles. J. Am. Chem. Soc. 2007;129:10358–10360. doi: 10.1021/ja072574a. [DOI] [PubMed] [Google Scholar]
  • 314.Parkinson G.S. Iron oxide surfaces. Surf. Sci. Rep. 2016;71:272–365. doi: 10.1016/j.surfrep.2016.02.001. [DOI] [Google Scholar]
  • 315.Gunkel F., Christensen D.V., Chen Y.Z., Pryds N. Oxygen vacancies: The (in)visible friend of oxide electronics. Appl. Phys. Lett. 2020;116,12050 doi: 10.1063/1.5143309. [DOI] [Google Scholar]
  • 316.Zhuang G., Chen Y., Zhuang Z., Yu Y., Yu J. Oxygen vacancies in metal oxides: recent progress towards advanced catalyst design. Sci. China Mater. 2020;63:2089–2118. doi: 10.1007/s40843-020-1305-6. [DOI] [Google Scholar]
  • 317.Giese B. In: Encyclopedia of Biophysics. Roberts G., Watts A., editors. Springer Berlin Heidelberg; 2018. Electron Hopping in Biomolecules; pp. 1–3. [DOI] [Google Scholar]
  • 318.Mostajabi Sarhangi S., Matyushov D.V. Electron Tunneling in Biology: When Does it Matter? ACS Omega. 2023;8:27355–27365. doi: 10.1021/acsomega.3c02719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Miller S.L. A Production of Amino Acids Under Possible Primitive Earth Conditions. Science. 1953;117:528–529. doi: 10.1126/science.117.3046.528. [DOI] [PubMed] [Google Scholar]
  • 320.Marshall W.L. Hydrothermal synthesis of amino acids. Geochim. Cosmochim. Acta. 1994;58:2099–2106. doi: 10.1016/0016-7037(94)90288-7. [DOI] [Google Scholar]
  • 321.Levy M., Miller S.L., Brinton K., Bada J.L. Prebiotic Synthesis of Adenine and Amino Acids Under Europa-like Conditions. Icarus. 2000;145:609–613. doi: 10.1006/icar.2000.6365. [DOI] [PubMed] [Google Scholar]
  • 322.OrÓ J. Mechanism of Synthesis of Adenine from Hydrogen Cyanide under Possible Primitive Earth Conditions. Nature. 1961;191:1193–1194. doi: 10.1038/1911193a0. [DOI] [PubMed] [Google Scholar]
  • 323.Ferus M., Pietrucci F., Saitta A.M., Knížek A., Kubelík P., Ivanek O., Shestivska V., Civiš S. Formation of nucleobases in a Miller–Urey reducing atmosphere. Proc. Natl. Acad. Sci. USA. 2017;114:4306–4311. doi: 10.1073/pnas.1700010114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Powner M.W., Gerland B., Sutherland J.D. Synthesis of activated pyrimidine ribonucleotides in prebiotically plausible conditions. Nature. 2009;459:239–242. doi: 10.1038/nature08013. [DOI] [PubMed] [Google Scholar]
  • 325.Yadav M., Kumar R., Krishnamurthy R. Chemistry of Abiotic Nucleotide Synthesis. Chem. Rev. 2020;120:4766–4805. doi: 10.1021/acs.chemrev.9b00546. [DOI] [PubMed] [Google Scholar]
  • 326.Schwartz A.W., de Graaf R.M. The prebiotic synthesis of carbohydrates: A reassessment. J. Mol. Evol. 1993;36:101–106. doi: 10.1007/BF00166245. [DOI] [Google Scholar]
  • 327.Furukawa Y., Chikaraishi Y., Ohkouchi N., Ogawa N.O., Glavin D.P., Dworkin J.P., Abe C., Nakamura T. Extraterrestrial ribose and other sugars in primitive meteorites. Proc. Natl. Acad. Sci. USA. 2019;116:24440–24445. doi: 10.1073/pnas.1907169116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Deamer D.W., Pashley R.M. Amphiphilic components of the murchison carbonaceous chondrite: Surface properties and membrane formation. Orig. Life Evol. Biosph. 1989;19:21–38. doi: 10.1007/BF01808285. [DOI] [PubMed] [Google Scholar]
  • 329.Fayolle D., Altamura E., D’Onofrio A., Madanamothoo W., Fenet B., Mavelli F., Buchet R., Stano P., Fiore M., Strazewski P. Crude phosphorylation mixtures containing racemic lipid amphiphiles self-assemble to give stable primitive compartments. Sci. Rep. 2017;7 doi: 10.1038/s41598-017-18053-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Frenkel-Pinter M., Bouza M., Fernández F.M., Leman L.J., Williams L.D., Hud N.V., Guzman-Martinez A. Thioesters provide a plausible prebiotic path to proto-peptides. Nat. Commun. 2022;13:2569. doi: 10.1038/s41467-022-30191-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Halpern J., Orgel L.E. The theory of electron transfer between metal ions in bridged systems. Discuss. Faraday Soc. 1960;29:32–41. doi: 10.1039/DF9602900032. [DOI] [Google Scholar]
  • 332.Hopfield J.J. Electron transfer between biological molecules by thermally activated tunneling. Proc. Natl. Acad. Sci. USA. 1974;71:3640–3644. doi: 10.1073/pnas.71.9.3640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Marcus R.A., Sutin N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta. 1985;811:265–322. doi: 10.1016/0304-4173(85)90014-X. [DOI] [Google Scholar]
  • 334.Warren J.J., Ener M.E., Vlček A., Jr., Winkler J.R., Gray H.B. Electron hopping through proteins. Coord. Chem. Rev. 2012;256:2478–2487. doi: 10.1016/j.ccr.2012.03.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Gray H.B., Winkler J.R. Hole hopping through tyrosine/tryptophan chains protects proteins from oxidative damage. Proc. Natl. Acad. Sci. USA. 2015;112:10920–10925. doi: 10.1073/pnas.1512704112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Gray H.B., Winkler J.R. Functional and protective hole hopping in metalloenzymes. Chem. Sci. 2021;12:13988–14003. doi: 10.1039/D1SC04286F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Berglund G.I., Carlsson G.H., Smith A.T., Szöke H., Henriksen A., Hajdu J. The catalytic pathway of horseradish peroxidase at high resolution. Nature. 2002;417:463–468. doi: 10.1038/417463a. [DOI] [PubMed] [Google Scholar]
