Next Article in Journal
Synthesis, Conformational Analysis and ctDNA Binding Studies of Flavonoid Analogues Possessing the 3,5-di-tert-butyl-4-hydroxyphenyl Moiety
Next Article in Special Issue
Protein Glutathionylation and Glutaredoxin: Role in Neurodegenerative Diseases
Previous Article in Journal
Formulation, Characterization and Evaluation of Innovative O/W Emulsions Containing Curcumin Derivatives with Enhanced Antioxidant Properties
Previous Article in Special Issue
Regulation of Retroviral and SARS-CoV-2 Protease Dimerization and Activity through Reversible Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Defining the S-Glutathionylation Proteome by Biochemical and Mass Spectrometric Approaches

Biological Sciences Division, Pacific Northwest National Laboratory, Richland, WA 99354, USA
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(11), 2272; https://doi.org/10.3390/antiox11112272
Submission received: 28 October 2022 / Accepted: 14 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue Glutaredoxin and Glutathione)

Abstract

:
Protein S-glutathionylation (SSG) is a reversible post-translational modification (PTM) featuring the conjugation of glutathione to a protein cysteine thiol. SSG can alter protein structure, activity, subcellular localization, and interaction with small molecules and other proteins. Thus, it plays a critical role in redox signaling and regulation in various physiological activities and pathological events. In this review, we summarize current biochemical and analytical approaches for characterizing SSG at both the proteome level and at individual protein levels. To illustrate the mechanism underlying SSG-mediated redox regulation, we highlight recent examples of functional and structural consequences of SSG modifications. Finally, we discuss the analytical challenges in characterizing SSG and the thiol PTM landscape, future directions for understanding of the role of SSG in redox signaling and regulation and its interplay with other PTMs, and the potential role of computational approaches to accelerate functional discovery.

1. Introduction

Cellular redox homeostasis is a dynamic and controlled process that continuously balances the generation and removal of electrophiles (e.g., reactive oxygen species, or ROS) and nucleophiles (antioxidant defense systems) under the physiological steady state [1]. At the homeostatic level, ROS such as superoxide anion radical (O2•−), hydrogen peroxide (H2O2), or hydroxyl radical (HO) are produced endogenously during normal aerobic metabolism and play an essential role in cell signaling and regulation [2,3]. A marked increase in ROS production under various stimuli can lead to the development of pathological conditions [4]. To maintain redox homeostasis, multiple types of antioxidants act to scavenge ROS, including antioxidation enzymes (e.g., superoxide dismutase, catalase, glutaredoxin, and thioredoxin), low-molecular weight antioxidants (e.g., glutathione and coenzyme Q), and dietary antioxidants (e.g., vitamin C) [5,6]. Among them, the highly abundant (~1–10 mM in cells) tripeptide glutathione (L-γ-glutamyl-L-cysteinylglycine, GSH) is a key component of the intracellular redox buffering system [7]. Two free GSH molecules can be reversibly oxidized to GSSG by forming a disulfide bond. The overall GSH/GSSG ratio in cells is typically greater than 30:1 or even ≥100:1 to provide a strong reducing cellular environment for normal physiological activities and tightly regulated to maintain a proper cellular redox state. A significant decrease in the GSH/GSSG ratio has been reported during oxidative stress with either over-production of ROS and/or deficient capacity of the antioxidant systems [8,9].
Protein cysteine (Cys) thiols have been increasingly recognized for their critical role in redox-dependent signaling and regulation through a diverse set of post-translational modifications (PTMs), including protein S-glutathionylation (SSG). The reactivity of protein Cys thiols is largely dependent on the pKa of their thiol group, as well as the protein microenvironment where the thiol group resides. While many protein Cys thiols retain pKa > 8, and thus, are almost completely protonated (SH) and stable under physiological conditions [10,11], reactive cysteines exhibit relatively low pKa and can easily form highly reactive thiolate anions (S) under neutral pH. The low thiol pKa can be attributed to neighboring amino acids with positive charges (e.g., arginine, lysine, histidine) or those that promote hydrogen bonds (e.g., serine, threonine, tyrosine) in protein three-dimensional structures [12,13]. It has also been reported that in the dipoles of α-helices, N-termini facing towards Cys residues lowers the thiol pKa and stabilizes thiolate anions [14]. For instance, the well-known redox-sensitive Cys site (Cys-152) in human glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is vicinal to His-179 and has an estimated pKa of 6.03 [15]. An X-ray crystal structure (PDB: 6yne) shows that C152 is located at the N-terminus of an α-helix that can stabilize the thiolate anion (Figure 1). SSG modification of this highly conserved catalytic Cys site is reported in an array of organisms and reversibly inhibits enzymatic activity [16,17,18,19]. The accessibility of Cys thiols to various molecules or solvents (i.e., whether the Cys thiol is buried in protein folds or exposed to the surfaces) is another factor determining Cys reactivity. For instance, the surface-located active site Cys-72 of recombinant human thioredoxin has greater susceptibility to form a SSG modification and was found to be the only site alkylated by iodoacetamide under non-denaturing conditions [20].
Reactive Cys thiols can be reversibly oxidized to several forms of PTMs, including S-sulfenylation (SOH), S-nitrosylation (SNO), SSG, S-polysulfidation (SSnH, n ≥ 1), and disulfides (S-S). These PTMs can lead to changes in protein function and conformation, subcellular location, and protein interaction with proteins or small molecules. Among the PTMs, SSG is recognized as one of the most prevalent forms with significant biological relevance [21,22]. Initially, SSG was considered to serve a protective role to avoid over-oxidation of thiols to irreversible forms, such as S-sulfinylation (SO2H) and S-sulfonylation (SO3H). However, the regulatory and signaling roles of SSG have been increasingly established in recent years. With an expanding list of protein targets being discovered, SSG has been observed as being broadly involved in the regulation of gene expression, signaling, metabolism, cell death and survival, as well as protein folding and degradation [21,23]. The significance of SSG in the physiology and pathology of various human diseases, such as neurodegenerative, cancer, lung, cardiovascular, and other inflammatory diseases, have also been extensively reported and reviewed [9,22,24,25,26].
Given the abundant cellular availability of GSH and GSSG, reactive cysteines can form SSG via a diverse set of non-enzymatic or enzymatic mechanisms. Briefly, SSG can be formed via several ROS-mediated pathways (Figure 2A), including: (1) thiol-disulfide exchange between protein thiols and GSSG; (2) GSH directly reacting with unstable oxidized sulfhydryl intermediates (protein SOH and thiyl radical S) induced by ROS; (3) protein thiols reacting with GSOH or GS [27,28,29,30,31,32]. Nitric oxide (NO)-induced SSG is another proposed mechanism via secondary generation of reactive nitrogen species (RNS). Nitrosoglutathione (GSNO) was reported to induce both SNO and SSG in many in vitro studies on papain, GAPDH, creatine kinase, actin, c-Jun, etc. [33,34]. Furthermore, SSG formation can be facilitated by enzymatic processes. Glutathione S-transferase (GST, primarily GST-pi or GSTP) has been the most studied enzyme promoting SSG formation on a wide array of proteins [35,36,37,38,39]. GSTP is proposed to bind GSH and lower its thiol pKa, thus forming a GS anion. GS can, then, be transferred to receptive cysteines in protein targets [40]. Other studies have demonstrated that oxidized 1-cys peroxiredoxin (Prx-SOH) can be reactivated by GSTP, which transfers GSH to facilitate Prx-SSG via formation of a GSTP(GSH)-Prx-SOH complex, followed by inter-disulfide formation and subsequent reduction (activation) of 1-CysPrx [41,42] (Figure 2B).
While the formation of SSG can occur via non-enzymatic or enzymatic mechanisms, de-glutathionylation is usually catalyzed by enzymes to precisely control the reversal process. Current known enzymes include glutaredoxin (Grx), sulfiredoxin (Srx), GST-omega (GSTO), and thioredoxin (Trx) [43,44,45,46]. Grx is the major player in reducing SSG to free thiols in mammalian cells by attacking the glutathionyl sulfur of SSG and forming SSG on its active-site cysteine of the CXXC motif. Grx-SSG can then be reduced by free GSH via thiol-disulfide exchange (because of the low pKa of the Grx active-site cysteine), while the product GSSG is then reduced by GSSG reductase and NADPH (Figure 2B, monothiol mechanism) [47]. On the other hand, it was also proposed that both cysteines in the active site of Grx can form a disulfide to reduce SSG on protein targets (dithiol mechanism). The monothiol mechanism is recognized as the prevalent mechanism [48]. GSTO shares similar functional structure to Grx, which performs deglutathionylation through a similar mechanism using GSH as co-substrate. By contrast, Srx and Trx were initially characterized to reduce protein-SO2H and disulfides (and other redox PTMs), respectively, but show broader substrate specificity [49]. The role and mechanism of Srx- and Trx-catalyzed deglutathionylation is still not fully understood.
Our knowledge of SSG modifications is largely dependent on analytical approaches for the characterization of this modification. Significant advances in biomedical and mass spectrometry (MS)-based proteomics approaches have transformed our understanding of the SSG proteome. While biochemical methods such as immunoblotting with anti-GSH antibody are generally easy to implement, MS-based proteomics has become an indispensable tool for enabling broad site-specific quantification of SSG modifications and for greatly expanding our knowledge of the SSG proteome. Herein, we present an overview on both biochemical methods and MS-based quantitative proteomics approaches for characterizing SSG with a focus on recent advances in SSG proteome characterization. We also showcase novel discoveries on the SSG-dependent regulation of protein functions. Lastly, limitations of current approaches and future perspectives on the SSG proteome characterization, including structural and functional aspects, are discussed.

2. General Strategies for the Detection of Protein S-Glutathionylation

There are three general strategies for SSG detection (Figure 3). The first one is to directly detect the SSG moieties (through preserving endogenous SSG) by either affinity reagents, such as anti-GSH antibodies, or MS. While this strategy seems to be straightforward, it has been generally challenging for such direct detection due to the lack of specificity of affinity reagents and the labile nature of the modification for MS detection. The second strategy involves selective reduction of SSG back to a free thiol followed by thiol-reactive tagging to enable subsequent detection. In this case, the protein cysteine thiols (SH) need to be initially blocked by alkylation reagents such as N-ethylmaleimide (NEM). Then, SSG is selectively reduced back to a free thiol (e.g., by Grx), and the nascent SH is tagged by various chemical approaches or enriched by a thiol-reactive resin [50], to facilitate subsequent detection either using biochemical assays or MS. The third strategy utilizes a chemoselective GSH probe (GSH analog) to metabolically label reactive protein Cys thiols in situ and form protein-SSG with a tag, which can then be enriched and detected. One should note that the third strategy detects “artificially” formed SSG with the GSH analogs, but not endogenous SSG modifications.
Regardless of the strategies, critical to successful analysis of SSG is the preservation of the endogenous protein thiol redox state and prevention of potential oxidative artifacts during sample processing. Thiol-reactive alkylation reagents such as NEM and iodoacetamide (IAM) are commonly used to block free thiols to minimize potential oxidation, thiol-disulfide exchange (e.g., between reactive thiols and SSG), and disulfide scrambling during sample processing. In the case of SSG, transient thiol PTMs such as SOH and SNO, as well as disulfides, could contribute to a false positive identification [51,52,53]. Acidic pH is also useful to protonate reactive thiols and stabilize the redox proteome. For example, trichloroacetic acid (TCA) precipitation is often used during cell lysis and protein extraction, where a buffer pH around 6–7 can be used in certain processing steps [54,55,56]. After blocking, protein SSG can be subjected to various strategies to facilitate the detection. In the following sections, we will describe various detection methods in detail.

