Skip to main content
Protein & Cell logoLink to Protein & Cell
. 2023 Aug 10;15(3):191–206. doi: 10.1093/procel/pwad048

Ergothioneine and its congeners: anti-ageing mechanisms and pharmacophore biosynthesis

Li Chen 1,2,#, Liping Zhang 3,#, Xujun Ye 4,, Zixin Deng 5,6,, Changming Zhao 7,8,
PMCID: PMC10903977  PMID: 37561026

Abstract

Ergothioneine, Ovothiol, and Selenoneine are sulfur/selenium-containing histidine-derived natural products widely distributed across different organisms. They exhibit significant antioxidant properties, making them as potential lead compounds for promoting health. Increasing evidence suggests that Ergothioneine is positively correlated with healthy ageing and longevity. The mechanisms underlying Ergothioneine's regulation of the ageing process at cellular and molecular levels are beginning to be understood. In this review, we provide an in-depth and extensive coverage of the anti-ageing studies on Ergothioneine and discuss its possible intracellular targeting pathways. In addition, we highlight the recent efforts in elucidating the biosynthetic details for Ergothioneine, Ovothiol, and Selenoneine, with a particular focus on the study of their pharmacophore-forming enzymology.

Keywords: Ergothioneine, antioxidant, anti-ageing, longevity, biosynthesis, enzymology

Introduction

Ergothioneine (EGT), a sulfur-containing histidine derivative, has received extensive attention as an efficient antioxidant for decades. Among the multitude of naturally occurring antioxidants, Ergothioneine is both abundant and unique. It is produced by a wide variety of species, including editable fungi and some prokaryotes. Mammals cannot synthesize Ergothioneine, but possess a highly specific organic cation transporter OCTN1 (now known as solute carrier family 22 member 4, SLC22A4) allowing for efficient absorption of Ergothioneine from one’s daily diet, leading to accumulation of Ergothioneine in tissues and organs of humans and other animals (Gründemann et al., 2005; Nikodemus et al., 2011). An ABC transporter specific for Ergothioneine has recently been discovered in gastrointestinal microbes. This indicates dietary Ergothioneine was competitively absorbed and metabolized by gut microbes, which may potentially affect its absorption in humans (Dumitrescu et al., 2022; Zhang et al., 2022).

Many literatures provide an overview of Ergothioneine’s diverse cytoprotective properties, with particular emphasis on its antioxidant activity (Halliwell et al., 2018; Borodina et al., 2020; Cheah and Halliwell, 2020; Fu and Shen, 2022). Ergothioneine scavenges reactive oxygen species (ROS) and reactive nitrogen species (RNS), which are known to damage DNA, proteins and lipids (Stoffels et al., 2017; Halliwell et al., 2018). Ergothioneine chelates metal cations such as Cu2+/Cu+ so as to inhibit the production of ROS (De Luna et al., 2013). On the other hand, it activates antioxidation enzymes via modulating cellular antioxidant defense systems (Colognato et al., 2006; Hseu et al., 2015; Dare et al., 2021). The antioxidation properties, together with its ultraviolet/infrared (UV/IR) radiation protective efficacy (Obayashi et al., 2005; Bazela et al., 2014; Hseu et al., 2015) and anti-inflammatory activities (Asahi et al., 2016; Cheah et al., 2017), render Ergothioneine an attractive choice for cosmetics, dietary supplements, and other applications.

During the last 2 decades, the anti-ageing research has been brought to the forefront of science. Evidence from model animal and human studies increasingly suggests that Ergothioneine is associated with healthy aging and actives against age-related diseases. Ergothioneine was found to have a positive age-prolonging effect in Drosophila melanogaster, and Caenorhabditis elegans exhibited a shorter lifespan when the Ergothioneine transporter was knocked out (Cheah et al., 2013; Pan et al., 2022). Low levels of Ergothioneine in blood and plasma are associated with frailty in the elderly population (Kameda et al., 2020; Teruya et al., 2021) as well as several age-related diseases, including neurodegenerative diseases (Jang et al., 2004; Yang et al., 2012; Cheah et al., 2016; Wu et al., 2021), chronic inflammation (Shimizu et al., 2015; Cheah et al., 2017), cardiovascular diseases (Smith et al., 2020; Lam-Sidun et al., 2021), and diabetes (Dare et al., 2021). On the other hand, the risk of several of these diseases is reduced when mushrooms, a major source of Ergothioneine, are consumed in increased quantities (Zhang et al., 2017; Ba et al., 2022), although the protective effects of Ergothioneine on the development of those diseases have not been established. Because of its anti-ageing effect, Ergothioneine was described as one of the “longevity vitamins” by Bruce Ames (2018). These findings, combined with people’s pressing requirement for healthy ageing, call for further research on Ergothioneine’s role in longevity and anti-senescence control, especially its molecular mechanisms underlying anti-ageing properties (Apparoo et al., 2022). In this review, we provide an overview of the antioxidant characteristics of Ergothioneine, which functions as both a ROS scavenger and an antioxidant defense regulator in modulating the Kelch-like ECH-associated protein 1 (KEAP1) and nuclear factor erythroid 2-related factor 2 (NRF2) signaling pathway. Furthermore, we highlight the possible involvement of Ergothioneine in ameliorating the development of genomic instability and ageing-associated epigenetic alterations, which are two of the primary hallmarks of ageing (López-Otín et al., 2013, 2023).

In addition to Ergothioneine’s remarkable anti-ageing properties, the biosynthesis of Ergothioneine and its congeners, Ovothiol and Selenoneine (Fig. 1), has attracted widespread attention. This is because nature has evolved unique biosynthetic strategies for the formation of their pharmacophores, including the thiol groups in Ergothioneine and Ovothiol, as well as the selenol group of Selenoneine, a selenium analogue of Ergothioneine. The non-heme iron enzyme, sulfoxide synthase, catalyzes unusual oxidative carbon-sulfur bond formation in the biosynthesis of Ergothioneine and Ovothiol. Strong efforts have been made to elucidate the catalytic mechanisms of these enzymes, including enzyme biochemistry, steady-state kinetics, protein crystallography, unnatural amino acid incorporation, and computational modeling. We herein highlight representative enzymatic models of oxygen-dependent sulfoxide synthases, which are the most popular enzymes responsible for C–S bond formation in Ergothioneine and Ovothiol biosynthesis. We summarize the recent progress on Ergothioneine oxygen-independent formation and Selenoneine biosynthesis as well.

Figure 1.

Figure 1.

Chemical structures of Ergothioneine, Ovothiol, and Selenoneine. The sulfur atom is respectively installed at the ε- and δ-carbon of the imidazole ring in Ergothioneine and Ovothiol, while it is a selenium attached at the ε-position in Selenoneine.

Ergothioneine and its proposed anti-ageing mechanisms

A direct antioxidant or a regulator of the antioxidant defense system

As an antioxidant, Ergothioneine was found to be an effective ROS scavenger in vitro. Ergothioneine reacts rapidly with singlet oxygen (1O2) (Stoffels et al., 2017), and scavenges other ROS species including superoxide radicals (O2·), hydrogen peroxides (H2O2), hydroxyl radicals (·OH), peroxynitrite (ONOO) and hypochlorite (ClO) (Fig. 2) (Akanmu et al., 1991; Aruoma et al., 1997; Servillo et al., 2015, 2017; Stoffels et al., 2017; Oumari et al., 2019; Ando and Morimitsu, 2021). Even at nanomolar levels, Ergothioneine has impressive cellular antioxidant properties (Hseu et al., 2015), although its reported redox potential (E0 = −0.06 V) is not as low as that of the classic reductants, such as glutathione (GSH) (E0 = −0.24 V), and certain antioxidant vitamins (Yadan, 2022; Hondal, 2023). As reviewed by Halliwell et al., in living cells, Ergothioneine may not principally react to ROS the same way as primary antioxidants do. It becomes important only when primary antioxidants, such as GSH, are exhausted during oxidative stress (Paul, 2022; Halliwell et al., 2023). In a review about antioxidant therapy, it has been claimed that antioxidant enzymes provide the predominant antioxidant defense because they react 103–106 of times more rapidly with ROS than small molecule antioxidants do (Forman and Zhang, 2021). Therefore, one hypothesis is that Ergothioneine may function as a regulator of antioxidant defense system more than a direct antioxidant.

Figure 2.

Figure 2.

Ergothioneine’s potential impacts on epigenetic methylation and demethylation via influencing cellular redox homeostasis. Ergothioneine targets cellular ROS, including O2·, H2O2, ·OH, and 1O2, in the manner of both a self-sacrificing antioxidant and a regulator of cellular antioxidant defense system, which modulates the cellular GSH redox state. When GSH levels are sufficiently high, it promotes the maximal activity of SAM synthase. Epigenetic methyltransferases, such as METTLs, DNMTs, and HMTs, utilize SAM as a substrate and their catalytic activity may be therefore influenced by Ergothioneine. On the other hand, cellular unliganded iron promotes hydroxyl radical production through the Fenton and Haber–Weiss reaction. Those iron from labile iron pool are also the essential co-factor of epigenetic demethyltransferases, including JmjC, TETs, FTO, and AlkBs. As ergothioneine scavenges ROS and maintains the labile iron pool, it ensures the activity of enzymes involved in epigenetic demethylation.

Several studies reported that Ergothioneine plays its role of antioxidation and anti-ageing by interacting with intercellular signaling cascades in vivo, such as the KEAP1 and NRF2 signaling pathway (Hseu et al., 2015, 2020; Dare et al., 2021; Salama et al., 2021). As part of the cellular antioxidant defense systems, the KEAP1–NRF2 signaling pathway plays critical roles in maintaining redox balance, defending against oxidative stress and inflammation (Surh, 2003). Under normal conditions, NRF2 is bound to Cullin3 (CUL3) and KEAP1 for its proteasomal degradation, ensuring the low abundance of cellular NRF2 (Fig. 3). But when exposed to oxidative stress, KEAP1/CUL3 polyubiquitination is hindered, thereby leading NRF2 to be released and then translocated to the nucleus, where NRF2 binds to the Antioxidant Response Element (ARE) and triggers the activation of a variety of antioxidant genes (Ganesh Yerra et al., 2013). Accumulating evidences have shown the significant contributions of the KEAP1–NRF2 system to the prevention and attenuation of ageing and ageing-related diseases (Zhang et al., 2015; Hiebert et al., 2018; Matsumaru and Motohashi, 2021).

Figure 3.

Figure 3.