  • 338.Uzun M. Nova Science Publishers; 2019. Horseradish Peroxidase: Structure, Functions and Applications; p. 186. [Google Scholar]
  • 339.Lah M.S., Dixon M.M., Pattridge K.A., Stallings W.C., Fee J.A., Ludwig M.L. Structure-Function in Escherichia coli Iron Superoxide Dismutase: Comparisons with the Manganese Enzyme from Thermus thermophilus. Biochemistry. 1995;34:1646–1660. doi: 10.1021/bi00005a021. [DOI] [PubMed] [Google Scholar]
  • 340.Loewen P.C., Carpena X., Rovira C., Ivancich A., Perez-Luque R., Haas R., Odenbreit S., Nicholls P., Fita I. Structure of Helicobacter pylori Catalase, with and Without Formic Acid Bound, at 1.6 Å Resolution. Biochemistry. 2004;43:3089–3103. doi: 10.1021/bi035663i. [DOI] [PubMed] [Google Scholar]
  • 341.Sheng Y., Abreu I.A., Cabelli D.E., Maroney M.J., Miller A.F., Teixeira M., Valentine J.S. Superoxide dismutases and superoxide reductases. Chem. Rev. 2014;114:3854–3918. doi: 10.1021/cr4005296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Shaik S., Kumar D., de Visser S.P., Altun A., Thiel W. Theoretical perspective on the structure and mechanism of cytochrome P450 enzymes. Chem. Rev. 2005;105:2279–2328. doi: 10.1021/cr030722j. [DOI] [PubMed] [Google Scholar]
  • 343.Groves, J.T. (2015). Cytochrome P450 Enzymes: Understanding the Biochemical Hieroglyphs. [version 1; peer review: 3 approved]. F1000Research 2015, 4(F1000 Faculty Rev):178 10.12688/f1000research.6314.1. [DOI]
  • 344.Bent D.V., Hayon E. Excited state chemistry of aromatic amino acids and related peptides. III. Tryptophan. J. Am. Chem. Soc. 1975;97:2612–2619. doi: 10.1021/ja00843a004. [DOI] [PubMed] [Google Scholar]
  • 345.Bent D.V., Hayon E. Excited State Chemistry of Aromatic Amino Acids and Related Peptides. I. Tyrosine. J. Am. Chem. Soc. 1975;97:2599–2606. doi: 10.1021/ja00843a002. [DOI] [PubMed] [Google Scholar]
  • 346.Sivaraja M., Goodin D.B., Smith M., Hoffman B.M. Identification by ENDOR of Trp191 as the free-radical site in cytochrome c peroxidase compound ES. Science. 1989;245:738–740. doi: 10.1126/science.2549632. [DOI] [PubMed] [Google Scholar]
  • 347.Fan K., Wang H., Xi J., Liu Q., Meng X., Duan D., Gao L., Yan X. Optimization of Fe3O4 nanozyme activity via single amino acid modification mimicking an enzyme active site. Chem. Commun. 2016;53:424–427. doi: 10.1039/c6cc08542c. [DOI] [PubMed] [Google Scholar]
  • 348.Liang H., Lin F., Zhang Z., Liu B., Jiang S., Yuan Q., Liu J. Multicopper laccase mimicking nanozymes with nucleotides as ligands. ACS Appl. Mater. Interfaces. 2017;9:1352–1360. doi: 10.1021/acsami.6b15124. [DOI] [PubMed] [Google Scholar]
  • 349.Niu X., Xu X., Li X., Pan J., Qiu F., Zhao H., Lan M. Surface charge engineering of nanosized CuS via acidic amino acid modification enables high peroxidase-mimicking activity at neutral pH for one-pot detection of glucose. Chem. Commun. 2018;54:13443–13446. doi: 10.1039/C8CC07800A. [DOI] [PubMed] [Google Scholar]
  • 350.Zhang Y., Wang F., Liu C., Wang Z., Kang L., Huang Y., Dong K., Ren J., Qu X. Nanozyme Decorated Metal-Organic Frameworks for Enhanced Photodynamic Therapy. ACS Nano. 2018;12:651–661. doi: 10.1021/acsnano.7b07746. [DOI] [PubMed] [Google Scholar]
  • 351.Wu W., Wang Q., Chen J., Huang L., Zhang H., Rong K., Dong S. Biomimetic design for enhancing the peroxidase mimicking activity of hemin. Nanoscale. 2019;11:12603–12609. doi: 10.1039/C9NR03506K. [DOI] [PubMed] [Google Scholar]
  • 352.Vallabani N.V.S., Vinu A., Singh S., Karakoti A. Tuning the ATP-triggered pro-oxidant activity of iron oxide-based nanozyme towards an efficient antibacterial strategy. J. Colloid Interface Sci. 2020;567:154–164. doi: 10.1016/j.jcis.2020.01.099. [DOI] [PubMed] [Google Scholar]
  • 353.Han J., Zou Q., Su W., Yan X. Minimal metallo-nanozymes constructed through amino acid coordinated self-assembly for hydrolase-like catalysis. Chem. Eng. J. 2020;394 doi: 10.1016/j.cej.2020.124987. [DOI] [Google Scholar]
  • 354.Niu X., Li X., Lyu Z., Pan J., Ding S., Ruan X., Zhu W., Du D., Lin Y. Metal–organic framework based nanozymes: promising materials for biochemical analysis. Chem. Commun. 2020;56:11338–11353. doi: 10.1039/D0CC04890A. [DOI] [PubMed] [Google Scholar]
  • 355.Chen J., Ma Q., Li M., Wu W., Huang L., Liu L., Fang Y., Dong S. Coenzyme-dependent nanozymes playing dual roles in oxidase and reductase mimics with enhanced electron transport. Nanoscale. 2020;12:23578–23585. doi: 10.1039/D0NR06605B. [DOI] [PubMed] [Google Scholar]
  • 356.Wang D., Jana D., Zhao Y. Metal-Organic Framework Derived Nanozymes in Biomedicine. Acc. Chem. Res. 2020;53:1389–1400. doi: 10.1021/acs.accounts.0c00268. [DOI] [PubMed] [Google Scholar]