3. Methods for Characterizing Protein S-Glutathionylation

3.1. Direct Detection

Radiolabeled (35S-) GSH was one of the earliest tools used for direct detection of SSG [20,57,58]. To analyze in vivo SSG, protein synthesis is first blocked by preincubating cells with cycloheximide and followed by supplement of 35S-cysteine in cell culture to radioactively label the intracellular GSH pool. Induced by oxidative stress, the synthesized 35S-GSH is employed to form radiolabeled SSG, which can be extracted and visualized on a non-reducing SDS-PAGE gel using autoradiography and phosphor imaging technology (after drying the gel or transferring to nitrocellulose). Using this method, a number of proteins were identified as modified by GSH in vivo, such as GAPDH [59], actin [60], protein kinase C-α [61], and thioredoxin [20]. This method was further coupled with peptide mass fingerprinting to identify more proteins in mammalian cells, yeast, and algae [57,58,62]. However, this method has several limitations, including the disturbed cell physiology resulting from the need to block protein synthesis, potential non-specific signals (e.g., radiolabeled cysteinylated proteins), and the inability to distinguish individual SSG sites.
Another type of popular tool for direct detection of SSG is through Western immunoblotting of glutathionylated proteins on one-dimensional (1D) or two-dimensional (2D) non-reducing gels using anti-GSH antibodies [63,64]. Similar to radiolabeled GSH, the antibody method has limited specificity and sensitivity and is unable to distinguish individual SSG sites within a target protein, which may have different functional consequences. Nevertheless, this method proves to be a useful tool to evaluate global SSG level changes. In conjunction with protein-specific antibodies, it also allows for the quantification of SSG of targeted proteins in complex samples [65,66,67].
Besides the radiolabeled and antibody approaches, mass spectrometry can also directly detect the SSG moiety on proteins or peptides. For example, top-down proteomics has been applied to identify sarcomeric protein PTMs, including SSG [68,69]. By detecting the glutathione conjugation (+305.03 Da) on proteoforms, SSG on troponin I (TnI) was identified as a potential marker that increases with aging. Alternatively, we recently applied bottom-up MS-based global proteome profiling to directly detect thiol PTMs without enrichment [70]. In this approach, proteins were extracted by blocking free thiols and polysulfides (while other thiol PTMs were preserved) and then digested. The resulting peptides were then subjected to fractionation and proteome profiling. In the end, only 81 unique Cys sites with SSG were identified, illustrating the challenge of direct SSG detection by MS proteomics.

3.2. Selective Reduction and Tagging Approaches

Given the challenges in direct detection of SSG, alternative methods are necessary to enable effective characterization. The selective reduction and tagging strategy has become the most applied approach for enabling enrichment and SSG measurement. This strategy was largely built upon the concept of differential alkylation, which was first introduced in the biotin switch technique (BST) for protein SNO detection [71]. BST was later modified to profile other redox PTMs including SSG [72]. The reversibility of the target PTMs is the key to this concept. The specifically reduced and tagged proteins or peptides can be subsequently enriched, thus improving the overall coverage of modified proteins significantly. Common concerns of the selective reduction and tagging strategy are potential false identifications caused by potential incomplete alkylation of free thiols and non-specific reduction of targeted PTMs. Evaluation of these critical steps should be considered during method development and optimization.
To selectively target SSG, highly specific glutaredoxin (Grx), such as Grx1C14S and Grx3C14S, C65Y from E. coli, has been used for the deglutathionylation reaction [72,73,74]. Lind et al. first employed the modified BST to enrich and detect SSG specifically reduced by the mutant Grx3 in the presence of GSH, NADPH, and glutathione reductase (GR) [72]. Free thiols are first alkylated by NEM, followed by the reduction of SSG by Grx3. The resulting nascent free thiols are then alkylated using, NEM-biotin, which can be captured by avidin affinity chromatography (Figure 4A). The captured proteins can be eluted, separated, and visualized on a gel, and further processed for MS-based proteomics analysis. Using this method, 43 novel SSG protein targets were identified including chaperones, cytoskeletal proteins, cell-cycle regulators, and metabolic enzymes in cultured human endothelial-like cells [72]. In addition, this method also allowed visualization of SSG in fixed cells by using N-(3-maleimidylpropionyl) biocytin and streptavidin-FITC [75]. However, these biotin-based methods were generally ineffective in site-specific SSG identifications.
Alternatively, the Qian group introduced a resin-assisted capture (RAC) technique for enriching thiol-containing proteins or peptides [50,76], which was later adapted for SSG proteome profiling [8,73,77]. In this workflow, following the blocking of free thiols by NEM, protein SSG is selectively reduced by Grx1C14S in the presence of GSSG, NADPH, and GR. The nascent protein thiols are then captured by a thiol-affinity resin, which contains a 2-thiopyridyl disulfide functional group to react with thiol-containing proteins via thiol-disulfide exchange, forming a mixed disulfide bond. The captured proteins can be eluted by DTT for Western blot or proteomics analyses. The stable covalent capture allows for on-resin protein digestion and subsequent multiplexed isobaric labeling (e.g., with tandem mass tags, TMT). Then, the TMT-labeled, Cys-containing peptides can be eluted from the resin by DTT (Figure 4B). This RAC-TMT approach offers several notable advantages. First, the covalent binding of Cys-containing proteins or peptides on the resin permits harsh washing conditions, thus allowing a high enrichment specificity (i.e., typically >95% of final peptides containing cysteines). Second, the multiplex quantification design enables not only comparative quantification of SSG across different biological samples, but also stoichiometric estimation of SSG occupancy at the Cys site level through incorporating both SSG channels and total thiol channels in the same TMT plex experiment. The total thiol samples are prepared with the reduction of all reversibly oxidized thiols by DTT (without initial blocking of the free thiols), thus representing the sum of all thiols on each given Cys site (except irreversible modifications such as SO2H and SO3H) (Figure 4B). Thus, the SSG occupancy at each Cys site can be calculated based on the ratio of TMT reporter ion intensities from the SSG channels to those of the total thiol channels [77,78]. The quantification of individual PTM occupancy at the site-specific level is a critical aspect to understand potential functional consequences of PTMs, which was often overlooked in many current studies. Interested readers are redirected to previous reviews on stoichiometric quantification in redox proteomics [79,80].
RAC-TMT and similar approaches have been applied to investigate protein SSG changes under physiological and pathological conditions in different cell types and tissues [8,77,78,81,82]. For example, in a recent study of redox regulation in hyperoxia-induced lung injury, over 7600 SSG sites were quantified by RAC-TMT, providing a landscape view of the SSG proteome in the mouse lung [83]. In another study, Duan et al. quantified both SSG and total oxidation occupancies for ~4000 Cys sites under basal conditions in mouse RAW macrophages, revealing a mean occupancy of 4.0% for SSG and 11.9% for total oxidation [78]. Interestingly, it was observed that the average occupancies of SSG and total oxidation for individual subcellular compartments was well correlated with their redox potentials.
Recently, an approach called GluICAT, a modified version of the original oxICAT method [84], also demonstrated stoichiometric quantification of SSG at the site level [85,86]. Using isotopically light or heavy thiol-reactive ICAT, free thiols and Grx1C14S-reduced protein SSG are differentially labeled, enriched, and quantified to determine SSG level vs. free thiol level (heavy-to-light ratio) of individual Cys sites (Figure 4C). This method was able to identify 2133 SSG-peptides (~2300 SSG sites) in Leber’s hereditary optic neuropathy (LHON) fibroblasts, where 439 peptides were quantified to show an increased SSG level [85]. One notable limitation of this method is that it quantifies the ratio of SSG vs. free thiols without considering other types of thiol PTMs (e.g., disulfides). Thus, the data need to be carefully interpreted with consideration of this important caveat.
Finally, there are several other approaches including the Cys-reactive phosphate tag (CPT)- TMT [87] or OxiTMT methods [88] that are potentially applicable, but have not yet been demonstrated for SSG analyses.

3.3. Chemoselective Probes

The concept of using chemoselective probes to label free thiols either in vivo or in vitro to form protein-SSG analogs is an interesting chemical biology strategy for enabling the characterization of protein-SSG products. In this strategy, glutathione analogs or tagged amino acids (to produce tagged GSH endogenously) were introduced either for in vivo reactions by using membrane permeable probes during cell culture or for in vitro reactions. Glutathione ethyl ester (BioGEE) [89] and GSSG-biotin [90,91] are the first class of glutathione analogs for in vivo and in vitro labeling, and the biotin tag facilitates the enrichment of protein-SSG products. The use of BioGEE or GSSG-biotin is usually accompanied by addition of pro-oxidants to induce oxidative stress, and the resulting protein-SSG products can be identified by Western blot using anti-biotin antibodies or streptavidin-horseradish peroxidase (HRP). The biotin tags also allow capture of glutathionylated proteins by streptavidin agarose beads, which can be released from beads by a reducing reagent (e.g., DTT), thereby cleaving the mixed disulfide bond within the SSG moiety. The eluted proteins can be further analyzed by gel or mass spectrometry [90,92]. Furthermore, Brennan et al. demonstrated the capacity to image biotinylated protein SSG stained with ExtrAvidin-FITC in cells using confocal fluorescence microscopy [91].
Metabolic labeling with clickable GSH probes has been developed as an alternative approach instead of biotin tags. Clickable GSH [93,94], including azido-GSH, allyl-GSH, and allyl-O-GSH, can be biosynthesized in vivo by expressing an engineered glutathione synthetase (GS M4, F152A/S151G), which catalyzes azido-Ala, allyl-Gly, and allyl-Ser respectively instead of the endogenous substrate Gly (Figure 5). A subsequent click reaction (azide-alkyne or tetrazine-alkene bioorthogonal chemistry) can then tag the glutathionylated proteins with biotin or a fluorophore after protein extraction, thus enabling biotin-streptavidin pull-down or fluorescence labeling for detection. This method has been further extended to LC-MS/MS profiling of glutathionylated proteins by using a cleavable biotin-DADPS-alkyne. Biotinylated proteins are enriched by streptavidin agarose and digested by trypsin, then released by an acidic cleavage of the DADPS linker for detection by LC-MS/MS (Figure 5) [95]. By developing light or heavy isotopic labeled clickable GSH, this method was applied to mouse cardiomyocyte cells which quantified 1398 SSG-peptides induced by H2O2 [96].
It should be noted that the addition of GSSG-biotin to cells in culture itself causes oxidative stress due to an increase in cellular GSSG [91]. On the other hand, intracellularly synthesized clickable GSH may not disturb the cellular GSH/GSSG ratio. Azido-GSH, azido-GSSG, and an azido-GSH-modified model peptide were reported to be efficient substrates of Grx, GR, and GST omega, which does not interfere with the reversibility of the glutathionylation process and GSH/GSSG cycling [93]. However, the application of such probes is limited to living cells that can exogenously express GS M4 and not other types of biological samples, such as tissues. Although the use of chemoselective probes overcomes the potential non-specificity issue of selective reduction and tagging approaches, there are caveats to consider. Introducing these probes themselves is likely to perturb the physiological state and native protein-SSGs are not analyzed.