Interaction of Ergothioneine with KEAP1–NRF2 and Sirtuin pathways. Ergothioneine regulates KEAP1–NRF2 complex, as well as SIRT1 and SIRT6 expression levels, to upregulate antioxidant genes, such as superoxide dismutase (SOD), catalase (CAT), heme oxygenase-1 (HO-1) and NAD(P)H:quinone oxidoreductase (NQO1). SIRT1 deacetylates NRF2, histone H1, H3, and H4, and modulates the activity of certain epigenetic enzymes, including methyltransferase SUV39H1 for histone H3K9me3 trimethylation, DNMT1 and DNMT3b for DNA 5mC methylation.

Experimental evidences have indicated that Ergothioneine regulates the KEAP1–NRF2 pathway in a dose-dependent manner, resulting in the upregulation of downstream antioxidant genes, including heme oxygenase-1 (HO-1), NAD(P)H: quinone oxidoreductase (NQO1), and superoxide dismutase (SOD) and catalase (CAT) (Hseu et al., 2015, 2020; Dare et al., 2021). NRF2 can detach itself from the KEAP1–NRF2 complex by self-modifications, for example, phosphorylation of its serine or threonine residues by cellular kinases, and (or) deacetylation by Sirtuins (Liu et al., 2021). Studies indicated that upon Ergothioneine treatment the protein levels of phosphatidylinositol 3-kinase (PI3K), serine/threonine kinase (AKT) and protein kinase C (PKC) are elevated. Additionally, it has also been revealed that NRF2 translocation was mediated by PI3K/AKT and PKC signaling pathways (Hseu et al., 2015, 2020). An in silico work indicated that Ergothioneine is an allosteric effector of NRF2, suggesting a direct interaction of Ergothioneine with NRF2 (Dare et al., 2022). Alternatively, Ergothioneine influences the KEAP1–NRF2 pathway possibly via its interaction with KEAP1, a thiol-rich protein. The abundant cysteine residue thiol groups presented on KEAP1 surface are specific sensors, which can interact with both redox-disrupting stimuli (such as electrophiles and ROS) (Zhang and Hannink, 2003), and endogenous or exogenous metabolites (e.g., Itaconate and Sulforaphane), thus resulting in the deactivation of KEAP1 and nuclear accumulation of NRF2 (Dinkova-Kostova et al., 2002; Hong et al., 2005; Mills et al., 2018). As for Ergothioneine, it is possible that it activates the KEAP1 cysteine sensors, leading to the dissociation of KEAP1–NRF2 complex and activation of antioxidant genes, however, this kind of possibility remains to be examined. It is worth investigating the interaction mechanisms between Ergothioneine and KEAP1–NRF2 signaling pathway.

Ergothioneine facilitates genome stability

Several hallmarks of human ageing have been identified and included as a significant part of the fundamental criteria for longevity intervention discoveries (López-Otín et al., 2013, 2023; Partridge et al., 2020). Among them, genomic instability, epigenetic alterations, telomere attrition, loss of proteostasis, and disabled macroautophagy are suggested as primary hallmarks. We explore Ergothioneine’s contributions toward genome stability as well as epigenetic modifications in this review.

DNA damage, largely in consequence of oxidative stress, plays a critical role in the ageing process and influences several key aspects of the ageing phenotype (Hasty et al., 2003; Kujoth et al., 2005; Schumacher et al., 2021). The integrity and stability of both nuclear DNA and mitochondrial DNA (mtDNA) are continuously declined due to exogenous damage, including chemicals and UV/IR irradiation, as well as endogenous damage, such as ROS and RNS (Hoeijmakers, 2009). As an effective antioxidant, Ergothioneine was shown to prevent DNA damage induced by ROS and RNS (Colognato et al., 2006; Markova et al., 2009; Paul and Snyder, 2010; Zhu et al., 2011; Hseu et al., 2015). In addition, Ergothioneine was known as a physiological protectant against UV rays-induced ROS generation and damage since it can absorb light in the UV range directly, as well as enable DNA repair in UV-irradiated cells (Carlsson et al., 1974; Markova et al., 2009). Pre-treatment of HaCaT cells with Ergothioneine suppresses the intercellular ROS level induced by UVA and protects DNA against oxidative damage (Hseu et al., 2015). Cells lacking the Ergothioneine transporter exhibited an increased level of DNA damage (Paul and Snyder, 2010).

Mitochondria, the energy factory of our body, produce the vast majority of cellular ATP, as well as ~90% of ROS, such as superoxide radicals and hydrogen peroxide. In the process of ageing, one of the main reasons that mtDNA shows a higher damage level than nuclear DNA is the proximity of ROS sources (Balaban et al., 2005). It has been reported that Ergothioneine may also reduce ROS production and inhibit oxidative damage to mtDNA (Paul and Snyder, 2010). In another study, Ergothioneine reduces mitochondria specific hydrogen peroxide production in the rat kidney (Williamson et al., 2020). Growing evidences suggested that OCTN1 is also present in mitochondria: An increased radioactivity was detected in the mitochondria of rat liver after injecting 3H labeled-Ergothioneine in rat (Kawano et al., 1982). Upon exposure to Ergothioneine, a significant accumulation of Ergothioneine in mitochondria was observed in both cells and tissues, as reviewed by Halliwell et al. (2023), from their unpublished data. However, at present, OCTN1’s mitochondrial location remains controversial due to an absence of conclusive evidence (Gründemann et al., 2022). While it would make sense for the Ergothioneine transporter to be found in mitochondria, more direct evidence may be needed.

Ergothioneine’s potential impacts on epigenetic methylation/demethylation

In addition to genomic instability, epigenetic alteration, including DNA and RNA methylation and histone modification, is another layer of age-related changes that harm the basic functions of cells and increases the risk of age-related diseases (Rando and Chang, 2012; López-Otín et al., 2013; Sen et al., 2016). C5 methylation of cytosine in DNA CpG dinucleotides (5mC) is the most abundant type of DNA epigenetic marker, and its pattern is altered with age (Horvath, 2013; Horvath and Raj, 2018). Post-translational modifications of histone, including methylation and acetylation, are crucial to chromatin function and vary with age. Many naturally occurring antioxidants have been found to exert their activity via epigenetic mechanisms, such as reversal of altered DNA methylation patterns (Kaufman-Szymczyk et al., 2015; Arora et al., 2019; Beetch et al., 2020). As reviewed by Hitchler et al., the intracellular redox chemistry has the potential to generate significant alterations in the epigenetic landscape through GSH and iron ions, which curtail the accessibility of epigenetic co-factors including S-adenosylmethionine (SAM), α-ketoglutarate, ascorbate, and nicotine adenine dinucleotide (NAD+) (Hitchler and Domann, 2021). Considering the close relationship between cellular redox metabolites and epigenetic modifications, it can be inferred that antioxidants and epigenomes are also potentially linked. The beneficial effects of Ergothioneine as a potent antioxidant and anti-ageing agent might also be mediated through epigenetic modifications on DNA, RNA, and Histone, although this possibility remains to be fully investigated.

Epigenetic methylation

DNA methyltransferases (DNMTs), using SAM as the methyl donor, are responsible for transferring methyl groups to the 5-position of cytosine residue in DNA to generate 5mC (Fig. 2). SAM availability is influenced by cellular GSH, which has been reported to play an essential role in the regulation of epigenetic methylations (Lertratanangkoon et al., 1996, 1997). First, GSH and SAM share a homocysteine intermediate in the de novo synthetic pathway, Ergothioneine therefore would contribute to SAM availability by affecting GSH level. Over recent years, growing evidence has indicated a link between Ergothioneine and GSH function. Ergothioneine eliminates ROS through direct or indirect mechanisms thereby modulating the cellular GSH redox state (Jacob, 2006; Servillo et al., 2015; Oumari et al., 2019; Hartmann et al., 2023). Second, GSH redox status influences the activity of SAM synthase (also named methionine adenosyl-transferase), in which a high [GSH]/[GSSG] (glutathione disulfide) ratio promotes SAM synthase to achieve its maximum activity (Pajares et al., 1992; Rahman et al., 2003; Hitchler and Domann, 2007). During oxidative stress, Ergothioneine interacts with the KEAP1–NRF2 signaling pathway, and NRF2 subsequently upregulates a multitude of antioxidation genes in the GSH-based system, including γ-glutamate cysteine ligase catalytic subunit (γ-GCLC), γ-glutamate cysteine ligase modifier subunit (γ-GCLM), glutathione reductase, glutathione peroxidase and glutathione S-transferase α2 (GSTA2) (Hayes and Dinkova-Kostova, 2014). Ergothioneine in the antioxidant system may therefore influence the cellular GSH amount and [GSH]/[GSSG] ratio, and those effects could further transmit to the availability of SAM, preventing the dysregulation of cellular methylation function. Additionally, as reviewed by Garcia-Gimenez and Pallardo (2014) and Garcia-Gimenez et al. (2014), GSH may influence epigenetic mechanisms in more ways than just regulating SAM levels. Perhaps Ergothioneine and GSH metabolic associations may lead to more complex epigenetic regulations, especially in redox imbalanced senescent cells.

Epigenetic demethylation

In addition to epigenetic methyltransferases, demethylases such as the ten–eleven translocation family demethylases (TETs) may also be influenced by Ergothioneine. The TETs are non-heme iron (NHFe) dependent monooxygenases, whose activities are severely affected by the availability of iron in the labile iron pool (Kakhlon and Cabantchik, 2002; Camarena et al., 2021). The abundance and oxidation state of cellular free irons are seriously affected by ROS, meanwhile, free ferrous (FeII) and ferric (FeIII) play critical roles in the development of fairly-benign ROS species into more toxic ones. Under oxidative stress, superoxide radicals release ferrous ions from certain iron-sulfur cluster proteins, ferritins, and transferrins. Furthermore, excessive hydrogen peroxides decompose heme and release free irons from heme proteins, including myoglobin, hemoglobin and cytochrome c (Borodina et al., 2020; Halliwell, 2020). The unliganded irons promote the formation of hydroxyl radicals from superoxide radicals and hydrogen peroxides via iron-catalyzed Haber–Weiss reactions (reaction 1). Reaction (1) is the sum of reactions (2) and (3), in which reaction (2) is also known as the Fenton reaction. The hypochlorous acid equivalent could be written as reaction (4) (Wardman and Candeias, 1996; Kehrer, 2000).