  • 357.Han J., Gong H., Ren X., Yan X. Supramolecular nanozymes based on peptide self-assembly for biomimetic catalysis. Nano Today. 2021;41 doi: 10.1016/j.nantod.2021.101295. [DOI] [Google Scholar]
  • 358.Geng R., Chang R., Zou Q., Shen G., Jiao T., Yan X. Biomimetic Nanozymes Based on Coassembly of Amino Acid and Hemin for Catalytic Oxidation and Sensing of Biomolecules. Small. 2021;17 doi: 10.1002/smll.202008114. [DOI] [PubMed] [Google Scholar]
  • 359.Li G., Ma W., Yang Y., Zhong C., Huang H., Ouyang D., He Y., Tian W., Lin J., Lin Z. Nanoscale Covalent Organic Frameworks with Donor-Acceptor Structures as Highly Efficient Light-Responsive Oxidase-like Mimics for Colorimetric Detection of Glutathione. ACS Appl. Mater. Interfaces. 2021;13:49482–49489. doi: 10.1021/acsami.1c13997. [DOI] [PubMed] [Google Scholar]
  • 360.Sun D., Liu T., Dong Y., Liu Q., Wu X., Wang G.-L. Coupling p-Hydroxybenzoate Hydroaxylase with the Photoresponsive Nanozyme for Universal Dehydrogenase-Based Bioassays. Sens. Actuators B Chem. 2021;327 doi: 10.1016/j.snb.2020.128859. [DOI] [Google Scholar]
  • 361.Xu W., He W., Du Z., Zhu L., Huang K., Lu Y., Luo Y. Functional Nucleic Acid Nanomaterials: Development, Properties, and Applications. Angew. Chem. Int. Ed. 2021;60:6890–6918. doi: 10.1002/anie.201909927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Castillo-Caceres C., Duran-Meza E., Nova E., Araya-Secchi R., Monasterio O., Diaz-Espinoza R. Functional characterization of the ATPase-like activity displayed by a catalytic amyloid. Biochim. Biophys. Acta. Gen. Subj. 2021;1865 doi: 10.1016/j.bbagen.2020.129729. [DOI] [PubMed] [Google Scholar]
  • 363.Makam P., Yamijala S.S.R.K.C., Bhadram V.S., Shimon L.J.W., Wong B.M., Gazit E. Single amino acid bionanozyme for environmental remediation. Nat. Commun. 2022;13:1505. doi: 10.1038/s41467-022-28942-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 364.Li S., Zhou Z., Tie Z., Wang B., Ye M., Du L., Cui R., Liu W., Wan C., Liu Q., et al. Data-informed discovery of hydrolytic nanozymes. Nat. Commun. 2022;13:827. doi: 10.1038/s41467-022-28344-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Han X., Li Y., Zhou Y., Song Z., Deng Y., Qin J., Jiang Z. Metal-organic frameworks-derived bimetallic nanozyme platform enhances cytotoxic effect of photodynamic therapy in hypoxic cancer cells. Mater. Des. 2021;204 doi: 10.1016/j.matdes.2021.109646. [DOI] [Google Scholar]
  • 366.Huang L., Sun D.W., Pu H. Photosensitized Peroxidase Mimicry at the Hierarchical 0D/2D Heterojunction-Like Quasi Metal-Organic Framework Interface for Boosting Biocatalytic Disinfection. Small. 2022;18 doi: 10.1002/smll.202200178. [DOI] [PubMed] [Google Scholar]
  • 367.Wang S., Wang X., Du B., Jin Y., Ai W., Zhang G., Zhou T., Wang F., Zhang Z. Hydrogen peroxide-assisted and histidine-stabilized copper-containing nanozyme for efficient degradation of various organic dyes. Spectrochim. Acta Mol. Biomol. Spectrosc. 2023;287 doi: 10.1016/j.saa.2022.122084. [DOI] [PubMed] [Google Scholar]
  • 368.de Castro I.A., Datta R.S., Ou J.Z., Castellanos-Gomez A., Sriram S., Daeneke T., Kalantar-zadeh K. Molybdenum Oxides – From Fundamentals to Functionality. Adv. Mater. 2017;29 doi: 10.1002/adma.201701619. [DOI] [PubMed] [Google Scholar]
  • 369.Ragg R., Natalio F., Tahir M.N., Janssen H., Kashyap A., Strand D., Strand S., Tremel W. Molybdenum trioxide nanoparticles with intrinsic sulfite oxidase activity. ACS Nano. 2014;8:5182–5189. doi: 10.1021/nn501235j. [DOI] [PubMed] [Google Scholar]
  • 370.Dixon D.A., Lipscomb W.N. Electronic structure and bonding of the amino acids containing first row atoms. J. Biol. Chem. 1976;251:5992–6000. doi: 10.1016/S0021-9258(17)33049-1. [DOI] [PubMed] [Google Scholar]
  • 371.Dwyer D.S. Nearest-neighbor effects and structural preferences in dipeptides are a function of the electronic properties of amino acid side-chains. Proteins. 2006;63:939–948. doi: 10.1002/prot.20906. [DOI] [PubMed] [Google Scholar]
  • 372.Berstis L., Beckham G.T., Crowley M.F. Electronic coupling through natural amino acids. J. Chem. Phys. 2015;143 doi: 10.1063/1.4936588. [DOI] [PubMed] [Google Scholar]
  • 373.Sepunaru L., Refaely-Abramson S., Lovrinčić R., Gavrilov Y., Agrawal P., Levy Y., Kronik L., Pecht I., Sheves M., Cahen D. Electronic Transport via Homopeptides: The Role of Side Chains and Secondary Structure. J. Am. Chem. Soc. 2015;137:9617–9626. doi: 10.1021/jacs.5b03933. [DOI] [PubMed] [Google Scholar]
  • 374.Grishin N.V. Fold change in evolution of protein structures. J. Struct. Biol. 2001;134:167–185. doi: 10.1006/jsbi.2001.4335. [DOI] [PubMed] [Google Scholar]
  • 375.Raanan H., Poudel S., Pike D.H., Nanda V., Falkowski P.G. Small protein folds at the root of an ancient metabolic network. Proc. Natl. Acad. Sci. USA. 2020;117:7193–7199. doi: 10.1073/pnas.1914982117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Holm R.H., Kennepohl P., Solomon E.I. Structural and functional aspects of metal sites in biology. Chem. Rev. 1996;96:2239–2314. doi: 10.1021/cr9500390. [DOI] [PubMed] [Google Scholar]
