Abstract
Advances in the field of redox biology have contributed to the understanding of the complexity of the thiol-based system in mediating signal transduction. The redox environment is the overall spatiotemporal balance of oxidation-reduction systems within the integrated compartments of the cell, tissues and whole organisms. The ratio of the reduced to disulfide glutathione redox couple (GSH:GSSG) is a key indicator of the redox environment and its associated cellular health. The reaction mechanisms of glutathione-dependent and related thiol-based enzymes play a fundamental role in the function of GSH as a redox regulator. Glutathione homeostasis is maintained by the balance of GSH synthesis (de novo and salvage pathways) and its utilization through its detoxification, thiol signalling, and antioxidant defence functions via GSH-dependent enzymes and free radical scavenging. As such, GSH acts in concert with the entire redox network to maintain reducing conditions in the cell. Caenorhabditis elegans offers a simple model to facilitate further understanding at the multicellular level of the physiological functions of GSH and the GSH-dependent redox network. This review discusses the C. elegans studies that have investigated glutathione and related systems of the redox network including; orthologs to the protein-encoding genes of GSH synthesis; glutathione peroxidases; glutathione-S-transferases; and the glutaredoxin, thioredoxin and peroxiredoxin systems.
Keywords: Caenorhabditis elegans, Glutathione, Glutathione peroxidase, Glutathione S-Transferase, Glutaredoxin, Thioredoxin
Abbreviations: C. elegans, Caenorhabditis elegans; Grx, glutaredoxin; GSH, reduced glutathione; GSSG, disulfide glutathione; GST, glutathione-S-transferase; Prx, peroxiredoxin; Trx, thioredoxin; TrxR, thioredoxin reductase
Graphical abstract
Highlights
-
•
Thiol-mediated signalling is a central node in the cellular redox network.
-
•
Caenorhabditis elegans possess orthologs for all key mammalian redox systems.
-
•
This nematode is a simple but robust multicellular model for redox studies.
1. Introduction
Redox systems are regulated under non-equilibrium steady states that maintain the reducing conditions of the cell via electron transfer [1,2]. They provide central energy currencies to support metabolism and coordinate the overall organization of living organisms, including cell structure and function [[3], [4], [5]]. The cellular redox state is determined by the activity of relative ratios of cellular and extracellular redox couples. The two major redox systems, the pyridine nucleotide (NADH/NAD+, NADPH/NADP+) and thiol systems differ considerably in organization but both co-ordinate the cellular redox environment [6]. The thiol system includes the low molecular-weight thiol/disulfide couples such as glutathione (GSH/GSSG) and cysteine/cystine [7], and the oxidized/reduced states of redox enzymes that have cysteine residues at their active sites such as thioredoxins, glutaredoxins, and peroxiredoxins [8]. All these play major roles in the maintenance of the cellular redox environment. The electron carriers NADH and NADPH support metabolic requirements and maintain the redox status of the cell [9]. The NADH/NAD+ redox couple assists in ATP synthesis, while the NADPH/NADP+ system supports the maintenance of the thiol system which includes recycling of GSSG back to GSH [6]. Current evidence for the central role of GSH in thiol signalling suggests that the cellular redox environment is governed by the enzymatically-controlled reactions that facilitate the turnover of GSH rather than by the thermodynamic control of the GSH/GSSG couple [10]. This position is supported by in vivo studies, which used genetically encoded redox sensors to investigate the spatial patterning of cellular redox potential in Caenorhabditis elegans [11]. It was concluded that small changes in the GSH/GSSG ratio is amplified to large effects on redox potential through the oxidation/reduction of cysteine residues within proteins, rather than changes in the GSH/GSSG ratio directly affecting the cellular redox potential [11].
In this review, we focus on studies that have investigated glutathione synthesis and the related thiol systems that support the use of C. elegans as a suitable model organism for investigations of the thiol redox network.
2. Glutathione
Glutathione (GSH) is the most abundant low molecular-weight thiol found in almost all eukaryotic organisms, most gram-negative bacteria and a few gram-positive bacteria [12,13]. GSH is often referred to as the major antioxidant of the cell, though the functions of GSH extend beyond its antioxidant properties to many other essential cellular processes, including detoxification of xenobiotic compounds, modulation of cell proliferation, transport and storage of cysteine and maintenance of redox status [[14], [15], [16], [17]]. Cellular GSH is maintained by the de novo and salvage synthesis pathways [18]. The GSH:GSSG ratio is used as the primary indicator of cellular redox status due to the GSH pool being approximately three to four orders of magnitude higher in abundance than other major redox couples, such as the pyridine nucleotide and related thiol couples [7,19]. Mechanisms that involve GSH as a cofactor in enzymatic reactions may be more crucial to the regulatory role GSH has on the redox environment than the capacity of the GSH/GSSG couple to act as a redox buffer [20].
3. Thiol-based redox signalling
Thiol-based signalling occurs as part of normal cellular processes in response to elevated free radical and ROS production, which can affect many essential cellular processes including; phosphorylation pathways, gene transcription, cytoskeletal organization and ion channel activity [[21], [22], [23]]. Protein thiols undergo reversible and irreversible post-translational oxidative modifications such as, via the direct reaction with H2O2 that results in various oxidation states of cysteine residues including, sulfenic (—SOH), sulfinic (—SO2H), or the irreversible sulfonic acid (—SO3H) [[24], [25], [26]]. Protein S-glutathionylation is recognized as an efficient regulator of redox signal transduction largely due to the high cellular concentrations of GSH and the reversibility of the reaction catalyzed by the glutaredoxin enzyme family [[27], [28], [29]]. Many intracellular redox responses involved in signalling and protein function are regulated by S-glutathionylation, which forms disulfide bonds between GSH and the cysteine residues of proteins [[30], [31], [32]]. Protein S-glutathionylation can either activate or inhibit the activity and function of a range of structural proteins and enzymes [27,28]. Some enzymes can be S-glutathionylated at more than one cysteine residue, allowing for their activities to be modulated in a site-selective manner [28]. S-Glutathionylation can also prevent the irreversible oxidation of cysteine residues and thereby conserve the functions of the redox signalling pathways [33].
The reversible oxidation of protein thiols can act as ‘redox switches’ that control a range of functions of the protein, including regulation of protein activity, stabilization of protein structure, formation of protein activation complexes, and control of protein distribution [2,34]. The majority of cysteine residues support redox-signalling, largely by non-radical two-electron transfer [35]. A recent report has also discussed the possible role that free-radicals have in signal transmission of thiol-based signalling [21]. In the thiol network, hydrogen peroxide (H2O2) acts as a second messenger in transmitting the signal irrespective of which oxidant was responsible for its initiation [21]. The specific activities of the GSH-utilising enzymes and other related antioxidant systems which include, glutaredoxins, thioredoxins and peroxiredoxin systems are critical to GSH's influence on the thiol signalling system (Fig. 1) [8,10,36,37].
Fig. 1.
The network of the cellular thiol redox system in multicellular eukaryotes. Glutathione (GSH) is synthesized in the cytosol via the de novo pathway (green box). The GSH salvage pathway (light blue boxes) occurs intracellularly via the reduction of GSSG by glutathione reductase (GR), and extracellularly through γ-glutamyl transferase (γ-GT) mediated degradation of exogenous GSH. The γ-GT activity generates cysteinylglycine (CysGly) and a gamma-glutamyl amino acid (γ-Glu-AA). CysGly is hydrolyzed by dipeptidases (DP) with the released cysteine (Cys) and glycine (Gly), which are taken up by peptide transporters (Pept) and become available for de novo GSH synthesis. The transsulfuration pathway provides an alternate source of cysteine via methionine. GSH is utilized by several major GSH-dependent enzymes; glutathione peroxidases (GPx) to reduce hydroperoxides; glutathione S-transferases (GST) to detoxify xenobiotic compounds by the formation of GSH conjugates (GS-x). These GSH-conjugates, and GSSG, can be exported from the cell via multidrug resistance proteins (MRP). Glutaredoxins (Grx) are primarily responsible for the de-glutathionylation of cysteine residues (PS-SG) present on proteins (P) and they can also reduce protein disulfide bonds, providing a backup for the thioredoxin system (not depicted in figure). The NADPH-dependent thioredoxin system (Trx/TrxR) functions by providing reducing equivalents for peroxiredoxins (Prx).
The abundance of cysteine residues in proteins increases with organism complexity, ranging from ∼0.50% in archaea to ∼2.25% in mammals [38], and is postulated to reflect the evolution of cysteine's signalling and control functions [39,40]. The percentage of cysteine in the Caenorhabditis elegans proteome is 1.97%, which is similar to the more complex eukaryotes [38]. Phylogenetic analysis of the protein-encoding genes of the major redox systems have shown that C. elegans thiol systems possess a considerable amount of similarity with human isoforms [41]. Accordingly, C. elegans has been used extensively to investigate processes related to redox biology, including the role of oxidative stress in aging and disease models (for reviews see Refs. [[42], [43], [44], [45]]). This significant level of evolutionary conservation with higher eukaryotes suggests that C. elegans should be a suitable model to explore signalling mechanisms in the thiol proteome.
4. Caenorhabditis elegans as a model organism
C. elegans has become a prominent model for the study of various aspects of biology due to the ease of genetic manipulation, its completely traced embryonic [46] and post-embryonic cell lineages [47], its short larval lifecycle (∼3 days), and median adult lifespan (∼2–3 weeks) [48]. After embryonic development and hatching, the life cycle of the worm includes a series of four larval stages (L1-L4) before reaching adulthood. Another intermediary stage (dauer) begins to form late in the L1 stage in response to environmental stresses, such as overcrowding or high temperatures which results in the worm entering a state of arrested development [49].
The majority of C. elegans are hermaphrodites with only a small proportion, generally around 0.1%, of males produced each generation [50]. The small size of the nematode (∼1 mm), along with the predominance of self-fertilizing hermaphrodites, makes it a valuable animal model for its ease in maintaining isogenic populations. Breeding males with hermaphrodites enables genetic crossing and is another key advantage of C. elegans as a genetic model. The nematode is optically transparent throughout all stages of its development and lifespan; from zygote to adult [51]. This has allowed the use of fluorescent reporters to investigate the localization patterns of gene expression and their protein products at tissue and subcellular levels in vivo [52]. RNA interference (RNAi) is another well-established method for the temporal silencing of genes to investigate gene function in C. elegans [53]. More recently, CRISPR/Cas9 gene editing has provided a tool for generating heritable modifications of the C. elegans genome [54]. The advent of this technique has allowed for targeted generation of small insertions or deletions (indels), or integration of larger DNA fragments into the genome including single copy exogenous genes, such as those that encode for fluorescent reporters [55]. Application of established and emerging genetic manipulation techniques are key to the success and the continued relevance of C. elegans as a valuable multicellular model organism.
5. Glutathione synthesis in C. elegans
5.1. De novo GSH synthesis pathway
The identification of the GSH cycle thiol intermediates in C. elegans indicated that it has a GSH metabolic network similar to that present in mammals [56]. In whole worm homogenates, the GSH concentrations have been reported to range between 10 and 40 nmol/mg protein [[57], [58], [59], [60], [61], [62]], with the GSH:GSSG ratio of young adult wild-types being approximately 70:1 [57,63]. In mammals, the first reaction of the de novo GSH synthesis pathway is catalyzed by glutamate cysteine ligase (GCL), also known as γ-glutamylcysteine synthetase (GCS), which contains all the substrate and cofactor binding sites to produce γ-glutamylcysteine (γ-GC) [64]. The second reaction is catalyzed by glutathione synthetase to condense γ-GC with glycine to produce GSH [65].
In C. elegans, the gcs-1 gene encodes for the ortholog of the mammalian GCLC (catalytic) subunit. Based on the predicted amino acid sequence homology, C. elegans GCS-1 has a 54% identity to the human GCLC ortholog [41]. The expression of gcs-1 is regulated by the SKN-1 transcription factor in both constitutive and stress-induced conditions in various C. elegans tissues [66]. SKN-1 is a functionally conserved ortholog of Nrf2 that controls the regulation of approximately 300 genes under non-stressed conditions [67,68]. It is also involved in many stress response pathways, and is a central regulator of the nematode's longevity and healthspan [67,69]. In response to oxidative stress, the p38 mitogen-activated protein kinase (p38 MAPK) pathway in C. elegans has been shown to upregulate gcs-1 expression in the intestinal cells via PMK-1-mediated phosphorylation of SKN-1, which leads to the accumulation of SKN-1 into the nuclei [70]. In the absence of oxidative stress, SKN-1 is phosphorylated by the activity of glycogen synthase kinase-3 (GSK-3), which prevents SKN-1 from accumulating in the intestinal cell nuclei, resulting in lower targeted expression of gcs-1 [71].
In the heterozygote mutant gcs-1(ok436) strain, GSH levels were shown to be ∼70% compared to the wild-type [72]. Homozygosity for gcs-1(ok436) null allele is lethal at the L2 larval stage [11]. The dynamic temporal impact of the gcs-1 gene has been shown in a study where RNAi targeting gcs-1 was conditionally induced in worms either immediately after hatching or at the start of adulthood [58]. When the gcs-1 gene was knocked down immediately after worm hatching, no difference was observed in median lifespan, though a higher incidence of vulval rupture and a significantly lower resistance against paraquat exposure was observed [58]. Conversely, knockdown of gcs-1 at the beginning of adulthood, showed an increase in lifespan and paraquat stress resistance without the vulval rupture phenotype. This observation of increased lifespan and stress resistance may have been afforded by the compensatory increase (∼2-fold) in expression levels of several genes of the thioredoxin system (trx-1, trxr-1, trxr-2) and glutathione S-transferase gene, gst-4 [58]. Other studies have shown that, whilst gcs-1 mRNA levels do not significantly change up to day 12 adults compared to young adults, induction of gcs-1 expression in response to paraquat exposure diminishes in older worms [73], suggesting that changes in skn-1 induction in response to oxidative stress may decline during aging.