O2  +H2O2 OH+OH+O2 (1)
FeII+H2O2FeIII+OH+OH (2)
FeIII+O2  FeII+O2 (3)
FeII+HOClFeIII+OH+Cl (4)

These iron-dependent ROS autoxidation reactions produce highly toxic hydroxyl radicals, which can quickly damage the biomacromolecules. To prevent such reactions, cells sequestrate transition metal ions, especially iron (Halliwell, 2020). The ROS-induced iron restriction had been supported by the observation that exogenous H2O2 attenuated the demethylation activity of TET demethylases and resulted in an epigenetic shift (Niu et al., 2015; Hitchler and Domann, 2021). Ergothioneine’s remarkable ability to rapidly neutralize hydroxyl radicals makes it a critical player in the maintenance of cellular redox homeostasis. Other ROS, such as hydrogen peroxide, superoxide and hypochlorites, could also be targeted by Ergothioneine, in the manner of both a self-sacrificing antioxidant and a regulator of cellular antioxidant defense system (Figs. 2 and 3). Therefore, cellular iron co-factors may indirectly but critically be affected by Ergothioneine.

Similar to DNA, cellular RNA and histone are also decorated with chemical modifications, and such modifications participate in many aspects of life processes, including ageing (Michalak et al., 2019). Methylation/demethylation of RNA and histone could also be affected by Ergothioneine because both the histone methyltransferases (HMTs) and methyltransferase-like proteins (METTLs) employ SAM as a methyl group donor (Fig. 2), as well as the Jumonjic (JmjC)-domain-containing histone demethylase and the m6A demethylase (FTO, AlkBs) are NHFe/α-KG-dependent dioxygenases (Shi, 2007; Jia et al., 2011; Markolovic et al., 2016). Collectively, the epigenetic modifications to DNA, RNA, and histone would be greatly affected by varying the access to epigenetic enzyme co-factors, such as SAM and ferrous, that may be influenced by Ergothioneine. Although the direct control of Ergothioneine in epigenetic regulation has not been revealed, it would make sense that Ergothioneine preserves cellular redox balance to influence the availability of SAM and ferrous, ultimately preventing epigenetic alterations. Further investigations are required to elucidate whether and how Ergothioneine impacts epigenetic modifications with ageing.

Ergothioneine regulates the Sirtuin pathways

Beyond the methylation/demethylation modifications on DNA, RNA, and histone, acetylation/deacetylation of histone plays important roles in cellular ageing. Sirtuins, a family of NAD+-dependent deacetylases, play critical roles in a range of biological processes, including but not limited to epigenetic reprogramming and epigenetic drift, DNA damage repair and genome stability, oxidative stress and antioxidant defense pathways, mitochondrial function, as well as healthy longevity (Finkel et al., 2009; Singh et al., 2018). In mammals, the Sirtuin family of seven enzymes has been linked to both epigenetic functions and metabolic regulation (Brunet et al., 2004; Chen et al., 2005; Kawahara et al., 2009; Kanfi et al., 2012). Several findings indicated that SIRT1 and SIRT6 provide both profound health benefits and potent longevity activities (Kanfi et al., 2010, 2012; Satoh et al., 2013).

Ergothioneine interacts with the Sirtuin pathways to regulate ageing. It has been reported that Ergothioneine protects against endothelial senescence by regulating a group of Sirtuins. A study by D’Onofrio et al., revealed that Ergothioneine protects endothelial cells against high-glucose treatment through the upregulation of SIRT1 and SIRT6. Additionally, Ergothioneine’s protective effect against endothelial senescence was reduced when SIRT1 activity is inhibited or the SIRT6 gene is silenced (D’onofrio et al., 2016). Another recent study from the same group showed that Ergothioneine induces necroptosis in colorectal cancer cells by upregulating SIRT6 (D’onofrio et al., 2022). Both of the studies revealed that Ergothioneine exerts anti-ageing and anti-cancer properties via Sirtuin signaling, suggesting that Ergothioneine has a dynamic regulatory role in ageing signaling pathways.

These findings, combined with the fact that Sirtuin’s deacetylase consumes NAD+, an essential redox signaling molecule, reminiscing the idea that Ergothioneine may affect Sirtuins’ activity through modulating NAD+ availability. [NAD+]/[NADH] ratio is a key redox indicator of the metabolic and physiological status of the cell (Imai and Guarente, 2014), and NAD+ depletion caused by oxidative stress is harmful for the proper functioning of Sirtuins. As an antioxidant, Ergothioneine could control the prooxidant-antioxidant balance and maintain the metabolic co-factor pool of NAD+, which affects Sirtuins’ function directly. In addition, as summarized by Kalous et al. (2021), one of the mechanisms that decrease the activity of Sirtuins could be the oxidative post-translational modification by ROS and RNS. In ageing cells, ROS and RNS increase, contributing to the loss of Sirtuin activity (Salminen et al., 2013). Ergothioneine, the scavenger of ROS and RNS, might also enhance Sirtuins’ activity by preventing oxidative modifications to Sirtuin enzymes.

In mammals, oxidative stress and epigenetic functions are closely interconnected (Guillaumet-Adkins et al., 2017). SIRT1 is the main deacetylase of histones H3 and H4, as well as a direct regulator of SUV39H1 methyltransferase, promoting SUV39H1 activity on H3K9 methylation (Fig. 3) (Jing and Lin, 2015). Oxidative stress in rat myocytes was shown to induce a rapid upregulation of SUV39H1 (Yang et al., 2017). Additionally, a genome-wide distribution study of histone marks showed that during oligodendrocyte differentiation, a large portion of H3K9Me3 modifications can be mapped to the gene body encoding functional proteins (Liu et al., 2015). These data suggest that SIRT1 and SUV39H1 may involve in euchromatin transcription regulation in distinct tissues when exposed to oxidative stress. In a 2022 review (Padeken et al., 2022), Padeken highlighted emerging evidence that H3K9me3 is upregulated by oxidative stress and proposed that the alternation of epigenetic landscape in specific tissues may result from a long-term adaptation to stress. With Ergothioneine’s significant contribution to antioxidative stress, it is an intriguing possibility that Ergothioneine’s anti-ageing effect could be partially attribute to its impact on epigenetic dynamics. Collectively, the beneficial effects of Ergothioneine on cellular epigenetics, metabolism, and ageing could be mediated through Sirtuin pathways, although the details remain to be thoroughly explored.

Studies have also revealed a close linkage between Sirtuins and the KEAP1–NRF2 pathways. NRF2 is one of the common targets for Sirtuins in the regulation of antioxidative genes (Pan et al., 2016; Xue et al., 2016; Singh et al., 2018). SIRT1 is known to deacetylate NRF2 and contribute to its stability as well as stimulate the transport of NRF2 to the nucleus (Huang et al., 2013; Yang et al., 2014; Singh and Ubaid, 2020). In addition to enhancing NRF2 transcription and translocation, SIRT1 also negatively impacts its polyubiquitination by reducing the expression of KEAP1/CUL3, as well as increasing the binding ability of NRF2 to ARE (Wang et al., 2019). Moreover, SIRT6 positively regulates the NRF2-ARE antioxidant pathway as a NRF2 coactivator. It reduces the acetylation level of H3K56 to facilitate chromatin looping. As a scaffold, SIRT6 recruits RNA polymerase II (RNAP II) to generate a NRF2–SIRT6–RNAP II complex, leading to transcriptional activation of NRF2-regulated antioxidant genes (Pan et al., 2016; Rezazadeh et al., 2019). Ergothioneine affects both Sirtuins and the KEAP1–NRF2 pathways, supporting the hypothesis that it is not only an antioxidant but also an anti-ageing agent.

During the last 2 decades, Ergothioneine has attracted considerable attention due to its potential as an antioxidant and anti-ageing compound to treat numerous age-related ailments and even extend lifespan. Ergothioneine is believed to modulate the level of epigenetic enzyme Sirtuins, as well as the supplementation of epigenetic enzyme co-factors, including ferrous ions, SAM, and NAD+, leading to an Ergothioneine–Epigenome–Longevity axis. Since the epigenetic dynamics play one of the most significant roles in development, ageing, disease, and longevity, it is important to explore the effects of Ergothioneine on epigenome toward histone acetylome, as well as DNA, RNA, and histone methylome. Uncovering the functional mechanism of Ergothioneine would be a challenging and invaluable task, and it would lead to great advances in the development of Ergothioneine as an anti-ageing agent.

Biosynthesis of Ergothioneine, Ovothiol, and Selenoneine

Accumulated discoveries have bolstered the evidence of Ergothioneine’s anti-ageing activity and its therapeutic potential against ageing-associated diseases, leading to further exploration into its biosynthesis and that of related compounds. Trans-sulfur reactions involved in Ergothioneine and Ovothiol biosynthesis, as well as the unique selenium metabolic pathway responsible for Selenoneine production, have attracted broad attention in the field of natural product biosynthesis and synthetic biology. This section summarizes the pharmacophore formation steps, particularly the enzymatic mechanism of C–S bond formation reactions.

Ergothioneine biosynthesis

For over 100 years, Ergothioneine has been known to process amazing biological activities, yet it was only in the last few decades that its biosynthetic pathway has been revealed. Generally, there are a few distinct pathways for Ergothioneine synthesis in nature: the most common mechanism includes the oxygen-dependent formation of C–S bond, catalyzed by iron-dependent sulfoxide synthases (EgtB and Egt1) (Seebeck, 2010; Hu et al., 2014). The alternative pathways involve oxygen-independent sulfur transformations, catalyzed by Ergothioneine synthases (EanB and MES) (Fig. 4A) (Burn et al., 2017; Beliaeva and Seebeck, 2022). In the first step of EgtB-pathway, methyltransferase EgtD catalyzes the trimethylation of histidine to form trimethylhistidine (TMH, 4) using SAM as methyl donors. The following step is then catalyzed by EgtB, coupling TMH with γ-Glu-Cys (γ-GC, which is formed by ligating glutamic acid and cysteine through EgtA). EgtC and EgtE then trim the sulfoxide intermediate (5) to eventually release Ergothioneine (Seebeck, 2010). A similar but simplified pathway was catalyzed by Egt1&2: cysteine is used as the sulfur-donor directly, Egt1 catalyzes sulfoxide formation and Egt2 assumes the functions of EgtE to generate Ergothioneine (Hu et al., 2014). On the other hand, the oxygen-independent EanB- and MES-pathways eschew sulfoxide intermediates, instead, the C–S bonds are formed by transferring sulfur from polysulfides (“Sn”) or cysteine persulfides (CysSSH) to TMH directly (Burn et al., 2017; Beliaeva and Seebeck, 2022).

Figure 4.

Figure 4.