3.4. Global Profiling of the SSG Proteome

Enabled by the advances described above, extensive studies for profiling the SSG proteome in various cells or tissues have been reported. Table 1 lists selected examples of recent studies aiming at the global profiling of the SSG proteome, illustrating the different levels of proteome coverage achieved in various samples using different approaches.

4. Functional Roles of S-Glutathionylation

The advances in analytical approaches for SSG characterization has resulted in a continuing expansion of the list of protein SSG modifications. Protein SSG has been reported to cause a decrease (in most cases) or increase in the activity of proteins that are involved in energy metabolism and glycolysis, gene expression, signaling, calcium-dependent channels, apoptosis and autophagy, cell structure and dynamics, and protein folding [48,99,100]. Dynamic changes of protein SSGs were also shown to be prevalent in oxidative stress-associated disease progression, thus providing insights in clinical applications, including discovery of biomarkers, molecular targets for drug development, or therapeutic interventions [48,77,101,102,103,104].
With the identification of SSG modifications, the next important step is to elucidate the structural and functional relevance by orthogonal methods for a mechanistic understanding of their regulatory roles. To pinpoint the regulatory roles of protein-SSG, detailed functional studies will be required. Table 2 lists a set of novel SSG regulated proteins discovered by proteomics followed by functional studies.
GAPDH is one of the most extensively studied redox-sensitive proteins due to its diverse cellular functions in glycolysis, cellular apoptosis, cellular survival, DNA repair, and nuclear tRNA export. The enzymatic activity of GAPDH is inhibited when the catalytic cysteine (in a highly conserved CXXC motif; Figure 1) is glutathionylated [19,113]. Recently, it was reported that glutathionylation of the catalytic cysteine in recombinant plant GAPDH results in partially misfolded proteins and promotes protein self-aggregation to generate insoluble oligomeric particles. The aggregation process can be accelerated by further oxidation of SSG to an intramolecular disulfide bond with the other cysteine in the CXXC motif [110]. It was also found that SSG on C247 is necessary for GAPDH nuclear translocation, while an intra-subunit disulfide bond was observed within the CXXC motif in HEK 293T cells. The nuclear GAPDH binds to Sirtuin-1 (SirT1) and leads to glutathionylation of SirT1 and inhibition of SirT1 deacetylase activity, which subsequently impairs SirT1-p53 interaction and induces apoptosis [114]. Thus, these studies significantly improved the knowledge on the structural and functional consequences of SSG modification and potential pathogenesis related to PTMs on GAPDH [115].
Site specificity and context dependence are important features of SSG regulation. Human Hsp70 SSG was well characterized using multiple techniques including modified BST, NMR, size-exclusion chromatography, and intrinsic fluorescence spectroscopy [105]. Among the five Cys residues, SSGs on C574 and C603 located in the C-terminal α-helical lid of the substrate-binding domain led to unfolding of the lid structure, thus blocking the substrate-binding site and decreasing peptide binding activity. Another example of site-specific regulation is fatty-acid binding protein (FABP5), a protein discovered from RAC-TMT SSG proteome profiling in microphages [9]. FABP5 SSG facilitated fatty acid binding and nuclear translocation under oxidative stress. Although four cysteines of FABP5 were identified with SSG, mutagenesis experiments elucidated that only C127 was required for these events. Molecular docking analysis revealed that C127 was within the ligand-binding pocket, which could perfectly accommodate GSH molecules in both monomeric or dimeric FABP5 [9]. The same study also showed that FABP5 SSG inhibited inflammation in macrophages by binding and activating nuclear receptor PPARβ/δ and its target genes.
Notably, glutathionylation of cryptic cysteines on the elastic protein titin in cardiac muscle was also reported. Cryptic cysteines are buried inside of a protein and not accessible to oxidants under basal conditions. Thus, they remain inert in the folding of the immunoglobulin (Ig) domains in I-band of titin. However, glutathionylation of these residues occurred upon unfolding of Ig domain by heating or mechanical force in the presence of GSSG, which was reversible by Grx [116]. Recently, in vivo characterization of SSG (or oxidation) in unfolded Ig domains suggested a role in regulating human cardiomyocyte passive force [106]. SSG of one Cys (C13585) in a distal Ig domain led to the weakening of titin-based stiffness by stabilizing the unfolded and misfolded domains and promoting aggregation. The novel finding of cryptic Cys modification contrasts common redox-sensitive cysteines with low pKa and good solvent accessibility, further underlining the structure-function relationship in understanding the regulatory role of PTMs.
Interestingly, various PTMs can occur simultaneously on the same protein, group of proteins, or even the same residues, which increases proteome complexity [117]. Crosstalk between PTMs is proposed to contribute to higher order regulation, such as the relationship between protein SSG and phosphorylation. SSG modification in the unfolded Ig domain of titin prevents the refolding of the domain, thus potentially exposing phosphorylation sites. Following thermal unfolding of the Ig domain, glutathionylation of the domain led to higher phosphorylation levels at a single serine residue by Ca2+/calmodulin-dependent protein kinase-II delta (CaMKIIδ), compared to the thermal unfolded domain without SSG [106]. The interplay between SSG and phosphorylation was also demonstrated in another work focusing on two myofilament proteins (myosin binding protein C (cMyBP-C) and cardiac troponin I (cTnI)) and titin in human failing hearts [118]. Along with increased levels of hydrogen peroxide and lipid peroxide and decreased levels of GSH, glutathionylation of cMyBP-C, cTnI, and titin were observed in end-stage human failing heart samples. Meanwhile, oxidized protein kinase G (PKG) was found forming dimers and polymers, which were associated with decreased activity, causing titin hypo-phosphorylation. Lower levels of protein kinase A (PKA)-dependent phosphorylation of cMyBP-C and cTnI was also observed in failing heart samples compared to non-failing donors. The dysregulated phosphorylation by PKA is potentially attributed to altered PKA affinity or accessibility of phosphorylation sites following oxidation of vicinal cysteines. This study demonstrates that the contractile dysfunction in human cardiomyocytes is co-regulated by both oxidative stress and phosphorylation, providing therapeutic targets to cardiovascular diseases.

5. Conclusions and Perspectives

Protein S-glutathionylation plays a critical role in ROS-mediated signaling and various cellular functions under physiological conditions, while abnormal regulation of SSG can lead to disease development. Recent biochemical and analytical advances have enabled in-depth SSG proteome profiling, and several studies have reported that the SSG modifications occur broadly across all subcellular compartments [77,78,83]. The coverage of the SSG proteome might be further enhanced by adopting recent workflows involving fractionation followed by LC-MS/MS, where a deep coverage of the thiol redox proteome was demonstrated for thiol oxidation [87,119]. In typical quantitative analyses of the SSG proteome between different biological conditions, it is possible to identify 100s or even 1000s SSG sites with statistically significant changes. Besides statistical significance and fold-changes, the stoichiometry information of Cys site SSG occupancy will provide a significant value in determining potential regulatory Cys sites [80]. Future studies investigating the roles of specific redox-regulating enzymes such as GSTP might help to elucidate the broad regulatory landscape of the SSG proteome [81].
One of the notable challenges in the SSG proteome profiling is the difficulty to directly identify in situ SSG moieties by MS approaches, partially due to the bulky tripeptide structure. The SSG moiety on Cys residue can undergo fragmentation in MS/MS mode, leading to neutral losses via cleavage of the peptide bond, carbon-sulfur bond, and even disulfide bond, depending on the dissociation techniques employed by tandem mass spectrometry. A recent LC-MS/MS analysis of in vitro cofilin glutathionylation reported neutral loss of the glutamic acid residue (−129 Da) from the SSG moiety on precursor ions following collision-induced dissociation (CID) [120]. In contrast, electron transfer dissociation (ETD) tended to break the disulfide bond as well as release the whole SSG moiety (−305 Da), which was detectable as a glutathione oxonium ion (308.4 m/z) in the same spectrum. Other neutral losses resulting from cleaved carbon-sulfur bond in Cys residues were observed as well. Peptide identifications of such complicated mass spectra represents a challenge in current database searching algorithms for shotgun proteomics.
Another significant challenge for SSG data interpretation is the complex landscape of the thiol PTMs. In a recent study involving direct detection of multiple types of thiol PTMs, many proteins were observed with ≥4 types of thiol PTMs on single Cys site [70]. These include well-established redox-sensitive protein Cys sites such as the Cys150 of GAPDH with 6 types of thiol PTMs: SSG, SSH, SSSH, SO2H, SO3H, and a disulfide with C154. This observation of the thiol PTM landscape shed light on the complexity of thiol redox regulation, which might not be exclusively dependent on a particular type of PTM such as SSG or SNO. Further development of techniques that quantify the stoichiometry of multiple types of thiol PTMs will greatly improve our understanding of the details of redox regulation.
The presence of multiple types of redox PTMs on the same protein/residue also implicate distinct regulatory roles or potential functional interplay among them, a largely untapped area in redox biology. Multiple types of thiol PTMs are hypothetically affecting protein functions or activities differently because of the diversity in their chemical structures. In addition, the interplay among distinct redox PTMs can serve as a redox-switch regulatory mechanism on proteins and biological processes. However, the crosstalk between different types of thiol PTMs has been long overlooked due to the lack of methods to characterize them. Recently, Gao et al. reported a switch from SSG to SSH in a group of 250 glutathionylated proteins when diamide-treated pancreatic beta cells were subsequently treated with a H2S donor, which restored metabolic flux in glycolysis and TCA cycle under oxidative stress [121]. Competitive actions by SNO and SSG were also reported on the same Cys site of troponin I exerting opposing effects on Ca2+ sensitivity in rat fast-twitch muscle fiber [112]. Dynamic characterization of individual or multiple types of thiol PTMs across timepoints and conditions at the site level are expected to facilitate an understanding of the synergetic or competing regulation networks in redox signaling. The crosstalk between redox PTMs and other types of PTMs such as phosphorylation is another interesting area to explore. For example, broad dynamic Cys thiol oxidation was observed during the epidermal growth factor signaling in cancer cells [122], suggesting interplay between redox and phosphorylation PTMs.
The ultimate goal of PTM profiling is to discover and validate novel functional PTM regulators including SSG. While MS-based proteomics can provide large-scale and broad characterization of thiol PTMs, functional studies that investigate the impact of these PTMs on protein structures and activities is a labor intensive and slow process. There is a possibility that advanced computational approaches to integrate these PTMs sites such as SSG can help to elucidate functional significance of SSG in more rapid fashion. One possible approach is to map the PTM sites in protein structures so that their structural localization can inform their effects on protein functions. With the recent advancement of protein structure prediction, AlphaFold 2 now can provide highly accurate predictive structures that are close to experimentally resolved ones [123]. The AlphaFold database now has predicted structures of almost all proteins available, with more than 200 million proteins from about one million species [124]. One can easily map the quantified SSG modification changes to proteins structures [125] and calculate the relative positions of the modified site to the protein surface or specific domains. These analyses can help infer the possible perturbations from SSG on proteins structure, function, or activity. Besides the static structural analysis, it is also promising to study the function of SSG sites using molecular dynamics simulations or Monte Carlo sampling of the conformation or energy landscapes [126,127,128]. Structure and molecular dynamics analyses can facilitate prioritization of functionally important SSG sites and guide the design of validating experiments.
Finally, other forms of GSH-mediated PTMs besides SSG were also reported and may involve in regulatory events. For example, it has been observed that β-elimination of serine, threonine, and cysteine in proteins or phosphoserine and phosphothreonine result in dehydro-amino acids, such as dehydroalanine (Dha) and didehydrobutyrine (Dhb) [129]. The Dha/Dhb can react with GSH via a Michael addition reaction to form an irreversible carbon-glutathionylation (C-SG) modification. The C-SG modifications have been reported in aging tissues [130]. More recently, the protein C-SG was reported on kinases such as MEK1 catalyzed by LanC-like enzymes (LanCLs), which inactivated the kinase activity [131]. The addition of glutathione to Dha/Dhb in proteins can be catalyzed by LanCLs, thus forming CSG. Interestingly, the damage product MEK1-Dha was observed to be aberrantly activated, and LanCL-catalyzed CSG was proposed as a repair mechanism for protein damage physiologically [131].