  • 377.Drennan C.L., Peters J.W. Surprising cofactors in metalloenzymes. Curr. Opin. Struct. Biol. 2003;13:220–226. doi: 10.1016/s0959-440x(03)00038-1. [DOI] [PubMed] [Google Scholar]
  • 378.Milner-White E.J. Protein three-dimensional structures at the origin of life. Interface Focus. 2019;9 doi: 10.1098/rsfs.2019.0057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.Waldron K.J., Rutherford J.C., Ford D., Robinson N.J. Metalloproteins and metal sensing. Nature. 2009;460:823–830. doi: 10.1038/nature08300. [DOI] [PubMed] [Google Scholar]
  • 380.Hong Enriquez R.P., Do T.N. Bioavailability of metal ions and evolutionary adaptation. Life. 2012;2:274–285. doi: 10.3390/life2040274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Dudev T., Lim C. Competition among metal ions for protein binding sites: Determinants of metal ion selectivity in proteins. Chem. Rev. 2014;114:538–556. doi: 10.1021/cr4004665. [DOI] [PubMed] [Google Scholar]
  • 382.Moore E.K., Jelen B.I., Giovannelli D., Raanan H., Falkowski P.G. Metal availability and the expanding network of microbial metabolisms in the Archaean eon. Nat. Geosci. 2017;10:629–636. doi: 10.1038/ngeo3006. [DOI] [Google Scholar]
  • 383.Eom H., Song W.J. Emergence of metal selectivity and promiscuity in metalloenzymes. J. Biol. Inorg. Chem. 2019;24:517–531. doi: 10.1007/s00775-019-01667-0. [DOI] [PubMed] [Google Scholar]
  • 384.Wade J., Byrne D.J., Ballentine C.J., Drakesmith H. Temporal variation of planetary iron as a driver of evolution. Proc. Natl. Acad. Sci. USA. 2021;118 doi: 10.1073/pnas.2109865118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Aulakh S.K., Varma S.J., Ralser M. Metal ion availability and homeostasis as drivers of metabolic evolution and enzyme function. Curr. Opin. Genet. Dev. 2022;77 doi: 10.1016/j.gde.2022.101987. [DOI] [PubMed] [Google Scholar]
  • 386.Eck R.V., Dayhoff M.O. Evolution of the structure of ferredoxin based on living relics of primitive amino Acid sequences. Science. 1966;152:363–366. doi: 10.1126/science.152.3720.363. [DOI] [PubMed] [Google Scholar]
  • 387.Scintilla S., Bonfio C., Belmonte L., Forlin M., Rossetto D., Li J., Cowan J.A., Galliani A., Arnesano F., Assfalg M., Mansy S.S. Duplications of an iron–sulphur tripeptide leads to the formation of a protoferredoxin. Chem. Commun. 2016;52:13456–13459. doi: 10.1039/C6CC07912A. [DOI] [PubMed] [Google Scholar]
  • 388.Bonfio C., Valer L., Scintilla S., Shah S., Evans D.J., Jin L., Szostak J.W., Sasselov D.D., Sutherland J.D., Mansy S.S. UV-light-driven prebiotic synthesis of iron–sulfur clusters. Nat. Chem. 2017;9:1229–1234. doi: 10.1038/nchem.2817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Hall D.O., Cammack R., Rao K.K. Role for Ferredoxins in the Origin of Life and Biological Evolution. Nature. 1971;233:136–138. doi: 10.1038/233136a0. [DOI] [PubMed] [Google Scholar]
  • 390.Beinert H. Iron-sulfur proteins: Ancient structures, still full of surprises. J. Biol. Inorg. Chem. 2000;5:2–15. doi: 10.1007/s007750050002. [DOI] [PubMed] [Google Scholar]
  • 391.Lill R. Function and biogenesis of iron–sulphur proteins. Nature. 2009;460:831–838. doi: 10.1038/nature08301. [DOI] [PubMed] [Google Scholar]
  • 392.Weiss M.C., Sousa F.L., Mrnjavac N., Neukirchen S., Roettger M., Nelson-Sathi S., Martin W.F. The physiology and habitat of the last universal common ancestor. Nat. Microbiol. 2016;1 doi: 10.1038/nmicrobiol.2016.116. [DOI] [PubMed] [Google Scholar]
  • 393.Baross J.A., Hoffman S.E. Submarine hydrothermal vents and associated gradient environments as sites for the origin and evolution of life. Orig. Life Evol. Biosph. 1985;15:327–345. doi: 10.1007/BF01808177. [DOI] [Google Scholar]
  • 394.Nitschke W., Russell M.J. Hydrothermal focusing of chemical and chemiosmotic energy, supported by delivery of catalytic Fe, Ni, Mo/W, Co, S and Se, forced life to emerge. J. Mol. Evol. 2009;69:481–496. doi: 10.1007/s00239-009-9289-3. [DOI] [PubMed] [Google Scholar]
  • 395.Russell M.J., Hall A.J. The emergence of life from iron monosulphide bubbles at a submarine hydrothermal redox and pH front. J. Geol. Soc. London. 1997;154:377–402. doi: 10.1144/gsjgs.154.3.0377. [DOI] [PubMed] [Google Scholar]
  • 396.Zhao D., Bartlett S., Yung Y.L. Quantifying mineral-ligand structural similarities: Bridging the geological world of minerals with the biological world of enzymes. Life. 2020;10:338–339. doi: 10.3390/life10120338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 397.McGuinness K.N., Klau G.W., Morrison S.M., Moore E.K., Seipp J., Falkowski P.G., Nanda V. Evaluating Mineral Lattices as Evolutionary Proxies for Metalloprotein Evolution. Orig. Life Evol. Biosph. 2022;52:263–275. doi: 10.1007/s11084-022-09630-x. [DOI] [PubMed] [Google Scholar]
  • 398.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. Coord. Chem. Rev. 2022;452 doi: 10.1016/j.ccr.2021.214287. [DOI] [Google Scholar]