Biosynthesis of Ergothioneine and Ovothiol A. (A) Ergothioneine and Ovothiol biosynthetic pathways. (B) MthEgtB (top row) and CthEgtB (bottom row) enzymatic proposals; (C) EanB proposals from Seebeck’s model (top row) and Liu’s model (bottom row); (D) MES proposal.

Oxygen-dependent C–S bond formation

The crystal structures of two sulfoxide synthases, MthEgtB from Mycobacterium thermoresistibile and CthEgtB from Chloracidobacterium thermophilum, which prefer γ-Glu-Cys and cysteine as sulfur-donors, have been resolved, respectively (Goncharenko et al., 2015; Naowarojna et al., 2019; Stampfli et al., 2019). MthEgtB and MthEgtB-substrate complex structures showed that the catalytic FeII is coordinated by three histidine residues in a facial geometry, and the three remaining coordination sites of FeII center are occupied by the sulfur-donor γ-GC, the sulfur acceptor TMH and a crystallographic water. Mutagenesis and kinetic experiments of MthEgtB showed that a conserved residue Tyr377 plays a critical role in C–S bond formation and sulfoxidation (Fig. 4B). Replacement of Tyr377 with phenylalanine dramatically reduced the sulfoxide synthase activity of EgtB while remained its side activity as a cysteine dioxygenase (Goncharenko and Seebeck, 2016).

Several mechanistic models have been proposed for MthEgtB on the basis of computational studies, which all stem from the reactive FeIII–superoxo complex: a generally believed intermediate in many non-heme iron oxidases formed by oxidative addition of O2 to the FeII center (Faponle et al., 2017; Wei et al., 2017; Tian et al., 2018). Three points have been raised with distinct opinions: (i) whether thioether formation (Cε–S bond formation) or sulfenic acid formation (hydroxylation of the sulfur atom) is the first half of EgtB reaction; (ii) Whether the Cε–H cleavage occurs before or after the Cε–S bond formation? (iii) Is the active site tyrosine Tyr377 a redox agent or an acid–base catalyst? The Liao and Liu models suggest that hydroxylation of the sulfur atom is the first half, while the Visser model prefers that Cε–S bond formation is the first one. The catalytic residue Tyr377 was proposed to play a redox-active role in the Visser model, but was not involved in the Liu model. Additionally, the Liao and Liu models have also predicted that Cε–H cleavage is the rate determining step, at least in part, and R-sulfoxide is the reaction product, yet experimental observations indicate a close to unity substrate KIE and S-sulfoxide product formation (Goncharenko and Seebeck, 2016).

In a recent review, Stampfli and Seebeck combined computational studies and experimental work to develop a MthEgtB proposal, suggesting that protonation of FeIII-superoxo (9) by Tyr377 induces a thiyl radical of γ-GC, and deprotonation of the TMH imidazole ring initiates its attack to the electron-deficient thiyl radical (10). Cε–S bond formation precedes hydroxylation of the sulfur atom (11), and heterolytic cleavage of the Cε–H bond (12) gives the sulfoxide intermediate (5) (Fig. 4B) (Goncharenko et al., 2020; Stampfli and Seebeck, 2020). It partially aligns with the Visser computational model, the main difference is that Tyr377 functions as a Lewis acid–base but not the redox agent (Faponle et al., 2017).

The overall structure of CthEgtB is similar to that of MthEgtB with a difference that there are two essential tyrosine residues, Tyr93 and Tyr94, in the active center. Computational study based on CthEgtB crystal structure has been carried out as well, and results indicate that CthEgtB Tyr93–Tyr94 is the counterpart of MthEgtB Tyr377, functioning as an acid-base catalyst (13). Hydroxylation of the sulfur atom occurred prior to Cε–S bond formation, Cε–H cleavage is no longer the rate determining step. Notably, a coordination switch of γ-GC sulfenic acid intermediate from sulfur (14) to oxygen atom (15) was proposed (Wu et al., 2022; Zhang et al., 2023). It well accommodates the EgtB experimental results, and sheds light on the mechanism of other sulfoxide synthases, including Egt1 and OvoA (Cδ–S bond formation enzyme involved in Ovothiol biosynthesis). Although significant progresses have been achieved in the study of EgtB chemistry, the EgtB catalytic mechanism is still not well understood. Trapping and characterizing the intermediates are needed to clarify the debates about the reaction sequence and the exact role of active site tyrosine residue.

Oxygen-independent sulfur transformation

The anaerobic biosynthetic pathways of Ergothioneine were revealed recently. Oxygen-independent Ergothioneine synthase EanB from the green sulfur bacterium Chlorobium limicola has been characterized, and the enzymatic mechanism was proposed on the basis of protein crystal structure and kinetic studies (Fig. 4C) (Burn et al., 2017; Leisinger et al., 2019). Active site residues Tyr353 and Cys412 were found to play critical roles in the TMH-sulfurization-reaction. Tyr353 protonates the imidazole ring of TMH (16) at first, then the nucleophilic cysteine persulfide anion (Cys412-SS) attacks the imidazole ring to form Cε–S bond (17), and deprotonation of the imidazole ring by Tyr353 base (18) releases Ergothioneine ultimately. Another independent work further identified that polysulfide could be used as sulfur source directly (Cheng et al., 2020, 2021). It has reached a consensus that protonation (19) and deprotonation (20) of the TMH imidazole ring by Tyr353 are key steps of EanB catalysis, but the Cε–S bond is formed via an imidazole-2-yl carbene intermediate (21) in the Liu model on the basis of detection of an ε-carbon deuterium/hydrogen exchange reaction in EanB catalysis and quantum mechanics/molecular mechanics calculations. Computational studies showed that the carbene pathway is energetically preferable and the Arg417 guanidinium group and Tyr353 phenol group play key roles in the stabilization of carbene intermediates (Cheng et al., 2021; Lai and Cui, 2022).

A very recent work based on genome mining identified another oxygen-independent enzyme MES for Ergothioneine synthesis from the anaerobic bacterium Caldithrix abyssi (Beliaeva and Seebeck, 2022). MES harbors two functional domains, in which the C-terminal one is a cysteine desulfurase and the N-terminal one is a metallopterin-dependent enzyme responsible for the sulfurization of TMH. It is distinct from oxygen-dependent sulfoxide synthases EgtB, Egt1, and OvoA, which are iron-dependent enzymes. MES desulfurizes free cysteine and then transfers the sulfur onto TMH with Cys1074 and Cys1135 as intramolecular sulfur-transport-chain in form of cysteine persulfide CysSSH. For the Cε–S bond formation, sulfur is transferred from protein-borne CysSSH to the MoIV co-factor (22), generating an active MoVI = S species (23), then a nucleophilic attack on the sulfido ligand of MoVI = S by TMH imidazole ring initiates the reaction. Base assisted intermediate tautomerization (24) and product dissociation (25) recycle the co-factor to a reduced state, which is ready for the next round of trans-sulfur reaction. This is the first example that the mononuclear molybdenum dependent enzyme catalyzes carbon-sulfur bond formation.

Ovothiol biosynthesis

Ovothiol, a homologue of Ergothioneine, is another histidine derivative containing sulfur substitutions at δ position carbon of the imidazole ring. Found in a variety of marine invertebrates, algae, and protozoa, it plays a significant role in protecting against oxidative stress in the fertilization and embryo-release processes in seawater. It shows antioxidant properties and amazing potential to treat chronic low-grade systemic inflammation and related diseases (Castellano and Seebeck, 2018; Brancaccio et al., 2022). Three types of ovothiols have been characterized, with Ovothiol A’s α-amino group being unmethylated, while B and C are respectively mono- and di-methylated. The biosynthetic pathway of Ovothiol A has been established and the trans-sulfur reaction was then deciphered. The N-terminal domain of OvoA, a homolog of EgtB and Egt1, catalyzes the coupling of cysteine and histidine to give a sulfoxide intermediate (7) (Braunshausen and Seebeck, 2011). The sulfoxide is then cleaved by OvoB to release pyruvate and ammonia, with the C-terminal domain of OvoA methylating the π-nitrogen of imidazole ring to produce Ovothiol A (Fig. 4A) (Naowarojna et al., 2018).

Although OvoA’s crystal structure has not been determined, homology modeling and site-directed mutagenesis studies have shown that Tyr417 in OvoA from Erwinia tasmaniensis (EtaOvoA) is the counterpart of Tyr377 in MthEgtB. To explore the role of Tyr417, an unnatural amino acid 2-amino-3-(4-hydroxy-3-(methoxyl) phenyl) propanoic acid (MeOTyr), a tyrosine analogue which has comparable pKa, but a ~200 mV lower reduction potential relative to that of tyrosine, has been incorporated into EtaOvoA. Analyses of the OvoA Y417MtTyr variant reaction with [U-2H5]-Histidine revealed a deuterium isotope effect of 0.86 ± 0.03 (Chen et al., 2019), which is consistent with the computational prediction of an inverse isotope effect (Faponle et al., 2017). Meanwhile, the wild type EtaOvoA enzyme reaction with [U-2H5]-his yielded a substrate KIE of 1.01 ± 0.02, departing from the large magnitude KIE (5.7) predicted by another computational study about EgtB (Wei et al., 2017; Chen et al., 2018). Collectively, our experimental results prefer the Visser model that active site Y417 is part of a proton-coupled electron transfer process, Cε–S bond formation precedes Cε–H homolytic cleavage (Faponle et al., 2017; Chen et al., 2018, 2019). This is not fully consistent with the MthEgtB enzymatic proposal, in which the corresponding tyrosine residue (Y377) functions as a Lewis base (Stampfli and Seebeck, 2020). It is important to note that even though EgtB and OvoA are similar, they are not identical, since they show distinct substrate preference and regioselectivity. Further, a recent study on Methyloversatilis thermotolerans OvoA (MthOvoA) has indicated that its sulfoxide synthetase activity is still maintained when the tyrosine residue at the active site was mutated to phenylalanine (Cheng et al., 2022), indicating that there might be certain overlooked interactions that direct the product formation in the reaction serials of sulfoxide synthesis.