In mammals, heterodimer formation between the GCLC and GCLM (modifier) subunits results in the GCL holoenzyme complex which has increased catalytic efficiency over the GCLC subunit alone [74]. The E01A2.1 gene in C. elegans encodes the ortholog of the mammalian GCLM regulatory subunit [68,75]. Under conditions of arsenite-induced stress, the E01A2.1 gene is upregulated by the SKN-1 transcription factor [68,76]. An early investigation of the protein-protein interactions of C. elegans using yeast two-hybrid (Y2H) screens, showed that the E01A2.1 protein interacts directly with the GCS-1 protein [77]. A later RNAi screen of a library of 11,511 genes determined that the silencing of a total of 37 C. elegans genes, one of which included E01A2.1, induced gcs-1 expression [78], suggesting that the down regulation of the modifier subunit increases the expression of the catalytic subunit (gcs-1) in vivo.
Others have shown that RNAi knockdown of the E01A2.1 gene in wild-type worms had no effect on lifespan under 1.5 mM paraquat challenge, whereas it decreased the lifespan of the long-lived daf-2(e1370) mutant [79]. The authors reported that the daf-2(e1370) L4 larvae had GSH levels 14-fold higher than the wild-type [79]. Microarray analysis of the genes involved in GSH synthesis, including the salvage synthesis pathway, did not show any significant changes in the abundance of mRNA levels in the daf-2(e1370) worms that could explain the high GSH levels [79]. It was postulated that the discrepancy between the high GSH levels and no observable difference in the transcript levels of the genes involved in GSH synthesis may be due to regulatory mechanisms at the post-transcriptional level [79].
The enzyme involved in the second step in the de novo GSH synthesis pathway, glutathione synthetase, is encoded by the gss-1 gene in C. elegans, which shares 39% identity with the human isoform [41]. Initial work predicted several SKN-1 binding sites to be present in the gss-1 gene promoter region [66]. Later studies have shown that the gss-1 gene is upregulated by SKN-1 in response to oxidative stress following arsenite [68] and benzo-α-pyrene exposure [80]. RNAi knockdown of gss-1 has no effect on the lifespan of the wild-type or daf-2 mutants under non-stressed conditions [79]. Knockdown of gss-1 has also been shown to increase the fluorescent intensity of the gcs-1::GFP reporter, indicating a regulatory control mechanism exists to increase gene expression for the first step of the GSH biosynthesis pathway when the second step becomes compromised [81].
5.2. GSH salvage synthesis pathways
The enzymatic recycling of GSH from GSSG by glutathione reductase maintains the cellular GSH:GSSG ratio [37]. In C. elegans, the gsr-1 gene encodes for the glutathione reductase enzyme, which produces two protein isoforms (GSR-1a and GSR-1b) [82]. Similar to the genes of the de novo GSH synthesis pathway, expression of gsr-1 is modulated by the SKN-1 transcription factor [57]. The GSR-1b isoform is located in the cytoplasm, whereas the GSR-1a isoform possesses an additional 14 amino acid N-terminal extension which is predicted to include a mitochondrial targeting sequence (MTS) [82]. Isoform-specific rescue of the embryonic lethal phenotype of gsr-1(tm3574) mutants, indicated that lethality is prevented by expressing the GSR-1b cytoplasmic form and not the GSR-1a mitochondrial form, demonstrating that GSR-1b is essential for embryonic development [82]. RNAi knockdown of the gsr-1 gene in wild-type worms has been reported to decrease lifespan under non-stressed conditions [57]. The potential role of gsr-1 in lifespan is supported by findings that demonstrate homozygous gsr-1 mutants with maternally contributed GSR-1 are relatively short-lived compared to wild-types [82]. However, other researchers have observed no difference in lifespan when the gsr-1 gene was knocked down [59]. The gsr-1 gene is vital in the stress response against several oxidants including juglone, cumene hydroperoxide, diamide, tert-butyl hydroperoxide and paraquat [57,79,82,83].
Extracellular degradation of GSH by the membrane-bound γ-glutamyl transferase (γ-GT) involves the transfer of GSH's γ-glutamyl moiety to a free amino acid, releasing the dipeptide cysteinylglycine in the process. The cysteinylglycine is hydrolyzed by dipeptidases located on the outer surface of the cell membrane to generate cysteine and glycine which are then taken up by the cell [15] where they contribute to the substrate pool for the GSH de novo synthesis pathway (Fig. 1). The detection of cysteinylglycine in C. elegans initially suggested the presence of the γ-GT ectoenzyme [56], with subsequent work reporting at least six γ-GT genes [57]. In eukaryotes, GSH can be exported from the cell as glutathione-S-conjugates and disulfide forms mainly via multidrug resistance proteins (MRPs) [84,85]. Experimental approaches are yet to determine if the mechanisms of cellular efflux of glutathione-S conjugates and GSSG in C. elegans are analogous to mammalian systems.
The conservation of GSH synthesis between C. elegans and mammals, including the transcriptional control by the Nrf2 ortholog, SKN-1, make C. elegans a useful multicellular model organism in studies aimed at exploring the mechanisms of glutathione homeostasis in humans (Fig. 2).
Fig. 2.
Homology of the GSH synthesis pathways in mammals and Caenorhabditis elegans. In mammals, the de novo synthesis pathway is comprised of the glutamate cysteine ligase catalytic (GCLC) and modifier subunits (GCLM), and glutathione synthetase (GS); with glutathione reductase (GR) involved in recycling intracellular GSSG to GSH. In worms, the respective orthologs are designated as glutamate cysteine synthetase heavy (GCS-1) and light (E01A2.1) subunits, glutathione synthetase (GSS-1) and glutathione reductase (GSR-1). Genes of the de novo (gcs-1, E01A2.1 and gss-1) and recycling pathway (gsr-1) are transcriptionally upregulated by the SKN-1 (Nrf2 ortholog) under constitutive and stress-induced conditions.
5.3. In vivo monitoring of glutathione homeostasis in C. elegans
In eukaryotes, the redox environment differs throughout the various organelles and sub-cellular compartments with the intracellular glutathione redox couple distributed heterogeneously throughout the cytosol [86,87]. Recent advances have seen the physical transparency of the nematode utilized in the development of redox biosensors that allow the visualization of redox status, which facilitates the resolution of tissue-specific differences in real-time [88]. The genetically encoded Grx1-roGFP2 fluorescent biosensor was the first redox-sensitive ratiometric reporter to enable in vivo measurement of the GSH:GSSG ratio in C. elegans [89]. Using the genetically encoded Grx1-roGFP2 sensor, it was shown that the GSH:GSSG ratio shifts to more reducing conditions during larval development [89]. In a later study, an alternative encoded redox sensor, roGFP1-R12 was used to investigate the oxidation state of protein thiols in several tissue types including, the intestine, pharynx and the PLM mechanosensory neurons [11]. Using roGFP1-R12 transgenic worms, the authors concluded that the GSH/GSSG couple amplifies small changes in its oxidation state to large changes in redox potential by the oxidation of cysteine residues in proteins, rather than having a direct buffering role on the cellular redox potential [11]. The authors show that significant differences in the glutathione redox potentials exist between isogenic worm populations and suggest that this may be due to the variation in abundance of GSSG [11].
Due to the high intracellular concentrations of GSH (1–10 mM) in aerobic organisms, GSH biosynthesis pathways must function efficiently to maintain adequate levels of this important low molecular-weight thiol. Reports describing the role of GSH in maintaining the cellular redox environment often understate the activities of the GSH-dependent enzyme network and related thiol systems. The two major GSH-dependent enzyme families present in C. elegans are the glutathione peroxidases and glutathione S-transferases. These and other key oxidoreductase enzymes, including glutaredoxins, thioredoxins and peroxiredoxins are also discussed in the context of GSH's role in redox signalling in the C. elegans model.
6. Glutathione peroxidases in C. elegans
The glutathione peroxidase (GPx) family utilize GSH for the reduction of H2O2 and a variety of organic hydroperoxides to water or to the corresponding alcohol [90]. They are categorized into two subfamilies; the selenium subfamily containing a selenocysteine (SeCys) in the catalytic triad of amino acids (Trp-Glu-SeCys) and the non-selenium subfamily containing a cysteine (Trp-Glu-Cys) [91]. Sequence analysis of the eight C. elegans GPx genes (gpx-1 to gpx-8) revealed that none contain a selenocysteine [92]. Selenocysteine-containing GPxs have been the subject of many mammalian studies, yet the physiological roles of the non-selenium GPxs remain largely unclear. Though the majority of C. elegans GPx isoforms contain the Trp-Glu-Cys catalytic triad, the predicted amino acid sequence of GPX-4 lacks the Trp residue, and GPX-8 lacks both the Trp and Cys [92]. It was concluded that the absence of cysteine at the catalytic site indicates that the GPX-8 protein is likely not acting as a glutathione peroxidase [92].
Glutathione peroxidase activity in worms has been shown to decline when exposed to high glucose concentrations [62], paraquat [93] and the neurotoxin 6-hydroxydopamine (6-ODHA) [94], yet increase in response to copper exposure (CuSO4) [95]. In addition to their notable absence of selenium, the structures of the GPx genes differ considerably in the worm compared to human isoforms. Sequence identity analysis of the predicted amino acid sequences has indicated that four of the C. elegans GPx isoforms (GPX-1, GPX-2, GPX-6 and GPX-7) share between 43 and 47% identity to the human phospholipid hydroperoxide glutathione peroxidase (GPx4) [92]. A single strain generated with deletions for the four phospholipid hydroperoxide GPx isoforms had no difference in median lifespan, though their age-specific mortality rates and levels of lipid hydroperoxides were higher when compared to the wild-types [92]. Each of the four C. elegans phospholipid hydroperoxides GPx genes are reported to be expressed primarily in the intestine, with gpx-2 also expressed in several neurons located in the head and tail [92]. The expression levels of gpx-6 were also observed to increase in starved worms at several of the larval stages tested, suggesting it may have a role in dietary restriction [92]. In an RNAi screening study of 162 genes that are exclusively expressed in the intestine of the worm, gpx-1 was the only RNAi-targeted gene shown to increase expression and activity of PEPT-1 [96]. PEPT-1 is a transporter, located on the apical membrane of the enterocyte, responsible for the uptake of di- and tripeptides [97]. This suggests that lower gpx-1 transcript levels may stimulate uptake of dietary glutathione and its precursors.
The non-phospholipid hydroperoxide GPx genes (gpx-3, gpx-4, gpx-5, gpx-8) encoded in the C. elegans genome are reported to share putative homology with the human GPx3 and GPx5 isoforms [92]. Few studies have reported on the function of the non-phospholipid hydroperoxide GPx genes in C. elegans. However, gpx-5 has been shown to play a role in arsenite stress defence [57], and is potentially involved in synaptic remodelling [98], and circadian-stress tolerance in C. elegans [99]. The absence of selenocysteine at the active site of any of the C. elegans GPx isoforms suggests that the major mechanistic peroxidase activity of this enzyme family differs to that in mammals.
7. Glutathione-S-transferases in C. elegans
Glutathione S-transferases (GSTs) are a functionally diverse family of enzymes that utilize GSH in conjugation reactions [100]. GSTs represent a major class of enzymes that are involved in phase II detoxification in C. elegans [101]. The C. elegans genome contains 56 validated and putative GST genes [102] with most categorized within the sigma subfamily, though alpha, kappa, pi, zeta, and omega subfamilies are also present [103,104]. Changes in expression levels of specific GSTs in response to certain xenobiotic compounds suggest substrate-specific roles of GSTs in xenobiotic detoxification [80,105,106]. The correlation between GST activity and longevity has been described among evolutionary divergent species, including C. elegans and Drosophila melanogaster [107,108]. Several proteomic studies have indicated that C. elegans GST isoforms have a multitude of functions, and are involved in cell signal pathways via direct protein-protein interactions under H2O2-stressed and non-stressed conditions [[109], [110], [111], [112]].
7.1. Sigma and pi GST subfamilies
The most rigorously investigated GST gene in C. elegans is gst-4, which shares sequence similarity with the human sigma subfamily. An early study demonstrated that gst-4 expression increases in response to paraquat challenge [113]. Overexpression of gst-4 was later shown to increase stress-resistance against juglone and paraquat exposure, without any effect on lifespan under normal growth conditions [114]. With the advent of gst-4::GFP reporter strains [115], many studies have investigated the effect that certain compounds have on the expression of gst-4 (see Table 1). Most compounds elicit an increase in the intensity of the gst-4::GFP reporter in a stress-dependent manner. In contrast, 25 μM doses of folic acid treatment have been reported to increase the expression of gst-4, resulting in an increase in stress resistance [116]. Expression patterns of the gst-4::GFP reporters have been shown to localize in the intestine [117], body wall muscles, hypodermis, cells surrounding the pharynx, and the pharynx in L3 larvae only [114]. Among these tissues, the intestine is a prominent site of increased gst-4::GFP induction in response to exposure to chemical and oxidative stress [[117], [118], [119], [120]].
Table 1.
Compounds that affect gst-4 (glutathione-S-transferase-4) expression in C. elegans.