  • 399.Holden H.M., Jacobson B.L., Hurley J.K., Tollin G., Oh B.H., Skjeldal L., Chae Y.K., Cheng H., Xia B., Markley J.L. Structure-function studies of [2Fe-2S] ferredoxins. J. Bioenerg. Biomembr. 1994;26:67–88. doi: 10.1007/bf00763220. [DOI] [PubMed] [Google Scholar]
  • 400.Gurbiel R.J., Doan P.E., Gassner G.T., Macke T.J., Case D.A., Ohnishi T., Fee J.A., Ballou D.P., Hoffman B.M. Active site structure of Rieske-type proteins: electron nuclear double resonance studies of isotopically labeled phthalate dioxygenase from Pseudomonas cepacia and Rieske protein from Rhodobacter capsulatus and molecular modeling studies of a Rieske center. Biochemistry. 1996;35:7834–7845. doi: 10.1021/bi960380u. [DOI] [PubMed] [Google Scholar]
  • 401.Imlay J.A. Iron-sulphur clusters and the problem with oxygen. Mol. Microbiol. 2006;59:1073–1082. doi: 10.1111/j.1365-2958.2006.05028.x. [DOI] [PubMed] [Google Scholar]
  • 402.Dupuy J., Volbeda A., Carpentier P., Darnault C., Moulis J.M., Fontecilla-Camps J.C. Crystal structure of human iron regulatory protein 1 as cytosolic aconitase. Structure. 2006;14:129–139. doi: 10.1016/j.str.2005.09.009. [DOI] [PubMed] [Google Scholar]
  • 403.Solomon E.I., Randall D.W., Glaser T. Electronic structures of active sites in electron transfer metalloproteins: contributions to reactivity. Coord. Chem. Rev. 2000;200–202:595–632. doi: 10.1016/S0010-8545(00)00332-5. [DOI] [Google Scholar]
  • 404.Breiter D.R., Meyer T.E., Rayment I., Holden H.M. The molecular structure of the high potential iron-sulfur protein isolated from Ectothiorhodospira halophila determined at 2.5-A resolution. J. Biol. Chem. 1991;266:18660–18667. doi: 10.2210/pdb2hip/pdb. [DOI] [PubMed] [Google Scholar]
  • 405.Bruschi M., Guerlesquin F. Structure, function and evolution of bacterial ferredoxins. FEMS Microbiol. Rev. 1988;4:155–175. doi: 10.1016/0378-1097(88)90175-9. [DOI] [PubMed] [Google Scholar]
  • 406.Stokkermans J.P., Pierik A.J., Wolbert R.B., Hagen W.R., Van Dongen W.M., Veeger C. The primary structure of a protein containing a putative [6Fe-6S] prismane cluster from Desulfovibrio vulgaris (Hildenborough) Eur. J. Biochem. 1992;208:435–442. doi: 10.1111/j.1432-1033.1992.tb17205.x. [DOI] [PubMed] [Google Scholar]
  • 407.Reyda M.R., Fugate C.J., Jarrett J.T. A complex between biotin synthase and the iron-sulfur cluster assembly chaperone HscA that enhances in vivo cluster assembly. Biochemistry. 2009;48:10782–10792. doi: 10.1021/bi901393t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 408.Castro L., Tórtora V., Mansilla S., Radi R. Aconitases: Non-redox Iron-Sulfur Proteins Sensitive to Reactive Species. Acc. Chem. Res. 2019;52:2609–2619. doi: 10.1021/acs.accounts.9b00150. [DOI] [PubMed] [Google Scholar]
  • 409.Span I., Wang K., Wang W., Zhang Y., Bacher A., Eisenreich W., Li K., Schulz C., Oldfield E., Groll M. Discovery of acetylene hydratase activity of the iron–sulphur protein IspH. Nat. Commun. 2012;3:1042. doi: 10.1038/ncomms2052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 410.Seino H., Hidai M. Catalytic functions of cubane-type M4S4 clusters. Chem. Sci. 2011;2:847–857. doi: 10.1039/C0SC00545B. [DOI] [Google Scholar]
  • 411.Brown A.C., Thompson N.B., Suess D.L.M. Evidence for Low-Valent Electronic Configurations in Iron–Sulfur Clusters. J. Am. Chem. Soc. 2022;144:9066–9073. doi: 10.1021/jacs.2c01872. [DOI] [PubMed] [Google Scholar]
  • 412.Lee C.C., Hu Y., Ribbe M.W. Vanadium Nitrogenase Reduces CO. Science. 2010;329:642. doi: 10.1126/science.1191455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 413.Rebelein J.G., Hu Y., Ribbe M.W. Widening the Product Profile of Carbon Dioxide Reduction by Vanadium Nitrogenase. Chembiochem. 2015;16:1993–1996. doi: 10.1002/cbic.201500305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 414.Waser V., Mukherjee M., Tachibana R., Igareta N.V., Ward T.R. An Artificial [Fe4S4]-Containing Metalloenzyme for the Reduction of CO2 to Hydrocarbons. J. Am. Chem. Soc. 2023;145:14823–14830. doi: 10.1021/jacs.3c03546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 415.De Duve C. Neil Patterson Publishers; 1991. Blueprint for a Cell: The Nature and Origin of Life. [Google Scholar]
  • 416.Berg I.A., Kockelkorn D., Ramos-Vera W.H., Say R.F., Zarzycki J., Hügler M., Alber B.E., Fuchs G. Autotrophic carbon fixation in archaea. Nat. Rev. Microbiol. 2010;8:447–460. doi: 10.1038/nrmicro2365. [DOI] [PubMed] [Google Scholar]
  • 417.Wang W., Yang B., Qu Y., Liu X., Su W. FeS/S/FeS2 Redox System and Its Oxidoreductase-like Chemistry in the Iron-Sulfur World. Astrobiology. 2011;11:471–476. doi: 10.1089/ast.2011.0624. [DOI] [PubMed] [Google Scholar]
  • 418.de Aldecoa A.L.I., Roldán F.V., Menor-Salván C. Natural Pyrrhotite as a Catalyst in Prebiotic Chemical Evolution. Life. 2013;3:502–517. doi: 10.3390/life3030502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 419.Mitchell C.E., Terranova U., Beale A.M., Jones W., Morgan D.J., Sankar M., de Leeuw N.H. A surface oxidised Fe–S catalyst for the liquid phase hydrogenation of CO2. Catal. Sci. Technol. 2021;11:779–784. doi: 10.1039/D0CY01779E. [DOI] [Google Scholar]