Selenoneine biosynthesis

Selenoneine, discovered from the blood of bluefin tuna and the culture broth of fission yeast Schizosaccharomyces pombe, is another homologue of Ergothioneine, featuring a selenium atom substitutedfor the sulfur atom. Compared to Ergothioneine, Selenoneine exhibits almost 1,000-fold stronger radical-scavenging ability (Yamashita and Yamashita, 2010; Pluskal et al., 2014). The source of Selenoneine in tuna cells remains inconclusive: it is uncertain whether Selenoneine is produced by tuna cells or if it is an exogenous nutrient enriched in their blood cells. Seneloneine could be generated from a divergent biosynthetic pathway of Ergothioneine in the environment of high Selenocysteine (SeCys) concentration. Introduction of S. pombe egt1 and egt2 genes into Aspergillus established an artificial pathway for Seneloneine biosynthesis (Fig. 5) (Pluskal et al., 2014; Turrini et al., 2018), and a hercynylselenocysteine intermediate (26) was proposed (Goncharenko et al., 2020). This suggests that SeCys may act as the selenium donor for Selenoneine synthesis. Free SeCys emerge from two distinct routes, one of which is seleno-protein ribosomal assembly line, and the other is mis-incorporation of selenium into the cysteine de novo biosynthetic pathway. To protect against oxidative stress and the mis-incorporation of selenium in ribosomal translation, cellular concentration of SeCys is limited, thereby explaining why both S. pombe and recombinant Aspergillus produce very little Selenoneine despite being exposed to exogenous selenates in their culture media.

Figure 5.

Figure 5.

Biosynthetic pathway of Selenoneine.

Recently, a Selenoneine biosynthetic pathway was identified in Variovorax paradoxus that involves N-acetyl-1-seleno-β-d-glucosamine (SeGlcNAc) as the direct selenium-donor (Fig. 5) (Kayrouz et al., 2022). SeGlcNAc is synthesized by similar initial steps with that of SeCys-tRNASel, which is a building block of seleno-protein. HSe- is activated by ATP, forming H2SePO3 through the catalysis of SelD (also named as SenC in Selenoneine pathway), then H2SePO3 is used to generate SeCys-tRNASel under the catalysis of SelA in seleno-protein assembly (Stadtman, 1996). Alternatively, H2SePO3 could be used to form SeGlcNAc (27) under the catalysis of SenB for Selenoneine biosynthesis. Similar to the hercynylcysteine sulfoxide intermediate involved in Ergothioneine biosynthesis, a hercynyl-SeGlcNAc selenoxide intermediate (28) formed by SenA has been identified in Selenoneine biosynthesis. This intermediate spontaneously fragmentates to generate Selenoneine, or is reduced by reductants such as thiols and ascorbate to give a selenoether (Goncharenko et al., 2020). Comparing to the Ergothioneine-divergent pathway, the novel enzyme SenA provides an effective rout for Selenoneine formation. This work filled the gap in our understanding of Seneloneine biosynthesis and expanded our knowledge of selenium metabolism.

Concluding remarks

Ergothioneine has been widely recognized to possess therapeutic potential in a variety of ageing-associated diseases, and promoting healthy longevity of humans. According to the free radical theory of ageing, the production of intracellular ROS is the major driving forces of ageing (Balaban et al., 2005). Ergothioneine-mediated ROS elimination has the potential to mitigate genomic instability and epigenetic alterations, two of the hallmarks of ageing. It is suggested that Ergothioneine has the capacity to regulate antioxidant defense KEAP1–NRF2 pathway, interact with Sirtuin-mediated epigenetic pathways, and influence epigenetic methylation/demethylation. Taken together, we suppose that Ergothioneine is more of a regulatory factor than a self-sacrificing antioxidant, and a possible Ergothioneine–Epigenome–Longevity axis was consequently proposed. The excellent biological activities of Ergothioneine, Ovothiol, and Seneloneine have prompted substantial study attention to the elucidation of their biosynthetic pathways. Detailed enzymatic mechanisms of key steps in these pathways, especially the pharmacophore-forming reactions, have been deciphered extensively. The oxygen-dependent sulfoxide synthases, EgtB, Egt1, and OvoA represent a novel sulfurtransferase branch distinct from the well-known Rhodanese, and SenA for C–Se bond formation have been identified and characterized, expanding our understanding in sulfur and selenium utilization in nature. The oxygen-independent Ergothioneine synthases, EanB and MES, exhibit two unprecedented biochemical reactions. These achievements pave the way for the development of metabolic engineering and synthetic biotechnologies for industrial production of Ergothioneine, Ovothiol, and Seneloneine.

Glossary

Abbreviations

5mC

C5 methylation of cytosine

γ-GC

γ-Glu-Cys

γ-GCLC

γ-glutamate cysteine ligase catalytic subunit

γ-GCLM

γ-glutamate cysteine ligase modifier subunit

ABC

ATP-binding cassette

AKT

serine/threonine kinase

ARE

antioxidant response element

ATP

adenosine 5ʹ-triphosphate

CAT

catalase

ClO

hypochlorite

CUL3

Cullin3

CysSSH

cysteine persulfides

DNMTs

DNA methyltransferases

DRACH

motif sequence, D = G/A/U, R = G/A, H = A/U/C

EGT

ergothioneine

FTO

fat mass and obesity associated protein

GSH

glutathione

GSSG

glutathione disulfide

GSTA2

glutathione S-transferase α2

HMTs

histone methyltransferases

HO-1

heme oxygenase-1

H1K26

histone 1 lysine 26

H3K9

histone 3 lysine 9

H3K14

histone 3 lysine 14

H3K56

histone 3 lysine 56

H3K9Me3

histone 3 with trimethylated lysine 9

H4K6

histone 4 lysine 6

JmjC

JumonjiC

KEAP1

Kelch-like ECH-associated protein 1

KIE

kinetic isotope effect

MeOTyr

2-amino-3-(4-hydroxy-3-(methoxyl) phenyl) propanoic acid

METTLs

methyltransferase-like proteins

mtDNA

mitochondrial DNA

NAD+

nicotine adenine dinucleotide

NHFe

non-heme iron

NQO1

NAD(P)H, quinone oxidoreductase

NRF2

nuclear factor erythroid 2-related factor 2

1O2

singlet oxygen

O2·−

superoxide radical

H2O2

hydrogen peroxide

OH

hydroxyl radical

OCTN1

organic cation transporter 1

ONOO

peroxynitrite

PI3K

phosphatidylinositol 3-kinase

PKC

protein kinase C

RNAP II

RNA polymerase II

RNS

reactive nitrogen species

ROS

reactive oxygen species

SAM

S-adenosylmethionine

SeCys

selenocysteine

sMAF

small Maf protein

SeGlcNAc

N-acetyl-1-seleno-β-d-glucosamine

SIRTs

sirtuins

SIRT1

sirtuin 1

SIRT6

sirtuin 6

SLC22A4

solute carrier family 22 member 4

SOD

superoxide dismutase

SUV39H1

suppressor of variegation 3–9 homolog 1

TF

transcription factor

TMH

trimethylhistidine

TETs

ten–eleven translocation family demethylases

Ub

ubiquitin

UV/IR

ultraviolet/infrared

Contributor Information

Li Chen, Department of Geriatrics, Zhongnan Hospital of Wuhan University, Wuhan University, Wuhan 430072, China; Key Laboratory of Combinatory Biosynthesis and Drug Discovery, School of Pharmaceutical Sciences, Ministry of Education, Wuhan University, Wuhan 430072, China.

Liping Zhang, Key Laboratory of Combinatory Biosynthesis and Drug Discovery, School of Pharmaceutical Sciences, Ministry of Education, Wuhan University, Wuhan 430072, China.

Xujun Ye, Department of Geriatrics, Zhongnan Hospital of Wuhan University, Wuhan University, Wuhan 430072, China.

Zixin Deng, Department of Geriatrics, Zhongnan Hospital of Wuhan University, Wuhan University, Wuhan 430072, China; Key Laboratory of Combinatory Biosynthesis and Drug Discovery, School of Pharmaceutical Sciences, Ministry of Education, Wuhan University, Wuhan 430072, China.

Changming Zhao, Department of Geriatrics, Zhongnan Hospital of Wuhan University, Wuhan University, Wuhan 430072, China; Key Laboratory of Combinatory Biosynthesis and Drug Discovery, School of Pharmaceutical Sciences, Ministry of Education, Wuhan University, Wuhan 430072, China.

Conflict of interest

All authors declare that they have no conflict of interest. This article does not contain any studies with human or animal subjects performed by the any of the authors.

Funding

This work was supported by grants from the National Key R&D Program of China (No. 2018YFA0903200 to C.Z.) and the National Natural Science Foundation of China (Nos. 32270032 and 32070038 to C.Z., No. 32000029 to L.C.).

Author contributions

All authors involved in this review article contributed to data research, manuscript content discussion, and oversaw the writing, reviewing and editing processes of the manuscript.