Compound | gst-4 expression | Refs |
---|---|---|
acrylamide | ↑ | [129,190] |
aspirin | ↑ | [191] |
fluoxetinea | suppressed | [192] |
folic acid | ↑ | [116] |
3β-hydroxy-urs-12-en-28-oic acid | ↑ | [106] |
juglone | ↑ | [119,193,194] |
lithium compounds (LiCl, Li2CO3)b | suppressed | [195] |
microplastic particles | ↑ | [120] |
mianserina | ↑ | [192] |
mirtazapine | ↑ | [192] |
3-nitropropionic acid | ↑ | [196] |
paraquat | ↑ | [113] |
quinolinic acid | ↑ | [196] |
tributyltin | ↑ | [117] |
Co-treatment with fluoxetine suppressed the mianserin-induced increase of the gst-4::GFP reporter, though had negligible effect on reporter activity when treated with fluoxetine alone.
Expression levels measured using microarray analysis.
Transcriptional activation of the gst-4 gene is often used as an indicator of SKN-1 activity; with several inducers of gst-4 including H2O2 and sodium azide being shown to be skn-1 dependent [118]. SKN-1 also controls the induction of several other GST genes [66,68,69]. Induction of gst-4 has also recently been shown to be upregulated independent of SKN-1, by the EOR-1 transcription factor that mediates the effects of the epidermal growth factor (EGF) pathway involved in regulating cell growth and differentiation [121]. The authors concluded that the discovery that gst-4 transcriptional regulation is not exclusively controlled by the SKN-1 transcription factor should serve as a cautionary note for work that utilizes the gst-4::GFP reporter as an indicator of SKN-1 activity [121].
An RNAi screening study of 27 of the 44 then identified C. elegans GST genes showed that knockdown of five, which included one pi (gst-10) and four sigma isoforms (gst-5, gst-6, gst-8, and gst-24), sensitized the worms to exposure to the lipid peroxidation product, 4-hydroxyneoneal (4-HNE) [122]. Of the five, knockdown of gst-5 and gst-10, resulted in a shortened lifespan [122]. The pi class gene, gst-10, is involved in catalyzing the conjugation of GSH to 4-HNE in C. elegans, suggesting a functional role of this pi class GST in dealing with cellular stress [[122], [123], [124]]. Another pi class GST homologue, gst-1, has also been shown to protect against dopaminergic neuron degeneration in C. elegans [125], which may have implications for understanding the progression of Parkinson's Disease. In a later RNAi screening study, over 40 GST genes were tested for survival against juglone and arsenite exposure [57]. For juglone, gsto-1 was the only GST gene that showed a significant decrease in survival when knocked down. Though the mean survival against arsenite treatment for the majority of the silenced GST genes showed a modest decrease, the knockdown of only gst-32 and gst-44 significantly decreased survival [57].
7.2. Kappa GST subfamily
Unlike most vertebrates, which encode only one GST kappa gene, C. elegans encodes for two; gstk-1 and gstk-2 [104,126]. GSTK-1 is expressed in the intestine, rectal gland cells, body wall muscles and epidermis and contains a C-terminal peroxisomal-targeting sequence which directs its localization to the peroxisome [126]. The GSTK-2 is expressed in the pharynx, body wall muscles and the intestine and is localized in the mitochondria of these tissues [126]. RNAi knockdown of gstk-1 and gstk-2 were reported to have no effect on several phenotypes including, reproduction, development, motility and lifespan [126]. However, simultaneous double RNAi knockdown of both genes resulted in a significant decline in oxygen consumption and cis-vaccenic acid content which did not occur when the genes were silenced individually [126]. The authors concluded that both genes are involved in respiration and lipid metabolism [126].
7.3. Omega GST subfamily
Several isoforms of the GST omega subfamily are encoded in the C. elegans genome, including gsto-1, gsto-2, gsto-3 and the putative GST omega genes gst-44 and C02D5.4 [102]. Overexpression and RNAi studies have shown gsto-1 to be implicated with increased stress resistance and to be exclusively expressed in the intestine [127]. Intestinal expression of gsto-1 was initially shown to be regulated by the ELT-2 transcription factor [127]. Gsto-1 is upregulated under transient hypoxic conditions leading to an extension of lifespan mediated by the mechanistic Target of Rapamycin (mTOR) signalling pathway; a longevity pathway associated with dietary restriction [128].
The GST superfamily performs a diverse range of functions and have considerable variation in expression patterns in C. elegans [129]. While there remains much to be elucidated about the substrate-specificities, expression patterns and the relative levels of the many GST isoforms in C. elegans, research into the dynamic roles of this enzyme superfamily in detoxification and signalling will further our understanding on how the redox environment is mediated by enzymatically-controlled GSH-dependent redox processes.
8. Glutaredoxins in C. elegans
Glutaredoxins can serve as a backup for the thioredoxin system by catalyzing the reversible reduction of target protein disulfides [130]. These relatively small enzymes, which range between 9 and 15 kDa [131], are also regarded as the main catalysts of reversible protein de-glutathionylation [28]. Depending on the number of cysteine residues in the active site of the enzyme, glutaredoxins are classified as either monothiol or dithiol forms [132]. For the C. elegans glutaredoxin genes, five have been annotated; glrx-3, glrx-5, glrx-10, glrx-21, glrx-22 [57]. Analysis of the glutaredoxin sequences indicates that the GLRX-3 and GLRX-5 are monothiol forms, while the GLRX-10, GLRX-21 and GLRX-22 are dithiols [41]. Interestingly, while GLRX-5, GLRX-10, GLRX-21 and GLRX-22 are formal glutaredoxins ranging between 100 and 140 amino acids, GLRX-3 is a much larger isoform at 345 amino acids. Based on phylogenetic analysis, the C. elegans GLRX-3 isoform has been postulated to be an ortholog of the mammalian GLRX3 PICOT (protein kinase C-interacting cousin of thioredoxin) [41]. Except for glrx-5 which is predicted to be a mitochondrial glutaredoxin [133], the other annotated glutaredoxins are expected to be located in the cytosol. This is particularly interesting as the National BioResource Project (NBRP) reports the glrx-5(tm3867) deletion allele to be the only lethal or sterile glutaredoxin mutant strain [134], while the cytoplasmic glutaredoxin mutants including, glrx-10(tm4634) [135], glrx-21(tm2921) [136], and glrx-22(tm3743) are all reported as viable.
Only a few published reports are available that investigate the role of the dithiol glrx-10 and glrx-21 genes in C. elegans. In a transgenic C. elegans model of Parkinson's Disease which expresses two pathogenic mutations in the human leucine-rich repeat kinase 2 (LRRK2) gene, loss of the glrx-10 gene has been shown to lead to an increase in the degeneration of dopaminergic neurons [135,137]. Rescue of the glrx-10 function mutants by re-expression of the active form of the wild-type C. elegans glrx-10 gene partially afforded protection in dopaminergic neurons [135]. Two separate studies have shown that loss-of-function of glrx-21 results in selenium-induced decline of motility [136] and reproduction [138]. Both studies suggested that GLRX-21 is required for the partial protection afforded by GSH during selenium-induced impairment [136,138].
To date, there remains only these limited reported studies of glutaredoxins in C. elegans, which provides opportunities for future work to characterize the function of these genes. A recent report has demonstrated that C. elegans exhibit changes in protein S-thiolation patterns (i.e. S-glutathionylation and S-cysteinylation) of targeted cysteine residues [139]. This evidence suggests that glutaredoxins may provide an evolutionary conserved mechanism for catalyzing the reversal of S-glutathionylation in C. elegans. Functional characterization of the C. elegans GLRX isoforms may offer new insights into the role of S-glutathionylation to determine how central this post-translational modification is in the context of cellular redox signalling in vivo.
9. Thioredoxin systems in C. elegans
The thioredoxin system is comprised of thioredoxins (Trx) and NADPH-dependent thioredoxin reductases (TrxR) [140]. Thioredoxins are small (∼12 kDa) ubiquitous oxidoreductases that contain a highly specific thiol-disulfide active site to co-ordinate the regulation of the cellular redox environment, which is largely achieved by the supply of reducing equivalents for peroxiredoxins [141]. Oxidized thioredoxins (Trxox) are reduced (Trxred) by thioredoxin reductases (TrxR), utilising NADPH as a cofactor [7]. The C. elegans thioredoxin system is composed of five thioredoxin genes (trx-1 to trx-5) and two thioredoxin reductase genes (trxr-1 and trxr-2) [142].
The trx-1 gene is specifically expressed in the ASJ neurons [143]. Discovery of an ASJ motif, a functional cis-regulatory promoter region, has been reported to regulate the ASJ-specific gene expression of trx-1 by binding SPTF-1, an ortholog to the Sp family zinc-finger transcription factor [144]. Mutant worms that lack the trx-1 gene are more vulnerable to paraquat-induced oxidative stress [145] and have decreased lifespans [143,145]. TRX-1 has been shown to regulate lifespan extension in genetic and nutrient-based models of dietary restriction possibly via its upregulation in ASJ neurons [146]. Interestingly, trx-1 was later shown to regulate SKN-1 nuclear localization in a cell non-autonomous manner, with loss of trx-1 from ASJ neurons promoting the nuclear localization of intestinal SKN-1 [147]. TRX-1 has also been implicated to potentially have a mechanistic role in dauer formation via the down-regulation of the insulin-like DAF-28 signalling neuropeptide in ASJ neurons [148]. TRX-1 has also been shown to have a role in C. elegans avoidance behavior by coordinating a dynamic trans-nitrosylation/de-nitrosylation response to nitric oxide produced by Pseudomonas aeruginosa [149]. Recent studies have reported that trx-1 is involved in protection against methylmercury toxicity of dopaminergic neurons in males, but not hermaphrodites [150]. However, trx-1, in combination with trxr-1, can provide protection against methylmercury in aging hermaphrodites [151].
The trx-2 and trxr-2 genes comprise the mitochondrial thioredoxin system and are upregulated upon induction of the mitochondrial unfolded protein response (UPRmt) [152,153]. Characterization of the intestine-specific TRX-3 have shown that the trx-3(tm2820) mutant has no significant difference in reproductive capacity, longevity, and resistance to stress (including heat-treatment, juglone and paraquat exposure) compared to the wild-type, though they exhibit a decrease in physical size and a shorter timing of their defecation cycle [154]. Overexpression of TRX-3 protected against pathogen infection, which suggested that TRX-3 may have a role in the worm's innate immune response [154]. The lesser studied C. elegans thioredoxin genes, trx-3, trx-4 and trx-5 have been shown to afford no significant protection in dopaminergic neurons against methylmercury toxicity [151]. However, there is evidence to suggest that trx-5 may have a possible protective role against dopaminergic neuron loss in worms treated with the neurotoxin, 6-ODHA [142].
In addition to the five annotated thioredoxin genes (trx-1 to trx-5), there are four other genes, which encode for proteins with highly conserved thioredoxin domains that contain the CGPC (Cys-Gly-Pro-Cys) active site [154]. These include the two annotated dnj-27 and png-1 genes and the uncharacterized Y55F3AR.2 and txl-1 genes [154]. DNJ-27 is an ortholog of the mammalian endoplasmic reticulum (ER) localized ERdj5 protein, with both containing four thioredoxin-like domains [155]. The dnj-27 gene is highly expressed in the pharynx and vulva, with lower levels of expression in body wall muscles, intestine, gonadal sheath cells, rectum and hypodermis [155]. The dnj-27 gene has been shown to provide a protective role against pathological phenotypes in several C. elegans neurodegenerative disease models [155]. The png-1 gene encodes for a bifunctional enzyme that possesses deglycosylation enzyme activity in addition to its oxidoreductase activity [156,157]. The N-terminal thioredoxin domain of the PNG-1 protein is understood to be unique to C. elegans [156]. Higher eukaryotic organisms possess a PUB domain at the N-terminus, which is implicated to mediate protein-protein interactions in the ubiquitin-proteasome pathway [158]. In vivo C. elegans studies have shown that mutations in the thioredoxin domain of the png-1 gene results in an increase in axon branching defects [159]. Further investigations of the lesser studied thioredoxin genes should be of interest for future C. elegans studies.
Protein structure studies of the two C. elegans thioredoxin reductases have determined the TRXR-1 (∼74 kDa) to be 667 amino acids and the TRXR-2 (∼55 kDa) to be 503 amino acids [160]. The TRXR-1 protein is located in the cytosol and is highly expressed in the pharynx, hypodermis, rectal epithelial cells, intestine, nervous system [161] and vulva [162], whilst the mitochondrial thioredoxin reductase, TRXR-2, is expressed in the intestine and in several neurons located in the head [152]. Both the trxr-1 and gsr-1 (glutathione reductase) genes work in conjunction during C. elegans larval development by the reduction of disulfide bonds to carry out precise removal of the cuticle from the surface of epidermal cells [161].
Interestingly, trxr-1 is the only selenoprotein that has been detected in C. elegans [161,163]. Having only one selenoprotein has made C. elegans a valuable in vivo model to investigate dose-dependent beneficial and toxicological properties of selenium-based compounds [63,[164], [165], [166]]. The CRISPR/Cas9 system has been used to generate targeted mutations of the trxr-1 selenocysteine [167]. Two such trxr-1 variants; one with a point mutation where the selenocysteine residue was replaced by a cysteine and the other with a premature stop codon, showed resistance during development when exposed to the chemotherapeutic agent, cisplatin [167]. The trxr-1 gene affords partial protection from degeneration of dopaminergic neurons against the neurotoxin 6-hydroxydopamine (6-OHDA) [142]. Studies using the nematode could provide insights into the physiological roles of selenocysteine-containing proteins based on investigations focusing on the trxr-1 selenoprotein.