  • 420.Wächtershäuser G. Groundworks for an evolutionary biochemistry: the iron-sulphur world. Prog. Biophys. Mol. Biol. 1992;58:85–201. doi: 10.1016/0079-6107(92)90022-x. [DOI] [PubMed] [Google Scholar]
  • 421.Wächtershäuser G. Before enzymes and templates: theory of surface metabolism. Microbiol. Rev. 1988;52:452–484. doi: 10.1128/mr.52.4.452-484.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 422.Wächtershäuser G. Evolution of the first metabolic cycles. Proc. Natl. Acad. Sci. USA. 1990;87:200–204. doi: 10.1073/pnas.87.1.200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 423.Shan Y., Lu W., Xi J., Qian Y. Biomedical applications of iron sulfide-based nanozymes. Front. Chem. 2022;10 doi: 10.3389/fchem.2022.1000709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 424.Ren X., Chen D., Wang Y., Li H., Zhang Y., Chen H., Li X., Huo M. Nanozymes-recent development and biomedical applications. J. Nanobiotechnol. 2022;20:92. doi: 10.1186/s12951-022-01295-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 425.Fan K., Cao C., Pan Y., Lu D., Yang D., Feng J., Song L., Liang M., Yan X. Magnetoferritin nanoparticles for targeting and visualizing tumour tissues. Nat. Nanotechnol. 2012;7:459–464. doi: 10.1038/nnano.2012.90. [DOI] [PubMed] [Google Scholar]
  • 426.Zhang Y., Wang Z., Li X., Wang L., Yin M., Wang L., Chen N., Fan C., Song H. Dietary Iron Oxide Nanoparticles Delay Aging and Ameliorate Neurodegeneration in Drosophila. Adv. Mater. 2016;28:1387–1393. doi: 10.1002/adma.201503893. [DOI] [PubMed] [Google Scholar]
  • 427.Jahantab E., Farzadmehr J., Matinkhah S., Mohammad Abadi N.T., Shafeiyan E., Yazdanshenas H. Effect of metal oxide nanoparticles on the activity of glutathione reductase, catalase, peroxidase and superoxide dismutase in plants under drought. Irrig. Drain. 2022;71:1351–1362. doi: 10.1002/ird.2739. [DOI] [Google Scholar]
  • 428.Zhao L., Lu L., Wang A., Zhang H., Huang M., Wu H., Xing B., Wang Z., Ji R. Nano-Biotechnology in Agriculture: Use of Nanomaterials to Promote Plant Growth and Stress Tolerance. J. Agric. Food Chem. 2020;68:1935–1947. doi: 10.1021/acs.jafc.9b06615. [DOI] [PubMed] [Google Scholar]
  • 429.Cui Z., Li Y., Zhang H., Qin P., Hu X., Wang J., Wei G., Chen C. Lighting Up Agricultural Sustainability in the New Era through Nanozymology: An Overview of Classifications and Their Agricultural Applications. J. Agric. Food Chem. 2022;70:13445–13463. doi: 10.1021/acs.jafc.2c04882. [DOI] [PubMed] [Google Scholar]
  • 430.Deng P., Gao Y., Mu L., Hu X., Yu F., Jia Y., Wang Z., Xing B. Development potential of nanoenabled agriculture projected using machine learning. Proc. Natl. Acad. Sci. USA. 2023;120 doi: 10.1073/pnas.2301885120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 431.Ma C., White J.C., Zhao J., Zhao Q., Xing B. Uptake of Engineered Nanoparticles by Food Crops: Characterization, Mechanisms, and Implications. Annu. Rev. Food Sci. Technol. 2018;9:129–153. doi: 10.1146/annurev-food-030117-012657. [DOI] [PubMed] [Google Scholar]
  • 432.Gajdardziska-Josifovska M., McClean R.G., Schofield M.A., Sommer C.V., Kean W.F. Discovery of nanocrystalline botanical magnetite. Eur. J. Mineral. 2001;13:863–870. doi: 10.1127/0935-1221/2001/0013-0863. [DOI] [Google Scholar]
  • 433.Rodríguez N., Menéndez N., Tornero J., Amils R., de la Fuente V. Internal iron biomineralization in Imperata cylindrica, a perennial grass: chemical composition, speciation and plant localization. New Phytol. 2005;165:781–789. doi: 10.1111/j.1469-8137.2004.01264.x. [DOI] [PubMed] [Google Scholar]
  • 434.Khalilov R.I., Nasibova A.N., Serezhenkov V.A., Ramazanov M.A., Kerimov M.K., Garibov A.A., Vanin A.F. Accumulation of magnetic nanoparticles in plants grown on soils of Apsheron peninsula. Biophysics. 2011;56:316–322. doi: 10.1134/S000635091102014X. [DOI] [PubMed] [Google Scholar]
  • 435.Zhu H., Han J., Xiao J.Q., Jin Y. Uptake, translocation, and accumulation of manufactured iron oxide nanoparticles by pumpkin plants. J. Environ. Monit. 2008;10:713–717. doi: 10.1039/b805998e. [DOI] [PubMed] [Google Scholar]
  • 436.Tombuloglu H., Slimani Y., Tombuloglu G., Almessiere M., Baykal A. Uptake and translocation of magnetite (Fe3O4) nanoparticles and its impact on photosynthetic genes in barley (Hordeum vulgare L.) Chemosphere. 2019;226:110–122. doi: 10.1016/j.chemosphere.2019.03.075. [DOI] [PubMed] [Google Scholar]
  • 437.Tombuloglu H., Slimani Y., AlShammari T.M., Bargouti M., Ozdemir M., Tombuloglu G., Akhtar S., Sabit H., Hakeem K.R., Almessiere M., et al. Uptake, translocation, and physiological effects of hematite (α-Fe2O3) nanoparticles in barley (Hordeum vulgare L.) Environ. Pollut. 2020;266 doi: 10.1016/j.envpol.2020.115391. [DOI] [PubMed] [Google Scholar]