Author Contributions

X.L. and W.-J.Q. drafted manuscript and prepared figures; X.L., T.Z., N.J.D., S.F., M.J.G. and W.-J.Q. edited and revised manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Portions of the work were supported by NIH Grants R01 DK122160, R01 HL139335, and U01 CA227544.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ursini, F.; Maiorino, M.; Forman, H.J. Redox homeostasis: The Golden Mean of healthy living. Redox Biol. 2016, 8, 205–215. [Google Scholar] [CrossRef] [PubMed]
  2. Snezhkina, A.V.; Kudryavtseva, A.V.; Kardymon, O.L.; Savvateeva, M.V.; Melnikova, N.V.; Krasnov, G.S.; Dmitriev, A.A. ROS Generation and Antioxidant Defense Systems in Normal and Malignant Cells. Oxid. Med. Cell. Longev. 2019, 2019, 6175804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhang, J.; Wang, X.; Vikash, V.; Ye, Q.; Wu, D.; Liu, Y.; Dong, W. ROS and ROS-Mediated Cellular Signaling. Oxid. Med. Cell. Longev. 2016, 2016, 4350965. [Google Scholar] [CrossRef] [Green Version]
  4. Juan, C.; de la Lastra, J.P.; Plou, F.; Pérez-Lebeña, E. The Chemistry of Reactive Oxygen Species (ROS) Revisited: Outlining Their Role in Biological Macromolecules (DNA, Lipids and Proteins) and Induced Pathologies. Int. J. Mol. Sci. 2021, 22, 4642. [Google Scholar] [CrossRef]
  5. Marengo, B.; Nitti, M.; Furfaro, A.L.; Colla, R.; Ciucis, C.D.; Marinari, U.M.; Pronzato, M.A.; Traverso, N.; Domenicotti, C. Redox Homeostasis and Cellular Antioxidant Systems: Crucial Players in Cancer Growth and Therapy. Oxid. Med. Cell. Longev. 2016, 2016, 6235641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. He, L.; He, T.; Farrar, S.; Ji, L.; Liu, T.; Ma, X. Antioxidants Maintain Cellular Redox Homeostasis by Elimination of Reactive Oxygen Species. Cell. Physiol. Biochem. 2017, 44, 532–553. [Google Scholar] [CrossRef] [PubMed]
  7. 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. [Google Scholar] [CrossRef]
  8. Duan, J.; Kodali, V.K.; Gaffrey, M.J.; Guo, J.; Chu, R.K.; Camp, D.G.; Smith, R.D.; Thrall, B.D.; Qian, W.J. Quantitative Profiling of Protein S-Glutathionylation Reveals Redox-Dependent Regulation of Macrophage Function during Nanoparticle-Induced Oxidative Stress. ACS Nano 2016, 10, 524–538. [Google Scholar] [CrossRef] [Green Version]
  9. Guo, Y.; Liu, Y.; Zhao, S.; Xu, W.; Li, Y.; Zhao, P.; Wang, D.; Cheng, H.; Ke, Y.; Zhang, X. Oxidative stress-induced FABP5 S-glutathionylation protects against acute lung injury by suppressing inflammation in macrophages. Nat. Commun. 2021, 12, 7094. [Google Scholar] [CrossRef]
  10. Kyte, J. Structure in Protein Chemistry; Garland Science: New York, NY, USA, 2006. [Google Scholar]
  11. Cumming, R.C.; Andon, N.L.; Haynes, P.A.; Park, M.; Fischer, W.H.; Schubert, D. Protein Disulfide Bond Formation in the Cytoplasm during Oxidative Stress. J. Biol. Chem. 2004, 279, 21749–21758. [Google Scholar] [CrossRef] [PubMed]
  12. Poole, L.B. The basics of thiols and cysteines in redox biology and chemistry. Free Radic. Biol. Med. 2015, 80, 148–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Matsui, R.; Ferran, B.; Oh, A.; Croteau, D.; Shao, D.; Han, J.; Pimentel, D.R.; Bachschmid, M.M. Redox Regulation via Glutaredoxin-1 and Protein S-Glutathionylation. Antioxid. Redox Signal. 2020, 32, 677–700. [Google Scholar] [CrossRef] [PubMed]
  14. Kortemme, T.; Creighton, T.E. Ionisation of Cysteine Residues at the Termini of Model α-Helical Peptides. Relevance to Unusual Thiol pKaValues in Proteins of the Thioredoxin Family. J. Mol. Biol. 1995, 253, 799–812. [Google Scholar] [CrossRef] [PubMed]
  15. Martyniuk, C.J.; Fang, B.; Koomen, J.M.; Gavin, T.; Zhang, L.; Barber, D.S.; Lopachin, R.M. Molecular mechanism of glyceraldehyde-3-phosphate dehydrogenase inactivation by α,β-unsaturated carbonyl derivatives. Chem. Res. Toxicol. 2011, 24, 2302–2311. [Google Scholar] [CrossRef] [Green Version]
  16. Ravichandran, V.; Seres, T.; Moriguchi, T.; Thomas, J.A.; Johnston, R.B. S-thiolation of glyceraldehyde-3-phosphate dehydrogenase induced by the phagocytosis-associated respiratory burst in blood monocytes. J. Biol. Chem. 1994, 269, 25010–25015. [Google Scholar] [CrossRef]
  17. Barinova, K.V.; Serebryakova, M.V.; Muronetz, V.I.; Schmalhausen, E.V. S-glutathionylation of glyceraldehyde-3-phosphate dehydrogenase induces formation of C150-C154 intrasubunit disulfide bond in the active site of the enzyme. Biochim. Biophys. Acta (BBA)—Gen. Subj. 2017, 1861, 3167–3177. [Google Scholar] [CrossRef]
  18. Bedhomme, M.; Adamo, M.; Marchand, C.H.; Couturier, J.; Rouhier, N.; Lemaire, S.D.; Zaffagnini, M.; Trost, P. Glutathionylation of cytosolic glyceraldehyde-3-phosphate dehydrogenase from the model plant Arabidopsis thaliana is reversed by both glutaredoxins and thioredoxins in vitro. Biochem. J. 2012, 445, 337–347. [Google Scholar] [CrossRef] [Green Version]
  19. Tossounian, M.-A.; Zhang, B.; Gout, I. The Writers, Readers, and Erasers in Redox Regulation of GAPDH. Antioxidants 2020, 9, 1288. [Google Scholar] [CrossRef]
  20. Casagrande, S.; Bonetto, V.; Fratelli, M.; Gianazza, E.; Eberini, I.; Massignan, T.; Salmona, M.; Chang, G.; Holmgren, A.; Ghezzi, P. Glutathionylation of human thioredoxin: A possible crosstalk between the glutathione and thioredoxin systems. Proc. Natl. Acad. Sci. USA 2002, 99, 9745–9749. [Google Scholar] [CrossRef] [Green Version]
  21. Mieyal, J.J.; Chock, P.B. Posttranslational modification of cysteine in redox signaling and oxidative stress: Focus on s-glutathionylation. Antioxid. Redox Signal. 2012, 16, 471–475. [Google Scholar] [CrossRef]
  22. Pal, D.; Rai, A.; Checker, R.; Patwardhan, R.S.; Singh, B.; Sharma, D.; Sandur, S.K. Role of protein S-Glutathionylation in cancer progression and development of resistance to anti-cancer drugs. Arch. Biochem. Biophys. 2021, 704, 108890. [Google Scholar] [CrossRef] [PubMed]
  23. Mieyal, J.J.; Gallogly, M.M.; Qanungo, S.; Sabens, E.A.; Shelton, M.D. Molecular mechanisms and clinical implications of reversible protein S-glutathionylation. Antioxid. Redox Signal. 2008, 10, 1941–1988. [Google Scholar] [CrossRef] [PubMed]
  24. Cha, S.J.; Kim, H.; Choi, H.-J.; Lee, S.; Kim, K. Protein Glutathionylation in the Pathogenesis of Neurodegenerative Diseases. Oxid. Med. Cell. Longev. 2017, 2017, 2818565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chia, S.B.; Elko, E.A.; Aboushousha, R.; Manuel, A.M.; Wetering, C.v.d.; Druso, J.E.; Velden, J.v.d.; Seward, D.J.; Anathy, V.; Irvin, C.G.; et al. Dysregulation of the glutaredoxin/S-glutathionylation redox axis in lung diseases. Am. J. Physiol.-Cell Physiol. 2020, 318, C304–C327. [Google Scholar] [CrossRef] [PubMed]
  26. Shelton, M.D.; Mieyal, J.J. Regulation by reversible S-glutathionylation: Molecular targets implicated in inflammatory diseases. Mol. Cells 2008, 25, 332–346. [Google Scholar] [PubMed]
  27. Gilbert, H.F. [2] Thiol/Disulfide Exchange Equilibria and Disulfidebond Stability. In Methods in Enzymology; Academic Press: Cambridge MA, USA, 1995; Volume 251, pp. 8–28. [Google Scholar]
  28. Gilbert, H.F. Molecular and Cellular Aspects of Thiol–Disulfide Exchange. In Advances in Enzymology and Related Areas of Molecular Biology; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1990; pp. 69–172. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  30. Klatt, P.; Molina, E.P.; de Lacoba, M.G.; Padilla, C.A.; Martínez-Galisteo, E.; Barcena, J.; Lamas, S. Redox regulation of c-Jun DNA binding by reversible S-glutathiolation. FASEB J. 1999, 13, 1481–1490. [Google Scholar] [CrossRef]
  31. Zhang, J.; Ye, Z.-w.; Singh, S.; Townsend, D.M.; Tew, K.D. An evolving understanding of the S-glutathionylation cycle in pathways of redox regulation. Free Radic. Biol. Med. 2018, 120, 204–216. [Google Scholar] [CrossRef]
  32. Kalinina, E.; Novichkova, M. Glutathione in Protein Redox Modulation through S-Glutathionylation and S-Nitrosylation. Molecules 2021, 26, 435. [Google Scholar] [CrossRef]
  33. Giustarini, D.; Milzani, A.; Aldini, G.; Carini, M.; Rossi, R.; Dalle-Donne, I. S-Nitrosation versus S-Glutathionylation of Protein Sulfhydryl Groups by S-Nitrosoglutathione. Antioxid. Redox Signal. 2005, 7, 930–939. [Google Scholar] [CrossRef]