10. Peroxiredoxins in C. elegans
Peroxiredoxins (Prx) are a class of thiol peroxidases that scavenge organic and inorganic peroxides. They are highly abundant proteins involved in the conversion of the majority of cellular H2O2 to water [168,169]. C. elegans contain three known peroxiredoxin genes, prdx-2, prdx-3, and prdx-6 [44]. Both prdx-2 and prdx-3 encode for typical 2-Cys peroxiredoxins, with prdx-6 encoding for a 1-Cys peroxiredoxin [170]. Expression patterns of prdx-2 have been reported in the I4 (pharyngeal) and I2 (sensory) interneurons [170], intestine, epithelial cells, muscle (pharyngeal, vulval and body wall), and various neurons in the head and tail [171].
In C. elegans, PRDX-2 is highly abundant, constituting ∼0.5% of the total protein and is involved in protecting protein thiol groups from H2O2-induced oxidative modifications [172]. In prdx-2(gk169) mutants, loss of peroxiredoxin activity leads to an increase in de novo GSH synthesis compared to wild-types, with an ∼2-fold increase in gcs-1 mRNA and in GSH levels [173], possibly indicating a compensatory response by the de novo GSH synthesis network. Loss-of-function prdx-2(gk169) mutants have a shorter lifespan compared to wild-types at 15 °C [172,173] and 20 °C [173], with no difference observed at 25 °C [172], indicating a dynamic temperature-dependent response for the PRDX-2 enzyme. This is supported by observations that transient changes in housing temperature during development increase lifespan through a PRDX-2-dependent stress response [174]. Moreover, survival rates of prdx-2(gk169) mutants exposed to 10 h heat-shock treatments at 35 °C show a marked decline in lifespan when compared to wild-types [175].
Over-oxidation of PRDX-2 has been contested to have a negligible effect on worm physiology, with the suggestion that C. elegans perhaps relies on degradation and subsequent clearance of over-oxidized PRDX-2; with de novo synthesis of PRDX-2 potentially playing a role in the physiological response [176]. The cyclic oxidation state of the PRDX-2 protein possibly represents a circadian-rhythm mechanism in the worm [177] which is conserved in other eukaryotic models [178].
Several studies have shown that PRDX-2 is involved in hormetic responses. Exposure of worms to low doses of H2O2 (0.01–1 μM) resulted in an increased sensory response of the ASH neurons which required the PRDX-2-mediated p38/PMK-1 signalling cascade [179]. PRDX-2 is also involved in the hormetic response leading to lifespan extension, where worms treated with metformin, the antihyperglycemic drug used in the treatment of type II diabetes, led to an increase in endogenous levels of H2O2 [180]. Interestingly, both increased and loss of expression of PRDX-2 increased stress resistance against 5 mM arsenite exposure [173]. A recent study has reported that PRDX-2 is required for insulin secretion (in the form of the DAF-28 neuropeptide), leading to higher activity of the DAF-2 insulin signalling pathway and inhibition of the nuclear localization of SKN-1 and DAF-16 [181]. This later study provides an explanation for the paradoxical observation seen in the loss-of-function prdx-2(gk169) mutants that exhibit increases in arsenite stress resistance [181].
The lesser studied prdx-3 gene is predicted to be localized in the mitochondria [182,183]. Though RNAi-targeted prdx-3 knockdown worms had no difference in lifespan, they exhibited increased mitochondrial uncoupling, with lower ATP levels, motility, and reproduction, indicating that while prdx-3 silencing does not affect lifespan under these conditions, it is important for healthspan [183]. Work on C. elegans peroxiredoxins have focused primarily on the typical 2-Cys peroxiredoxin, prdx-2. This is perhaps due to an early report that showed no significant difference in the longevity, development and progeny production phenotypes in the RNAi knockdown of the prdx-3 (2-Cys) and the prdx-6 (1-Cys) genes [170].
11. Conclusions
The anatomical simplicity of C. elegans, along with the ease of genetic modification and the ability to perform in vivo studies, has made it a successful research tool for investigating complex biological processes in multicellular organisms. In addition to the advantages of the nematode model itself, several open access resources including, WormBase [184], WormBook [185], and WormAtlas [186], provide comprehensive information on C. elegans genetics, biology and structural anatomy. Moreover, the two main strain repositories, the Caenorhabditis Genetics Center (CGC) and the National BioResource Project (NBRP), offer an extensive collection of mutant strains to service the C. elegans research community.
The advantage of studying redox systems in the simple nematode model is the capacity to investigate tissue and cell-specific differences in expression patterns in vivo. Most of what is understood about the gene expression patterns in C. elegans has been achieved through the generation of transcriptional and translational fluorescent reporter transgenes. Understanding the tissue and organelle-specific localization patterns of the thiol-related systems in C. elegans may provide insight into their roles in maintaining the redox environment in other multicellular organisms (Fig. 3).
Fig. 3.
Expression patterns and subcellular localization of select thiol-related systems in Caenorhabditis elegans. The tissue- and subcellular-specific differences in expression patterns of thiol-related genes in the worm. Generation of GFP reporter strains, allows the investigation of tissue, cell and organelle-specific gene expression patterns. Within the same tissue types and subcellular compartments (e.g. mitochondria), localized gene expression of the thiol-related system can control the redox environment of respective compartments. Note: the listed genes are not exhaustive of all known localization expression patterns of the thiol-related systems (see WormBase resource for comprehensive descriptions of all identified expression patterns).
It has recently been suggested that C. elegans could potentially be the subject of the first complete animal model of redox signal transduction [45]. Advances in proteomics techniques hold promise to further our understanding of how post-translational modifications influence signalling events of the thiol proteome [187]. Techniques such as OxiCat, which measures the reversible oxidation of cysteine residues have demonstrated the suitability of C. elegans for redox proteomic studies [172,188]. Others have shown that proteomic approaches that measure cysteine reactivity can be used as a predictor of protein functionality in C. elegans [189]. In terms of the current knowledge of the reaction mechanisms of glutathione-dependent enzymes (for review see Ref. [37]), future efforts should employ C. elegans to gain further understanding of the in vivo functional role of GSH in the context of thiol-mediated redox signalling.
Declaration of interest
GDF and WJB declares no conflict of interest.
Acknowledgements
GDF has been the recipient of an Australian Postgraduate Award Scholarship. The authors would like to acknowledge Myung-Jin Kang and Nicholas Bentley for their comments on the manuscript.
References
- 1.Schafer F.Q., Buettner G.R. Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radic. Biol. Med. 2001;30:1191–1212. doi: 10.1016/s0891-5849(01)00480-4. [DOI] [PubMed] [Google Scholar]
- 2.Kemp M., Go Y.M., Jones D.P. Nonequilibrium thermodynamics of thiol/disulfide redox systems: a perspective on redox systems biology. Free Radic. Biol. Med. 2008;44:921–937. doi: 10.1016/j.freeradbiomed.2007.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Jones D.P. Redox theory of aging. Redox Biol. 2015;5:71–79. doi: 10.1016/j.redox.2015.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Go Y.M., Jones D.P. Thiol/disulfide redox states in signaling and sensing. Crit. Rev. Biochem. Mol. Biol. 2013;48:173–181. doi: 10.3109/10409238.2013.764840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Herrmann J.M., Dick T.P. Redox biology on the rise. Biol. Chem. 2012;393:999–1004. doi: 10.1515/hsz-2012-0111. [DOI] [PubMed] [Google Scholar]
- 6.Jones D.P., Sies H. The redox code. Antioxidants Redox Signal. 2015;23:734–746. doi: 10.1089/ars.2015.6247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Rebrin I., Sohal R.S. Pro-oxidant shift in glutathione redox state during aging. Adv. Drug Deliv. Rev. 2008;60:1545–1552. doi: 10.1016/j.addr.2008.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Hanschmann E.M., Godoy J.R., Berndt C., Hudemann C., Lillig C.H. Thioredoxins, glutaredoxins, and peroxiredoxins--molecular mechanisms and health significance: from cofactors to antioxidants to redox signaling. Antioxidants Redox Signal. 2013;19:1539–1605. doi: 10.1089/ars.2012.4599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Pollak N., Dolle C., Ziegler M. The power to reduce: pyridine nucleotides--small molecules with a multitude of functions. Biochem. J. 2007;402:205–218. doi: 10.1042/BJ20061638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Berndt C., Lillig C.H., Flohe L. Redox regulation by glutathione needs enzymes. Front. Pharmacol. 2014;5:168. doi: 10.3389/fphar.2014.00168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Romero-Aristizabal C., Marks D.S., Fontana W., Apfeld J. Regulated spatial organization and sensitivity of cytosolic protein oxidation in Caenorhabditis elegans. Nat. Commun. 2014;5:5020. doi: 10.1038/ncomms6020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Fahey R.C., Brown W.C., Adams W.B., Worsham M.B. Occurrence of glutathione in bacteria. J. Bacteriol. 1978;133:1126–1129. doi: 10.1128/jb.133.3.1126-1129.1978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Copley S.D., Dhillon J.K. Lateral gene transfer and parallel evolution in the history of glutathione biosynthesis genes. Genome Biol. 2002;3 doi: 10.1186/gb-2002-3-5-research0025. research0025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Pompella A., Corti A. Editorial: the changing faces of glutathione, a cellular protagonist. Front. Pharmacol. 2015;6 doi: 10.3389/fphar.2015.00098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Lu S.C. Regulation of glutathione synthesis. Mol. Aspect. Med. 2009;30:42–59. doi: 10.1016/j.mam.2008.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Forman H.J., Zhang H., Rinna A. Glutathione: overview of its protective roles, measurement, and biosynthesis. Mol. Aspect. Med. 2009;30:1–12. doi: 10.1016/j.mam.2008.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Zhang H., Forman H.J. Glutathione synthesis and its role in redox signaling. Semin. Cell Dev. Biol. 2012;23:722–728. doi: 10.1016/j.semcdb.2012.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Townsend D.M., Tew K.D., Tapiero H. The importance of glutathione in human disease. Biomed. Pharmacother. 2003;57:145–155. doi: 10.1016/s0753-3322(03)00043-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Aquilano K., Baldelli S., Ciriolo M.R. Glutathione: new roles in redox signaling for an old antioxidant. Front. Pharmacol. 2014;5:196. doi: 10.3389/fphar.2014.00196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Flohe L. The fairytale of the GSSG/GSH redox potential. Biochim. Biophys. Acta. 2013;1830:3139–3142. doi: 10.1016/j.bbagen.2012.10.020. [DOI] [PubMed] [Google Scholar]
- 21.Winterbourn C.C. Are free radicals involved in thiol-based redox signaling? Free Radic. Biol. Med. 2015;80:164–170. doi: 10.1016/j.freeradbiomed.2014.08.017. [DOI] [PubMed] [Google Scholar]
- 22.Schieber M., Chandel N.S. ROS function in redox signaling and oxidative stress. Curr. Biol. 2014;24:R453–R462. doi: 10.1016/j.cub.2014.03.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Winterbourn C.C., Hampton M.B. Thiol chemistry and specificity in redox signaling. Free Radic. Biol. Med. 2008;45:549–561. doi: 10.1016/j.freeradbiomed.2008.05.004. [DOI] [PubMed] [Google Scholar]
- 24.Jacob C., Giles G.I., Giles N.M., Sies H. Sulfur and selenium: the role of oxidation state in protein structure and function. Angew Chem. Int. Ed. Engl. 2003;42:4742–4758. doi: 10.1002/anie.200300573. [DOI] [PubMed] [Google Scholar]