References

  1. Akanmu D, Cecchini R, Aruoma OIet al. The antioxidant action of ergothioneine. Arch Biochem Biophys 1991;288:10–16. [DOI] [PubMed] [Google Scholar]
  2. Ames BN. Prolonging healthy aging: longevity vitamins and proteins. Proc Natl Acad Sci USA 2018;115:10836–10844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Ando C, Morimitsu Y.. A proposed antioxidation mechanism of ergothioneine based on the chemically derived oxidation product hercynine and further decomposition products. Biosci Biotechnol Biochem 2021;85:1175–1182. [DOI] [PubMed] [Google Scholar]
  4. Apparoo Y, Phan CW, Kuppusamy URet al. Ergothioneine and its prospects as an anti-ageing compound. Exp Gerontol 2022;170:111982. [DOI] [PubMed] [Google Scholar]
  5. Arora I, Sharma M, Tollefsbol TO.. Combinatorial epigenetics impact of polyphenols and phytochemicals in cancer prevention and therapy. Int J Mol Sci 2019;20:4567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Aruoma OI, Whiteman M, England TGet al. Antioxidant action of ergothioneine: assessment of its ability to scavenge peroxynitrite. Biochem Biophys Res Commun 1997;231:389–391. [DOI] [PubMed] [Google Scholar]
  7. Asahi T, Wu X, Shimoda Het al. A mushroom-derived amino acid, ergothioneine, is a potential inhibitor of inflammation-related DNA halogenation. Biosci Biotechnol Biochem 2016;80:313–317. [DOI] [PubMed] [Google Scholar]
  8. Ba DM, Gao X, Al-Shaar Let al. Mushroom intake and cognitive performance among US older adults: the National Health and Nutrition Examination Survey, 2011–2014. Br J Nutr 2022;128:2241–2248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Balaban RS, Nemoto S, Finkel T.. Mitochondria, oxidants, and aging. Cell 2005;120:483–495. [DOI] [PubMed] [Google Scholar]
  10. Bazela K, Solyga-Zurek A, Debowska Ret al. l-Ergothioneine protects skin cells against UV-induced damage—a preliminary study. Cosmetics 2014;1:51–60. [Google Scholar]
  11. Beetch M, Harandi-Zadeh S, Shen Ket al. Dietary antioxidants remodel DNA methylation patterns in chronic disease. Br J Pharmacol 2020;177:1382–1408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Beliaeva MA, Seebeck FP.. Discovery and characterization of the metallopterin-dependent ergothioneine synthase from Caldithrix abyssi. JACS Au 2022;2:2098–2107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Borodina I, Kenny LC, Mccarthy CMet al. The biology of ergothioneine, an antioxidant nutraceutical. Nutr Res Rev 2020;33:190–217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Brancaccio M, Milito A, Viegas CAet al. First evidence of dermo-protective activity of marine sulfur-containing histidine compounds. Free Radic Biol Med 2022;192:224–234. [DOI] [PubMed] [Google Scholar]
  15. Braunshausen A, Seebeck FP.. Identification and characterization of the first ovothiol biosynthetic enzyme. J Am Chem Soc 2011;133:1757–1759. [DOI] [PubMed] [Google Scholar]
  16. Brunet A, Sweeney LB, Sturgill JFet al. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 2004;303:2011–2015. [DOI] [PubMed] [Google Scholar]
  17. Burn R, Misson L, Meury Met al. Anaerobic origin of ergothioneine. Angew Chem Int Ed Engl 2017;56:12508–12511. [DOI] [PubMed] [Google Scholar]
  18. Camarena V, Huff TC, Wang G.. Epigenomic regulation by labile iron. Free Radic Biol Med 2021;170:44–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Carlsson J, Kierstan MP, Brocklehurst K.. Reactions of l-ergothioneine and some other aminothiones with2,2ʹ-and 4,4ʹ-dipyridyl disulphides and of l-ergothioneine with iodoacetamide. 2-Mercaptoimidazoles, 2- and 4-thiopyridones, thiourea and thioacetamide as highly reactive neutral sulphur nucleophils. Biochem J 1974;139:221–235. [Google Scholar]
  20. Castellano I, Seebeck FP.. On ovothiol biosynthesis and biological roles: from life in the ocean to therapeutic potential. Nat Prod Rep 2018;35:1241–1250. [DOI] [PubMed] [Google Scholar]
  21. Cheah IK, Halliwell B.. Could ergothioneine aid in the treatment of coronavirus patients? Antioxidants 2020;9:595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cheah IK, Ong RL, Gruber Jet al. Knockout of a putative ergothioneine transporter in Caenorhabditis elegans decreases lifespan and increases susceptibility to oxidative damage. Free Radic Res 2013;47:1036–1045. [DOI] [PubMed] [Google Scholar]
  23. Cheah IK, Feng L, Tang RMYet al. Ergothioneine levels in an elderly population decrease with age and incidence of cognitive decline; a risk factor for neurodegeneration? Biochem Biophys Res Commun 2016;478:162–167. [DOI] [PubMed] [Google Scholar]
  24. Cheah IK, Tang RM, Yew TSet al. Administration of pure ergothioneine to healthy human subjects: uptake, metabolism, and effects on biomarkers of oxidative damage and inflammation. Antioxid Redox Signal 2017;26:193–206. [DOI] [PubMed] [Google Scholar]
  25. Chen D, Steele AD, Lindquist Set al. Increase in activity during calorie restriction requires Sirt1. Science 2005;310:1641. [DOI] [PubMed] [Google Scholar]
  26. Chen L, Naowarojna N, Song Het al. Use of a tyrosine analogue to modulate the two activities of a nonheme iron enzyme OvoA in ovothiol biosynthesis, cysteine oxidation versus oxidative C–S bond formation. J Am Chem Soc 2018;140:4604–4612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Chen L, Naowarojna N, Chen Bet al. Mechanistic studies of a nonheme iron enzyme OvoA in ovothiol biosynthesis using a tyrosine analogue, 2-Amino-3-(4-hydroxy-3-(methoxyl) phenyl) Propanoic Acid (MeOTyr). ACS Catal 2019;9:253–258. [Google Scholar]
  28. Cheng R, Wu L, Lai Ret al. Single-step replacement of an unreactive C–H bond by a C–S bond using polysulfide as the direct sulfur source in anaerobic ergothioneine biosynthesis. ACS Catal 2020;10:8981–8994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Cheng R, Lai R, Peng Cet al. Implications for an imidazol-2-yl carbene intermediate in the rhodanase-catalyzed C–S bond formation reaction of anaerobic ergothioneine biosynthesis. ACS Catal 2021;11:3319–3334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Cheng R, Weitz AC, Paris Jet al. OvoA(Mtht) from Methyloversatilis thermotolerans ovothiol biosynthesis is a bifunction enzyme: thiol oxygenase and sulfoxide synthase activities. Chem Sci 2022;13:3589–3598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Colognato R, Laurenza I, Fontana Iet al. Modulation of hydrogen peroxide-induced DNA damage, MAPKs activation and cell death in PC12 by ergothioneine. Clin Nutr 2006;25:135–145. [DOI] [PubMed] [Google Scholar]
  32. D’onofrio N, Servillo L, Giovane Aet al. Ergothioneine oxidation in the protection against high-glucose induced endothelial senescence: involvement of SIRT1 and SIRT6. Free Radic Biol Med 2016;96:211–222. [DOI] [PubMed] [Google Scholar]
  33. D’onofrio N, Martino E, Balestrieri Aet al. Diet-derived ergothioneine induces necroptosis in colorectal cancer cells by activating the SIRT3/MLKL pathway. FEBS Lett 2022;596:1313–1329. [DOI] [PubMed] [Google Scholar]
  34. Dare A, Channa ML, Nadar A.. l-Ergothioneine and its combination with metformin attenuates renal dysfunction in type-2 diabetic rat model by activating Nrf2 antioxidant pathway. Biomed Pharmacother 2021;141:111921. [DOI] [PubMed] [Google Scholar]
  35. Dare A, Elrashedy AA, Channa MLet al. Cardioprotective effects and in-silico antioxidant mechanism of l-ergothioneine in experimental type-2 diabetic rats. Cardiovasc Hematol Agents Med Chem 2022;20:133–147. [DOI] [PubMed] [Google Scholar]
  36. De Luna P, Bushnell EA, Gauld JW.. A density functional theory investigation into the binding of the antioxidants ergothioneine and ovothiol to copper. J Phys Chem A 2013;117:4057–4065. [DOI] [PubMed] [Google Scholar]
  37. Dinkova-Kostova AT, Holtzclaw WD, Cole RNet al. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc Natl Acad Sci USA 2002;99:11908–11913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Dumitrescu DG, Gordon EM, Kovalyova Yet al. A microbial transporter of the dietary antioxidant ergothioneine. Cell 2022;185:4526–4540.e18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Faponle AS, Seebeck FP, De Visser SP.. Sulfoxide synthase versus cysteine dioxygenase reactivity in a nonheme iron enzyme. J Am Chem Soc 2017;139:9259–9270. [DOI] [PubMed] [Google Scholar]
  40. Finkel T, Deng CX, Mostoslavsky R.. Recent progress in the biology and physiology of sirtuins. Nature 2009;460:587–591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Forman HJ, Zhang H.. Targeting oxidative stress in disease: promise and limitations of antioxidant therapy. Nat Rev Drug Discov 2021;20:689–709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Fu TT, Shen L.. Ergothioneine as a natural antioxidant against oxidative stress-related diseases. Front Pharmacol 2022;13:850813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Ganesh Yerra V, Negi G, Sharma SSet al. Potential therapeutic effects of the simultaneous targeting of the Nrf2 and NF-kappaB pathways in diabetic neuropathy. Redox Biol 2013;1:394–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Garcia-Gimenez JL, Pallardo FV.. Maintenance of glutathione levels and its importance in epigenetic regulation. Front Pharmacol 2014;5:88. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Garcia-Gimenez JL, Ibanez-Cabellos JS, Seco-Cervera Met al. Glutathione and cellular redox control in epigenetic regulation. Free Radic Biol Med 2014;75:S3. [DOI] [PubMed] [Google Scholar]
  46. Goncharenko KV, Seebeck FP.. Conversion of a non-heme iron-dependent sulfoxide synthase into a thiol dioxygenase by a single point mutation. Chem Commun (Camb) 2016;52:1945–1948. [DOI] [PubMed] [Google Scholar]
  47. Goncharenko KV, Vit A, Blankenfeldt Wet al. Structure of the sulfoxide synthase EgtB from the ergothioneine biosynthetic pathway. Angew Chem Int Ed Engl 2015;54:2821–2824. [DOI] [PubMed] [Google Scholar]
  48. Goncharenko KV, Flückiger S, Liao Cet al. Selenocysteine as a substrate, an inhibitor and a mechanistic probe for bacterial and fungal iron-dependent sulfoxide synthases. Chemistry 2020;26:1328–1334. [DOI] [PubMed] [Google Scholar]
  49. Gründemann D, Harlfinger S, Golz Set al. Discovery of the ergothioneine transporter. Proc Natl Acad Sci USA 2005;102:5256–5261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Gründemann D, Hartmann L, Flogel S.. The ergothioneine transporter (ETT): substrates and locations, an inventory. FEBS Lett 2022;596:1252–1269. [DOI] [PubMed] [Google Scholar]