  • 438.Chen G., Taherymoosavi S., Cheong S., Yin Y., Akter R., Marjo C.E., Rich A.M., Mitchell D.R.G., Fan X., Chew J., et al. Advanced characterization of biomineralization at plaque layer and inside rice roots amended with iron- and silica-enhanced biochar. Sci. Rep. 2021;11:159. doi: 10.1038/s41598-020-80377-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 439.Ghafariyan M.H., Malakouti M.J., Dadpour M.R., Stroeve P., Mahmoudi M. Effects of magnetite nanoparticles on soybean chlorophyll. Environ. Sci. Technol. 2013;47:10645–10652. doi: 10.1021/es402249b. [DOI] [PubMed] [Google Scholar]
  • 440.Kim J.H., Oh Y., Yoon H., Hwang I., Chang Y.S. Iron nanoparticle-induced activation of plasma membrane H+-ATPase promotes stomatal opening in Arabidopsis thaliana. Environ. Sci. Technol. 2015;49:1113–1119. doi: 10.1021/es504375t. [DOI] [PubMed] [Google Scholar]
  • 441.Li J., Hu J., Ma C., Wang Y., Wu C., Huang J., Xing B. Uptake, translocation and physiological effects of magnetic iron oxide (γ-Fe2O3) nanoparticles in corn (Zea mays L.) Chemosphere. 2016;159:326–334. doi: 10.1016/j.chemosphere.2016.05.083. [DOI] [PubMed] [Google Scholar]
  • 442.Al-Amri N., Tombuloglu H., Slimani Y., Akhtar S., Barghouthi M., Almessiere M., Alshammari T., Baykal A., Sabit H., Ercan I., Ozcelik S. Size effect of iron (III) oxide nanomaterials on the growth, and their uptake and translocation in common wheat (Triticum aestivum L.) Ecotoxicol. Environ. Saf. 2020;194 doi: 10.1016/j.ecoenv.2020.110377. [DOI] [PubMed] [Google Scholar]
  • 443.Cai L., Cai L., Jia H., Liu C., Wang D., Sun X. Foliar exposure of Fe3O4 nanoparticles on Nicotiana benthamiana: Evidence for nanoparticles uptake, plant growth promoter and defense response elicitor against plant virus. J. Hazard Mater. 2020;393 doi: 10.1016/j.jhazmat.2020.122415. [DOI] [PubMed] [Google Scholar]
  • 444.Kasote D.M., Lee J.H., Jayaprakasha G.K., Patil B.S. Seed Priming with Iron Oxide Nanoparticles Modulate Antioxidant Potential and Defense-Linked Hormones in Watermelon Seedlings. ACS Sustain. Chem. Eng. 2019;7:5142–5151. doi: 10.1021/acssuschemeng.8b06013. [DOI] [Google Scholar]
  • 445.Zheng T., Zhou Q., Tao Z., Ouyang S. Magnetic iron-based nanoparticles biogeochemical behavior in soil-plant system: A critical review. Sci. Total Environ. 2023;904 doi: 10.1016/j.scitotenv.2023.166643. [DOI] [PubMed] [Google Scholar]
  • 446.Palmqvist N.G.M., Seisenbaeva G.A., Svedlindh P., Kessler V.G. Maghemite Nanoparticles Acts as Nanozymes, Improving Growth and Abiotic Stress Tolerance in Brassica napus. Nanoscale Res. Lett. 2017;12:631. doi: 10.1186/s11671-017-2404-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 447.Yan L., Li P., Zhao X., Ji R., Zhao L. Physiological and metabolic responses of maize (Zea mays) plants to Fe3O4 nanoparticles. Sci. Total Environ. 2020;718 doi: 10.1016/j.scitotenv.2020.137400. [DOI] [PubMed] [Google Scholar]
  • 448.Liu Y., Xiao Z., Chen F., Yue L., Zou H., Lyu J., Wang Z. Metallic oxide nanomaterials act as antioxidant nanozymes in higher plants: Trends, meta-analysis, and prospect. Sci. Total Environ. 2021;780 doi: 10.1016/j.scitotenv.2021.146578. [DOI] [PubMed] [Google Scholar]
  • 449.Singh D., Sillu D., Kumar A., Agnihotri S. Dual nanozyme characteristics of iron oxide nanoparticles alleviate salinity stress and promote the growth of an agroforestry tree, Eucalyptus tereticornis Sm. Environ. Sci. Nano. 2021;8:1308–1325. doi: 10.1039/d1en00040c. [DOI] [Google Scholar]
  • 450.Ashraf H., Anjum T., Riaz S., Batool T., Naseem S., Li G. Sustainable synthesis of microwave-assisted IONPs using Spinacia oleracea L. for control of fungal wilt by modulating the defense system in tomato plants. J. Nanobiotechnol. 2022;20 doi: 10.1186/s12951-021-01204-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 451.Rizwan M., Noureen S., Ali S., Anwar S., Rehman M.Z.u., Qayyum M.F., Hussain A. Influence of biochar amendment and foliar application of iron oxide nanoparticles on growth, photosynthesis, and cadmium accumulation in rice biomass. J. Soils Sediments. 2019;19:3749–3759. doi: 10.1007/s11368-019-02327-1. [DOI] [Google Scholar]
  • 452.Sarraf M., Vishwakarma K., Kumar V., Arif N., Das S., Johnson R., Janeeshma E., Puthur J.T., Aliniaeifard S., Chauhan D.K., et al. Metal/Metalloid-Based Nanomaterials for Plant Abiotic Stress Tolerance: An Overview of the Mechanisms. Plants. 2022;11:316. doi: 10.3390/plants11030316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 453.Lu T., Wang X., Cui X., Li J., Xu J., Xu P., Wan J. Physiological and metabolomic analyses reveal that Fe3O4 nanoparticles ameliorate cadmium and arsenic toxicity in Panax notoginseng. Environ. Pollut. 2023;337 doi: 10.1016/j.envpol.2023.122578. [DOI] [PubMed] [Google Scholar]
  • 454.Manzoor N., Ahmed T., Noman M., Shahid M., Nazir M.M., Ali L., Alnusaire T.S., Li B., Schulin R., Wang G. Iron oxide nanoparticles ameliorated the cadmium and salinity stresses in wheat plants, facilitating photosynthetic pigments and restricting cadmium uptake. Sci. Total Environ. 2021;769 doi: 10.1016/j.scitotenv.2021.145221. [DOI] [PubMed] [Google Scholar]