- 25.Lo Conte M., Carroll K.S. The redox biochemistry of protein sulfenylation and sulfinylation. J. Biol. Chem. 2013;288:26480–26488. doi: 10.1074/jbc.R113.467738. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Paulsen C.E., Carroll K.S. Orchestrating redox signaling networks through regulatory cysteine switches. ACS Chem. Biol. 2010;5:47–62. doi: 10.1021/cb900258z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Dalle-Donne I., Rossi R., Giustarini D., Colombo R., Milzani A. S-glutathionylation in protein redox regulation. Free Radic. Biol. Med. 2007;43:883–898. doi: 10.1016/j.freeradbiomed.2007.06.014. [DOI] [PubMed] [Google Scholar]
- 28.Shelton M.D., Chock P.B., Mieyal J.J. Glutaredoxin: role in reversible protein s-glutathionylation and regulation of redox signal transduction and protein translocation. Antioxidants Redox Signal. 2005;7:348–366. doi: 10.1089/ars.2005.7.348. [DOI] [PubMed] [Google Scholar]
- 29.Gallogly M.M., Mieyal J.J. Mechanisms of reversible protein glutathionylation in redox signaling and oxidative stress. Curr. Opin. Pharmacol. 2007;7:381–391. doi: 10.1016/j.coph.2007.06.003. [DOI] [PubMed] [Google Scholar]
- 30.Brigelius-Flohe R. Mixed results with mixed disulfides. Arch. Biochem. Biophys. 2016;595:81–87. doi: 10.1016/j.abb.2015.11.011. [DOI] [PubMed] [Google Scholar]
- 31.Ghezzi P., Bonetto V., Fratelli M. Thiol-disulfide balance: from the concept of oxidative stress to that of redox regulation. Antioxidants Redox Signal. 2005;7:964–972. doi: 10.1089/ars.2005.7.964. [DOI] [PubMed] [Google Scholar]
- 32.Klatt P., Lamas S. Regulation of protein function by S-glutathiolation in response to oxidative and nitrosative stress. Eur. J. Biochem. 2000;267:4928–4944. doi: 10.1046/j.1432-1327.2000.01601.x. [DOI] [PubMed] [Google Scholar]
- 33.Forman H.J. Glutathione - from antioxidant to post-translational modifier. Arch. Biochem. Biophys. 2016;595:64–67. doi: 10.1016/j.abb.2015.11.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Jones D.P. Redox sensing: orthogonal control in cell cycle and apoptosis signalling. J. Intern. Med. 2010;268:432–448. doi: 10.1111/j.1365-2796.2010.02268.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Jones D.P. Radical-free biology of oxidative stress. Am. J. Physiol. Cell Physiol. 2008;295:C849–C868. doi: 10.1152/ajpcell.00283.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Flohe L. The impact of thiol peroxidases on redox regulation. Free Radic. Res. 2016;50:126–142. doi: 10.3109/10715762.2015.1046858. [DOI] [PubMed] [Google Scholar]
- 37.Deponte M. Glutathione catalysis and the reaction mechanisms of glutathione-dependent enzymes. Biochim. Biophys. Acta. 2013;1830:3217–3266. doi: 10.1016/j.bbagen.2012.09.018. [DOI] [PubMed] [Google Scholar]
- 38.Miseta A., Csutora P. Relationship between the occurrence of cysteine in proteins and the complexity of organisms. Mol. Biol. Evol. 2000;17:1232–1239. doi: 10.1093/oxfordjournals.molbev.a026406. [DOI] [PubMed] [Google Scholar]
- 39.Deponte M., Lillig C.H. Enzymatic control of cysteinyl thiol switches in proteins. Biol. Chem. 2015;396:401–413. doi: 10.1515/hsz-2014-0280. [DOI] [PubMed] [Google Scholar]
- 40.Mohanasundaram K.A., Haworth N.L., Grover M.P., Crowley T.M., Goscinski A., Wouters M.A. Potential role of glutathione in evolution of thiol-based redox signaling sites in proteins. Front. Pharmacol. 2015;6:1. doi: 10.3389/fphar.2015.00001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Johnston A.D., Ebert P.R. The redox system in C. elegans, a phylogenetic approach. J. Toxicol. 2012;2012 doi: 10.1155/2012/546915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Van Raamsdonk J.M., Hekimi S. Reactive oxygen species and aging in Caenorhabditis elegans: causal or casual relationship? Antioxidants Redox Signal. 2010;13:1911–1953. doi: 10.1089/ars.2010.3215. [DOI] [PubMed] [Google Scholar]
- 43.Brys K., Vanfleteren J.R., Braeckman B.P. Testing the rate-of-living/oxidative damage theory of aging in the nematode model Caenorhabditis elegans. Exp. Gerontol. 2007;42:845–851. doi: 10.1016/j.exger.2007.02.004. [DOI] [PubMed] [Google Scholar]
- 44.Gruber J., Chen C.B., Fong S., Ng L.F., Teo E., Halliwell B. Caenorhabditis elegans: what we can and cannot learn from aging worms. Antioxidants Redox Signal. 2015;23:256–279. doi: 10.1089/ars.2014.6210. [DOI] [PubMed] [Google Scholar]
- 45.Miranda-Vizuete A., Veal E.A. Caenorhabditis elegans as a model for understanding ROS function in physiology and disease. Redox Biol. 2016;11:708–714. doi: 10.1016/j.redox.2016.12.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Sulston J.E., Schierenberg E., White J.G., Thomson J.N. The embryonic cell lineage of the nematode Caenorhabditis elegans. Dev. Biol. 1983;100:64–119. doi: 10.1016/0012-1606(83)90201-4. [DOI] [PubMed] [Google Scholar]
- 47.Sulston J.E., Horvitz H.R. Post-embryonic cell lineages of the nematode, Caenorhabditis elegans. Dev. Biol. 1977;56:110–156. doi: 10.1016/0012-1606(77)90158-0. [DOI] [PubMed] [Google Scholar]
- 48.Klass M.R. Aging in the nematode Caenorhabditis elegans: major biological and environmental factors influencing life span. Mech. Ageing Dev. 1977;6:413–429. doi: 10.1016/0047-6374(77)90043-4. [DOI] [PubMed] [Google Scholar]
- 49.Golden J.W., Riddle D.L. The Caenorhabditis elegans dauer larva: developmental effects of pheromone, food, and temperature. Dev. Biol. 1984;102:368–378. doi: 10.1016/0012-1606(84)90201-x. [DOI] [PubMed] [Google Scholar]
- 50.Hodgkin J., Doniach T. Natural variation and copulatory plug formation in Caenorhabditis elegans. Genetics. 1997;146:149–164. doi: 10.1093/genetics/146.1.149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Corsi A.K., Wightman B., Chalfie M. A transparent window into biology: a primer on Caenorhabditis elegans. Genetics. 2015;200:387–407. doi: 10.1534/genetics.115.176099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Hobert O., Loria P. Uses of GFP in Caenorhabditis elegans. Methods Biochem. Anal. 2006;47:203–226. doi: 10.1002/0471739499.ch10. [DOI] [PubMed] [Google Scholar]
- 53.Timmons L., Fire A. Specific interference by ingested dsRNA. Nature. 1998;395:854. doi: 10.1038/27579. [DOI] [PubMed] [Google Scholar]
- 54.Friedland A.E., Tzur Y.B., Esvelt K.M., Colaiacovo M.P., Church G.M., Calarco J.A. Heritable genome editing in C. elegans via a CRISPR-Cas9 system. Nat. Methods. 2013;10:741–743. doi: 10.1038/nmeth.2532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Dickinson D.J., Goldstein B. CRISPR-based methods for Caenorhabditis elegans genome engineering. Genetics. 2016;202:885–901. doi: 10.1534/genetics.115.182162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Asamoto H., Ichibangase T., Saimaru H., Uchikura K., Imai K. Existence of low-molecular-weight thiols in Caenorhabditis elegans demonstrated by HPLC-fluorescence detection utilizing 7-chloro-N-[2-(dimethylamino)ethyl]-2,1,3-benzoxadiazole-4-sulfonamide. Biomed. Chromatogr. 2007;21:999–1004. doi: 10.1002/bmc.814. [DOI] [PubMed] [Google Scholar]
- 57.Luersen K., Stegehake D., Daniel J., Drescher M., Ajonina I., Ajonina C., Hertel P., Woltersdorf C., Liebau E. The glutathione reductase GSR-1 determines stress tolerance and longevity in Caenorhabditis elegans. PLoS One. 2013;8 doi: 10.1371/journal.pone.0060731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Urban N., Tsitsipatis D., Hausig F., Kreuzer K., Erler K., Stein V., Ristow M., Steinbrenner H., Klotz L.O. Non-linear impact of glutathione depletion on C. elegans life span and stress resistance. Redox Biol. 2016;11:502–515. doi: 10.1016/j.redox.2016.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Tang L., Choe K.P. Characterization of skn-1/wdr-23 phenotypes in Caenorhabditis elegans; pleiotrophy, aging, glutathione, and interactions with other longevity pathways. Mech. Ageing Dev. 2015;149:88–98. doi: 10.1016/j.mad.2015.06.001. [DOI] [PubMed] [Google Scholar]
- 60.Arkblad E.L., Tuck S., Pestov N.B., Dmitriev R.I., Kostina M.B., Stenvall J., Tranberg M., Rydstrom J. A Caenorhabditis elegans mutant lacking functional nicotinamide nucleotide transhydrogenase displays increased sensitivity to oxidative stress. Free Radic. Biol. Med. 2005;38:1518–1525. doi: 10.1016/j.freeradbiomed.2005.02.012. [DOI] [PubMed] [Google Scholar]
- 61.Heidler T., Hartwig K., Daniel H., Wenzel U. Caenorhabditis elegans lifespan extension caused by treatment with an orally active ROS-generator is dependent on DAF-16 and SIR-2.1. Biogerontology. 2010;11:183–195. doi: 10.1007/s10522-009-9239-x. [DOI] [PubMed] [Google Scholar]
- 62.Yan F., Chen X., Zheng X. Protective effect of mulberry fruit anthocyanin on human hepatocyte cells (LO2) and Caenorhabditis elegans under hyperglycemic conditions. Food Res. Int. 2017;102:213–224. doi: 10.1016/j.foodres.2017.10.009. [DOI] [PubMed] [Google Scholar]
- 63.Salgueiro W.G., Goldani B.S., Peres T.V., Miranda-Vizuete A., Aschner M., da Rocha J.B.T., Alves D., Avila D.S. Insights into the differential toxicological and antioxidant effects of 4-phenylchalcogenil-7-chloroquinolines in Caenorhabditis elegans. Free Radic. Biol. Med. 2017;110:133–141. doi: 10.1016/j.freeradbiomed.2017.05.020. [DOI] [PubMed] [Google Scholar]
- 64.Griffith O.W. Biologic and pharmacologic regulation of mammalian glutathione synthesis. Free Radic. Biol. Med. 1999;27:922–935. doi: 10.1016/s0891-5849(99)00176-8. [DOI] [PubMed] [Google Scholar]
- 65.Meister A., Anderson M.E. Glutathione. Annu. Rev. Biochem. 1983;52:711–760. doi: 10.1146/annurev.bi.52.070183.003431. [DOI] [PubMed] [Google Scholar]
- 66.An J.H., Blackwell T.K. SKN-1 links C. elegans mesendodermal specification to a conserved oxidative stress response. Genes Dev. 2003;17:1882–1893. doi: 10.1101/gad.1107803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Blackwell T.K., Steinbaugh M.J., Hourihan J.M., Ewald C.Y., Isik M. SKN-1/Nrf, stress responses, and aging in Caenorhabditis elegans. Free Radic. Biol. Med. 2015;88:290–301. doi: 10.1016/j.freeradbiomed.2015.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Oliveira R.P., Porter Abate J., Dilks K., Landis J., Ashraf J., Murphy C.T., Blackwell T.K. Condition-adapted stress and longevity gene regulation by Caenorhabditis elegans SKN-1/Nrf. Aging Cell. 2009;8:524–541. doi: 10.1111/j.1474-9726.2009.00501.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Park S.K., Tedesco P.M., Johnson T.E. Oxidative stress and longevity in Caenorhabditis elegans as mediated by SKN-1. Aging Cell. 2009;8:258–269. doi: 10.1111/j.1474-9726.2009.00473.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Inoue H., Hisamoto N., An J.H., Oliveira R.P., Nishida E., Blackwell T.K., Matsumoto K. The C. elegans p38 MAPK pathway regulates nuclear localization of the transcription factor SKN-1 in oxidative stress response. Genes Dev. 2005;19:2278–2283. doi: 10.1101/gad.1324805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.An J.H., Vranas K., Lucke M., Inoue H., Hisamoto N., Matsumoto K., Blackwell T.K. Regulation of the Caenorhabditis elegans oxidative stress defense protein SKN-1 by glycogen synthase kinase-3. Proc. Natl. Acad. Sci. U.S.A. 2005;102:16275–16280. doi: 10.1073/pnas.0508105102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Liao V.H.C., Yu C.W. Caenorhabditis elegans gcs-1 confers resistance to arsenic-induced oxidative stress. Biometals. 2005;18:519–528. doi: 10.1007/s10534-005-2996-3. [DOI] [PubMed] [Google Scholar]