  51. Guillaumet-Adkins A, Yanez Y, Peris-Diaz MDet al. Epigenetics and oxidative stress in aging. Oxid Med Cell Longev 2017;2017:9175806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Halliwell B. Reflections of an aging free radical. Free Radic Biol Med 2020;161:234–245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Halliwell B, Cheah IK, Tang RMY.. Ergothioneine—a diet-derived antioxidant with therapeutic potential. FEBS Lett 2018;592:3357–3366. [DOI] [PubMed] [Google Scholar]
  54. Halliwell B, Tang RMY, Cheah IK.. Diet-derived antioxidants: the special case of ergothioneine. Annu Rev Food Sci Technol 2023;14:323–345. [DOI] [PubMed] [Google Scholar]
  55. Hartmann L, Seebeck FP, Schmalz HGet al. Isotope-labeled ergothioneine clarifies the mechanism of reaction with singlet oxygen. Free Radic Biol Med 2023;198:12–26. [DOI] [PubMed] [Google Scholar]
  56. Hasty P, Campisi J, Hoeijmakers Jet al. Aging and genome maintenance: lessons from the mouse? Science 2003;299:1355–1359. [DOI] [PubMed] [Google Scholar]
  57. Hayes JD, Dinkova-Kostova AT.. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem Sci 2014;39:199–218. [DOI] [PubMed] [Google Scholar]
  58. Hiebert P, Wietecha MS, Cangkrama Met al. Nrf2-mediated fibroblast reprogramming drives cellular senescence by targeting the matrisome. Dev Cell 2018;46:145–161.e10. [DOI] [PubMed] [Google Scholar]
  59. Hitchler MJ, Domann FE.. An epigenetic perspective on the free radical theory of development. Free Radic Biol Med 2007;43:1023–1036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Hitchler MJ, Domann FE.. The epigenetic and morphogenetic effects of molecular oxygen and its derived reactive species in development. Free Radic Biol Med 2021;170:70–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Hoeijmakers JH. DNA damage, aging, and cancer. N Engl J Med 2009;361:1475–1485. [DOI] [PubMed] [Google Scholar]
  62. Hondal RJ. Selenium vitaminology: the connection between selenium, vitamin C, vitamin E, and ergothioneine. Curr Opin Chem Biol 2023;75:102328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Hong F, Freeman ML, Liebler DC.. Identification of sensor cysteines in human Keap1 modified by the cancer chemopreventive agent sulforaphane. Chem Res Toxicol 2005;18:1917–1926. [DOI] [PubMed] [Google Scholar]
  64. Horvath S. DNA methylation age of human tissues and cell types. Genome Biol 2013;14:R115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Horvath S, Raj K.. DNA methylation-based biomarkers and the epigenetic clock theory of ageing. Nat Rev Genet 2018;19:371–384. [DOI] [PubMed] [Google Scholar]
  66. Hseu YC, Lo HW, Korivi Met al. Dermato-protective properties of ergothioneine through induction of Nrf2/ARE-mediated antioxidant genes in UVA-irradiated human keratinocytes. Free Radic Biol Med 2015;86:102–117. [DOI] [PubMed] [Google Scholar]
  67. Hseu YC, Vudhya Gowrisankar Y, Chen XZet al. The antiaging activity of ergothioneine in UVA-irradiated human dermal fibroblasts via the inhibition of the AP-1 pathway and the activation of Nrf2-mediated antioxidant genes. Oxid Med Cell Longev 2020;2020:2576823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Hu W, Song H, Sae Her Aet al. Bioinformatic and biochemical characterizations of C–S bond formation and cleavage enzymes in the fungus Neurospora crassa ergothioneine biosynthetic pathway. Org Lett 2014;16:5382–5385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Huang K, Huang J, Xie Xet al. Sirt1 resists advanced glycation end products-induced expressions of fibronectin and TGF-beta1 by activating the Nrf2/ARE pathway in glomerular mesangial cells. Free Radic Biol Med 2013;65:528–540. [DOI] [PubMed] [Google Scholar]
  70. Imai S, Guarente L.. NAD+ and sirtuins in aging and disease. Trends Cell Biol 2014;24:464–471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Jacob C. A scent of therapy: pharmacological implications of natural products containing redox-active sulfur atoms. Nat Prod Rep 2006;23:851–863. [DOI] [PubMed] [Google Scholar]
  72. Jang JH, Aruoma OI, Jen LSet al. Ergothioneine rescues PC12 cells from beta-amyloid-induced apoptotic death. Free Radic Biol Med 2004;36:288–299. [DOI] [PubMed] [Google Scholar]
  73. Jia G, Fu Y, Zhao Xet al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat Chem Biol 2011;7:885–887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Jing H, Lin H.. Sirtuins in epigenetic regulation. Chem Rev 2015;115:2350–2375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Kakhlon O, Cabantchik ZI.. The labile iron pool: characterization, measurement, and participation in cellular processes(1). Free Radic Biol Med 2002;33:1037–1046. [DOI] [PubMed] [Google Scholar]
  76. Kalous KS, Wynia-Smith SL, Smith BC.. Sirtuin oxidative post-translational modifications. Front Physiol 2021;12:763417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Kameda M, Teruya T, Yanagida Met al. Frailty markers comprise blood metabolites involved in antioxidation, cognition, and mobility. Proc Natl Acad Sci USA 2020;117:9483–9489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Kanfi Y, Peshti V, Gil Ret al. SIRT6 protects against pathological damage caused by diet-induced obesity. Aging Cell 2010;9:162–173. [DOI] [PubMed] [Google Scholar]
  79. Kanfi Y, Naiman S, Amir Get al. The sirtuin SIRT6 regulates lifespan in male mice. Nature 2012;483:218–221. [DOI] [PubMed] [Google Scholar]
  80. Kaufman-Szymczyk A, Majewski G, Lubecka-Pietruszewska Ket al. The role of sulforaphane in epigenetic mechanisms, including interdependence between histone modification and DNA methylation. Int J Mol Sci 2015;16:29732–29743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Kawahara TL, Michishita E, Adler ASet al. SIRT6 links histone H3 lysine 9 deacetylation to NF-kappaB-dependent gene expression and organismal life span. Cell 2009;136:62–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Kawano H, Otani M, Takeyama Ket al. Studies on ergothioneine. VI. Distribution and fluctuations of ergothioneine in rats. Chem Pharm Bull (Tokyo) 1982;30:1760–1765. [DOI] [PubMed] [Google Scholar]
  83. Kayrouz CM, Huang J, Hauser Net al. Biosynthesis of selenium-containing small molecules in diverse microorganisms. Nature 2022;610:199–204. [DOI] [PubMed] [Google Scholar]
  84. Kehrer JP. The Haber–Weiss reaction and mechanisms of toxicity. Toxicology 2000;149:43–50. [DOI] [PubMed] [Google Scholar]
  85. Kujoth GC, Hiona A, Pugh TDet al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science 2005;309:481–484. [DOI] [PubMed] [Google Scholar]
  86. Lai R, Cui Q.. How to stabilize carbenes in enzyme active sites without metal ions. J Am Chem Soc 2022;144:20739–20751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Lam-Sidun D, Peters KM, Borradaile NM.. Mushroom-derived medicine? Preclinical studies suggest potential benefits of ergothioneine for cardiometabolic health. Int J Mol Sci 2021;22:3246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Leisinger F, Burn R, Meury Met al. Structural and mechanistic basis for anaerobic ergothioneine biosynthesis. J Am Chem Soc 2019;141:6906–6914. [DOI] [PubMed] [Google Scholar]
  89. Lertratanangkoon K, Orkiszewski RS, Scimeca JM.. Methyl-donor deficiency due to chemically induced glutathione depletion. Cancer Res 1996;56:995–1005. [PubMed] [Google Scholar]
  90. Lertratanangkoon K, Wu CJ, Savaraj Net al. Alterations of DNA methylation by glutathione depletion. Cancer Lett 1997;120:149–156. [DOI] [PubMed] [Google Scholar]
  91. Liu J, Magri L, Zhang Fet al. Chromatin landscape defined by repressive histone methylation during oligodendrocyte differentiation. J Neurosci 2015;35:352–365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Liu T, Lv YF, Zhao JLet al. Regulation of Nrf2 by phosphorylation: consequences for biological function and therapeutic implications. Free Radic Biol Med 2021;168:129–141. [DOI] [PubMed] [Google Scholar]
  93. López-Otín C, Blasco MA, Partridge Let al. The hallmarks of aging. Cell 2013;153:1194–1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. López-Otín C, Blasco MA, Partridge Let al. Hallmarks of aging: an expanding universe. Cell 2023;186:243–278. [DOI] [PubMed] [Google Scholar]
  95. Markolovic S, Leissing TM, Chowdhury Ret al. Structure–function relationships of human JmjC oxygenases-demethylases versus hydroxylases. Curr Opin Struct Biol 2016;41:62–72. [DOI] [PubMed] [Google Scholar]
  96. Markova NG, Karaman-Jurukovska N, Dong KKet al. Skin cells and tissue are capable of using l-ergothioneine as an integral component of their antioxidant defense system. Free Radic Biol Med 2009;46:1168–1176. [DOI] [PubMed] [Google Scholar]
  97. Matsumaru D, Motohashi H.. The KEAP1–NRF2 system in healthy aging and longevity. Antioxidants (Basel) 2021;10:1929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Michalak EM, Burr ML, Bannister AJet al. The roles of DNA, RNA and histone methylation in ageing and cancer. Nat Rev Mol Cell Biol 2019;20:573–589. [DOI] [PubMed] [Google Scholar]
  99. Mills EL, Ryan DG, Prag HAet al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature 2018;556:113–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Naowarojna N, Huang P, Cai Yet al. In vitro reconstitution of the remaining steps in ovothiol A biosynthesis: C–S lyase and methyltransferase reactions. Org Lett 2018;20:5427–5430. [DOI] [PubMed] [Google Scholar]
  101. Naowarojna N, Irani S, Hu Wet al. Crystal structure of the ergothioneine sulfoxide synthase from Candidatus chloracidobacterium thermophilum and structure-guided engineering to modulate its substrate selectivity. ACS Catal 2019;9:6955–6961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Nikodemus D, Lazic D, Bach Met al. Paramount levels of ergothioneine transporter SLC22A4 mRNA in boar seminal vesicles and cross-species analysis of ergothioneine and glutathione in seminal plasma. J Physiol Pharmacol 2011;62:411–419. [PubMed] [Google Scholar]
  103. Niu Y, Desmarais TL, Tong Zet al. Oxidative stress alters global histone modification and DNA methylation. Free Radic Biol Med 2015;82:22–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Obayashi K, Kurihara K, Okano Yet al. l-Ergothioneine scavenges superoxide and singlet oxygen and suppresses TNF-alpha and MMP-1 expression in UV-irradiated human dermal fibroblasts. J Cosmet Sci 2005;56:17–27. [PubMed] [Google Scholar]