  34. Klatt, P.; Molina, E.P.; Lamas, S. Nitric Oxide Inhibits c-Jun DNA Binding by Specifically TargetedS-Glutathionylation. J. Biol. Chem. 1999, 274, 15857–15864. [Google Scholar] [CrossRef] [Green Version]
  35. Uemura, T.; Tsaprailis, G.; Gerner, E.W. GSTΠ stimulates caveolin-1-regulated polyamine uptake via actin remodeling. Oncotarget 2019, 10, 5713–5723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhang, J.; Ye, Z.-w.; Chen, W.; Manevich, Y.; Mehrotra, S.; Ball, L.; Janssen-Heininger, Y.; Tew, K.D.; Townsend, D.M. S-Glutathionylation of estrogen receptor α affects dendritic cell function. J. Biol. Chem. 2018, 293, 4366–4380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wetzelberger, K.; Baba, S.P.; Thirunavukkarasu, M.; Ho, Y.-S.; Maulik, N.; Barski, O.A.; Conklin, D.J.; Bhatnagar, A. Postischemic Deactivation of Cardiac Aldose Reductase: ROLE OF GLUTATHIONE S-TRANSFERASE P AND GLUTAREDOXIN IN REGENERATION OF REDUCED THIOLS FROM SULFENIC ACIDS. J. Biol. Chem. 2010, 285, 26135–26148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ye, Z.-W.; Zhang, J.; Ancrum, T.; Manevich, Y.; Townsend, D.M.; Tew, K.D. Glutathione S-Transferase P-Mediated Protein S-Glutathionylation of Resident Endoplasmic Reticulum Proteins Influences Sensitivity to Drug-Induced Unfolded Protein Response. Antioxid. Redox Signal. 2017, 26, 247–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Townsend, D.M.; Manevich, Y.; He, L.; Hutchens, S.; Pazoles, C.J.; Tew, K.D. Novel Role for Glutathione S-Transferase π: Regulator of Protein S-Glutathionylation Following Oxidative and Nitrosative Stress. J. Biol. Chem. 2009, 284, 436–445. [Google Scholar] [CrossRef] [Green Version]
  40. Graminski, G.F.; Kubo, Y.; Armstrong, R.N. Spectroscopic and kinetic evidence for the thiolate anion of glutathione at the active site of glutathione S-transferase. Biochemistry 1989, 28, 3562–3568. [Google Scholar] [CrossRef]
  41. Manevich, Y.; Feinstein, S.I.; Fisher, A.B. Activation of the antioxidant enzyme 1-CYS peroxiredoxin requires glutathionylation mediated by heterodimerization with πGST. Proc. Natl. Acad. Sci. USA 2004, 101, 3780–3785. [Google Scholar] [CrossRef] [Green Version]
  42. Ralat, L.A.; Manevich, Y.; Fisher, A.B.; Colman, R.F. Direct Evidence for the Formation of a Complex between 1-Cysteine Peroxiredoxin and Glutathione S-Transferase π with Activity Changes in Both Enzymes. Biochemistry 2006, 45, 360–372. [Google Scholar] [CrossRef]
  43. Mannervik, B.; Axelsson, K. Role of cytoplasmic thioltransferase in cellular regulation by thiol-disulphide interchange. Biochem. J. 1980, 190, 125–130. [Google Scholar] [CrossRef] [Green Version]
  44. Gladyshev, V.N.; Liu, A.; Novoselov, S.V.; Krysan, K.; Sun, Q.-A.; Kryukov, V.M.; Kryukov, G.V.; Lou, M.F. Identification and Characterization of a New Mammalian Glutaredoxin (Thioltransferase), Grx2. J. Biol. Chem. 2001, 276, 30374–30380. [Google Scholar] [CrossRef] [Green Version]
  45. Menon, D.; Board, P.G. A Role for Glutathione Transferase Omega 1 (GSTO1-1) in the Glutathionylation Cycle. J. Biol. Chem. 2013, 288, 25769–25779. [Google Scholar] [CrossRef] [Green Version]
  46. Park, J.W.; Mieyal, J.J.; Rhee, S.G.; Chock, P.B. Deglutathionylation of 2-Cys Peroxiredoxin Is Specifically Catalyzed by Sulfiredoxin*. J. Biol. Chem. 2009, 284, 23364–23374. [Google Scholar] [CrossRef] [Green Version]
  47. Gallogly, M.M.; Starke, D.W.; Leonberg, A.K.; Ospina, S.M.E.; Mieyal, J.J. Kinetic and Mechanistic Characterization and Versatile Catalytic Properties of Mammalian Glutaredoxin 2: Implications for Intracellular Roles. Biochemistry 2008, 47, 11144–11157. [Google Scholar] [CrossRef] [Green Version]
  48. Musaogullari, A.; Chai, Y.-C. Redox Regulation by Protein S-Glutathionylation: From Molecular Mechanisms to Implications in Health and Disease. Int. J. Mol. Sci. 2020, 21, 8113. [Google Scholar] [CrossRef]
  49. Wu, C.; Parrott, A.M.; Fu, C.; Liu, T.; Marino, S.M.; Gladyshev, V.N.; Jain, M.R.; Baykal, A.T.; Li, Q.; Oka, S.; et al. Thioredoxin 1-Mediated Post-Translational Modifications: Reduction, Transnitrosylation, Denitrosylation, and Related Proteomics Methodologies. Antioxid. Redox Signal. 2011, 15, 2565–2604. [Google Scholar] [CrossRef] [Green Version]
  50. Guo, J.; Gaffrey, M.J.; Su, D.; Liu, T.; Camp, D.G., II; Smith, R.D.; Qian, W.J. Resin-assisted enrichment of thiols as a general strategy for proteomic profiling of cysteine-based reversible modifications. Nat. Protoc. 2014, 9, 64–75. [Google Scholar] [CrossRef] [Green Version]
  51. Carroll, L.; Jiang, S.; Irnstorfer, J.; Beneyto, S.; Ignasiak, M.T.; Rasmussen, L.M.; Rogowska-Wrzesinska, A.; Davies, M.J. Oxidant-induced glutathionylation at protein disulfide bonds. Free Radic. Biol. Med. 2020, 160, 513–525. [Google Scholar] [CrossRef]
  52. Jiang, S.; Carroll, L.; Rasmussen, L.M.; Davies, M.J. Oxidation of protein disulfide bonds by singlet oxygen gives rise to glutathionylated proteins. Redox Biol. 2021, 38, 101822. [Google Scholar] [CrossRef]
  53. Wolhuter, K.; Whitwell, H.J.; Switzer, C.H.; Burgoyne, J.R.; Timms, J.F.; Eaton, P. Evidence against Stable Protein S-Nitrosylation as a Widespread Mechanism of Post-translational Regulation. Mol. Cell 2018, 69, 438–450.e435. [Google Scholar] [CrossRef]
  54. Guo, J.; Nguyen, A.Y.; Dai, Z.; Su, D.; Gaffrey, M.J.; Moore, R.J.; Jacobs, J.M.; Monroe, M.E.; Smith, R.D.; Koppenaal, D.W.; et al. Proteome-wide Light/Dark Modulation of Thiol Oxidation in Cyanobacteria Revealed by Quantitative Site-specific Redox Proteomics. Mol. Cell. Proteom. 2014, 13, 3270–3285. [Google Scholar] [CrossRef] [Green Version]
  55. Gaffrey, M.J.; Day, N.J.; Li, X.; Qian, W.J. Resin-Assisted Capture Coupled with Isobaric Tandem Mass Tag Labeling for Multiplexed Quantification of Protein Thiol Oxidation. J. Vis. Exp. 2021, 172, e62671. [Google Scholar] [CrossRef]
  56. Lu, S.; Fan, S.-B.; Yang, B.; Li, Y.-X.; Meng, J.-M.; Wu, L.; Li, P.; Zhang, K.; Zhang, M.-J.; Fu, Y.; et al. Mapping native disulfide bonds at a proteome scale. Nat. Methods 2015, 12, 329–331. [Google Scholar] [CrossRef]
  57. Michelet, L.; Zaffagnini, M.; Vanacker, H.; Le Maréchal, P.; Marchand, C.; Schroda, M.; Lemaire, S.D.; Decottignies, P. In Vivo Targets of S-Thiolation in Chlamydomonas reinhardtii. J.Bio. Chem. 2008, 283, 21571–21578. [Google Scholar] [CrossRef]
  58. Fratelli, M.; Demol, H.; Puype, M.; Casagrande, S.; Villa, P.; Eberini, I.; Vandekerckhove, J.; Gianazza, E.; Ghezzi, P. Identification of proteins undergoing glutathionylation in oxidatively stressed hepatocytes and hepatoma cells. Proteomics 2003, 3, 1154–1161. [Google Scholar] [CrossRef]
  59. Schuppe-Koistinen, I.; Moldeus, P.; Bergman, T.; Cotgreave, I.A. S-Thiolation of human endothelial cell glyceraldehyde-3-phosphate dehydrogenase after hydrogen peroxide treatment. Eur. J. Biochem. 1994, 221, 1033–1037. [Google Scholar] [CrossRef]
  60. Rokutan, K.; Johnston, R.B.; Kawai, K. Oxidative stress induces S-thiolation of specific proteins in cultured gastric mucosal cells. Am. J. Physiol.-Gastrointest. Liver Physiol. 1994, 266, G247–G254. [Google Scholar] [CrossRef]
  61. Ward, N.E.; Stewart, J.R.; Ioannides, C.G.; O’Brian, C.A. Oxidant-Induced S-Glutathiolation Inactivates Protein Kinase C-α (PKC-α):  A Potential Mechanism of PKC Isozyme Regulation. Biochemistry 2000, 39, 10319–10329. [Google Scholar] [CrossRef]
  62. Shenton, D.; Grant, C.M. Protein S-thiolation targets glycolysis and protein synthesis in response to oxidative stress in the yeast Saccharomyces cerevisiae. Biochem. J. 2003, 374, 513–519. [Google Scholar] [CrossRef]
  63. Dalle-Donne, I.; Giustarini, D.; Colombo, R.; Milzani, A.; Rossi, R. S-glutathionylation in human platelets by a thiol-disulfide exchange-independent mechanism. Free Radic. Biol. Med. 2005, 38, 1501–1510. [Google Scholar] [CrossRef]
  64. Fratelli, M.; Gianazza, E.; Ghezzi, P. Redox proteomics: Identification and functional role of glutathionylated proteins. Expert Rev. Proteom. 2004, 1, 365–376. [Google Scholar] [CrossRef]