- 73.Meng J., Lv Z., Qiao X., Li X., Li Y., Zhang Y., Chen C. The decay of Redox-stress Response Capacity is a substantive characteristic of aging: revising the redox theory of aging. Redox Biol. 2016;11:365–374. doi: 10.1016/j.redox.2016.12.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Franklin C.C., Backos D.S., Mohar I., White C.C., Forman H.J., Kavanagh T.J. Structure, function, and post-translational regulation of the catalytic and modifier subunits of glutamate cysteine ligase. Mol. Aspect. Med. 2009;30:86–98. doi: 10.1016/j.mam.2008.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Przybysz A.J., Choe K.P., Roberts L.J., Strange K. Increased age reduces DAF-16 and SKN-1 signaling and the hormetic response of Caenorhabditis elegans to the xenobiotic juglone. Mech. Ageing Dev. 2009;130:357–369. doi: 10.1016/j.mad.2009.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Choe K.P., Leung C.K., Miyamoto M.M. Unique structure and regulation of the nematode detoxification gene regulator, SKN-1: implications to understanding and controlling drug resistance. Drug Metab. Rev. 2012;44:209–223. doi: 10.3109/03602532.2012.684799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Li S., Armstrong C.M., Bertin N., Ge H., Milstein S., Boxem M., Vidalain P.O., Han J.D., Chesneau A., Hao T., Goldberg D.S., Li N., Martinez M., Rual J.F., Lamesch P., Xu L., Tewari M., Wong S.L., Zhang L.V., Berriz G.F., Jacotot L., Vaglio P., Reboul J., Hirozane-Kishikawa T., Li Q., Gabel H.W., Elewa A., Baumgartner B., Rose D.J., Yu H., Bosak S., Sequerra R., Fraser A., Mango S.E., Saxton W.M., Strome S., Van Den Heuvel S., Piano F., Vandenhaute J., Sardet C., Gerstein M., Doucette-Stamm L., Gunsalus K.C., Harper J.W., Cusick M.E., Roth F.P., Hill D.E., Vidal M. A map of the interactome network of the metazoan C. elegans. Science. 2004;303:540–543. doi: 10.1126/science.1091403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Wang J., Robida-Stubbs S., Tullet J.M., Rual J.F., Vidal M., Blackwell T.K. RNAi screening implicates a SKN-1-dependent transcriptional response in stress resistance and longevity deriving from translation inhibition. PLoS Genet. 2010;6 doi: 10.1371/journal.pgen.1001048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Spanier B., Rubio-Aliaga I., Hu H., Daniel H. Altered signalling from germline to intestine pushes daf-2;pept-1 Caenorhabditis elegans into extreme longevity. Aging Cell. 2010;9:636–646. doi: 10.1111/j.1474-9726.2010.00591.x. [DOI] [PubMed] [Google Scholar]
- 80.Wu H., Huang C., Taki F.A., Zhang Y., Dobbins D.L., Li L., Yan H., Pan X. Benzo-alpha-pyrene induced oxidative stress in Caenorhabditis elegans and the potential involvements of microRNA. Chemosphere. 2015;139:496–503. doi: 10.1016/j.chemosphere.2015.08.031. [DOI] [PubMed] [Google Scholar]
- 81.Crook-McMahon H.M., Olahova M., Button E.L., Winter J.J., Veal E.A. Genome-wide screening identifies new genes required for stress-induced phase 2 detoxification gene expression in animals. BMC Biol. 2014;12:64. doi: 10.1186/s12915-014-0064-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Mora-Lorca J.A., Saenz-Narciso B., Gaffney C.J., Naranjo-Galindo F.J., Pedrajas J.R., Guerrero-Gomez D., Dobrzynska A., Askjaer P., Szewczyk N.J., Cabello J., Miranda-Vizuete A. Glutathione reductase gsr-1 is an essential gene required for Caenorhabditis elegans early embryonic development. Free Radic. Biol. Med. 2016;96:446–461. doi: 10.1016/j.freeradbiomed.2016.04.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Gusarov I., Pani B., Gautier L., Smolentseva O., Eremina S., Shamovsky I., Katkova-Zhukotskaya O., Mironov A., Nudler E. Glycogen controls Caenorhabditis elegans lifespan and resistance to oxidative stress. Nat. Commun. 2017;8:15868. doi: 10.1038/ncomms15868. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Homolya L., Varadi A., Sarkadi B. Multidrug resistance-associated proteins: export pumps for conjugates with glutathione, glucuronate or sulfate. Biofactors. 2003;17:103–114. doi: 10.1002/biof.5520170111. [DOI] [PubMed] [Google Scholar]
- 85.Ballatori N., Krance S.M., Marchan R., Hammond C.L. Plasma membrane glutathione transporters and their roles in cell physiology and pathophysiology. Mol. Aspect. Med. 2009;30:13–28. doi: 10.1016/j.mam.2008.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Montero D., Tachibana C., Rahr Winther J., Appenzeller-Herzog C. Intracellular glutathione pools are heterogeneously concentrated. Redox Biol. 2013;1:508–513. doi: 10.1016/j.redox.2013.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Go Y.M., Jones D.P. Redox compartmentalization in eukaryotic cells. Biochim. Biophys. Acta Gen. Subj. 2008;1780:1271–1290. doi: 10.1016/j.bbagen.2008.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Braeckman B.P., Smolders A., Back P., De Henau S. In vivo detection of reactive oxygen species and redox status in Caenorhabditis elegans. Antioxidants Redox Signal. 2016;25:577–592. doi: 10.1089/ars.2016.6751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Back P., De Vos W.H., Depuydt G.G., Matthijssens F., Vanfleteren J.R., Braeckman B.P. Exploring real-time in vivo redox biology of developing and aging Caenorhabditis elegans. Free Radic. Biol. Med. 2012;52:850–859. doi: 10.1016/j.freeradbiomed.2011.11.037. [DOI] [PubMed] [Google Scholar]
- 90.Margis R., Dunand C., Teixeira F.K., Margis-Pinheiro M. Glutathione peroxidase family - an evolutionary overview. FEBS J. 2008;275:3959–3970. doi: 10.1111/j.1742-4658.2008.06542.x. [DOI] [PubMed] [Google Scholar]
- 91.Herbette S., Roeckel-Drevet P., Drevet J.R. Seleno-independent glutathione peroxidases. More than simple antioxidant scavengers. FEBS J. 2007;274:2163–2180. doi: 10.1111/j.1742-4658.2007.05774.x. [DOI] [PubMed] [Google Scholar]
- 92.Sakamoto T., Maebayashi K., Nakagawa Y., Imai H. Deletion of the four phospholipid hydroperoxide glutathione peroxidase genes accelerates aging in Caenorhabditis elegans. Genes Cells. 2014;19:778–792. doi: 10.1111/gtc.12175. [DOI] [PubMed] [Google Scholar]
- 93.Xu Z., Feng S.L., Shen S.A., Wang H.D., Yuan M., Liu J., Huang Y., Ding C.B. The antioxidant activities effect of neutral and acidic polysaccharides from Epimedium acuminatum Franch. on Caenorhabditis elegans. Carbohydr. Polym. 2016;144:122–130. doi: 10.1016/j.carbpol.2016.02.041. [DOI] [PubMed] [Google Scholar]
- 94.Li H., Shi R., Ding F., Wang H., Han W., Ma F., Hu M., Ma C.W., Huang Z. Astragalus polysaccharide suppresses 6-hydroxydopamine-induced neurotoxicity in Caenorhabditis elegans. Oxid. Med. Cell Longev. 2016;2016 doi: 10.1155/2016/4856761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Song S., Zhang X., Wu H., Han Y., Zhang J., Ma E., Guo Y. Molecular basis for antioxidant enzymes in mediating copper detoxification in the nematode Caenorhabditis elegans. PLoS One. 2014;9 doi: 10.1371/journal.pone.0107685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Benner J., Daniel H., Spanier B. A glutathione peroxidase, intracellular peptidases and the TOR complexes regulate peptide transporter PEPT-1 in C. elegans. PLoS One. 2011;6 doi: 10.1371/journal.pone.0025624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Spanier B. Transcriptional and functional regulation of the intestinal peptide transporter PEPT1. J. Physiol. 2014;592:871–879. doi: 10.1113/jphysiol.2013.258889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Petersen S.C., Watson J.D., Richmond J.E., Sarov M., Walthall W.W., Miller D.M., 3rd A transcriptional program promotes remodeling of GABAergic synapses in Caenorhabditis elegans. J. Neurosci. 2011;31:15362–15375. doi: 10.1523/JNEUROSCI.3181-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Simonetta S.H., Romanowski A., Minniti A.N., Inestrosa N.C., Golombek D.A. Circadian stress tolerance in adult Caenorhabditis elegans. J. Comp. Physiol. A Neuroethol. Sens Neural Behav. Physiol. 2008;194:821–828. doi: 10.1007/s00359-008-0353-z. [DOI] [PubMed] [Google Scholar]
- 100.Sheehan D., Meade G., Foley V.M., Dowd C.A. Structure, function and evolution of glutathione transferases: implications for classification of non-mammalian members of an ancient enzyme superfamily. Biochem. J. 2001;360:1–16. doi: 10.1042/0264-6021:3600001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Lindblom T.H., Dodd A.K. Xenobiotic detoxification in the nematode Caenorhabditis elegans. J. Exp. Zool. A Comp. Exp. Biol. 2006;305:720–730. doi: 10.1002/jez.a.324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Dancy B.M., Brockway N., Ramadasan-Nair R., Yang Y., Sedensky M.M., Morgan P.G. Glutathione S-transferase mediates an ageing response to mitochondrial dysfunction. Mech. Ageing Dev. 2016;153:14–21. doi: 10.1016/j.mad.2015.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Hasegawa K., Miwa S., Tsutsumiuchi K., Miwa J. Allyl Isothiocyanate that induces GST and UGT expression confers oxidative stress resistance on C. elegans, as demonstrated by nematode biosensor. PLoS One. 2010;5 doi: 10.1371/journal.pone.0009267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Morel F., Aninat C. The glutathione transferase kappa family. Drug Metab. Rev. 2011;43:281–291. doi: 10.3109/03602532.2011.556122. [DOI] [PubMed] [Google Scholar]
- 105.Lee J.H., Kim Y.G., Kim M., Kim E., Choi H., Kim Y., Lee J. Indole-associated predator-prey interactions between the nematode Caenorhabditis elegans and bacteria. Environ. Microbiol. 2017;19:1776–1790. doi: 10.1111/1462-2920.13649. [DOI] [PubMed] [Google Scholar]
- 106.Negi H., Saikia S.K., Kanaujia R., Jaiswal S., Pandey R. 3beta-Hydroxy-urs-12-en-28-oic acid confers protection against ZnONPs induced adversity in Caenorhabditis elegans. Environ. Toxicol. Pharmacol. 2017;53:105–110. doi: 10.1016/j.etap.2017.05.004. [DOI] [PubMed] [Google Scholar]
- 107.McElwee J.J., Schuster E., Blanc E., Piper M.D., Thomas J.H., Patel D.S., Selman C., Withers D.J., Thornton J.M., Partridge L., Gems D. Evolutionary conservation of regulated longevity assurance mechanisms. Genome Biol. 2007;8:R132. doi: 10.1186/gb-2007-8-7-r132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.McElwee J.J., Schuster E., Blanc E., Thomas J.H., Gems D. Shared transcriptional signature in Caenorhabditis elegans dauer larvae and long-lived daf-2 mutants implicates detoxification system in longevity assurance. J. Biol. Chem. 2004;279:44533–44543. doi: 10.1074/jbc.M406207200. [DOI] [PubMed] [Google Scholar]
- 109.Jones L.M., Staffa K., Perally S., LaCourse E.J., Brophy P.M., Hamilton J.V. Proteomic analyses of Caenorhabditis elegans dauer larvae and long-lived daf-2 mutants implicates a shared detoxification system in longevity assurance. J. Proteome Res. 2010;9:2871–2881. doi: 10.1021/pr9009639. [DOI] [PubMed] [Google Scholar]
- 110.Perally S., Lacourse E.J., Campbell A.M., Brophy P.M. Heme transport and detoxification in nematodes: subproteomics evidence of differential role of glutathione transferases. J. Proteome Res. 2008;7:4557–4565. doi: 10.1021/pr800395x. [DOI] [PubMed] [Google Scholar]
- 111.van Rossum A.J., Brophy P.M., Tait A., Barrett J., Jefferies J.R. Proteomic identification of glutathione S-transferases from the model nematode Caenorhabditis elegans. Proteomics. 2001;1:1463–1468. doi: 10.1002/1615-9861(200111)1:11<1463::AID-PROT1463>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
- 112.Greetham D., Morgan C., Campbell A.M., van Rossum A.J., Barrett J., Brophy P.M. Evidence of glutathione transferase complexing and signaling in the model nematode Caenorhabditis elegans using a pull-down proteomic assay. Proteomics. 2004;4:1989–1995. doi: 10.1002/pmic.200300719. [DOI] [PubMed] [Google Scholar]
- 113.Tawe W.N., Eschbach M.L., Walter R.D., Henkle-Duhrsen K. Identification of stress-responsive genes in Caenorhabditis elegans using RT-PCR differential display. Nucleic Acids Res. 1998;26:1621–1627. doi: 10.1093/nar/26.7.1621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Leiers B., Kampkotter A., Grevelding C.G., Link C.D., Johnson T.E., Henkle-Duhrsen K. A stress-responsive glutathione S-transferase confers resistance to oxidative stress in Caenorhabditis elegans. Free Radic. Biol. Med. 2003;34:1405–1415. doi: 10.1016/s0891-5849(03)00102-3. [DOI] [PubMed] [Google Scholar]
- 115.Link C.D., Johnson C.J. Reporter transgenes for study of oxidant stress in Caenorhabditis elegans. Methods Enzymol. 2002;353:497–505. doi: 10.1016/s0076-6879(02)53072-x. [DOI] [PubMed] [Google Scholar]
- 116.Rathor L., Akhoon B.A., Pandey S., Srivastava S., Pandey R. Folic acid supplementation at lower doses increases oxidative stress resistance and longevity in Caenorhabditis elegans. Age. 2015;37:113. doi: 10.1007/s11357-015-9850-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Wang Y., Jian F.L., Wu J.Y., Wang S.C. Stress-response protein expression and DAF-16 translocation were induced in tributyltin-exposed Caenorhabditis elegans. Bull. Environ. Contam. Toxicol. 2012;89:704–711. doi: 10.1007/s00128-012-0760-2. [DOI] [PubMed] [Google Scholar]