  • 455.Cao X., Yue L., Wang C., Luo X., Zhang C., Zhao X., Wu F., White J.C., Wang Z., Xing B. Foliar Application with Iron Oxide Nanomaterials Stimulate Nitrogen Fixation, Yield, and Nutritional Quality of Soybean. ACS Nano. 2022;16:1170–1181. doi: 10.1021/acsnano.1c08977. [DOI] [PubMed] [Google Scholar]
  • 456.Silva S., Dias M.C., Pinto D.C.G.A., Silva A.M.S. Metabolomics as a Tool to Understand Nano-Plant Interactions: The Case Study of Metal-Based Nanoparticles. Plants. 2023;12:491. doi: 10.3390/plants12030491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 457.Guo H., Barnard A.S. Naturally occurring iron oxide nanoparticles: Morphology, surface chemistry and environmental stability. J. Mater. Chem. A. 2013;1:27–42. doi: 10.1039/c2ta00523a. [DOI] [Google Scholar]
  • 458.Sharma V.K., Filip J., Zboril R., Varma R.S. Natural inorganic nanoparticles--formation, fate, and toxicity in the environment. Chem. Soc. Rev. 2015;44:8410–8423. doi: 10.1039/c5cs00236b. [DOI] [PubMed] [Google Scholar]
  • 459.Findlay A.J., Estes E.R., Gartman A., Yücel M., Kamyshny A., Luther G.W. Iron and sulfide nanoparticle formation and transport in nascent hydrothermal vent plumes. Nat. Commun. 2019;10:1597. doi: 10.1038/s41467-019-09580-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 460.Yücel M., Sevgen S., Le Bris N. Soluble, Colloidal, and Particulate Iron Across the Hydrothermal Vent Mixing Zones in Broken Spur and Rainbow, Mid-Atlantic Ridge. Front. Microbiol. 2021;12 doi: 10.3389/fmicb.2021.631885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 461.Rickard D., Luther G.W. Chemistry of Iron Sulfides. Chem. Rev. 2007;107:514–562. doi: 10.1021/cr0503658. [DOI] [PubMed] [Google Scholar]
  • 462.Subramani T., Lilova K., Abramchuk M., Leinenweber K.D., Navrotsky A. Greigite (Fe3S4) is thermodynamically stable: Implications for its terrestrial and planetary occurrence. Proc. Natl. Acad. Sci. USA. 2020;117:28645–28648. doi: 10.1073/pnas.2017312117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 463.Gu X., Heaney P.J., Reis F.D.A.A., Brantley S.L. Deep abiotic weathering of pyrite. Science. 2020;370 doi: 10.1126/science.abb8092. [DOI] [PubMed] [Google Scholar]
  • 464.Holland H.D. Treatise on Geochemistry. Elsevier Inc; 2007. The Geologic History of Seawater; pp. 1–46. [DOI] [Google Scholar]
  • 465.Halevy I., Alesker M., Schuster E., Popovitz-Biro R., Feldman Y. A key role for green rust in the Precambrian oceans and the genesis of iron formations. Nat. Geosci. 2017;10:135–139. doi: 10.1038/ngeo2878. [DOI] [Google Scholar]
  • 466.Catling D.C. Treatise on Geochemistry. Second Edition. Elsevier Inc; 2013. The Great Oxidation Event Transition; pp. 177–195. [DOI] [Google Scholar]
  • 467.Bekker A., Planavsky N.J., Krapež B., Rasmussen B., Hofmann A., Slack J.F., Rouxel O.J., Konhauser K.O. Treatise on Geochemistry. Second Edition. Elsevier Inc; 2013. Iron Formations: Their Origins and Implications for Ancient Seawater Chemistry; pp. 561–628. [DOI] [Google Scholar]
  • 468.Nitschke W., McGlynn S.E., Milner-White E.J., Russell M.J. On the antiquity of metalloenzymes and their substrates in bioenergetics. Biochim. Biophys. Acta. 2013;1827:871–881. doi: 10.1016/j.bbabio.2013.02.008. [DOI] [PubMed] [Google Scholar]
  • 469.Russell M.J. Green rust: The simple organizing ‘seed’ of all life? Life. 2018;8:35. doi: 10.3390/life8030035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 470.Li Y., Kitadai N., Nakamura R. Chemical diversity of metal sulfide minerals and its implications for the origin of life. Life. 2018;8:46. doi: 10.3390/life8040046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 471.Russell M.J., Ponce A. Six ‘Must-Have’ Minerals for Life’s Emergence: Olivine, Pyrrhotite, Bridgmanite, Serpentine, Fougerite and Mackinawite. Life. 2020;10:291. doi: 10.3390/life10110291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 472.Duval S., Branscomb E., Trolard F., Bourrié G., Grauby O., Heresanu V., Schoepp-Cothenet B., Zuchan K., Russell M.J., Nitschke W. On the why's and how's of clay minerals' importance in life's emergence. Appl. Clay Sci. 2020;195 doi: 10.1016/j.clay.2020.105737. [DOI] [Google Scholar]
  • 473.Duval S., Zuchan K., Baymann F., Schoepp-Cothenet B., Branscomb E., Russell M.J., Nitschke W. In: Metals, Microbes, and Minerals - The Biogeochemical Side of Life. Kroneck P., Sosa Torres M., editors. De Gruyter; 2021. Minerals and the emergence of life. [DOI] [Google Scholar]
  • 474.Russell M.J. A self-sustaining serpentinization mega-engine feeds the fougerite nanoengines implicated in the emergence of guided metabolism. Front. Microbiol. 2023;14 doi: 10.3389/fmicb.2023.1145915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 475.Cairns-Smith A.G. Cambridge University Press; 1985. Seven Clues to the Origin of Life, A Scientific Detective Story. [Google Scholar]
  • 476.Bernal J.D. Routledge and Kegan Paul; 1951. The Physical Basis of Life. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Document S1. Figure S1–S16
mmc1.pdf (3.3MB, pdf)

Articles from iScience are provided here courtesy of Elsevier

RESOURCES