  65. Domenico, F.D.; Cenini, G.; Sultana, R.; Perluigi, M.; Uberti, D.; Memo, M.; Butterfield, D.A. Glutathionylation of the Pro-apoptotic Protein p53 in Alzheimer’s Disease Brain: Implications for AD Pathogenesis. Neurochem. Res. 2009, 34, 727–733. [Google Scholar] [CrossRef]
  66. Kil, I.S.; Kim, S.Y.; Park, J.-W. Glutathionylation regulates IκB. Biochem. Biophys. Res. Commun. 2008, 373, 169–173. [Google Scholar] [CrossRef]
  67. Carvalho, A.N.; Marques, C.; Guedes, R.C.; Castro-Caldas, M.; Rodrigues, E.; van Horssen, J.; Gama, M.J. S-Glutathionylation of Keap1: A new role for glutathione S-transferase pi in neuronal protection. FEBS Lett. 2016, 590, 1455–1466. [Google Scholar] [CrossRef] [Green Version]
  68. Jin, Y.; Diffee, G.M.; Colman, R.J.; Anderson, R.M.; Ge, Y. Top-down Mass Spectrometry of Sarcomeric Protein Post-translational Modifications from Non-human Primate Skeletal Muscle. J. Am. Soc. Mass Spectrom. 2019, 30, 2460–2469. [Google Scholar] [CrossRef]
  69. Wei, L.; Gregorich, Z.R.; Lin, Z.; Cai, W.; Jin, Y.; McKiernan, S.H.; McIlwain, S.; Aiken, J.M.; Moss, R.L.; Diffee, G.M.; et al. Novel Sarcopenia-related Alterations in Sarcomeric Protein Post-translational Modifications (PTMs) in Skeletal Muscles Identified by Top-down Proteomics. Mol. Cell. Proteom. 2018, 17, 134–145. [Google Scholar] [CrossRef] [Green Version]
  70. Li, X.; Day, N.J.; Feng, S.; Gaffrey, M.J.; Lin, T.D.; Paurus, V.L.; Monroe, M.E.; Moore, R.J.; Yang, B.; Xian, M.; et al. Mass spectrometry-based direct detection of multiple types of protein thiol modifications in pancreatic beta cells under endoplasmic reticulum stress. Redox Biol. 2021, 46, 102111. [Google Scholar] [CrossRef]
  71. Jaffrey, S.R.; Snyder, S.H. The Biotin Switch Method for the Detection of S-Nitrosylated Proteins. Sci. STKE 2001, 2001, pl1. [Google Scholar] [CrossRef]
  72. Lind, C.; Gerdes, R.; Hamnell, Y.; Schuppe-Koistinen, I.; von Löwenhielm, H.B.; Holmgren, A.; Cotgreave, I.A. Identification of S-glutathionylated cellular proteins during oxidative stress and constitutive metabolism by affinity purification and proteomic analysis. Arch. Biochem. Biophys. 2002, 406, 229–240. [Google Scholar] [CrossRef]
  73. Su, D.; Gaffrey, M.J.; Guo, J.; Hatchell, K.E.; Chu, R.K.; Clauss, T.R.W.; Aldrich, J.T.; Wu, S.; Purvine, S.; Camp, D.G.; et al. Proteomic identification and quantification of S-glutathionylation in mouse macrophages using resin-assisted enrichment and isobaric labeling. Free Radic. Biol. Med. 2014, 67, 460–470. [Google Scholar] [CrossRef] [Green Version]
  74. Bushweller, J.H.; Aaslund, F.; Wuethrich, K.; Holmgren, A. Structural and functional characterization of the mutant Escherichia coli glutaredoxin (C14.fwdarw.S) and its mixed disulfide with glutathione. Biochemistry 1992, 31, 9288–9293. [Google Scholar] [CrossRef] [PubMed]
  75. Reynaert, N.L.; Ckless, K.; Guala, A.S.; Wouters, E.F.M.; van der Vliet, A.; Janssen-Heininger, Y.M.W. In situ detection of S-glutathionylated proteins following glutaredoxin-1 catalyzed cysteine derivatization. Biochim. Biophys. Acta (BBA)—Gen. Subj. 2006, 1760, 380–387. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, T.; Qian, W.-J.; Strittmatter, E.F.; Camp, D.G.; Anderson, G.A.; Thrall, B.D.; Smith, R.D. High-Throughput Comparative Proteome Analysis Using a Quantitative Cysteinyl-peptide Enrichment Technology. Anal. Chem. 2004, 76, 5345–5353. [Google Scholar] [CrossRef]
  77. Kramer, P.A.; Duan, J.; Gaffrey, M.J.; Shukla, A.K.; Wang, L.; Bammler, T.K.; Qian, W.J.; Marcinek, D.J. Fatiguing contractions increase protein S-glutathionylation occupancy in mouse skeletal muscle. Redox Biol. 2018, 17, 367–376. [Google Scholar] [CrossRef]
  78. Duan, J.; Zhang, T.; Gaffrey, M.J.; Weitz, K.K.; Moore, R.J.; Li, X.; Xian, M.; Thrall, B.D.; Qian, W.J. Stochiometric quantification of the thiol redox proteome of macrophages reveals subcellular compartmentalization and susceptibility to oxidative perturbations. Redox Biol. 2020, 36, 101649. [Google Scholar] [CrossRef]
  79. Zhang, T.; Gaffrey, M.J.; Li, X.; Qian, W.J. Characterization of cellular oxidative stress response by stoichiometric redox proteomics. Am. J. Physiol. Cell Physiol. 2021, 320, C182–C194. [Google Scholar] [CrossRef]
  80. Day, N.J.; Gaffrey, M.J.; Qian, W.J. Stoichiometric Thiol Redox Proteomics for Quantifying Cellular Responses to Perturbations. Antioxidants 2021, 10, 499. [Google Scholar] [CrossRef]
  81. McGarry, D.J.; Chen, W.; Chakravarty, P.; Lamont, D.L.; Wolf, C.R.; Henderson, C.J. Proteome-wide identification and quantification of S-glutathionylation targets in mouse liver. Biochem. J. 2015, 469, 25–32. [Google Scholar] [CrossRef]
  82. Campbell, M.D.; Duan, J.; Samuelson, A.T.; Gaffrey, M.J.; Merrihew, G.E.; Egertson, J.D.; Wang, L.; Bammler, T.K.; Moore, R.J.; White, C.C.; et al. Improving mitochondrial function with SS-31 reverses age-related redox stress and improves exercise tolerance in aged mice. Free Radic. Biol. Med. 2019, 134, 268–281. [Google Scholar] [CrossRef]
  83. Zhang, T.; Day, N.J.; Gaffrey, M.; Weitz, K.K.; Attah, K.; Mimche, P.N.; Paine, R., III; Qian, W.J.; Helms, M.N. Regulation of hyperoxia-induced neonatal lung injury via post-translational cysteine redox modifications. Redox Biol. 2022, 55, 102405. [Google Scholar] [CrossRef]
  84. Leichert, L.I.; Gehrke, F.; Gudiseva, H.V.; Blackwell, T.; Ilbert, M.; Walker, A.K.; Strahler, J.R.; Andrews, P.C.; Jakob, U. Quantifying changes in the thiol redox proteome upon oxidative stress in vivo. Proc. Natl. Acad. Sci. USA 2008, 105, 8197–8202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zhou, L.; Chan, J.C.Y.; Chupin, S.; Gueguen, N.; Desquiret-Dumas, V.; Koh, S.K.; Li, J.; Gao, Y.; Deng, L.; Verma, C.; et al. Increased Protein S-Glutathionylation in Leber’s Hereditary Optic Neuropathy (LHON). Int. J. Mol. Sci. 2020, 21, 3027. [Google Scholar] [CrossRef] [PubMed]
  86. Chan, J.C.Y.; Soh, A.C.K.; Kioh, D.Y.Q.; Li, J.; Verma, C.; Koh, S.K.; Beuerman, R.W.; Zhou, L.; Chan, E.C.Y. Reactive Metabolite-induced Protein Glutathionylation: A Potentially Novel Mechanism Underlying Acetaminophen Hepatotoxicity. Mol. Cell. Proteom. 2018, 17, 2034–2050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Xiao, H.; Jedrychowski, M.P.; Schweppe, D.K.; Huttlin, E.L.; Yu, Q.; Heppner, D.E.; Li, J.; Long, J.; Mills, E.L.; Szpyt, J.; et al. A Quantitative Tissue-Specific Landscape of Protein Redox Regulation during Aging. Cell 2020, 180, 968–983.e24. [Google Scholar] [CrossRef]
  88. Shakir, S.; Vinh, J.; Chiappetta, G. Quantitative analysis of the cysteine redoxome by iodoacetyl tandem mass tags. Anal. Bioanal. Chem. 2017, 409, 3821–3830. [Google Scholar] [CrossRef]
  89. Demasi, M.; Silva, G.M.; Netto, L.E. 20 S proteasome from Saccharomyces cerevisiae is responsive to redox modifications and is S-glutathionylated. J. Biol. Chem. 2003, 278, 679–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Dixon, D.P.; Skipsey, M.; Grundy, N.M.; Edwards, R. Stress-Induced Protein S-Glutathionylation in Arabidopsis. Plant Physiol. 2005, 138, 2233–2244. [Google Scholar] [CrossRef] [Green Version]
  91. Brennan, J.P.; Miller, J.I.A.; Fuller, W.; Wait, R.; Begum, S.; Dunn, M.J.; Eaton, P. The Utility of N,N-Biotinyl Glutathione Disulfide in the Study of Protein S-Glutathiolation. Mol. Cell. Proteom. 2006, 5, 215–225. [Google Scholar] [CrossRef] [Green Version]
  92. Sullivan, D.M.; Wehr, N.B.; Fergusson, M.M.; Levine, R.L.; Finkel, T. Identification of Oxidant-Sensitive Proteins:  TNF-α Induces Protein Glutathiolation. Biochemistry 2000, 39, 11121–11128. [Google Scholar] [CrossRef]
  93. Samarasinghe, K.T.G.; Munkanatta Godage, D.N.P.; Zhou, Y.; Ndombera, F.T.; Weerapana, E.; Ahn, Y.-H. A clickable glutathione approach for identification of protein glutathionylation in response to glucose metabolism. Mol. BioSystems 2016, 12, 2471–2480. [Google Scholar] [CrossRef] [Green Version]
  94. Kekulandara, D.N.; Samarasinghe, K.T.G.; Godage, D.N.P.M.; Ahn, Y.-H. Clickable glutathione using tetrazine-alkene bioorthogonal chemistry for detecting protein glutathionylation. Org. Biomol. Chem. 2016, 14, 10886–10893. [Google Scholar] [CrossRef] [PubMed]