- 118.Kahn N.W., Rea S.L., Moyle S., Kell A., Johnson T.E. Proteasomal dysfunction activates the transcription factor SKN-1 and produces a selective oxidative-stress response in Caenorhabditis elegans. Biochem. J. 2008;409:205–213. doi: 10.1042/BJ20070521. [DOI] [PubMed] [Google Scholar]
- 119.Buchter C., Zhao L., Havermann S., Honnen S., Fritz G., Proksch P., Watjen W. TSG (2,3,5,4'-tetrahydroxystilbene-2-O- beta -D-glucoside) from the Chinese herb polygonum multiflorum increases life span and stress resistance of Caenorhabditis elegans. Oxid. Med. Cell Longev. 2015;2015 doi: 10.1155/2015/124357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Lei L., Wu S., Lu S., Liu M., Song Y., Fu Z., Shi H., Raley-Susman K.M., He D. Microplastic particles cause intestinal damage and other adverse effects in zebrafish Danio rerio and nematode Caenorhabditis elegans. Sci. Total Environ. 2017;619–620:1–8. doi: 10.1016/j.scitotenv.2017.11.103. [DOI] [PubMed] [Google Scholar]
- 121.Detienne G., Van de Walle P., De Haes W., Schoofs L., Temmerman L. SKN-1-independent transcriptional activation of glutathione S-transferase 4 (GST-4) by EGF signaling. Worm. 2016;5 doi: 10.1080/21624054.2016.1230585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Ayyadevara S., Dandapat A., Singh S.P., Siegel E.R., Shmookler Reis R.J., Zimniak L., Zimniak P. Life span and stress resistance of Caenorhabditis elegans are differentially affected by glutathione transferases metabolizing 4-hydroxynon-2-enal. Mech. Ageing Dev. 2007;128:196–205. doi: 10.1016/j.mad.2006.11.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Ayyadevara S., Engle M.R., Singh S.P., Dandapat A., Lichti C.F., Benes H., Shmookler Reis R.J., Liebau E., Zimniak P. Lifespan and stress resistance of Caenorhabditis elegans are increased by expression of glutathione transferases capable of metabolizing the lipid peroxidation product 4-hydroxynonenal. Aging Cell. 2005;4:257–271. doi: 10.1111/j.1474-9726.2005.00168.x. [DOI] [PubMed] [Google Scholar]
- 124.Singh S.P., Niemczyk M., Zimniak L., Zimniak P. Fat accumulation in Caenorhabditis elegans triggered by the electrophilic lipid peroxidation product 4-Hydroxynonenal (4-HNE) Aging (Albany NY) 2009;1:68–80. doi: 10.18632/aging.100005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Settivari R., VanDuyn N., LeVora J., Nass R. The Nrf2/SKN-1-dependent glutathione S-transferase pi homologue GST-1 inhibits dopamine neuron degeneration in a Caenorhabditis elegans model of manganism. Neurotoxicology. 2013;38:51–60. doi: 10.1016/j.neuro.2013.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Petit E., Michelet X., Rauch C., Bertrand-Michel J., Terce F., Legouis R., Morel F. Glutathione transferases kappa 1 and kappa 2 localize in peroxisomes and mitochondria, respectively, and are involved in lipid metabolism and respiration in Caenorhabditis elegans. FEBS J. 2009;276:5030–5040. doi: 10.1111/j.1742-4658.2009.07200.x. [DOI] [PubMed] [Google Scholar]
- 127.Burmeister C., Luersen K., Heinick A., Hussein A., Domagalski M., Walter R.D., Liebau E. Oxidative stress in Caenorhabditis elegans: protective effects of the Omega class glutathione transferase (GSTO-1) FASEB J. 2008;22:343–354. doi: 10.1096/fj.06-7426com. [DOI] [PubMed] [Google Scholar]
- 128.Schieber M., Chandel N.S. TOR signaling couples oxygen sensing to lifespan in C. elegans. Cell Rep. 2014;9:9–15. doi: 10.1016/j.celrep.2014.08.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Hasegawa K., Miwa S., Isomura K., Tsutsumiuchi K., Taniguchi H., Miwa J. Acrylamide-responsive genes in the nematode Caenorhabditis elegans. Toxicol. Sci. 2008;101:215–225. doi: 10.1093/toxsci/kfm276. [DOI] [PubMed] [Google Scholar]
- 130.Fernandes A.P., Holmgren A. Glutaredoxins: glutathione-dependent redox enzymes with functions far beyond a simple thioredoxin backup system. Antioxidants Redox Signal. 2004;6:63–74. doi: 10.1089/152308604771978354. [DOI] [PubMed] [Google Scholar]
- 131.Berndt C., Lillig C.H., Holmgren A. Thioredoxins and glutaredoxins as facilitators of protein folding. Biochim. Biophys. Acta. 2008;1783:641–650. doi: 10.1016/j.bbamcr.2008.02.003. [DOI] [PubMed] [Google Scholar]
- 132.Lillig C.H., Berndt C., Holmgren A. Glutaredoxin systems. Biochim. Biophys. Acta. 2008;1780:1304–1317. doi: 10.1016/j.bbagen.2008.06.003. [DOI] [PubMed] [Google Scholar]
- 133.Jeong D.E., Lee D., Hwang S.Y., Lee Y., Lee J.E., Seo M., Hwang W., Seo K., Hwang A.B., Artan M., Son H.G., Jo J.H., Baek H., Oh Y.M., Ryu Y., Kim H.J., Ha C.M., Yoo J.Y., Lee S.V. Mitochondrial chaperone HSP-60 regulates anti-bacterial immunity via p38 MAP kinase signaling. EMBO J. 2017;36:1046–1065. doi: 10.15252/embj.201694781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Yamazaki Y., Akashi R., Banno Y., Endo T., Ezura H., Fukami-Kobayashi K., Inaba K., Isa T., Kamei K., Kasai F., Kobayashi M., Kurata N., Kusaba M., Matuzawa T., Mitani S., Nakamura T., Nakamura Y., Nakatsuji N., Naruse K., Niki H., Nitasaka E., Obata Y., Okamoto H., Okuma M., Sato K., Serikawa T., Shiroishi T., Sugawara H., Urushibara H., Yamamoto M., Yaoita Y., Yoshiki A., Kohara Y. NBRP databases: databases of biological resources in Japan. Nucleic Acids Res. 2010;38:D26–D32. doi: 10.1093/nar/gkp996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Johnson W.M., Yao C., Siedlak S.L., Wang W.Z., Zhu X.W., Caldwell G.A., Wilson-Delfosse A.L., Mieyal J.J., Chen S.G. Glutaredoxin deficiency exacerbates neurodegeneration in C. elegans models of Parkinson's disease. Hum. Mol. Genet. 2015;24:1322–1335. doi: 10.1093/hmg/ddu542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Morgan K.L., Estevez A.O., Mueller C.L., Cacho-Valadez B., Miranda-Vizuete A., Szewczyk N.J., Estevez M. The glutaredoxin GLRX-21 functions to prevent selenium-induced oxidative stress in Caenorhabditis elegans. Toxicol. Sci. 2010;118:530–543. doi: 10.1093/toxsci/kfq273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Johnson W.M., Golczak M., Choe K., Curran P.L., Miller O.G., Yao C., Wang W., Lin J., Milkovic N.M., Ray A., Ravindranath V., Zhu X., Wilson M.A., Wilson-Delfosse A.L., Chen S.G., Mieyal J.J. Regulation of DJ-1 by glutaredoxin 1 in vivo: implications for Parkinson's disease. Biochemistry. 2016;55:4519–4532. doi: 10.1021/acs.biochem.5b01132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Estevez A.O., Mueller C.L., Morgan K.L., Szewczyk N.J., Teece L., Miranda-Vizuete A., Estevez M. Selenium induces cholinergic motor neuron degeneration in Caenorhabditis elegans. Neurotoxicology. 2012;33:1021–1032. doi: 10.1016/j.neuro.2012.04.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Henze A., Homann T., Rohn I., Aschner M., Link C.D., Kleuser B., Schweigert F.J., Schwerdtle T., Bornhorst J. Caenorhabditis elegans as a model system to study post-translational modifications of human transthyretin. Sci. Rep. 2016;6:37346. doi: 10.1038/srep37346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Arner E.S., Holmgren A. Physiological functions of thioredoxin and thioredoxin reductase. Eur. J. Biochem. 2000;267:6102–6109. doi: 10.1046/j.1432-1327.2000.01701.x. [DOI] [PubMed] [Google Scholar]
- 141.Lu J., Holmgren A. The thioredoxin antioxidant system. Free Radic. Biol. Med. 2014;66:75–87. doi: 10.1016/j.freeradbiomed.2013.07.036. [DOI] [PubMed] [Google Scholar]
- 142.Arodin L., Miranda-Vizuete A., Swoboda P., Fernandes A.P. Protective effects of the thioredoxin and glutaredoxin systems in dopamine-induced cell death. Free Radic. Biol. Med. 2014;73:328–336. doi: 10.1016/j.freeradbiomed.2014.05.011. [DOI] [PubMed] [Google Scholar]
- 143.Miranda-Vizuete A., Fierro Gonzalez J.C., Gahmon G., Burghoorn J., Navas P., Swoboda P. Lifespan decrease in a Caenorhabditis elegans mutant lacking TRX-1, a thioredoxin expressed in ASJ sensory neurons. FEBS Lett. 2006;580:484–490. doi: 10.1016/j.febslet.2005.12.046. [DOI] [PubMed] [Google Scholar]
- 144.Gonzalez-Barrios M., Fierro-Gonzalez J.C., Krpelanova E., Mora-Lorca J.A., Pedrajas J.R., Penate X., Chavez S., Swoboda P., Jansen G., Miranda-Vizuete A. Cis- and trans-regulatory mechanisms of gene expression in the ASJ sensory neuron of Caenorhabditis elegans. Genetics. 2015;200:123–134. doi: 10.1534/genetics.115.176172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Jee C., Vanoaica L., Lee J., Park B.J., Ahnn J. Thioredoxin is related to life span regulation and oxidative stress response in Caenorhabditis elegans. Genes Cells. 2005;10:1203–1210. doi: 10.1111/j.1365-2443.2005.00913.x. [DOI] [PubMed] [Google Scholar]
- 146.Fierro-Gonzalez J.C., Gonzalez-Barrios M., Miranda-Vizuete A., Swoboda P. The thioredoxin TRX-1 regulates adult lifespan extension induced by dietary restriction in Caenorhabditis elegans. Biochem. Biophys. Res. Commun. 2011;406:478–482. doi: 10.1016/j.bbrc.2011.02.079. [DOI] [PubMed] [Google Scholar]
- 147.McCallum K.C., Liu B., Fierro-Gonzalez J.C., Swoboda P., Arur S., Miranda-Vizuete A., Garsin D.A. TRX-1 regulates SKN-1 nuclear localization cell non-autonomously in Caenorhabditis elegans. Genetics. 2016;203:387–402. doi: 10.1534/genetics.115.185272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Fierro-Gonzalez J.C., Cornils A., Alcedo J., Miranda-Vizuete A., Swoboda P. The thioredoxin TRX-1 modulates the function of the insulin-like neuropeptide DAF-28 during dauer formation in Caenorhabditis elegans. PLoS One. 2011;6 doi: 10.1371/journal.pone.0016561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Hao Y., Yang W., Ren J., Hall Q., Zhang Y., Kaplan J.M. Thioredoxin shapes the C. elegans sensory response to Pseudomonas produced nitric oxide. elife. 2018;7 doi: 10.7554/eLife.36833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Ruszkiewicz J.A., Teixeira de Macedo G., Miranda-Vizuete A., Bowman A.B., Bornhorst J., Schwerdtle T., Antunes Soares F.A., Aschner M. Sex-specific response of Caenorhabditis elegans to methylmercury toxicity. Neurotox. Res. 2019;35:208–216. doi: 10.1007/s12640-018-9949-4. [DOI] [PubMed] [Google Scholar]
- 151.Ruszkiewicz J.A., Teixeira de Macedo G., Miranda-Vizuete A., Teixeira da Rocha J.B., Bowman A.B., Bornhorst J., Schwerdtle T., Aschner M. The cytoplasmic thioredoxin system in Caenorhabditis elegans affords protection from methylmercury in an age-specific manner. Neurotoxicology. 2018;68:189–202. doi: 10.1016/j.neuro.2018.08.007. [DOI] [PubMed] [Google Scholar]
- 152.Cacho-Valadez B., Munoz-Lobato F., Pedrajas J.R., Cabello J., Fierro-Gonzalez J.C., Navas P., Swoboda P., Link C.D., Miranda-Vizuete A. The characterization of the Caenorhabditis elegans mitochondrial thioredoxin system uncovers an unexpected protective role of thioredoxin reductase 2 in beta-amyloid peptide toxicity. Antioxidants Redox Signal. 2012;16:1384–1400. doi: 10.1089/ars.2011.4265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Lacey B.M., Hondal R.J. Characterization of mitochondrial thioredoxin reductase from C. elegans. Biochem. Biophys. Res. Commun. 2006;346:629–636. doi: 10.1016/j.bbrc.2006.05.095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Jiménez-Hidalgo M., Kurz C.L., Pedrajas J.R., Naranjo-Galindo F.J., Gonzalez-Barrios M., Cabello J., Saez A.G., Lozano E., Button E.L., Veal E.A., Fierro-Gonzalez J.C., Swoboda P., Miranda-Vizuete A. Functional characterization of thioredoxin 3 (TRX-3), a Caenorhabditis elegans intestine-specific thioredoxin. Free Radic. Biol. Med. 2014;68:205–219. doi: 10.1016/j.freeradbiomed.2013.11.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Munoz-Lobato F., Rodriguez-Palero M.J., Naranjo-Galindo F.J., Shephard F., Gaffney C.J., Szewczyk N.J., Hamamichi S., Caldwell K.A., Caldwell G.A., Link C.D., Miranda-Vizuete A. Protective role of DNJ-27/ERdj5 in Caenorhabditis elegans models of human neurodegenerative diseases. Antioxidants Redox Signal. 2014;20:217–235. doi: 10.1089/ars.2012.5051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Suzuki T., Tanabe K., Hara I., Taniguchi N., Colavita A. Dual enzymatic properties of the cytoplasmic peptide: N-glycanase in C. elegans. Biochem. Biophys. Res. Commun. 2007;358:837–841. doi: 10.1016/j.bbrc.2007.04.199. [DOI] [PubMed] [Google Scholar]