  105. Oumari M, Goldfuss B, Stoffels Cet al. Regeneration of ergothioneine after reaction with singlet oxygen. Free Radic Biol Med 2019;134:498–504. [DOI] [PubMed] [Google Scholar]
  106. Padeken J, Methot SP, Gasser SM.. Establishment of H3K9-methylated heterochromatin and its functions in tissue differentiation and maintenance. Nat Rev Mol Cell Biol 2022;23:623–640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Pajares MA, Durán C, Corrales Fet al. Modulation of rat liver S-adenosylmethionine synthetase activity by glutathione. J Biol Chem 1992;267:17598–17605. [PubMed] [Google Scholar]
  108. Pan H, Guan D, Liu Xet al. SIRT6 safeguards human mesenchymal stem cells from oxidative stress by coactivating NRF2. Cell Res 2016;26:190–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Pan HY, Ye ZW, Zheng QWet al. Ergothioneine exhibits longevity-extension effect in Drosophila melanogaster via regulation of cholinergic neurotransmission, tyrosine metabolism, and fatty acid oxidation. Food Funct 2022;13:227–241. [DOI] [PubMed] [Google Scholar]
  110. Partridge L, Fuentealba M, Kennedy BK.. The quest to slow ageing through drug discovery. Nat Rev Drug Discov 2020;19:513–532. [DOI] [PubMed] [Google Scholar]
  111. Paul BD. Ergothioneine: a stress vitamin with antiaging, vascular, and neuroprotective roles? Antioxid Redox Signal 2022;36:1306–1317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Paul BD, Snyder SH.. The unusual amino acid l-ergothioneine is a physiologic cytoprotectant. Cell Death Differ 2010;17:1134–1140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Pluskal T, Ueno M, Yanagida M.. Genetic and metabolomic dissection of the ergothioneine and selenoneine biosynthetic pathway in the fission yeast, S. pombe, and construction of an overproduction system. PLoS One 2014;9:e97774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Rahman I, Gilmour PS, Jimenez LAet al. Ergothioneine inhibits oxidative stress- and TNF-alpha-induced NF-kappa B activation and interleukin-8 release in alveolar epithelial cells. Biochem Biophys Res Commun 2003;302:860–864. [DOI] [PubMed] [Google Scholar]
  115. Rando TA, Chang HY.. Aging, rejuvenation, and epigenetic reprogramming: resetting the aging clock. Cell 2012;148:46–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Rezazadeh S, Yang D, Tombline Get al. SIRT6 promotes transcription of a subset of NRF2 targets by mono-ADP-ribosylating BAF170. Nucleic Acids Res 2019;47:7914–7928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Salama SA, Abd-Allah GM, Mohamadin AMet al. Ergothioneine mitigates cisplatin-evoked nephrotoxicity via targeting Nrf2, NF-κB, and apoptotic signaling and inhibiting γ-glutamyl transpeptidase. Life Sci 2021;278:119572. [DOI] [PubMed] [Google Scholar]
  118. Salminen A, Kaarniranta K, Kauppinen A.. Crosstalk between oxidative stress and SIRT1: impact on the aging process. Int J Mol Sci 2013;14:3834–3859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Satoh A, Brace CS, Rensing Net al. Sirt1 extends life span and delays aging in mice through the regulation of Nk2 homeobox 1 in the DMH and LH. Cell Metab 2013;18:416–430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Schumacher B, Pothof J, Vijg Jet al. The central role of DNA damage in the ageing process. Nature 2021;592:695–703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Seebeck FP. In vitro reconstitution of Mycobacterial ergothioneine biosynthesis. J Am Chem Soc 2010;132:6632–6633. [DOI] [PubMed] [Google Scholar]
  122. Sen P, Shah PP, Nativio Ret al. Epigenetic mechanisms of longevity and aging. Cell 2016;166:822–839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Servillo L, Castaldo D, Casale Ret al. An uncommon redox behavior sheds light on the cellular antioxidant properties of ergothioneine. Free Radic Biol Med 2015;79:228–236. [DOI] [PubMed] [Google Scholar]
  124. Servillo L, D’onofrio N, Casale Ret al. Ergothioneine products derived by superoxide oxidation in endothelial cells exposed to high-glucose. Free Radic Biol Med 2017;108:8–18. [DOI] [PubMed] [Google Scholar]
  125. Shi Y. Histone lysine demethylases: emerging roles in development, physiology and disease. Nat Rev Genet 2007;8:829–833. [DOI] [PubMed] [Google Scholar]
  126. Shimizu T, Masuo Y, Takahashi Set al. Organic cation transporter Octn1-mediated uptake of food-derived antioxidant ergothioneine into infiltrating macrophages during intestinal inflammation in mice. Drug Metab Pharmacokinet 2015;30:231–239. [DOI] [PubMed] [Google Scholar]
  127. Singh V, Ubaid S.. Role of silent information regulator 1 (SIRT1) in regulating oxidative stress and inflammation. Inflammation 2020;43:1589–1598. [DOI] [PubMed] [Google Scholar]
  128. Singh CK, Chhabra G, Ndiaye MAet al. The role of sirtuins in antioxidant and redox signaling. Antioxid Redox Signal 2018;28:643–661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Smith E, Ottosson F, Hellstrand Set al. Ergothioneine is associated with reduced mortality and decreased risk of cardiovascular disease. Heart 2020;106:691–697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Stadtman TC. Selenocysteine. Annu Rev Biochem 1996;65:83–100. [DOI] [PubMed] [Google Scholar]
  131. Stampfli AR, Seebeck FP.. The catalytic mechanism of sulfoxide synthases. Curr Opin Chem Biol 2020;59:111–118. [DOI] [PubMed] [Google Scholar]
  132. Stampfli AR, Goncharenko KV, Meury Met al. An alternative active site architecture for O(2) activation in the ergothioneine biosynthetic EgtB from Chloracidobacterium thermophilum. J Am Chem Soc 2019;141:5275–5285. [DOI] [PubMed] [Google Scholar]
  133. Stoffels C, Oumari M, Perrou Aet al. Ergothioneine stands out from hercynine in the reaction with singlet oxygen: resistance to glutathione and TRIS in the generation of specific products indicates high reactivity. Free Radic Biol Med 2017;113:385–394. [DOI] [PubMed] [Google Scholar]
  134. Surh YJ. Cancer chemoprevention with dietary phytochemicals. Nat Rev Cancer 2003;3:768–780. [DOI] [PubMed] [Google Scholar]
  135. Teruya T, Chen YJ, Kondoh Het al. Whole-blood metabolomics of dementia patients reveal classes of disease-linked metabolites. Proc Natl Acad Sci USA 2021;118:e2022857118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Tian G, Su H, Liu Y.. Mechanism of sulfoxidation and C–S Bond formation involved in the biosynthesis of ergothioneine catalyzed by ergothioneine synthase (EgtB). ACS Catal 2018;8:5875–5889. [Google Scholar]
  137. Turrini NG, Kroepfl N, Jensen KBet al. Biosynthesis and isolation of selenoneine from genetically modified fission yeast. Metallomics 2018;10:1532–1538. [DOI] [PubMed] [Google Scholar]
  138. Wang W, Sun W, Cheng Yet al. Role of sirtuin-1 in diabetic nephropathy. J Mol Med (Berl) 2019;97:291–309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Wardman P, Candeias LP.. Fenton chemistry: an introduction. Radiat Res 1996;145:523–531. [PubMed] [Google Scholar]
  140. Wei WJ, Siegbahn PE, Liao RZ.. Theoretical study of the mechanism of the nonheme iron enzyme egtB. Inorg Chem 2017;56:3589–3599. [DOI] [PubMed] [Google Scholar]
  141. Williamson RD, Mccarthy FP, Manna Set al. l-(+)-Ergothioneine significantly improves the clinical characteristics of preeclampsia in the reduced uterine perfusion pressure rat model. Hypertension 2020;75:561–568. [DOI] [PubMed] [Google Scholar]
  142. Wu LY, Cheah IK, Chong JRet al. Low plasma ergothioneine levels are associated with neurodegeneration and cerebrovascular disease in dementia. Free Radic Biol Med 2021;177:201–211. [DOI] [PubMed] [Google Scholar]
  143. Wu P, Gu Y, Liao Let al. Coordination switch drives selective C–S bond formation by the non-heme sulfoxide synthases. Angew Chem Int Ed Engl 2022;61:e202214235. [DOI] [PubMed] [Google Scholar]
  144. Xue F, Huang JW, Ding PYet al. Nrf2/antioxidant defense pathway is involved in the neuroprotective effects of Sirt1 against focal cerebral ischemia in rats after hyperbaric oxygen preconditioning. Behav Brain Res 2016;309:1–8. [DOI] [PubMed] [Google Scholar]
  145. Yadan JC. Matching chemical properties to molecular biological activities opens a new perspective on l-ergothioneine. FEBS Lett 2022;596:1299–1312. [DOI] [PubMed] [Google Scholar]
  146. Yamashita Y, Yamashita M.. Identification of a novel selenium-containing compound, selenoneine, as the predominant chemical form of organic selenium in the blood of bluefin tuna. J Biol Chem 2010;285:18134–18138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Yang NC, Lin HC, Wu JHet al. Ergothioneine protects against neuronal injury induced by beta-amyloid in mice. Food Chem Toxicol 2012;50:3902–3911. [DOI] [PubMed] [Google Scholar]
  148. Yang Y, Li W, Liu Yet al. Alpha-lipoic acid improves high-fat diet-induced hepatic steatosis by modulating the transcription factors SREBP-1, FoxO1 and Nrf2 via the SIRT1/LKB1/AMPK pathway. J Nutr Biochem 2014;25:1207–1217. [DOI] [PubMed] [Google Scholar]
  149. Yang G, Weng X, Zhao Yet al. The histone H3K9 methyltransferase SUV39H links SIRT1 repression to myocardial infarction. Nat Commun 2017;8:14941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Zhang DD, Hannink M.. Distinct cysteine residues in Keap1 are required for Keap1-dependent ubiquitination of Nrf2 and for stabilization of Nrf2 by chemopreventive agents and oxidative stress. Mol Cell Biol 2003;23:8137–8151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Zhang H, Davies K JA, Forman HJ.. Oxidative stress response and Nrf2 signaling in aging. Free Radic Biol Med 2015;88:314–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Zhang S, Tomata Y, Sugiyama Ket al. Mushroom consumption and incident dementia in elderly Japanese: the Ohsaki Cohort 2006 Study. J Am Geriatr Soc 2017;65:1462–1469. [DOI] [PubMed] [Google Scholar]
  153. Zhang Y, Gonzalez-Gutierrez G, Legg KAet al. Discovery and structure of a widespread bacterial ABC transporter specific for ergothioneine. Nat Commun 2022;13:7586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Zhang J, Wu P, Zhang Xet al. Coordination dynamics of iron is a key player in the catalysis of Non-heme enzymes. ChemBioChem 2023;24:e202300119. [DOI] [PubMed] [Google Scholar]
  155. Zhu BZ, Mao L, Fan RMet al. Ergothioneine prevents copper-induced oxidative damage to DNA and protein by forming a redox-inactive ergothioneine–copper complex. Chem Res Toxicol 2011;24:30–34. [DOI] [PubMed] [Google Scholar]

Articles from Protein & Cell are provided here courtesy of Oxford University Press

RESOURCES