  95. VanHecke, G.C.; Abeywardana, M.Y.; Ahn, Y.-H. Proteomic Identification of Protein Glutathionylation in Cardiomyocytes. J. Proteome Res. 2019, 18, 1806–1818. [Google Scholar] [CrossRef] [PubMed]
  96. VanHecke, G.C.; Yapa Abeywardana, M.; Huang, B.; Ahn, Y.-H. Isotopically Labeled Clickable Glutathione to Quantify Protein S-Glutathionylation. ChemBioChem 2020, 21, 853–859. [Google Scholar] [CrossRef]
  97. Chardonnet, S.; Sakr, S.; Cassier-Chauvat, C.; Le Maréchal, P.; Chauvat, F.; Lemaire, S.D.; Decottignies, P. First Proteomic Study of S-Glutathionylation in Cyanobacteria. J. Proteome Res. 2015, 14, 59–71. [Google Scholar] [CrossRef]
  98. Li, Z.; Zhang, C.; Li, C.; Zhou, J.; Xu, X.; Peng, X.; Zhou, X. S-glutathionylation proteome profiling reveals a crucial role of a thioredoxin-like protein in interspecies competition and cariogenecity of Streptococcus mutans. PLOS Pathog. 2020, 16, e1008774. [Google Scholar] [CrossRef]
  99. Pastore, A.; Piemonte, F. S-Glutathionylation signaling in cell biology: Progress and prospects. Eur. J. Pharm. Sci. 2012, 46, 279–292. [Google Scholar] [CrossRef]
  100. Lermant, A.; Murdoch, C.E. Cysteine Glutathionylation Acts as a Redox Switch in Endothelial Cells. Antioxidants 2019, 8, 315. [Google Scholar] [CrossRef] [Green Version]
  101. Yi, M.; Ma, Y.; Zhu, S.; Luo, C.; Chen, Y.; Wang, Q.; Deng, H. Comparative proteomic analysis identifies biomarkers for renal aging. Aging 2020, 12, 21890–21903. [Google Scholar] [CrossRef]
  102. Anathy, V.; Lahue, K.G.; Chapman, D.G.; Chia, S.B.; Casey, D.T.; Aboushousha, R.; van der Velden, J.L.J.; Elko, E.; Hoffman, S.M.; McMillan, D.H.; et al. Reducing protein oxidation reverses lung fibrosis. Nat. Med. 2018, 24, 1128–1135. [Google Scholar] [CrossRef]
  103. Tsukahara, Y.; Ferran, B.; Minetti, E.T.; Chong, B.S.H.; Gower, A.C.; Bachschmid, M.M.; Matsui, R. Administration of Glutaredoxin-1 Attenuates Liver Fibrosis Caused by Aging and Non-Alcoholic Steatohepatitis. Antioxidants 2022, 11, 867. [Google Scholar] [CrossRef]
  104. Newman, S.F.; Sultana, R.; Perluigi, M.; Coccia, R.; Cai, J.; Pierce, W.M.; Klein, J.B.; Turner, D.M.; Butterfield, D.A. An increase in S-glutathionylated proteins in the Alzheimer’s disease inferior parietal lobule, a proteomics approach. J. Neurosci. Res. 2007, 85, 1506–1514. [Google Scholar] [CrossRef]
  105. Yang, J.; Zhang, H.; Gong, W.; Liu, Z.; Wu, H.; Hu, W.; Chen, X.; Wang, L.; Wu, S.; Chen, C.; et al. S-Glutathionylation of human inducible Hsp70 reveals a regulatory mechanism involving the C-terminal α-helical lid. J. Biol. Chem. 2020, 295, 8302–8324. [Google Scholar] [CrossRef] [Green Version]
  106. Loescher, C.M.; Breitkreuz, M.; Li, Y.; Nickel, A.; Unger, A.; Dietl, A.; Schmidt, A.; Mohamed, B.A.; Kötter, S.; Schmitt, J.P.; et al. Regulation of titin-based cardiac stiffness by unfolded domain oxidation (UnDOx). Proc. Natl. Acad. Sci. USA 2020, 117, 24545–24556. [Google Scholar] [CrossRef]
  107. Li, S.; Wang, L.; Xu, Z.; Huang, Y.; Xue, R.; Yue, T.; Xu, L.; Gong, F.; Bai, S.; Wu, Q.; et al. ASC deglutathionylation is a checkpoint for NLRP3 inflammasome activation. J. Exp. Med. 2021, 218, e20202637. [Google Scholar] [CrossRef]
  108. Watanabe, Y.; Watanabe, K.; Fujioka, D.; Nakamura, K.; Nakamura, T.; Uematsu, M.; Bachschmid, M.M.; Matsui, R.; Kugiyama, K. Protein S-glutathionylation stimulate adipogenesis by stabilizing C/EBPβ in 3T3L1 cells. FASEB J. 2020, 34, 5827–5837. [Google Scholar] [CrossRef] [Green Version]
  109. Zhang, J.; Ye, Z.-w.; Chen, W.; Culpepper, J.; Jiang, H.; Ball, L.E.; Mehrotra, S.; Blumental-Perry, A.; Tew, K.D.; Townsend, D.M. Altered redox regulation and S-glutathionylation of BiP contribute to bortezomib resistance in multiple myeloma. Free Radic. Biol. Med. 2020, 160, 755–767. [Google Scholar] [CrossRef]
  110. Zaffagnini, M.; Marchand, C.H.; Malferrari, M.; Murail, S.; Bonacchi, S.; Genovese, D.; Montalti, M.; Venturoli, G.; Falini, G.; Baaden, M.; et al. Glutathionylation primes soluble glyceraldehyde-3-phosphate dehydrogenase for late collapse into insoluble aggregates. Proc. Natl. Acad. Sci. USA 2019, 116, 26057–26065. [Google Scholar] [CrossRef]
  111. Kaus-Drobek, M.; Mücke, N.; Szczepanowski, R.H.; Wedig, T.; Czarnocki-Cieciura, M.; Polakowska, M.; Herrmann, H.; Wysłouch-Cieszyńska, A.; Dadlez, M. Vimentin S-glutathionylation at Cys328 inhibits filament elongation and induces severing of mature filaments in vitro. FEBS J. 2020, 287, 5304–5322. [Google Scholar] [CrossRef]
  112. Dutka, T.L.; Mollica, J.P.; Lamboley, C.R.; Weerakkody, V.C.; Greening, D.W.; Posterino, G.S.; Murphy, R.M.; Lamb, G.D. S-nitrosylation and S-glutathionylation of Cys134 on troponin I have opposing competitive actions on Ca2+ sensitivity in rat fast-twitch muscle fibers. Am. J. Physiol.-Cell Physiol. 2017, 312, C316–C327. [Google Scholar] [CrossRef]
  113. Mohr, S.; Hallak, H.; de Boitte, A.; Lapetina, E.G.; Brüne, B. Nitric Oxide-induced S-Glutathionylation and Inactivation of Glyceraldehyde-3-phosphate Dehydrogenase. J. Biol. Chem. 1999, 274, 9427–9430. [Google Scholar] [CrossRef] [Green Version]
  114. Mustafa Rizvi, S.H.; Shao, D.; Tsukahara, Y.; Pimentel, D.R.; Weisbrod, R.M.; Hamburg, N.M.; McComb, M.E.; Matsui, R.; Bachschmid, M.M. Oxidized GAPDH transfers S-glutathionylation to a nuclear protein Sirtuin-1 leading to apoptosis. Free Radic. Biol. Med. 2021, 174, 73–83. [Google Scholar] [CrossRef] [PubMed]
  115. Sirover, M.A. Moonlighting glyceraldehyde-3-phosphate dehydrogenase: Posttranslational modification, protein and nucleic acid interactions in normal cells and in human pathology. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 354–371. [Google Scholar] [CrossRef] [PubMed]
  116. Alegre-Cebollada, J.; Kosuri, P.; Giganti, D.; Eckels, E.; Rivas-Pardo, J.A.; Hamdani, N.; Warren, C.M.; Solaro, R.J.; Linke, W.A.; Fernández, J.M. S-Glutathionylation of Cryptic Cysteines Enhances Titin Elasticity by Blocking Protein Folding. Cell 2014, 156, 1235–1246. [Google Scholar] [CrossRef] [Green Version]
  117. Leutert, M.; Entwisle, S.W.; Villén, J. Decoding Post-Translational Modification Crosstalk With Proteomics. Mol. Cell. Proteom. 2021, 20. [Google Scholar] [CrossRef]
  118. Budde, H.; Hassoun, R.; Tangos, M.; Zhazykbayeva, S.; Herwig, M.; Varatnitskaya, M.; Sieme, M.; Delalat, S.; Sultana, I.; Kolijn, D.; et al. The Interplay between S-Glutathionylation and Phosphorylation of Cardiac Troponin I and Myosin Binding Protein C in End-Stage Human Failing Hearts. Antioxidants 2021, 10, 1134. [Google Scholar] [CrossRef]
  119. Day, N.J.; Zhang, T.; Gaffrey, M.J.; Zhao, R.; Fillmore, T.L.; Moore, R.J.; Rodney, G.G.; Qian, W.-J. A deep redox proteome profiling workflow and its application to skeletal muscle of a Duchenne Muscular Dystrophy model. Free Radic. Biol. Med. 2022, 193, 373–384. [Google Scholar] [CrossRef]
  120. Kruyer, A.; Ball, L.E.; Townsend, D.M.; Kalivas, P.W.; Uys, J.D. Post-translational S-glutathionylation of cofilin increases actin cycling during cocaine seeking. PLoS ONE 2019, 14, e0223037. [Google Scholar] [CrossRef] [Green Version]
  121. Gao, X.; Li, L.; Parisien, M.; Wu, J.; Bederman, I.; Gao, Z.; Krokowski, D.; Chirieleison, S.M.; Abbott, D.W.; Wang, B.; et al. Discovery of a redox thiol switch: Implications for cellular energy metabolism. Mol. Cell. Proteom. 2020, 19, 852–870. [Google Scholar] [CrossRef] [Green Version]
  122. Behring, J.B.; van der Post, S.; Mooradian, A.D.; Egan, M.J.; Zimmerman, M.I.; Clements, J.L.; Bowman, G.R.; Held, J.M. Spatial and temporal alterations in protein structure by EGF regulate cryptic cysteine oxidation. Sci. Signal. 2020, 13, eaay7315. [Google Scholar] [CrossRef]
  123. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  124. Tunyasuvunakool, K.; Adler, J.; Wu, Z.; Green, T.; Zielinski, M.; Zidek, A.; Bridgland, A.; Cowie, A.; Meyer, C.; Laydon, A.; et al. Highly accurate protein structure prediction for the human proteome. Nature 2021, 596, 590–596. [Google Scholar] [CrossRef]
  125. Huang, K.L.; Scott, A.D.; Zhou, D.C.; Wang, L.B.; Weerasinghe, A.; Elmas, A.; Liu, R.; Wu, Y.; Wendl, M.C.; Wyczalkowski, M.A.; et al. Spatially interacting phosphorylation sites and mutations in cancer. Nat. Commun. 2021, 12, 2313. [Google Scholar] [CrossRef] [PubMed]