- 157.Kato T., Kawahara A., Ashida H., Yamamoto K. Unique peptide : N-glycanase of Caenorhabditis elegans has activity of protein disulphide reductase as well as of deglycosylation. J. Biochem. 2007;142:175–181. doi: 10.1093/jb/mvm117. [DOI] [PubMed] [Google Scholar]
- 158.Suzuki T., Park H., Till E.A., Lennarz W.J. The PUB domain: a putative protein-protein interaction domain implicated in the ubiquitin-proteasome pathway. Biochem. Biophys. Res. Commun. 2001;287:1083–1087. doi: 10.1006/bbrc.2001.5688. [DOI] [PubMed] [Google Scholar]
- 159.Habibi-Babadi N., Su A., de Carvalho C.E., Colavita A. The N-glycanase png-1 acts to limit axon branching during organ formation in Caenorhabditis elegans. J. Neurosci. 2010;30:1766–1776. doi: 10.1523/JNEUROSCI.4962-08.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Gladyshev V.N., Krause M., Xu X.M., Korotkov K.V., Kryukov G.V., Sun Q.A., Lee B.J., Wootton J.C., Hatfield D.L. Selenocysteine-containing thioredoxin reductase in C. elegans. Biochem. Biophys. Res. Commun. 1999;259:244–249. doi: 10.1006/bbrc.1999.0765. [DOI] [PubMed] [Google Scholar]
- 161.Stenvall J., Fierro-Gonzalez J.C., Swoboda P., Saamarthy K., Cheng Q., Cacho-Valadez B., Arner E.S., Persson O.P., Miranda-Vizuete A., Tuck S. Selenoprotein TRXR-1 and GSR-1 are essential for removal of old cuticle during molting in Caenorhabditis elegans. Proc. Natl. Acad. Sci. U.S.A. 2011;108:1064–1069. doi: 10.1073/pnas.1006328108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Li W., Bandyopadhyay J., Hwaang H.S., Park B.J., Cho J.H., Il Lee J., Ahnn J., Lee S.K. Two thioredoxin reductases, trxr-1 and trxr-2, have differential physiological roles in Caenorhabditis elegans. Mol. Cell. 2012;34:209–218. doi: 10.1007/s10059-012-0155-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Taskov K., Chapple C., Kryukov G.V., Castellano S., Lobanov A.V., Korotkov K.V., Guigo R., Gladyshev V.N. Nematode selenoproteome: the use of the selenocysteine insertion system to decode one codon in an animal genome? Nucleic Acids Res. 2005;33:2227–2238. doi: 10.1093/nar/gki507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Li W.H., Shi Y.C., Chang C.H., Huang C.W., Liao V.H.C. Selenite protects Caenorhabditis elegans from oxidative stress via DAF-16 and TRXR-1. Mol. Nutr. Food Res. 2014;58:863–874. doi: 10.1002/mnfr.201300404. [DOI] [PubMed] [Google Scholar]
- 165.Boehler C.J., Raines A.M., Sunde R.A. Toxic-selenium and low-selenium transcriptomes in Caenorhabditis elegans: toxic selenium up-regulates oxidoreductase and down-regulates cuticle-associated genes. PLoS One. 2014;9 doi: 10.1371/journal.pone.0101408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Boehler C.J., Raines A.M., Sunde R.A. Deletion of thioredoxin reductase and effects of selenite and selenate toxicity in Caenorhabditis elegans. PLoS One. 2013;8 doi: 10.1371/journal.pone.0071525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Garcia-Rodriguez F.J., Martinez-Fernandez C., Brena D., Kukhtar D., Serrat X., Nadal E., Boxem M., Honnen S., Miranda-Vizuete A., Villanueva A., Ceron J. Genetic and cellular sensitivity of Caenorhabditis elegans to the chemotherapeutic agent cisplatin. Dis. Model Mech. 2018;11:dmm03350. doi: 10.1242/dmm.033506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Wood Z.A., Schroder E., Robin Harris J., Poole L.B. Structure, mechanism and regulation of peroxiredoxins. Trends Biochem. Sci. 2003;28:32–40. doi: 10.1016/s0968-0004(02)00003-8. [DOI] [PubMed] [Google Scholar]
- 169.Winterbourn C.C. Biological production, detection, and fate of hydrogen peroxide. Antioxidants Redox Signal. 2018;29:541–551. doi: 10.1089/ars.2017.7425. [DOI] [PubMed] [Google Scholar]
- 170.Isermann K., Liebau E., Roeder T., Bruchhaus I. A peroxiredoxin specifically expressed in two types of pharyngeal neurons is required for normal growth and egg production in Caenorhabditis elegans. J. Mol. Biol. 2004;338:745–755. doi: 10.1016/j.jmb.2004.03.021. [DOI] [PubMed] [Google Scholar]
- 171.Bhatla N., Horvitz H.R. Light and hydrogen peroxide inhibit C. elegans feeding through gustatory receptor orthologs and pharyngeal neurons. Neuron. 2015;85:804–818. doi: 10.1016/j.neuron.2014.12.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Kumsta C., Thamsen M., Jakob U. Effects of oxidative stress on behavior, physiology, and the redox thiol proteome of Caenorhabditis elegans. Antioxidants Redox Signal. 2011;14:1023–1037. doi: 10.1089/ars.2010.3203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Olahova M., Taylor S.R., Khazaipoul S., Wang J.L., Morgan B.A., Matsumoto K., Blackwell T.K., Veal E.A. A redox-sensitive peroxiredoxin that is important for longevity has tissue- and stress-specific roles in stress resistance. Proc. Natl. Acad. Sci. U.S.A. 2008;105:19839–19844. doi: 10.1073/pnas.0805507105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Henderson D., Huebner C., Markowitz M., Taube N., Harvanek Z.M., Jakob U., Knoefler D. Do developmental temperatures affect redox level and lifespan in C. elegans through upregulation of peroxiredoxin? Redox Biol. 2018;14:386–390. doi: 10.1016/j.redox.2017.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Spiro Z., Arslan M.A., Somogyvari M., Nguyen M.T., Smolders A., Dancso B., Nemeth N., Elek Z., Braeckman B.P., Csermely P., Soti C. RNA interference links oxidative stress to the inhibition of heat stress adaptation. Antioxidants Redox Signal. 2012;17:890–901. doi: 10.1089/ars.2011.4161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Thamsen M., Kumsta C., Li F., Jakob U. Is overoxidation of peroxiredoxin physiologically significant? Antioxidants Redox Signal. 2011;14:725–730. doi: 10.1089/ars.2010.3717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Olmedo M., O'Neill J.S., Edgar R.S., Valekunja U.K., Reddy A.B., Merrow M. Circadian regulation of olfaction and an evolutionarily conserved, nontranscriptional marker in Caenorhabditis elegans. Proc. Natl. Acad. Sci. U.S.A. 2012;109:20479–20484. doi: 10.1073/pnas.1211705109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Edgar R.S., Green E.W., Zhao Y., van Ooijen G., Olmedo M., Qin X., Xu Y., Pan M., Valekunja U.K., Feeney K.A., Maywood E.S., Hastings M.H., Baliga N.S., Merrow M., Millar A.J., Johnson C.H., Kyriacou C.P., O'Neill J.S., Reddy A.B. Peroxiredoxins are conserved markers of circadian rhythms. Nature. 2012;485:459–464. doi: 10.1038/nature11088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Li G., Gong J., Lei H., Liu J., Xu X.Z. Promotion of behavior and neuronal function by reactive oxygen species in C. elegans. Nat. Commun. 2016;7:13234. doi: 10.1038/ncomms13234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.De Haes W., Frooninckx L., Van Assche R., Smolders A., Depuydt G., Billen J., Braeckman B.P., Schoofs L., Temmerman L. Metformin promotes lifespan through mitohormesis via the peroxiredoxin PRDX-2. Proc. Natl. Acad. Sci. U.S.A. 2014;111:E2501–E2509. doi: 10.1073/pnas.1321776111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Olahova M., Veal E.A. A peroxiredoxin, PRDX-2, is required for insulin secretion and insulin/IIS-dependent regulation of stress resistance and longevity. Aging Cell. 2015;14:558–568. doi: 10.1111/acel.12321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Back P., Braeckman B.P., Matthijssens F. ROS in aging Caenorhabditis elegans: damage or signaling? Oxid. Med. Cell Longev. 2012;2012 doi: 10.1155/2012/608478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Ranjan M., Gruber J., Ng L.F., Halliwell B. Repression of the mitochondrial peroxiredoxin antioxidant system does not shorten life span but causes reduced fitness in Caenorhabditis elegans. Free Radic. Biol. Med. 2013;63:381–389. doi: 10.1016/j.freeradbiomed.2013.05.025. [DOI] [PubMed] [Google Scholar]
- 184.Lee R.Y.N., Howe K.L., Harris T.W., Arnaboldi V., Cain S., Chan J., Chen W.J., Davis P., Gao S., Grove C., Kishore R., Muller H.M., Nakamura C., Nuin P., Paulini M., Raciti D., Rodgers F., Russell M., Schindelman G., Tuli M.A., Van Auken K., Wang Q.H., Williams G., Wright A., Yook K., Berriman M., Kersey P., Schedl T., Stein L., Sternberg P.W. WormBase 2017: molting into a new stage. Nucleic Acids Res. 2018;46:D869–D874. doi: 10.1093/nar/gkx998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Girard L.R., Fiedler T.J., Harris T.W., Carvalho F., Antoshechkin I., Han M., Sternberg P.W., Stein L.D., Chalfie M. WormBook: the online review of Caenorhabditis elegans biology. Nucleic Acids Res. 2007;35:D472–D475. doi: 10.1093/nar/gkl894. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Altun Z.F., Herndon L.A., Wolkow C.A., Crocker C., Lints R., Hall D.H. 2002-2019. WormAtlas.http://www.wormatlas.org Available at: [Google Scholar]
- 187.Yang J., Carroll K.S., Liebler D.C. The expanding landscape of the thiol redox proteome. Mol. Cell. Proteomics. 2016;15:1–11. doi: 10.1074/mcp.O115.056051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Knoefler D., Thamsen M., Koniczek M., Niemuth N.J., Diederich A.K., Jakob U. Quantitative in vivo redox rensors uncover oxidative stress as an early event in life. Mol. Cell. 2012;47:767–776. doi: 10.1016/j.molcel.2012.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Martell J., Seo Y., Bak D.W., Kingsley S.F., Tissenbaum H.A., Weerapana E. Global cysteine-reactivity profiling during impaired insulin/IGF-1 signaling in C. elegans identifies uncharacterized mediators of longevity. Cell Chem. Biol. 2016;23:955–966. doi: 10.1016/j.chembiol.2016.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Hasegawa K., Miwa S., Tajima T., Tsutsumiuchi K., Taniguchi H., Miwa J. A rapid and inexpensive method to screen for common foods that reduce the action of acrylamide, a harmful substance in food. Toxicol. Lett. 2007;175:82–88. doi: 10.1016/j.toxlet.2007.09.013. [DOI] [PubMed] [Google Scholar]
- 191.Ayyadevara S., Bharill P., Dandapat A., Hu C., Khaidakov M., Mitra S., Shmookler Reis R.J., Mehta J.L. Aspirin inhibits oxidant stress, reduces age-associated functional declines, and extends lifespan of Caenorhabditis elegans. Antioxidants Redox Signal. 2013;18:481–490. doi: 10.1089/ars.2011.4151. [DOI] [PubMed] [Google Scholar]
- 192.Rangaraju S., Solis G.M., Andersson S.I., Gomez-Amaro R.L., Kardakaris R., Broaddus C.D., Niculescu A.B., 3rd, Petrascheck M. Atypical antidepressants extend lifespan of Caenorhabditis elegans by activation of a non-cell-autonomous stress response. Aging Cell. 2015;14:971–981. doi: 10.1111/acel.12379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Stefanello S.T., Gubert P., Puntel B., Mizdal C.R., de Campos M.M., Salman S.M., Dornelles L., Avila D.S., Aschner M., Soares F.A. Protective effects of novel organic selenium compounds against oxidative stress in the nematode Caenorhabditis elegans. Toxicol. Rep. 2015;2:961–967. doi: 10.1016/j.toxrep.2015.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Kampkotter A., Pielarski T., Rohrig R., Timpel C., Chovolou Y., Watjen W., Kahl R. The Ginkgo biloba extract EGb761 reduces stress sensitivity, ROS accumulation and expression of catalase and glutathione S-transferase 4 in Caenorhabditis elegans. Pharmacol. Res. 2007;55:139–147. doi: 10.1016/j.phrs.2006.11.006. [DOI] [PubMed] [Google Scholar]
- 195.Inokuchi A., Yamamoto R., Morita F., Takumi S., Matsusaki H., Ishibashi H., Tominaga N., Arizono K. Effects of lithium on growth, maturation, reproduction and gene expression in the nematode Caenorhabditis elegans. J. Appl. Toxicol. 2015;35:999–1006. doi: 10.1002/jat.3058. [DOI] [PubMed] [Google Scholar]
- 196.Kotlar I., Colonnello A., Aguilera-Gonzalez M.F., Avila D.S., de Lima M.E., Garcia-Contreras R., Ortiz-Plata A., Soares F.A.A., Aschner M., Santamaria A. Comparison of the toxic effects of quinolinic acid and 3-nitropropionic acid in C. elegans: involvement of the SKN-1 pathway. Neurotox. Res. 2017;33:259–267. doi: 10.1007/s12640-017-9794-x. [DOI] [PubMed] [Google Scholar]