  126. Alford, R.F.; Leaver-Fay, A.; Jeliazkov, J.R.; O’Meara, M.J.; DiMaio, F.P.; Park, H.; Shapovalov, M.V.; Renfrew, P.D.; Mulligan, V.K.; Kappel, K.; et al. The Rosetta All-Atom Energy Function for Macromolecular Modeling and Design. J. Chem. Theory Comput. 2017, 13, 3031–3048. [Google Scholar] [CrossRef] [PubMed]
  127. Audagnotto, M.; Dal Peraro, M. Protein post-translational modifications: In silico prediction tools and molecular modeling. Comput. Struct. Biotechnol. J. 2017, 15, 307–319. [Google Scholar] [CrossRef]
  128. Zhu, F.; Yang, S.; Meng, F.; Zheng, Y.; Ku, X.; Luo, C.; Hu, G.; Liang, Z. Leveraging Protein Dynamics to Identify Functional Phosphorylation Sites using Deep Learning Models. J. Chem. Inf. Model. 2022, 62, 3331–3345. [Google Scholar] [CrossRef] [PubMed]
  129. Siodlak, D. alpha, beta-Dehydroamino acids in naturally occurring peptides. Amino Acids 2015, 47, 1–17. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, Z.; Lyons, B.; Truscott, R.J.; Schey, K.L. Human protein aging: Modification and crosslinking through dehydroalanine and dehydrobutyrine intermediates. Aging Cell 2014, 13, 226–234. [Google Scholar] [CrossRef] [Green Version]
  131. Lai, K.-Y.; Galan, S.R.G.; Zeng, Y.; Zhou, T.H.; He, C.; Raj, R.; Riedl, J.; Liu, S.; Chooi, K.P.; Garg, N.; et al. LanCLs add glutathione to dehydroamino acids generated at phosphorylated sites in the proteome. Cell 2021, 184, 2680–2695.e2626. [Google Scholar] [CrossRef]
Figure 1. Electrostatic microenvironment lowers cysteine thiol pKa and stabilizes thiolate anion in human GAPDH (PDB: 6yne).
Figure 1. Electrostatic microenvironment lowers cysteine thiol pKa and stabilizes thiolate anion in human GAPDH (PDB: 6yne).
Antioxidants 11 02272 g001
Figure 2. (A) Potential molecular mechanisms of protein glutathionylation. (B) Protein glutathionylation (GSTP, adapted from [41]) and de-glutathionylation by enzymatic reaction (Grx).
Figure 2. (A) Potential molecular mechanisms of protein glutathionylation. (B) Protein glutathionylation (GSTP, adapted from [41]) and de-glutathionylation by enzymatic reaction (Grx).
Antioxidants 11 02272 g002
Figure 3. General strategies for detecting protein glutathionylation.
Figure 3. General strategies for detecting protein glutathionylation.
Antioxidants 11 02272 g003
Figure 4. Proteomic approaches for SSG detection and quantification using the selective reduction and tagging strategy. (A) Modified BST, (B) RAC-TMT. and (C) GluICAT approches.
Figure 4. Proteomic approaches for SSG detection and quantification using the selective reduction and tagging strategy. (A) Modified BST, (B) RAC-TMT. and (C) GluICAT approches.
Antioxidants 11 02272 g004
Figure 5. Proteomics approaches for SSG detection using clickable analogs of GSH.
Figure 5. Proteomics approaches for SSG detection using clickable analogs of GSH.
Antioxidants 11 02272 g005
Table 1. Selected MS-based profiling studies of the SSG proteome.
Table 1. Selected MS-based profiling studies of the SSG proteome.
Biological SystemMethodCoverage of the SSG ProteomeExperiment ConditionSignificanceRef.
Mouse lung RAC-TMT~7600 SSG sites. Hyperoxia vs. basal;
Neonatal wild-type vs. overexpression (β-ENaC)
Landscape view of SSG-modified proteome in mouse lung and the impact of hypoxia on the SSG proteome [83]
RAW 264.7 mouse macrophagesRAC-TMTOccupancy for 4099 SSG sites Basal conditionProteome-wide quantification of both SSG and total oxidation occupancy under physiological conditions, revealing cellular compartmentation of redox homeostasis. [78]
Mouse skeletal muscleRAC-TMTOccupancy for >2200 SSG sites and changes due to muscle fatiguing Gastrocnemius muscle with and without fatiguing contractionsIncreased muscle protein SSGs identified following a single bout of fatiguing contraction[77]
HL-1 cardiomyocyteClickable GSH1398 SSG-peptides in isotopic duplex experimentH2O2 treatment (1 mM)In vivo isotopic tagging of protein SSGs for direct enrichment and quantitative detection; Validation (by Western blot and site mutation) of two structural proteins of interest.[96]
Mouse liverModified RAC-TMT724 SSG-modified proteins Basal condition, GSTP-nulled miceThe SSG proteome mediated enzymatically by GSTP[81]
Synechocystis sp. PCC6803GSSG-Biotin349 proteins with SSG (protein level enrichment);
145 SSG sites (peptide level enrichment)
Lysate treated with GSSG-biotinFirst SSG proteome profiling in cyanobacteria by GSSG-biotin and LC-MS/MS. [97]
Streptococcus mutans UA159IodoTMT switch strategy357 SSG sites Wild-type vs. mutantsSSG profiling in in bacteria and mutants; Site mutagenesis validation for SSG function on a thioredoxin-like protein.[98]
Human skin fibroblastsGluICAT2307 SSG sitesLHON patients vs. healthy controlsQuantify the ratio of SSG and free thiols with heavy or light ICAT [85]
Table 2. List of novel SSG-regulated proteins discovered by proteomics followed by functional studies.
Table 2. List of novel SSG-regulated proteins discovered by proteomics followed by functional studies.
ProteinCys SiteBiological SystemApproachesStructural/Functional ChangePotential Physiological ConsequenceRef.
Hsp70, humanC574, C603HeLa cellsModified BSTUnfolding of the α-helical lid structure;
Blocking substrate-binding site;
Increasing ATPase activity
[105]
FABPC127Primary macrophages (mouse)RAC-TMT, Anti-GSH antibodies (Co-IP)Promote fatty acid binding function and nuclear translocation;
Active PPARβ/δ
Inhibition of LPS-induced inflammation[9]
Titin distal I-band (82Ig83 domain)C13585 (cryptic cysteines)Mouse heartsOxICAT, Anti-GSH antibodies (Western blot)Unfolded domain oxidation; Enhance titin phosphorylation Decrease titin-based stiffness[106]
ASC (apoptosis-associated speck-like protein containing a CARD)C171Bone marrow–derived macrophages (mouse)Anti-GSH antibodies (Western blot, proximity ligation assay)Domain rotation leading to reduced area of CARD-CARD binding interface; Preventing oligomerization Repress NLRP3 inflammasome activation[107]
C/EBPβC201 and C2963T3L1 preadipocyteAnti-GSH antibodies (Western blot)Decreased interaction with PIAS1; Stabilizing C/EBPβPromotion of adipogenesis[108]
BiPC41, C420Multiple myeloma cellsAnti-GSH antibodiesDecreased α-helix and increased β-sheets; Enhancing foldase activity; Decreasing ATPase activityProteasome inhibitor resistance[109]
GAPDHC149Arabidopsis thalianaLC-MS/MS Enzyme inactivation; Misfolding and destabilized conformation; Inducing aggregation and disulfide bond formation with C153 [110]
SMYD2C13H9c2 myocytesClickable GSH (Western blot) Dissociation; Disrupts SMYD2–Hsp90–N2A(titin) interactionsSarcomere destabilization[93]
VimentinC328 LC-MS/MSStabilizing pep2B region of vimentin tetramers; Impaired assembly into long filament [111]
Troponin IC134Skinned mammalian muscle fiberBioGEE, anti-GSH antibodies, LC-MS/MSCompetitive action with SNO on the same Cys siteIncreased Ca2+ sensitivity[112]
Estrogen receptor αC221, C245, C417, and C447Primary macrophages (mouse)Anti-GSH antibodies, LC-MS/MS Reduced binding potential: receptor density and affinity for 17β-estradiol [36]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Zhang, T.; Day, N.J.; Feng, S.; Gaffrey, M.J.; Qian, W.-J. Defining the S-Glutathionylation Proteome by Biochemical and Mass Spectrometric Approaches. Antioxidants 2022, 11, 2272. https://doi.org/10.3390/antiox11112272

AMA Style

Li X, Zhang T, Day NJ, Feng S, Gaffrey MJ, Qian W-J. Defining the S-Glutathionylation Proteome by Biochemical and Mass Spectrometric Approaches. Antioxidants. 2022; 11(11):2272. https://doi.org/10.3390/antiox11112272

Chicago/Turabian Style

Li, Xiaolu, Tong Zhang, Nicholas J. Day, Song Feng, Matthew J. Gaffrey, and Wei-Jun Qian. 2022. "Defining the S-Glutathionylation Proteome by Biochemical and Mass Spectrometric Approaches" Antioxidants 11, no. 11: 2272. https://doi.org/10.3390/antiox11112272

APA Style

Li, X., Zhang, T., Day, N. J., Feng, S., Gaffrey, M. J., & Qian, W.-J. (2022). Defining the S-Glutathionylation Proteome by Biochemical and Mass Spectrometric Approaches. Antioxidants, 11(11), 2272. https://doi.org/10.3390/antiox11112272

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop